Next Article in Journal
Green Valorization Strategies of Pleurotus ostreatus and Its By-Products: A Critical Review of Emerging Technologies and Sustainable Applications
Next Article in Special Issue
The Action of Cannabidiol on Doxycycline Cytotoxicity in Human Cells—In Vitro Study
Previous Article in Journal
Interacting Quantum Atoms Analysis of Covalent and Collective Interactions in Single Elongated Carbon–Carbon Bonds
Previous Article in Special Issue
Unlocking the Ilex guayusa Potential: Volatile Composition, Antioxidant, Antidiabetic, and Hemolytic Activities, with In Silico Molecular Docking and ADMET Analysis of Hydroethanolic Extracts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Phytochemicals as Epigenetic Modulators in Chronic Diseases: Molecular Mechanisms

1
Medical Doctoral School, Titu Maiorescu University, 040317 Bucharest, Romania
2
Faculty of Medicine, Titu Maiorescu University, 031593 Bucharest, Romania
3
Ilfov County Clinical Emergency Hospital, 022104 Bucharest, Romania
4
National Institute for Chemical-Pharmaceutical R&D, 031299 Bucharest, Romania
5
“Victor Babeș” National Institute of Pathology, 050096 Bucharest, Romania
6
Department of Cell Biology and Histology, “Carol Davila” University of Medicine and Pharmacy, 050474 Bucharest, Romania
*
Author to whom correspondence should be addressed.
Molecules 2025, 30(21), 4317; https://doi.org/10.3390/molecules30214317
Submission received: 30 September 2025 / Revised: 28 October 2025 / Accepted: 5 November 2025 / Published: 6 November 2025
(This article belongs to the Special Issue Phytochemistry, Human Health and Molecular Mechanisms)

Abstract

Phytochemicals are plant-derived bioactive compounds with antioxidant, anti-inflammatory, and epigenetic modulatory effects that may contribute to the prevention and management of chronic diseases. This review synthesizes recent evidence on the molecular mechanisms through which phytochemicals influence oxidative stress, inflammatory signaling, and epigenetic regulation. A targeted literature search of the PubMed and Web of Science databases (2015–2025) identified over 400 experimental and review studies investigating phytochemicals with documented antioxidant and epigenetic activities. Eligible articles were selected based on relevance to oxidative stress, inflammation, and DNA or histone modification pathways in chronic diseases. Data were qualitatively analyzed to highlight mechanistic links between redox balance, transcriptional regulation, and disease modulation. The results indicate that several phytochemicals, including hesperidin, phloretin, lycopene, and silybin, modulate signaling cascades—NF-κB, Nrf2, and PI3K/Akt—while also influencing DNA methylation and histone acetylation to restore gene expression homeostasis. Despite strong in vitro and in vivo evidence, translation to clinical practice remains limited by low bioavailability, lack of standardized formulations, and insufficient human trials. Future research should prioritize integrative study designs linking molecular mechanisms to clinical endpoints. Understanding the epigenetic actions of phytochemicals may guide the development of nutraceutical strategies for chronic disease prevention.

1. Introduction

Phytochemicals are plant secondary metabolites produced as defense molecules against pathogens and environmental stress factors [1]. They display extensive chemical diversity, ranging from low-molecular-weight compounds, such as phenolic acids, to complex structures, such as proanthocyanidins and ellagitannins [2,3]. These naturally occurring compounds possess antiviral, antifungal, and antibiotic properties, as well as photoprotective functions, contributing to plant metabolism and survival under biotic and abiotic stress [4].
For centuries, plant-derived treatments have been used to alleviate acute and chronic diseases, including diabetes, depression, hypertension, hypercholesterolemia, insomnia, and inflammation [5,6]. According to the World Health Organization, nearly 70–80% of the global population still relies on traditional medicine, mainly based on plant extracts [7].
Recent scientific evidence indicates that bioactive compounds from fruits, vegetables, and medicinal plants are biocompatible with cellular and molecular targets and generally less toxic than synthetic drugs [8,9]. Preclinical studies have shown that many phytochemicals with anti-inflammatory, antioxidant, and antiproliferative properties can prevent cancer initiation and progression by activating antioxidant enzymes and triggering apoptosis [10,11]. Moreover, phytochemicals act as substrates, cofactors, or inhibitors in biochemical reactions, as ligands that activate or block receptors, scavengers of reactive oxygen species, and enhancers of nutrient absorption and stability [10,12,13].
There is increasing evidence that chronic diseases such as cancer, metabolic, neurodegenerative, and autoimmune disorders are triggered by epigenetic dysregulation [14]. Epigenetic alterations may result from exposure to environmental factors, including pollutants, radiation, and lifestyle habits such as alcohol consumption, poor nutrition, smoking, and lack of physical activity [15]. Epigenetics refers to heritable changes in gene expression that occur without alteration of the DNA sequence, mainly mediated by DNA methylation, histone modifications, and non-coding RNAs [16,17].
Bioactive dietary compounds can induce protective epigenetic modifications that influence gene expression and health outcomes throughout life. Early identification and correction of epigenetic abnormalities using natural bioactive compounds represent a promising preventive approach [18].
Therefore, understanding the molecular mechanisms through which plant-derived compounds exert antioxidant and anti-inflammatory effects is essential to clarify their impact on human health [19]. This review focuses on the role of phytochemicals as natural bioactive molecules that modulate molecular and epigenetic mechanisms involved in the prevention and management of chronic diseases. A targeted literature search of the PubMed and Web of Science databases (2015–2025) identified over 400 experimental and review studies investigating phytochemicals with documented antioxidant/anti-inflammatory and epigenetic activities. Eligible articles were selected based on relevance to oxidative stress, inflammation, and DNA or histone modification pathways in chronic diseases. Data were qualitatively analyzed to highlight mechanistic links between redox balance, transcriptional regulation, and disease modulation.

2. Epigenetic Dysregulation in Chronic Diseases

Over the past decades, numerous studies have underscored the pivotal role of altered expression and activity of epigenetic regulatory proteins in the initiation and progression of various chronic diseases [20].
Epigenetic mechanisms regulate gene expression by modifying DNA, histones, and RNA. For example, the transformation of a condensed chromatin region (heterochromatin), which is transcriptionally inactive, into euchromatin, a transcriptionally active structure, is achieved by adding methyl or acetyl groups to histone proteins. The addition of these chemical radicals, along with the gene’s promoter demethylation, can loosen the chromatin structure, making DNA more accessible to the transcriptional machinery [21].
Enzymes that regulate epigenetic modifications are categorized into “writers,” “erasers,” “readers,” and “remodelers” based on their functions [22]. Writers are adding the specific chemical groups on DNA or RNA bases or histone’s amino acids, whereas erasers are acting in reverse by removing these modifications. For instance, DNA methyltransferase (DNMT) catalyzes the addition of methyl groups to form 5-methylcytosine (m5C) in a gene’s promoter DNA sequence [23]. In contrast, the ten-eleven translocation (TET) enzymes initiate the DNA demethylation process by converting m5C into other derivatives, such as 5-hydroxymethylcytosine (5hmC) [24]. Typically, transcriptionally active genes exhibit lower methylation levels, whereas inactive genes or those with lower expression levels are heavily methylated. Readers are proteins that recognize and bind to these chemical modifications, such as the methyl-CpG-binding domain (MBD), which recognizes 5mC [25]. The reader proteins dynamically influence chromatin status and, by recruiting or collaborating with other protein complexes, regulate gene expression [26].
Remodelers such as the SWI/SNF family, which possess DNA-stimulated ATPase activity, play a crucial role in nucleosome disruption and chromatin accessibility. They achieve this by disrupting nucleosomes and either moving or removing them from key regulatory regions, such as enhancers and promoters, thereby influencing gene expression [26]. Furthermore, there are unique classes of epigenetic regulators, represented by non-coding RNAs (ncRNAs), which directly bind to various genomic regions or specific RNA sequences to modulate gene expression [27].
The long non-coding RNAs (lncRNAs) have a variety of functions and, importantly, can interact with proteins, DNA sequences, and RNAs, regulating the epigenetic machinery directly or indirectly by affecting gene expression and chromatin organization [28,29].
When epigenetic regulators, either epigenetic enzymes or ncRNAs expression, become imbalanced—by being excessively active or insufficiently functional—they disrupt cellular homeostasis and could induce the development of diseases [30]. Consequently, investigating these individual epigenetic alterations, such as DNA methylation, histone modifications, and dysfunctional expression of ncRNAs, represents a promising opportunity for chronic disease therapy, fueling the development of targeted treatments [31,32]. For example, abnormal epigenetic mechanisms are present at various stages of cancer development, including initiation, progression, invasion, migration, and chemotherapy resistance. DNA methylation patterns are changed with hypermethylation of specific tumor suppressors and hypomethylation of oncogene promoters, which drives tumor development [33]. The silencing of tumor suppressor and DNA repair genes disrupts normal cell proliferation and differentiation, fostering the malignant phenotype of tumor cells [34]. Moreover, DNA methylation loss in oncogene promoter regions and extensive demethylation of DNA repetitive sequences compromise chromatin stability, facilitating tumor development [35].
Metabolic disorders, physiological conditions, pre-existing genetic mutations, and exposure to environmental factors, such as dietary patterns, can significantly influence the epigenome [36]. Several studies have shown that in mice fed a high-fat diet, hypermethylation of the promoters of genes such as Rac family small GTPase (Rac1), PPARα, and others is linked to retinal damage and microangiopathy in diabetic retinopathy models [21,37].
Diet-induced epigenetic modifications can affect future generations, heightening their susceptibility to metabolic diseases, including diabetes, non-alcoholic fatty liver disease, and others. Furthermore, diet-related epigenetic changes have been observed to impact the function of epigenetic regulators, enzymes, and their cofactors, such as TET enzymes and α-ketoglutarate (α-KG) derived from the tricarboxylic acid (TCA) cycle, in metabolic disorders [20,38]. This disruption further impairs epigenetic regulation and accelerates disease progression.
Thus, the reversible nature of epigenetic modifications that regulate enzyme activation or inhibition may serve as promising targets and valuable tools for understanding cellular and biological processes.

2.1. Oxidative Stress Impacting Epigenetic Mechanisms

The most studied property of plant-derived natural compounds is their antioxidant capacity. Scavenging reactive oxygen species (ROS) is thought to be an effective way to reduce oxidative stress in cells.
ROS are represented by free radicals such as hydrogen peroxide, hydroxyl radical (OH), peroxy radical (ROO), superoxide anion radical (O2), and singlet oxygen, among others, which are necessarily produced in the human body as a result of normal physiological processes occurring, and also by exposure to environmental stressors. Oxidative stress arises when the balance between ROS levels is disrupted and the cellular antioxidant defense system is overwhelmed [39].
Oxidative stress is a key modulator of the cellular environment, often altering gene expression through epigenetic modifications. The imbalance between pro-oxidant species and antioxidant defenses can lead to changes in DNA methylation patterns, histone modifications, and non-coding RNA expressions, which collectively impact chromatin structure and gene transcription. For instance, oxidative stress induces the formation of 8-oxo-2′-deoxyguanosine (8-oxo-dG), a DNA lesion that recruits repair machinery but also affects the local methylation status, potentially silencing tumor suppressor genes or activating oncogenes. Such oxidative DNA damage has been linked to disrupted DNMT activity and subsequent global DNA hypomethylation [40].
Multiple lines of scientific evidence have linked the oxidative stress process to pathophysiology associated with various diseases, including cancer [41,42]. ROS can effectively induce DNA damage through nucleotide alterations, DNA strand breakage, and mutations in the genome of normal cells, thereby promoting further disease-related transformation. Moreover, ROS can activate transcription factors, including NF-κB, cyclooxygenases (COX), p53, HIF-1α, PPAR-γ, β-catenin/Wnt, and Nrf2, leading to the expression of over 500 genes, including growth factors, pro- and anti-inflammatory cytokines, and cell cycle regulatory molecules. [43,44]. Moreover, Nrf2, a crucial transcription factor that acts as a master regulator of cellular defense mechanisms against oxidative stress, electrophiles, and other harmful substances, can influence epigenetic mechanisms by modulating histone acetylation and deacetylation [45]. This activity reshapes the cell’s epigenetic landscape, impacting gene expression in response to oxidative stress [46].
In addition, the ROS can influence the DNA methylation pattern by directly oxidizing the catalytic site of DNMTs or by reducing the availability of the DNMT cofactor S-adenosyl methionine, SAM [46]. Cellular ROS levels themselves have been suggested to be the subject of epigenetic modulation. Thereby, ROS-generating systems in mitochondria, including various subunits of the NOX complex, as well as antioxidant enzymes such as SODs and catalase, have been reported to be epigenetically regulated by distinct mechanisms [47]. ROS can modulate the activity of other epigenetic regulators such as histone acetyltransferases (HATs) and histone deacetylases (HDACs), influencing the histone acetylation epigenetic mechanism and thus, such as histone acetyltransferases (HATs) and histone deacetylases (HDACs), thereby altering the histone acetylation epigenetic mechanism and inducing dysregulated gene expression in cancer and cardiovascular disorders [46].
Phytochemicals are highly bioactive molecules; however, a particular class, namely phenolics (phenolic acids and flavonoids), is the most abundant and exhibits the highest antioxidant activity in the cellular environment [47,48].
Different mechanisms underlie the antioxidant capacity of phytochemicals, including free-radical scavenging, hydrogen atom donation, chelation of ferric or cupric ions, single-oxygen quenching, and acting as a substrate for superoxide and hydroxyl radicals (Figure 1) [10,49]. The antioxidant activity of phytochemicals also relies on their capacity to inhibit essential enzymes responsible for the endogenous production of reactive oxygen species (ROS), such as COXs, cytochrome P450 (CP450), lipoxygenase (LO), NADPH oxidase (NOX), nitric oxide synthases (NOSs), and xanthine oxidase (XO) [50]. The production of these enzymes is increased during inflammatory conditions; thus, inhibiting them can prevent ROS overproduction, thus mitigating the associated cellular damage and inflammation [51]. Since mediators of inflammation play a key role in various pathologies, such as cancer, metabolic, and neurodegenerative disorders, natural bioactive compounds might exert both antioxidant and anti-inflammatory activities [51].
According to recent scientific reports, the main molecular mechanism by which phytochemicals reduce cellular oxidative stress is by activating the Nrf2 pathway, which boosts antioxidant defense systems and diminishes inflammation [52,53]. Phytochemicals activate the Nrf2 pathway to combat oxidative stress primarily by modifying Keap1, the central negative regulator of Nrf2, or by promoting phosphorylation of Nrf2. Under normal conditions, Nrf2 is bound to Keap1 in the cytoplasm, which leads to its ubiquitination and degradation [54]. When phytochemicals induce oxidative or electrophilic stress, they cause conformational changes in Keap1 or directly modify it, leading to the release of Nrf2. Freed Nrf2 then translocates to the nucleus, where it forms heterodimers with Maf proteins and binds to antioxidant response elements (AREs) in the DNA. This binding initiates transcription of various antioxidant and detoxification genes, such as heme oxygenase-1 (HO-1), glutathione S-transferase (GST), and NAD(P)H: quinone oxidoreductase 1 (NQO1), which enhance cellular antioxidant defenses and reduce oxidative damage [55].
At the same time, phytochemicals influence epigenetic processes, acting as epigenetic modulators that primarily inhibit key epigenetic enzymes, including DNMTs, HDACs, HATs, and BETs [56].
Moreover, phytochemicals can donate the methyl group directly as a co-substrate in DNA, RNA, or histone methylation or indirectly by affecting the methyl pool [23].
As antioxidants, anti-inflammatory molecules, and epigenetic modulators, the phytochemicals represent valuable candidates for disease prevention and treatment through coordinated regulation of redox balance, the inflammatory process, and epigenetic mechanisms (Figure 1).

2.2. Epigenetic Dysregulation in the Inflammation Process

Inflammation is a highly dynamic and tightly regulated biological response, essential for defending the body against infections, injuries, and other harmful stimuli [57]. However, when the mechanism of controlling inflammation becomes dysregulated, the results can be chronic inflammatory states that contribute to a wide range of diseases, from autoimmune disorders to cancer and cardiovascular pathologies [58]. Epigenetic mechanisms are increasingly recognized for their essential role in inflammation. They act as modulators of normal inflammatory responses by fine-tuning the protein expression and also as drivers of pathological inflammation when disrupted [59].
Epigenetic dysregulation is increasingly recognized as a critical factor in modulating inflammation and immune responses. DNA methylation plays a dual role, either suppressing or activating key pro-inflammatory genes. Aberrant activity of DNA methyltransferases (DNMTs) can lead to inappropriate methylation patterns that sustain chronic inflammation by silencing anti-inflammatory genes. For example, DNMT3B has been shown to hypermethylate promoter regions of tumor suppressor genes, thereby contributing to sustained inflammatory signaling [40]. Recent studies indicate that DNMT1-mediated methylation of IL-10 and SOCS3 promoters also maintains pro-inflammatory macrophage polarization, thus reinforcing chronic inflammatory loops [60].
During an inflammatory response, immune cells such as macrophages, dendritic cells, and lymphocytes undergo rapid changes in gene expression. This transcriptional reprogramming is orchestrated in part by epigenetic modifications, which enable these cells to rapidly activate or repress large sets of genes in response to signals such as cytokines, pathogen-associated molecular patterns (PAMPs), or damage-associated molecular patterns (DAMPs) [61,62,63]. Such transcriptional shifts are supported by histone acetylation at promoters of TNF-α, IL-1β, and IL-6, mediated by p300/CBP histone acetyltransferases, which cooperate with NF-κB to enhance transcriptional output during acute inflammation. Conversely, HDAC3 recruitment by NF-κB p50 homodimers helps reduce inflammatory signaling [64]. Epigenetic mechanisms not only determine which genes are expressed, but also regulate the intensity of cellular response to inflammatory stressors [65].
Multiple epigenetic modifications may occur in a wide range of diseases associated with or related to inflammatory processes. For instance, the most investigated class of diseases, cancers, are often associated with epigenetic deregulation, which can occur in tumoral stroma or in immune cells. Epigenetics encompasses a series of reversible modifications that do not involve gene mutations and are strongly modulated by the internal environment, particularly that located close to tumoral processes. In this way, the pro-inflammatory environment surrounding the tumor can induce epigenetic modifications [66,67]. For example, persistent IL-6/STAT3 signaling within the tumor microenvironment induces DNMT1 and EZH2 expression, thereby reinforcing the epigenetic repression of tumor suppressor genes and maintaining a pro-oncogenic inflammatory state [68].
Yang et al. [69] analyzed the implications of epigenetic modifications—DNA methylation, chromatin remodeling, regulation of non-coding RNAs, and histone changes—during the progression from chronic inflammation to colorectal cancer. The study also outlined the accelerating effects of such changes on the activation of cancer signaling pathways. Among these, the NF-kB and STAT3 pathways appear to be of notable impact. A similar analysis was provided by Pandareesh et al. for prostate cancer, with an almost identical distribution and interaction among epigenetic factors, the tumor environment, and the pro-inflammatory environment [70]. Mechanistically, NF-κB activation enhances histone acetylation at promoters of inflammatory cytokines (IL-8, COX-2), while STAT3 can recruit DNMT1 to hypermethylate SOCS gene promoters, reinforcing a self-sustaining inflammatory state in cancer [71].
When epigenetic control mechanisms malfunction, the balance between pro-inflammatory and anti-inflammatory pathways can be disrupted [61,72]. For instance, aberrant DNA methylation patterns may silence anti-inflammatory genes (e.g., IL-10) or abnormally activate pro-inflammatory ones, such as IL-6 or TNF-α [73,74,75]. Moreover, histone modifications can alter chromatin accessibility and maintain certain genes in a persistently active state, even in the absence of inflammatory stimuli. For example, increased histone acetylation of NF-κB target gene promoters has been linked to sustained inflammation in rheumatoid arthritis and inflammatory bowel disease [76,77,78]. In these contexts, HDAC inhibitors have been shown to restore the balance of histone acetylation and suppress NF-κB-dependent transcription, partly by enhancing the acetylation and stability of IκBα, thereby preventing NF-κB nuclear translocation [79].
Notably, non-coding RNAs, such as miR-155 or miR-21, are often overexpressed in chronic inflammation and can modulate immune cell differentiation, cytokine production, and apoptosis resistance [80,81]. The eraser proteins HDACs are inhibited, leading to increased histone acetylation, opening chromatin and facilitating the expression of anti-proliferative and regulatory genes like p21 [82,83]. HDAC1 and HDAC2 repression has also been linked to enhanced p21 and GADD45β expression, which attenuate NF-κB-mediated inflammatory transcription [84].
The p21 gene plays a multifaceted role in cellular processes. It acts as a cyclin-dependent kinase (CDK) inhibitor, primarily inhibiting CDK2 and CDK1, which are crucial for cell cycle progression [85]. This inhibition leads to cell cycle arrest, particularly at the G1/S and G2/M transitions, preventing uncontrolled cell division and promoting genomic stability after DNA damage. Beyond cell cycle regulation, p21 also influences the cellular environment by secreting bioactive molecules, contributing to immunomodulatory functions and potentially aiding in the resolution of inflammation [86].
Histone modifications further contribute to inflammation-driven gene expression. Pro-inflammatory cytokines like TNF-α and IL-6 are regulated by dynamic histone acetylation and methylation at their promoter regions, with histone acetyltransferases (HATs) enhancing expression and histone deacetylases (HDACs) serving as brakes on transcription. Natural polyphenols such as resveratrol and quercetin have been shown to modulate these enzymes, reducing pro-inflammatory histone marks while enhancing the expression of anti-inflammatory genes [87]. For example, resveratrol promotes deacetylation of p65/RelA via SIRT1 activation, reducing NF-κB transcriptional activity, while quercetin inhibits p300-mediated H3K27 acetylation at IL-1β promoters [88].
Non-coding RNAs, especially microRNAs (miRNAs), play a pivotal role in fine-tuning inflammation. For instance, miR-146a suppresses NF-κB signaling, a master regulator of inflammation. Under oxidative stress conditions, miR-146a expression can be epigenetically downregulated, leading to sustained inflammation [89]. Additionally, decreased miR-146a expression has been associated with reduced negative feedback on TRAF6 and IRAK1, amplifying NF-κB activity in chronic inflammatory states [90]. Interestingly, dietary polyphenols may also exert protective effects on inflammation by modulating Nrf2, which coordinates antioxidant responses and regulates inflammatory pathways. By activating the Nrf2-ARE pathway, compounds like curcumin, sulforaphane, and EGCG (epigallocatechin gallate) can upregulate antioxidant genes while simultaneously downregulating pro-inflammatory mediators [91]. Curcumin, for example, directly promotes Nrf2 nuclear translocation and enhances HO-1 and NQO1 expression, while simultaneously repressing NF-κB p65 phosphorylation, providing dual antioxidant and anti-inflammatory benefits [92,93]. Moreover, recent evidence indicates that dysregulated Nrf2 signaling contributes to cancer cell survival under oxidative stress, positioning Nrf2 both as a protective and potentially oncogenic factor [94]. Furthermore, recent evidence from a clinical study suggests that polyphenol-rich diets can alter epigenetic markers of aging in immune cells, thereby indirectly influencing inflammation status [91]. Such interventions might slow down immunosenescence and mitigate inflammation-associated disorders.
Moreover, such epigenetic changes can affect not only immune cells but stromal and epithelial cells within the inflamed microenvironment, contributing to persistent tissue remodeling and fibrosis [66].
Another field of major interest is coronary artery disease, where, on the one hand, disease progression involves epigenetic changes, such as histone methylation, phosphorylation, and acetylation, playing key roles. In addition to these, the lncRNAs, miRNAs, and circRNAs are also involved in epigenetic processes and help regulate gene expression in response to environmental factors and lifestyle. Such modifications in the deregulation of genes regulating vascular function, inflammation, lipid metabolism, or oxidative stress [95,96].
Collectively, this epigenetic modification can act as a form of inflammatory memory, locking cells into a pro-inflammatory state even after the initial trigger has passed [62]. This phenomenon is especially evident in chronic diseases such as atherosclerosis, rheumatoid arthritis, and asthma, in which immune cells exhibit persistent activation profiles [97].
A defining feature of epigenetic regulation is the plasticity of epigenetic marks, which are not static and can be influenced by environmental exposures [98]. Factors such as smoking, diet, psychological stress, pollution, infections, and gut microbiota-derived metabolites can all modify the epigenetic landscape [99,100].
For instance, short-chain fatty acids, such as butyrate, produced by commensal bacteria, can inhibit HDACs and thereby alter gene expression in immune cells [101]. Similarly, oxidative stress induced by environmental pollutants can trigger histone modifications that amplify transcription of inflammatory genes [102,103]. These influences suggest that chronic exposure to harmful environmental factors may not only trigger inflammation but may epigenetically fix it into the biology of the tissues [104,105].
Even subtle lifestyle factors, such as circadian rhythm disruption or chronic low-grade stress, may modulate the expression of inflammatory genes through epigenetic routes [106]. These influences suggest that chronic exposure to harmful environmental factors may not only trigger inflammation but also epigenetically fix it into the tissue’s biology [107].
Of interest is also the fact that natural compounds, for instance naringenin and naringin (pharmacologically active in anticancer and cardiovascular diseases, with antioxidant and anti-inflammatory activities), are involved in epigenetic mechanisms, thus broadening the mechanisms supporting the restoration of health [108,109].
Overall, the integration of dietary components, oxidative stress, and epigenetic alterations forms a complex network that modulates inflammation. Targeting these pathways through nutraceuticals offers a promising avenue for managing chronic inflammation-related diseases.
The reversibility of epigenetic marks makes them attractive targets for novel therapeutic strategies. Drugs that inhibit DNMT1 or HDACs are already being explored in cancer therapy, but they show promise in treating chronic inflammatory diseases as well. By reprogramming immune cells epigenetically, either to reduce their inflammatory potential or to restore tolerance, such an intervention could offer a more precise and durable solution than broad-spectrum immunosuppressants [110]. For example, HDAC inhibitors may promote the expression of p21 and other regulatory genes, helping to dampen persistent inflammation. Similarly, DNMT inhibitors could reactivate silenced anti-inflammatory genes or detoxifying enzymes such as GSTP1, restoring homeostasis in chronically inflamed tissues [111].

3. Phytochemicals as Antioxidant and Anti-Inflammatory Agents with Epigenetic Modulator Capacities

Chronic inflammation and oxidative stress are fundamental contributors to the pathogenesis of numerous non-communicable diseases, including cardiovascular disorders, cancer, neurodegenerative conditions, diabetes, and autoimmune syndromes. Oxidative stress arises from an imbalance between reactive oxygen species and the antioxidant defense system, which may also induce inflammation by elevating the levels of pro-inflammatory mediators and cytokines. These two pathological states are tightly interconnected and often exacerbate one another, creating a self-perpetuating cycle of tissue damage and immune dysregulation [112].
In recent years, phytochemicals found in fruit, vegetables, herbs, and medicinal plants have gained significant attention for their ability to counteract oxidative and inflammatory damage. These compounds include flavonoids, phenolic acids, terpenoids, alkaloids, and organosulfur compounds, many of which possess strong radical-scavenging capabilities and modulate inflammatory signaling pathways such as NF-κB, Nrf2, PI3K/Akt, and MAPKs (Table 1) [113,114].
Unlike conventional pharmaceuticals, phytochemicals often have multitarget potential, modulating cellular processes simultaneously and often with lower toxicity profiles. For instance, flavonoids enhance antioxidant enzyme activities, such as SOD, CAT, and GSH, while simultaneously inhibiting pro-inflammatory mediators, such as TNF-α, IL-6, and COX-2, through the suppression of transcription factors and kinases [115,116,117]. These mechanisms not only alleviate oxidative stress but also reduce systemic and local inflammation, contributing to disease prevention and adjunctive therapy. Notably, flavonoids represent a structurally diverse class of compounds, including flavonols, flavones, flavanones, and anthocyanidins, each with distinct bioactivities and tissue selectivity [118]. This diversity allows them to modulate a broad range of molecular targets implicated in inflammation, oxidative damage, and metabolic imbalance [119]. The phytochemical multitarget capacity supports their potential for disease prevention and as adjunctive therapy.
At the mechanistic level, NF-κB activation typically involves phosphorylation and degradation of its inhibitor IκBα by the IKK complex, allowing the p65/p50 heterodimer to translocate into the nucleus and transcribe pro-inflammatory genes such as COX-2, iNOS, and IL-6 [120,121]. Several phytochemicals, including genistein and phloretin, have been shown to inhibit this phosphorylation cascade, thereby attenuating NF-κB nuclear translocation and cytokine production [122].
Conversely, the Nrf2 pathway acts as a central defense mechanism against oxidative stress [53,123]. Under basal conditions, Nrf2 is sequestered in the cytoplasm by Keap1; oxidative or electrophilic stress disrupts this complex, enabling Nrf2 to translocate into the nucleus and induce transcription of antioxidant and cytoprotective genes such as HO-1, NQO1, and GCLC [53,54,124].
Compounds such as curcumin, artemisinin, and lycopene can activate Nrf2 either by modifying Keap1 cysteine residues or by enhancing Nrf2 stability [125,126].
Recent studies have provided further evidence of curcumin’s role as a potent activator of the Nrf2/ARE pathway [92]. These works highlight curcumin’s ability to promote Nrf2 nuclear accumulation and upregulation of its downstream targets, including HO-1 and NQO1, which together mitigate ROS-induced cellular damage [127]. However, the emerging literature also emphasizes the dual role of Nrf2 in cancer [94]: while transient activation supports antioxidant defense and cytoprotection, chronic or constitutive activation in tumor cells may promote survival, chemoresistance, and metabolic adaptation [126]. Therefore, the biological outcome of Nrf2 activation depends on the cellular and pathological context, underscoring the need for balanced modulation of this pathway [128,129].
Beyond these canonical signaling routes, phytochemicals also modulate epigenetic mechanisms that regulate gene expression without altering DNA sequence [130,131,132,133,134]. Several compounds inhibit DNA methyltransferases (DNMTs) and histone deacetylases (HDACs), thereby reactivating silenced tumor suppressor genes such as p16, p21, SOCS3, and RARB [133,134,135]. For instance, lycopene and silibinin have been shown to reduce DNMT1 and HDAC2 expression in hepatic and prostate cell lines, restoring normal histone acetylation patterns and promoting apoptosis [134,136]. Similarly, phloretin and caffeic acid increase histone acetylation and DNA hypomethylation, thereby facilitating transcription of antioxidant and anti-inflammatory genes [137,138].
Table 1 presents the phytochemicals with in vitro-demonstrated antioxidant, anti-inflammatory, and potential epigenetic modulator capacities.
Table 1. Example of phytochemicals with antioxidant, anti-inflammatory activities, and epigenetic modulator capacity, investigated in vitro.
Table 1. Example of phytochemicals with antioxidant, anti-inflammatory activities, and epigenetic modulator capacity, investigated in vitro.
Class of CompoundPhytochemicalsStudy Design:
Cell Culture Type; Phytochemicals Concentration
Key OutcomesReferences
AntioxidantAnti-InflammatoryEpigenetic Modulator
FlavonoidsHesperidinMCF-7
(50 µM)
Increases activities of SOD, CAT, GSH; reduces lipid peroxidation;
activates Nrf2
Reduces the expression of TNF-α and IL-6Inhibits DNMT1; hypo-methylation of p16 promoter[54,55,56]
PhloretinRAW 264.7 macrophages
A549 cells
(25 µM);
Inhibits ROS production and increases GSH activitySuppresses NF-κB activation and reduces iNOS and COX-2 expressionInhibits HDAC; increases histone acetylation, p21 expression[139,140,141,142]
GenisteinMCF-7,
(10 µM)
PC3
(25 µM–100 µM),
HL-60 cells
Increases antioxidant enzymes; reduces lipid peroxidationReduces pro-inflammatory cytokines (IL-6, IL-1β) expression via NF-κBInhibits DNMT1, demethylates, and reactivates tumor suppressor genes[143,144,145,146]
Phenolic acidsCaffeic acidCaco-2,
HeLa cells
(20 µM)
Neutralizes free radicals and boosts antioxidant enzymesInhibits pro-inflammatory cytokines and NF-κB activationInhibits DNMT; DNA hypomethylation; gene reactivation[147,148,149]
Coumaric acidLPS-stimulated RAW 264.7 cells; MDA-MB-231 cells
(30 µM)
Free radical scavenger and increases SOD activityReduces NO production and COX-2 expressionInhibits HDAC; increases histone acetylation[150,151,152]
TerpenoidsLycopeneHepG2,
PC3 cells
(10 µM)
Reduces ROS and increases antioxidant enzyme activityInhibits COX-2 expression and lowers prostaglandin E2 levelsInhibits DNMT1 and HDAC2 expression; modulates miR-let-7f-1/AKT2 axis; restores tumor suppressor genes (p21, RARB, SOCS3).[136,153]
SilibininHepG2, DU145 cells (50 µM)Increases SOD and CAT activity; reduces lipid peroxidationInhibits pro-inflammatory cytokines and NF-κB activationInhibits DNMT and HDAC; hypomethylation and histone acetylation[154,155,156]
ArtemisininHeLa cells
(50 µM)
Reduces oxidative stress by lowering ROS and activating Nrf2Inhibits IL-1β, TNF-α, and COX-2 expression[157]
GeraniolRAW 264.7 macrophages (50 µM)Reduces ROS production and increases GSH activitySuppresses NO, IL-6, and TNF-α production by inhibiting NF-κB[158]
Organo-sulfur com-poundsAllyl mercaptanHUVEC, HT-29 cells
(15 µM)
Enhances antioxidant enzyme activity and reduces ROSInhibits adhesion molecules and reduces vascular inflammationInhibits HDAC; increases histone acetylation; gene reactivation[159,160]
In summary, phytochemicals exert their antioxidant and anti-inflammatory actions through interconnected pathways that encompass the inhibition of NF-κB–driven transcription, activation of the Nrf2/ARE antioxidant response, and epigenetic reprogramming of gene expression. The convergence of these molecular mechanisms provides a comprehensive perspective on their therapeutic potential.

3.1. Flavonoids

Hesperidin, a bioactive flavonoid predominantly found in citrus such as oranges and lemons, has demonstrated significant antioxidant and anti-inflammatory properties, particularly in the context of breast cancer [161,162]. In vitro studies on MCF-7 breast cancer cells treated with 50 µM hesperidin showed increased antioxidant enzymes (SOD, CAT, and glutathione) and reduced lipid peroxidation [145]. These effects are attributed to the activation of the Nrf2 signaling pathway, which plays a crucial role in cellular defense against oxidative stress. Furthermore, hesperidin significantly decreased the expression of pro-inflammatory cytokines such as TNF-alpha and IL-6 by inhibiting the pathway, thereby modulating inflammation within the tumor microenvironment [163,164]. These findings underscore hesperidin’s potential as a therapeutic agent managing oxidative stress and inflammation in breast cancer models [123]. Beyond its antioxidant and anti-inflammatory activities, hesperidin also exerts epigenetic modulatory effects. Studies have demonstrated that hesperidin can reduce global DNA methylation levels by inhibiting DNA methyltransferases (DNMTs), particularly in cancer models [165]. In vitro and in vivo cancer models have shown that hesperidin downregulates hypermethylation of repetitive DNA elements, such as KINE-1 and ALU sequences, and modulates microRNA expression, including the upregulation of tumor-suppressive miR-34a and the downregulation of oncogenic miR-221 [166]. These epigenetic modifications contribute to the reactivation of tumor suppressor genes and the inhibition of oncogenic pathways, further supporting hesperidin’s role in cancer and chemoprevention. These findings suggest that hesperidin supports human health through several interconnected mechanisms. By enhancing antioxidant defenses, reducing harmful inflammation, and influencing epigenetic processes that control gene activity, hesperidin shows great promise as a natural compound that could help prevent or support the integrative treatment of cancer.
Genistein is a naturally occurring isoflavone; it is predominantly found in leguminous plants, especially soybeans (Glycine max), where it occurs mainly as the glycoside genistin [167]. Genistein exhibits a spectrum of biological activities beyond its well-known estrogenic and anticancer effects, including moderate antioxidant and anti-inflammatory actions, as well as epigenetic mechanisms regulator [168,169]. This compound exerts moderate but significant antioxidant effects, primarily by reducing ROS [39] and enhances the activities of antioxidant enzymes such as SOD, catalase CAT and glutathione peroxidase, largely via the Nrf2 signaling pathway [170]. These effects stabilize redox balance and protect cellular components from oxidative damage, especially under inflammation or tumorigenic conditions. In addition to its redox balancing effects, genistein demonstrated anti-inflammatory activity by downregulating the NF-kB signaling pathway and suppressing the expression of pro-inflammatory cytokines [171] such as TNF-alpha, IL-6, and Il-1ß [172].
Genistein also inhibits the activity of enzymes such as COX-2 and iNOS, thereby further reducing the inflammatory response [173]. This is particularly relevant in chronic diseases where low-grade inflammation plays a pathogenic role in atherosclerosis and arthritis. One of genistein’s most compelling biological features is its ability to modulate epigenetic mechanisms by inhibiting DNMTs, especially DNMT1, thereby reactivating silenced tumor suppressor genes such as p16, RASSF1A, and BRCA1 [174,175]. Recent findings also suggest that genistein can modulate sirtuin activity and potentially interact with TET enzymes, thereby impacting DNA demethylation and metabolic-epigenetic crosstalk [176].
Phloretin, a dihydrochalcone flavonoid predominantly found in apples, has demonstrated notable antioxidant and anti-inflammatory properties. In vitro studies using RAQ 264.7 macrophages treated with 25 µM phloretin have shown a significant reduction in reactive oxygen species (ROS) production and an increase in intracellular GSH levels, indicating its potential to mitigate oxidative stress [177].
Furthermore, phloretin has been shown to suppress nuclear factor kappa B (NF-κB) activation, a pivotal transcription factor in the inflammatory response. This suppression leads to the downregulation of pro-inflammatory enzymes such as inducible nitric oxide synthase (iNOS) and cyclooxygenase-2 (COX-2). Additionally, phloretin modulates the mitogen-activated protein kinase (MAPK) pathways, including extracellular signal-regulated kinase (ERK), c-Jun N-terminal kinase (JNK), and p38 MAPK [178]. These findings are supported by Leet et al., who demonstrated that phloretin attenuates lipopolysaccharide-induced acute lung injury in mice by blocking the NF-kB and MAPK pathways [179]. These pathways are integral to the production of pro-inflammatory mediators [177]. The combined modulation of these signaling pathways by phloretin suggests its potential application in managing chronic inflammatory conditions and cardiovascular diseases [180,181].
In addition to its well-established antioxidant and anti-inflammatory actions, emerging evidence suggests that phloretin may act as an epigenetic modulator. According to recent findings, phloretin can inhibit DNA methyltransferases (DNMTs), enzymes that add methyl groups to DNA and are often involved in the silencing of tumor suppressor genes [40]. By modulating DNA methylation, phloretin may help restore normal gene expression profiles disrupted in various chronic diseases, including cancer. Although further research is needed to clarify its impact on histone modification and microRNA regulation, these initial insights suggest that phloretin’s health benefits extend beyond redox and inflammatory pathways to include epigenetic regulation of gene expression. Altogether, these multiple molecular actions, combining antioxidant, anti-inflammatory, and epigenetic modulatory effects, position phloretin as a promising candidate for managing chronic inflammatory diseases, metabolic disorders, and cancer prevention.
Silybin, a flavonolignan extracted from Silybum marianum (milk thistle), is widely recognized for its hepatoprotective, antioxidant, and anti-inflammatory effects. In vitro studies on HepG2 cells have shown that silybin increases the activity of endogenous antioxidant enzymes, including SOD and CAT, while significantly decreasing lipid peroxidation, thereby alleviating oxidative damage [182,183]. The compound also inhibits NF-Kb activation, the key transcription factor that regulates inflammatory cytokines [184]. As a result, treatment with silybin leads to reduced expressions of TNF-α, IL-6, and iNOS, contributing to a pronounced anti-inflammatory effect. Moreover, it modulates JAK/STAT and TGF-βsignaling pathways, both of which play central roles in liver fibrosis and hepatocarcinogenesis [185].
Beyond its cellular effects, silibinin is under clinical investigation for its antifibrotic and chemoprotective potential, particularly in the context of chronic liver diseases such as non-alcoholic fatty liver disease (NAFLD) and hepatocellular carcinoma (HCC) [186]. Its multifunctional properties, spanning antioxidation, inflammation suppression, and interference with the fibrotic pathway, highlight its promise as a therapeutic agent in hepatic disorders [186].
In addition to these biological effects, silybin has also been shown to exert epigenetic regulatory activities. Silybin modulates the activity of DNA methyltransferases (DNMTs) and histone deacetylases (HDACs), leading to changes in DNA methylation and histone acetylation status in liver cells. These epigenetic modifications reactivate tumor suppressor genes and suppress oncogenic pathways, thereby contributing to their chemopreventive potential in hepatocarcinogenesis. Therefore, the therapeutic role of silybin in liver diseases may also involve epigenetic regulation of gene expression, providing a broader mechanistic basis for its hepatoprotective and anticancer effects.

3.2. Carotenoids

Lycopene, a lipophilic carotenoid found in tomatoes and watermelon, acts as a potent scavenger of ROS in HepG2 cells [187]. At a concentration of 10 µM, lycopene reduced ROS levels and increased antioxidant enzyme activity [188]. Additionally, it inhibited COX-2 expression and decreased prostaglandin E2 synthesis, indicating an anti-inflammatory effect. Lycopene also modulates transcription factors such as AP-1 and STAT3, suggesting its potential for liver cancer prevention and for reducing hepatic inflammation [189]. In addition to these well-established effects, lycopene also exerts epigenetic modulatory activity. Recent findings indicate that lycopene can influence DNA methylation patterns by inhibiting DNA methyltransferases (DNMTs). This inhibition may contribute to the reactivation of silenced tumor suppressor genes and the correction of aberrant gene expression associated with carcinogenesis. While further research is needed to fully elucidate its effects on histone modification and non-coding RNAs, these early insights suggest that lycopene contributes to the regulation of gene activity through epigenetic pathways, adding a new layer to its chemopreventive potential. Overall, these findings suggest that lycopene supports liver health through several complementary actions. By reducing oxidative stress, calming inflammation, and influencing gene activity via epigenetic mechanisms, lycopene shows promise in cells and may reduce the risk of liver cancer [40].

3.3. Phenolic Compounds

Caffeic acid, a hydroxycinnamic acid prevalent in coffee, fruits, and vegetables, has demonstrated significant antioxidant and anti-inflammatory properties, particularly within intestinal epithelial cells [190,191]. In vitro studies utilizing Caco-2 cells treated with 20 µM caffeic acid have shown a notable reduction in ROS production and an enhancement of endogenous antioxidant enzyme activities, including SOD, CAT, and GSH [192,193]. These effects are primarily mediated through the activation of the Nrf2 signaling pathway, which plays a crucial role in cellular defense against oxidative stress [194,195].
Moreover, caffeic acid has been observed to inhibit the activation of nuclear factor kappa B NF-kB, a key transcription factor involved in the inflammatory response [196]. This inhibition reduces the secretion of pro-inflammatory cytokines, such as interleukin-8 (IL-8), thereby mitigating intestinal inflammation [197]. Additionally, caffeic acid modulates the phosphoinositide 3-kinase/protein kinase B (PI3K/Akt) and mitogen-activated protein kinase (MAPK) signaling pathways, including extracellular signal-regulated kinases (ERK), c-Jun N-terminal kinase (JNK), and p38 MAPK [198]. These pathways are integral to the regulation of cellular responses to oxidative stress and inflammation [199].
Collectively, these findings suggest that caffeic acid exerts protective effects against oxidative stress and inflammatory conditions in intestinal epithelial cells, highlighting its potential therapeutic applications in managing intestinal inflammation and related disorders [200].
Through these antioxidant and anti-inflammatory mechanisms, caffeic acid has also been identified as an epigenetic modulator.
Recent research highlights its capacity to influence DNA methylation and histone modifications, two key mechanisms controlling gene expression without altering the underlying DNA sequence [201]. Specifically, caffeic acid has been shown to inhibit DNA methyltransferase (DNMT) activity, potentially reversing abnormal gene silencing associated with chronic diseases such as cancer and inflammatory disorders. Moreover, caffeic acid may affect histone acetylation and methylation, thereby further regulating genes involved in oxidative stress responses, inflammation, and cellular homeostasis.
These findings show that caffeic acid protects intestinal cells not only by reducing oxidative stress and inflammation but also by helping restore healthy gene activity, making it a promising natural compound for managing intestinal disorders.
P-Coumaric acid, a naturally occurring hydroxycinnamic acid found in foods like red wine and balsamic vinegar, exhibits significant antioxidant and anti-inflammatory properties [202,203]. In vitro studies using lipopolysaccharide (LPS)-simulated RAW 264.7 macrophages have demonstrated that treatment with p-coumaric acid at concentration ranging from 10 to 100 μg/mL effectively reduces the production of nitric oxide (NO) and suppresses the expression of pro-inflammatory mediators such as inducible nitric oxide synthase (iNOS) and cyclooxygenase 2 (COX-2) at both mRNA and protein levels [151,204].
The anti-inflammatory effects of p-coumaric acid are primarily mediated through the inhibition of the nuclear factor kappa B (NF-kB) signaling pathway. Specifically, p-coumaric acid prevents the phosphorylation and subsequent degradation of IκBα, an inhibitor of NF-kB, thereby blocking the translocation of the NF-kB p65 subunit into the nucleus and reducing the transcription of pro-inflammatory genes [100].
Additionally, p-coumaric acid modulates the mitogen-activated protein kinase (MAPK) pathways by inhibiting the phosphorylation of extracellular signal-regulated kinase (ERK1/2) and c-Jun N-terminal kinase (JNK), which are crucial in the expression of inflammatory mediators [205].
Beyond its effects on macrophages, p-coumaric acid has been shown to protect against atherosclerosis. In studies involving THP-1 macrophages treated with oxidized low-density lipoprotein (ox-LDL) and LPS, p-coumaric acid at 20 μM significantly inhibited lipid accumulation and foam cell formation. This effect is attributed to the upregulation of cholesterol efflux-related genes, such as ATP binding cassette transporter A1 (ABCA1), liver X receptor alpha (LXRα), and peroxisome proliferator-activated receptor gamma (PPARγ), and the downregulation of lipid uptake related genes, including lecithin-like oxidized low-density lipoprotein receptor-1 (LOX-1), cluster of differentiation 36 (CD36), and scavenger receptor class A1 (SR-A1). Furthermore, p-coumaric acid suppressed the expression of NF-kB, COX-2, tumor necrosis factor-alpha (TNF-α), and interleukin-6 (IL-6), suggesting its potential to mitigate the inflammatory response associated with atherosclerosis. These findings suggest that p-coumaric acid holds promise as a therapeutic agent in managing oxidative stress-related inflammatory diseases, such as atherosclerosis, by modulating key. Inflammatory signaling pathways and promoting cholesterol homeostasis [206].
Anacardic acid, a phenolic lipid predominantly found in cashew nuts (Anacardium occidentale) and mango seeds, has gained attention for its potent neuroprotective properties, particularly in the context of neurodegenerative diseases such as Parkinson’s and Alzheimer’s [207].
In a study by Augusto et al. (2020) [208], purified anacardic acid (AAs) were administered at a dose of 50 mg/kg in a rotenone-induced mouse model of Parkinson’s disease. The treatment led to a significant reduction in lipid peroxidation and NO levels, alongside an increase in the glutathione (GSH)/oxidized glutathione (GSSH) ratio in both the substantia nigra and striatum regions of the brain. These changes indicate a restoration of redox balance and mitigation of oxidative stress-induced neuronal damage [208].
Furthermore, the study observed that AAs attenuated nuclear kappa B (NF-kB) activation and reduced the expression of pro-inflammatory cytokines, such as interleukin-1β (IL-1β). This anti-inflammatory effect was complemented by the downregulation of matrix metalloproteinase-9 (MMP-9) and glial fibrillary acidic protein (GFAP), markers of neuroinflammation and astrogliosis.
In addition to its neuroprotective and anti-inflammatory effects, anacardic acid also shows potential as a natural modulator of gene activity. Recent studies suggest that it can block the activity of histone acetyltransferases (HATs), particularly p300/CBP, enzymes that normally loosen chromatin structure to activate gene expression [209]. By limiting histone acetylation, anacardic acid may help turn down the expression of pro-inflammatory genes and other harmful pathways involved in neurodegenerative diseases. This ability to influence gene regulation adds another important layer to its protective action, complementing its antioxidant and anti-inflammatory effects. Together, these findings highlight anacardic acid as a promising natural compound that works on multiple levels to protect neurons and slow the progression of neurodegeneration. Allyl mercaptan, a sulfur-containing compound derived from garlic (Allium sativum), has been shown to enhance the activity of key antioxidant enzymes such as superoxide dismutase (SOD) and catalase (CAT), while also reducing intracellular reactive oxygen species (ROS) levels in HUVEC cells treated at a concentration of 15 μM [210]. This antioxidant enhancement helps preserve redox homeostasis in vascular endothelial cells [211].

3.4. Organosulfur Compounds

In addition to its antioxidant effects, allyl mercaptan also inhibits the expression of adhesion molecules, such as ICAM-1 and VCAM-1. It reduces vascular inflammation by preventing NF-kB nuclear translocation, a key event in endothelial activation. Although the precise molecular pathway remains under investigation, some studies have demonstrated that garlic-derived compounds, such as diallyl disulfide and diallyl trisulfide, may act synergistically or more potently by inhibiting NF-kB [212].
These findings support the therapeutic potential of allyl mercaptan as a candidate for endothelial protection and cardiovascular disease prevention, particularly in the context of oxidative stress and inflammation [213].
Moreover, beyond its antioxidant and anti-inflammatory action, allyl mercaptan has been reported to exert epigenetic regulatory effects. According to findings on carcinogenesis [214], allyl mercaptan can inhibit DNMT activity, leading to hypomethylation of promoter regions in key genes involved in cell regulation, such as p21 WAF1. This epigenetic modulation reactivates tumor suppressor pathways and contributes to the regulation of cellular proliferation and apoptosis. Therefore, the protective role of allyl mercaptan in vascular endothelial cells may not only involve redox homeostasis and inflammation suppression but also the regulation of gene expression through epigenetic modifications, highlighting its complex mechanism of action in maintaining cardiovascular health.

3.5. Terpenoids

Artemisinin, a sesquiterpene lactone extracted from Artemisia annua, has shown strong antioxidant, anti-inflammatory, and anticancer properties [215,216]. In HeLa cells treated with 50 μM artemisinin, studies have demonstrated a significant reduction in ROS and enhanced expression of antioxidant enzymes, such as heme oxygenase-1 (HO-1), via activation of the Nrf2 signaling pathway [217,218].
In addition to its antioxidant effects, artemisinin effectively inhibits NF-kB activation, leading to downregulation of pro-inflammatory cytokines such as IL-1β, TNF-α, and COX-2 [219,220]. These molecular effects contribute to the compound’s ability to suppress the tumor microenvironment’s inflammation and modulate immune responses.
Artemisinin also exerts pro-apoptotic activity, inducing loss of mitochondrial membrane potential, caspase activation, and cell cycle arrest in various cancer cell lines, including breast, cervical, and colorectal cancers [221]. Moreover, its antiangiogenic properties, achieved through inhibition of VEGF-mediated endothelial cell migration and proliferation, make it a valuable candidate in anti-tumor therapy, particularly for limiting tumor neovascularization [222].
Due to its multimodal activity profile—antioxidant, anti-inflammatory, apoptotic, and anti-angiogenic—artemisinin is increasingly regarded as a promising adjunct in oncological treatment and is currently under investigation in both preclinical and early-phase clinical studies [223].
Geraniol, a monoterpene found in rose and citronella essential oils, reduces ROS levels and increases GSH activity in RAW 264.7 macrophages (50 μM). Inhibiting NF-kB decreases NO, IL-6, and TNF-α [224]. Geraniol also modulates the PI3K/Akt and MAPK pathways, with potential applications in managing chronic inflammatory and metabolic disorders [225].
Recent research has uncovered an emerging role for geraniol in epigenetic regulation. A 2025 study demonstrates that geraniol can influence gene expression by modulating DNA methylation and histone acetylation, processes essential for controlling gene expression [226]. Specifically, geraniol was found to inhibit the expression of DNA methyltransferases (DNMTs) and histone deacetylases (HDACs), enzymes that play key roles in maintaining aberrant gene silencing in inflammation and metabolic dysfunction. Through this action, geraniol may help restore normal gene expression patterns, further supporting its protective effects at the molecular level. This evidence suggests that geraniol supports health by reducing oxidative stress and inflammation while also helping to restore healthy gene activity, making it a promising natural option for preventing chronic diseases.
Various classes of bioactive compounds, including flavonoids (hesperidin, phloretin, silybin), carotenoids (lycopene), phenolic acids (caffeic acid, p-coumaric acid, anacardic acid), sulfur-containing molecules (allyl mercaptan), terpenes (geraniol), and sesquiterpene lactones (artemisinin), have demonstrated strong antioxidant and anti-inflammatory activities in multiple in vitro studies.
Most of these phytochemicals enhance the activity of endogenous antioxidant enzymes such as SOD, CAT, and GSH, while reducing lipid peroxidation and ROS accumulation. These mechanisms contribute to cellular protection against oxidative damage.
Additionally, several compounds, such as hesperidin, lycopene, artemisinin, and silybin, exhibit antiproliferative effects by inhibiting NF-kB signaling and suppressing pro-inflammatory mediators, including TNF-α, IL-6, COX-2, and iNOS. These bioactive compounds support potential roles as chemopreventive and therapeutic agents in inflammation-associated disease, including cancer, cardiovascular, and neurodegenerative disorders.
Moreover, several compounds, such as hesperidin, lycopene, artemisinin, and silibinin, exhibit antiproliferative effects by inhibiting NF-κB signaling and suppressing pro-inflammatory mediators, including TNF-α, IL-6, COX-2, and iNOS. These bioactivities support their potential roles as chemopreventive and therapeutic agents in inflammation-associated diseases, including cancer, cardiovascular, and neurodegenerative disorders.
Interestingly, some phytochemicals can epigenetically regulate NRF2 levels, which act as a master regulator of the antioxidant response. As such, the coupling of NRF2 to antioxidant response element sequences in gene promoters results in mechanisms of antioxidant detoxification. Yates et al. Phytochemicals documented to modulate NRF2 signaling act by reversing hypermethylated states in the CpG islands of NFE2L2 or Nfe2l2, via the inhibition of DNA methyltransferases (DNMTs) and histone deacetylases (HDACs), through the induction of ten-eleven translocation (TET) enzymes, or by inducing miRNA to target the 3′-UTR of the corresponding mRNA transcripts. To date, fewer than twenty phytochemicals have been reported as NRF2 epigenetic modifiers, including curcumin, sulforaphane, resveratrol, reserpine, and ursolic acid. Bhattachaarjee et al. [227] provided a broader view of the action of phytochemicals on a broad array of disorders, including cancer and chronic diseases. A more focused insight is provided [228], which focuses on natural compounds acting as epigenetic regulators in diabetes mellitus, with curcumin, resveratrol, sulforaphane, genistein, and quercetin identified as major actors.
Naturally occurring phytochemicals, such as those mentioned above, along with other well-studied compounds like resveratrol [229], quercetin [230], and apigenin [231], are commonly used in herbal extracts and have potential as complementary therapies for managing inflammatory diseases, cancer, and other chronic conditions.
When these bioactive compounds are consumed from natural sources, they are typically present alongside other related phytochemicals. This combination allows them to act on multiple regulatory pathways simultaneously [232]. Additionally, these natural compounds can be combined with synthetic medicines to create more complex therapeutic strategies that target several disease mechanisms simultaneously [233,234].

3.6. Synergistic Effects of Phytochemical Combinations

Among the various phytochemicals examined, the combination of genistein and lycopene emerges as one of the most promising candidates for exploring synergistic interactions between natural compounds. Both exhibit potent antioxidant, anti-inflammatory, and epigenetic modulatory properties, yet they act through complementary molecular pathways that may synergize to enhance their bioactivity and therapeutic potential.
Genistein, a soy-derived isoflavone, functions as a multifaceted modulator of cellular signaling. It reduces oxidative stress by stimulating antioxidant enzymes (SOD, CAT, GPx) via the Nrf2 pathway, while simultaneously suppressing pro-inflammatory mediators by inhibiting NF-κB signaling [235]. At the epigenetic level, genistein inhibits DNA methyltransferases (DNMTs) and histone deacetylases (HDACs), leading to reactivation of silenced tumor suppressor genes such as p16, RASSF1A, and BRCA1 [236]. This dual activity—antioxidant and epigenetic—positions genistein as a model nutraceutical for cancer prevention and modulation of chronic disease [237].
Lycopene, a carotenoid abundant in tomatoes and other red fruits, is characterized by its exceptional lipophilic antioxidant capacity [238]. It effectively quenches singlet oxygen and scavenges reactive oxygen species (ROS), protecting membrane lipids from peroxidation [239]. Lycopene also modulates transcription factors involved in inflammation and proliferation, including AP-1, STAT3, and COX-2, thus exerting anti-inflammatory and antiproliferative effects [240]. Moreover, lycopene has been reported to indirectly influence epigenetic processes, such as modulating DNMT activity and microRNA expression [153].
The synergistic rationale for combining genistein and lycopene lies in their complementary mechanisms: genistein acts primarily at the nuclear level, regulating chromatin and gene expression, while lycopene stabilizes redox and membrane integrity, reducing the oxidative triggers that often initiate epigenetic aberrations [241,242].
This cross-interaction between epigenetic reprogramming and oxidative balance could yield enhanced protective effects against carcinogenesis, metabolic inflammation, and age-related disorders [243]. In addition, lycopene’s lipid-soluble nature may improve genistein’s bioavailability by facilitating its absorption and protecting it from intestinal degradation, addressing one of the main challenges in translating genistein’s benefits to clinical settings [244,245].
Emerging evidence from preclinical studies supports the combined use of polyphenols and carotenoids as a novel strategy in nutraceutical formulations [246,247,248]. However, studies specifically investigating the synergy between genistein and lycopene remain limited. Future research should employ co-administration models in vitro and in vivo, focusing on oxidative stress–related diseases such as breast, liver, or prostate cancer. Key endpoints should include measurements of oxidative stress markers, inflammatory cytokine profiles, DNMT/HDAC activity, and downstream gene expression changes associated with cellular homeostasis [249,250].
Furthermore, advanced delivery systems, such as nanoemulsions, liposomes, or polymeric nanoparticles, could be explored to optimize their pharmacokinetic profiles and ensure co-delivery to target tissues. The integration of gut microbiota analysis in these models may also help elucidate how microbial metabolism influences the combined bioactivity of genistein and lycopene [251,252,253].
Altogether, this phytochemical pair exemplifies how strategic combinations of compounds can act on distinct yet interconnected molecular networks—redox regulation, inflammation, and epigenetic remodeling—paving the way for more effective dietary interventions and adjuvant therapies in chronic diseases and cancer prevention.

3.7. Challenges and Strategies for Clinical Translation

Despite the promising antioxidant, anti-inflammatory, and epigenetic-modulating capacities of specific phytochemicals, translating these findings from preclinical to clinical applications remains a significant challenge.
The flavonoids (e.g., hesperidin, genistein, phloretin, and silybin) exhibit strong biological effects in vitro but have low oral bioavailability due to poor solubility and extensive first-pass metabolism [254,255]. For instance, hesperidin and genistein require transformation by gut microbiota into more active metabolites, introducing high interindividual variability. Strategies such as glycosylation modification, nanoencapsulation, and phospholipid complexes have been investigated to improve their absorption and stability [256].
Lycopene (carotenoid) exhibits potent antioxidant and anti-inflammatory activity; however, it is lipophilic and unstable, limiting its systemic availability [257]. Current approaches include liposomal formulation and emulsified delivery systems to enhance its stability and intestinal uptake [258].
Phenolic compounds are rapidly metabolized and eliminated, resulting in a short plasma half-life and reduced clinical efficacy. Moreover, their metabolism is significantly modulated by gut microbiota composition, which can differentiate between individuals. To overcome this, prodrug strategies and targeted delivery systems have been proposed, though evidence remains scarce [259].
Allyl mercaptan (organosulfur compounds) derived from garlic is strongly affected by gut metabolism and instability in circulation. Clinical translation requires stabilized formulation or co-administration with other garlic-derived sulfur compounds to preserve its bioactivity [260].
Artemisinin (a terpenoid) has potent anti-inflammatory and anticancer activity, but rapid clearance and poor water solubility limit its clinical utility. Semi-synthetic derivatives and nanocarrier systems are being explored to overcome these pharmacokinetic barriers [261,262]. Geraniol also suffers from limited bioavailability, and lipid-based carrier encapsulation has been proposed as a solution [254,263].
While small-scale clinical trials with flavonoids and carotenoids have shown beneficial effects on oxidative stress and metabolic regulation, results remain heterogeneous and not sufficient for clinical recommendations [264]. For other compounds reviewed (phloretin, caffeic acid, p-coumaric acid, anacardic acid, allyl mercaptan, geraniol), clinical evidence is very limited or absent, emphasizing the urgent need for rigorous, well-designed clinical trials. This gap is consistent with previous findings on the anticancer and epigenetic modulation activity of natural compounds [265].

4. Conclusions and Future Perspectives

The preventive potential of phytochemicals is a rapidly growing area of research, with in vitro and in vivo studies demonstrating their ability to simultaneously target multiple cellular and molecular pathways, thereby enhancing biological effects. These natural bioactive compounds are increasingly recognized for their roles in promoting health and preventing chronic disease. The majority of chronic diseases are characterized by early epigenetic dysregulation, which may disrupt normal patterns of gene expression. Intrinsic factors such as aging and genetic background, along with extrinsic environmental factors (e.g., diet, pollution, radiation, stress), can significantly influence epigenetic mechanisms, ultimately impacting human health.
The reversibility of epigenetic modifications and the ability to reset cellular homeostasis through compounds with epigenetic modulator capacity have opened new avenues for prevention interventions, particularly in aging, cancer, cardiovascular, and metabolic disorders. Targeting these reversible changes with dietary bioactive compounds offers a promising approach to prevent diseases where early epigenetic abnormalities play a key role.
This review emphasizes that certain phytochemicals, including flavonoids, which possess both antioxidant and anti-inflammatory properties, can directly or indirectly inhibit the aberrant activities of epigenetic regulators. The underlying molecular mechanisms of phytochemicals’ multifaceted actions are incompletely understood. Still, their capacity to act through various epigenetic pathways, including DNA methylation, histone modifications, and chromatin remodeling, is likely to impact a range of cellular processes.
The translation of preclinical data into preventive medicine is a challenging process that requires rigorous clinical trials while accounting for individual variability in the gut microbiota and its impact on phytochemical metabolism. Long-term trials, which should consider the synergistic effects of dietary phytochemicals, are also needed to evaluate overall efficacy and safety in humans, as the diet provides a combination of phytochemicals with specific cellular and organ-level actions.
Despite promising laboratory results, a key challenge remains the low bioavailability of many phytochemicals, which limits their therapeutic translation. Strategies such as nanocarrier delivery systems, liposomal encapsulation, structural modification of parent compounds, and co-administration with bioenhancers are currently being investigated to improve absorption stability and tissue targeting.
Another critical aspect is the inter-individual variability in gut microbiota composition, which profoundly affects phytochemical metabolism and biological activity. Personalized approaches, including microbiome profiling, may be required in the further design of trials to better capture this variability.
While preclinical evidence is extensive, clinical studies remain limited and often heterogeneous. Some small-scale trials have shown beneficial effects of polyphenols (e.g., genistein) on biomarkers of oxidative stress, inflammation, or metabolic regulation, but larger, standardized, placebo-controlled trials are lacking.
Therefore, future clinical trials should be designed with adequate sample sizes, stratification of individuals by microbiota/metabolic phenotype, and long-term follow-up, while considering dietary context and potential synergistic effects with conventional therapies.

Author Contributions

Conceptualization: D.C., M.C.R. and M.P.I.; methodology: R.A., S.P., C.T. and M.-L.P.; software: D.C. and M.P.I.; validation: M.C.R., R.A. and S.P.; formal analysis: R.A., S.P., C.T. and M.-L.P.; investigation: D.C., M.C.R., M.P.I. and S.P.; resources: D.C., M.C.R., M.P.I., R.A., S.P., C.T. and M.-L.P.; data curation: R.A., C.T. and M.-L.P.; writing—original draft preparation: D.C., M.C.R. and M.P.I.; writing—review and editing: R.A., C.T. and M.-L.P.; visualization: S.P., C.T. and M.-L.P.; supervision: R.A., C.T. and M.-L.P.; project administration: C.T. and M.-L.P.; funding acquisition: D.C., M.C.R., M.P.I., R.A., S.P., C.T. and M.-L.P. All authors have read and agreed to the published version of the manuscript.

Funding

This study did not receive any specific grant from funding agencies in the public, commercial, or not-for-profit sectors.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this study.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Anjali; Kumar, S.; Korra, T.; Thakur, R.; Arutselvan, R.; Kashyap, A.S.; Nehela, Y.; Chaplygin, V.; Minkina, T.; Keswani, C. Role of Plant Secondary Metabolites in Defence and Transcriptional Regulation in Response to Biotic Stress. Plant Stress 2023, 8, 100154. [Google Scholar] [CrossRef]
  2. Rabizadeh, F.; Mirian, M.S.; Doosti, R.; Kiani-Anbouhi, R.; Eftekhari, E. Phytochemical Classification of Medicinal Plants Used in the Treatment of Kidney Disease Based on Traditional Persian Medicine. Evid.-Based Complement. Altern. Med. 2022, 2022, 8022599. [Google Scholar] [CrossRef]
  3. Rudrapal, M.; Rakshit, G.; Singh, R.P.; Garse, S.; Khan, J.; Chakraborty, S. Dietary Polyphenols: Review on Chemistry/Sources, Bioavailability/Metabolism, Antioxidant Effects, and Their Role in Disease Management. Antioxidants 2024, 13, 429. [Google Scholar] [CrossRef]
  4. Veiga, M.; Costa, E.M.; Silva, S.; Pintado, M. Impact of Plant Extracts upon Human Health: A Review. Crit. Rev. Food Sci. Nutr. 2020, 60, 873–886. [Google Scholar] [CrossRef] [PubMed]
  5. Chihomvu, P.; Ganesan, A.; Gibbons, S.; Woollard, K.; Hayes, M.A. Phytochemicals in Drug Discovery—A Confluence of Tradition and Innovation. Int. J. Mol. Sci. 2024, 25, 8792. [Google Scholar] [CrossRef]
  6. Panossian, A.G.; Efferth, T.; Shikov, A.N.; Pozharitskaya, O.N.; Kuchta, K.; Mukherjee, P.K.; Banerjee, S.; Heinrich, M.; Wu, W.; Guo, D.; et al. Evolution of the Adaptogenic Concept from Traditional Use to Medical Systems: Pharmacology of Stress- and Aging-related Diseases. Med. Res. Rev. 2021, 41, 630–703. [Google Scholar] [CrossRef]
  7. Bhat, R.A.; Hakeem, K.R.; Dervash, M.A. Phytomedicine: A Treasure of Pharmacologically Active Products from Plants; Academic Press: London, UK, 2021; ISBN 978-0-12-824109-7. [Google Scholar]
  8. Howes, M.R.; Perry, N.S.L.; Vásquez-Londoño, C.; Perry, E.K. Role of Phytochemicals as Nutraceuticals for Cognitive Functions Affected in Ageing. Br. J. Pharmacol. 2020, 177, 1294–1315. [Google Scholar] [CrossRef]
  9. Dar, R.A.; Shahnawaz, M.; Ahanger, M.A.; Majid, I.U. Exploring the Diverse Bioactive Compounds from Medicinal Plants: A Review. J. Phytopharm. 2023, 12, 189–195. [Google Scholar] [CrossRef]
  10. Albulescu, L.; Suciu, A.; Neagu, M.; Tanase, C.; Pop, S. Differential Biological Effects of Trifolium Pratense Extracts—In Vitro Studies on Breast Cancer Models. Antioxidants 2024, 13, 1435. [Google Scholar] [CrossRef]
  11. George, B.P.; Chandran, R.; Abrahamse, H. Role of Phytochemicals in Cancer Chemoprevention: Insights. Antioxidants 2021, 10, 1455. [Google Scholar] [CrossRef]
  12. Alarabei, A.A.; Abd Aziz, N.A.L.; Ab Razak, N.I.; Abas, R.; Bahari, H.; Abdullah, M.A.; Hussain, M.K.; Abdul Majid, A.M.S.; Basir, R. Immunomodulating Phytochemicals: An Insight into Their Potential Use in Cytokine Storm Situations. Adv. Pharm. Bull. 2023, 14, 105–119. [Google Scholar] [CrossRef]
  13. Ionescu, V.S.; Popa, A.; Alexandru, A.; Manole, E.; Neagu, M.; Pop, S. Dietary Phytoestrogens and Their Metabolites as Epigenetic Modulators with Impact on Human Health. Antioxidants 2021, 10, 1893. [Google Scholar] [CrossRef] [PubMed]
  14. Samanta, S.; Rajasingh, S.; Cao, T.; Dawn, B.; Rajasingh, J. Epigenetic Dysfunctional Diseases and Therapy for Infection and Inflammation. Biochim. Biophys. Acta (BBA)—Mol. Basis Dis. 2017, 1863, 518–528. [Google Scholar] [CrossRef]
  15. Tiffon, C. The Impact of Nutrition and Environmental Epigenetics on Human Health and Disease. Int. J. Mol. Sci. 2018, 19, 3425. [Google Scholar] [CrossRef]
  16. Kundu, T.K. Epigenetics: Development and Disease; Subcellular Biochemistry; Springer: Dordrecht, The Netherlands, 2013; ISBN 978-94-007-4524-7. [Google Scholar]
  17. Bure, I.V.; Nemtsova, M.V.; Kuznetsova, E.B. Histone Modifications and Non-Coding RNAs: Mutual Epigenetic Regulation and Role in Pathogenesis. Int. J. Mol. Sci. 2022, 23, 5801. [Google Scholar] [CrossRef]
  18. Lorenzo, P.M.; Izquierdo, A.G.; Rodriguez-Carnero, G.; Fernández-Pombo, A.; Iglesias, A.; Carreira, M.C.; Tejera, C.; Bellido, D.; Martinez-Olmos, M.A.; Leis, R.; et al. Epigenetic Effects of Healthy Foods and Lifestyle Habits from the Southern European Atlantic Diet Pattern: A Narrative Review. Adv. Nutr. 2022, 13, 1725–1747. [Google Scholar] [CrossRef]
  19. Nakadate, K.; Ito, N.; Kawakami, K.; Yamazaki, N. Anti-Inflammatory Actions of Plant-Derived Compounds and Prevention of Chronic Diseases: From Molecular Mechanisms to Applications. Int. J. Mol. Sci. 2025, 26, 5206. [Google Scholar] [CrossRef]
  20. Dai, W.; Qiao, X.; Fang, Y.; Guo, R.; Bai, P.; Liu, S.; Li, T.; Jiang, Y.; Wei, S.; Na, Z.; et al. Epigenetics-Targeted Drugs: Current Paradigms and Future Challenges. Signal Transduct. Target. Ther. 2024, 9, 332. [Google Scholar] [CrossRef]
  21. Zhu, Y.; Wang, X.; Zhou, X.; Ding, L.; Liu, D.; Xu, H. DNMT1-Mediated PPARα Methylation Aggravates Damage of Retinal Tissues in Diabetic Retinopathy Mice. Biol. Res. 2021, 54, 25. [Google Scholar] [CrossRef] [PubMed]
  22. Gillette, T.G.; Hill, J.A. Readers, Writers, and Erasers: Chromatin as the Whiteboard of Heart Disease. Circ. Res. 2015, 116, 1245–1253. [Google Scholar] [CrossRef] [PubMed]
  23. Pop, S.; Enciu, A.M.; Tarcomnicu, I.; Gille, E.; Tanase, C. Phytochemicals in Cancer Prevention: Modulating Epigenetic Alterations of DNA Methylation. Phytochem. Rev. 2019, 18, 1005–1024. [Google Scholar] [CrossRef]
  24. Zhang, X.; Zhang, Y.; Wang, C.; Wang, X. TET (Ten-Eleven Translocation) Family Proteins: Structure, Biological Functions and Applications. Signal Transduct. Target. Ther. 2023, 8, 297. [Google Scholar] [CrossRef]
  25. Moore, L.D.; Le, T.; Fan, G. DNA Methylation and Its Basic Function. Neuropsychopharmacol. Rev. 2013, 38, 23–38. [Google Scholar] [CrossRef]
  26. Becker, P.B.; Workman, J.L. Nucleosome Remodeling and Epigenetics. Cold Spring Harb. Perspect. Biol. 2013, 5, a017905. [Google Scholar] [CrossRef]
  27. Kaikkonen, M.U.; Lam, M.T.Y.; Glass, C.K. Non-Coding RNAs as Regulators of Gene Expression and Epigenetics. Cardiovasc. Res. 2011, 90, 430–440. [Google Scholar] [CrossRef]
  28. Pop, S.; Enciu, A.; Necula, L.G.; Tanase, C. Long Non-coding RNA s in Brain Tumours: Focus on Recent Epigenetic Findings in Glioma. J. Cell. Mol. Med. 2018, 22, 4597–4610. [Google Scholar] [CrossRef]
  29. Statello, L.; Guo, C.-J.; Chen, L.-L.; Huarte, M. Gene Regulation by Long Non-Coding RNAs and Its Biological Functions. Nat. Rev. Mol. Cell Biol. 2021, 22, 96–118. [Google Scholar] [CrossRef] [PubMed]
  30. Wang, K.; Liu, H.; Hu, Q.; Wang, L.; Liu, J.; Zheng, Z.; Zhang, W.; Ren, J.; Zhu, F.; Liu, G.-H. Epigenetic Regulation of Aging: Implications for Interventions of Aging and Diseases. Signal Transduct. Target. Ther. 2022, 7, 374. [Google Scholar] [CrossRef] [PubMed]
  31. Hsieh, H.-H.; Kuo, M.-Z.; Chen, I.-A.; Lin, C.-J.; Hsu, V.; HuangFu, W.-C.; Wu, T.-Y. Epigenetic Modifications as Novel Therapeutic Strategies of Cancer Chemoprevention by Phytochemicals. Pharm. Res. 2025, 42, 69–78. [Google Scholar] [CrossRef] [PubMed]
  32. Copeland, R.A.; Olhava, E.J.; Scott, M.P. Targeting Epigenetic Enzymes for Drug Discovery. Curr. Opin. Chem. Biol. 2010, 14, 505–510. [Google Scholar] [CrossRef]
  33. Casalino, L.; Verde, P. Multifaceted Roles of DNA Methylation in Neoplastic Transformation, from Tumor Suppressors to EMT and Metastasis. Genes 2020, 11, 922. [Google Scholar] [CrossRef]
  34. Lee, A.V.; Nestler, K.A.; Chiappinelli, K.B. Therapeutic Targeting of DNA Methylation Alterations in Cancer. Pharmacol. Ther. 2024, 258, 108640. [Google Scholar] [CrossRef] [PubMed]
  35. Brückmann, N.H.; Pedersen, C.B.; Ditzel, H.J.; Gjerstorff, M.F. Epigenetic Reprogramming of Pericentromeric Satellite DNA in Premalignant and Malignant Lesions. Mol. Cancer Res. 2018, 16, 417–427. [Google Scholar] [CrossRef]
  36. Sales, V.M.; Ferguson-Smith, A.C.; Patti, M.-E. Epigenetic Mechanisms of Transmission of Metabolic Disease across Generations. Cell Metab. 2017, 25, 559–571. [Google Scholar] [CrossRef]
  37. Keleher, M.R.; Zaidi, R.; Hicks, L.; Shah, S.; Xing, X.; Li, D.; Wang, T.; Cheverud, J.M. A High-Fat Diet Alters Genome-Wide DNA Methylation and Gene Expression in SM/J Mice. BMC Genom. 2018, 19, 888. [Google Scholar] [CrossRef]
  38. Paul, B.; Barnes, S.; Demark-Wahnefried, W.; Morrow, C.; Salvador, C.; Skibola, C.; Tollefsbol, T.O. Influences of Diet and the Gut Microbiome on Epigenetic Modulation in Cancer and Other Diseases. Clin. Epigenet. 2015, 7, 112. [Google Scholar] [CrossRef]
  39. Jomova, K.; Raptova, R.; Alomar, S.Y.; Alwasel, S.H.; Nepovimova, E.; Kuca, K.; Valko, M. Reactive Oxygen Species, Toxicity, Oxidative Stress, and Antioxidants: Chronic Diseases and Aging. Arch. Toxicol. 2023, 97, 2499–2574. [Google Scholar] [CrossRef] [PubMed]
  40. Aanniz, T.; Bouyahya, A.; Balahbib, A.; El Kadri, K.; Khalid, A.; Makeen, H.A.; Alhazmi, H.A.; El Omari, N.; Zaid, Y.; Wong, R.S.-Y.; et al. Natural Bioactive Compounds Targeting DNA Methyltransferase Enzymes in Cancer: Mechanisms Insights and Efficiencies. Chem.-Biol. Interact. 2024, 392, 110907. [Google Scholar] [CrossRef]
  41. Zheng, X.; Sawalha, A.H. The Role of Oxidative Stress in Epigenetic Changes Underlying Autoimmunity. Antioxid. Redox Signal. 2022, 36, 423–440. [Google Scholar] [CrossRef]
  42. Pizzino, G.; Irrera, N.; Cucinotta, M.; Pallio, G.; Mannino, F.; Arcoraci, V.; Squadrito, F.; Altavilla, D.; Bitto, A. Oxidative Stress: Harms and Benefits for Human Health. Oxidative Med. Cell. Longev. 2017, 2017, 8416763. [Google Scholar] [CrossRef] [PubMed]
  43. Chakraborti, S.; Ray, B.K.; Roychoudhury, S. Handbook of Oxidative Stress in Cancer; Springer: Singapore, 2022; ISBN 978-981-15-9410-6. [Google Scholar]
  44. Reuter, S.; Gupta, S.C.; Chaturvedi, M.M.; Aggarwal, B.B. Oxidative Stress, Inflammation, and Cancer: How Are They Linked? Free Radic. Biol. Med. 2010, 49, 1603–1616. [Google Scholar] [CrossRef] [PubMed]
  45. Silva-Llanes, I.; Shin, C.H.; Jiménez-Villegas, J.; Gorospe, M.; Lastres-Becker, I. The Transcription Factor NRF2 Has Epigenetic Regulatory Functions Modulating HDACs, DNMTs, and miRNA Biogenesis. Antioxidants 2023, 12, 641. [Google Scholar] [CrossRef]
  46. Kietzmann, T.; Petry, A.; Shvetsova, A.; Gerhold, J.M.; Görlach, A. The Epigenetic Landscape Related to Reactive Oxygen Species Formation in the Cardiovascular System. Br. J. Pharmacol. 2017, 174, 1533–1554. [Google Scholar] [CrossRef]
  47. Madugundu, G.S.; Cadet, J.; Wagner, J.R. Hydroxyl-Radical-Induced Oxidation of 5-Methylcytosine in Isolated and Cellular DNA. Nucleic Acids Res. 2014, 42, 7450–7460. [Google Scholar] [CrossRef]
  48. Roy, A.; Khan, A.; Ahmad, I.; Alghamdi, S.; Rajab, B.S.; Babalghith, A.O.; Alshahrani, M.Y.; Islam, S.; Islam, M.R. Flavonoids a Bioactive Compound from Medicinal Plants and Its Therapeutic Applications. BioMed Res. Int. 2022, 2022, 5445291. [Google Scholar] [CrossRef]
  49. Lü, J.; Lin, P.H.; Yao, Q.; Chen, C. Chemical and Molecular Mechanisms of Antioxidants: Experimental Approaches and Model Systems. J. Cell. Mol. Med. 2010, 14, 840–860. [Google Scholar] [CrossRef]
  50. Andrés, C.M.C.; Pérez De La Lastra, J.M.; Juan, C.A.; Plou, F.J.; Pérez-Lebeña, E. Antioxidant Metabolism Pathways in Vitamins, Polyphenols, and Selenium: Parallels and Divergences. Int. J. Mol. Sci. 2024, 25, 2600. [Google Scholar] [CrossRef] [PubMed]
  51. Mucha, P.; Skoczyńska, A.; Małecka, M.; Hikisz, P.; Budzisz, E. Overview of the Antioxidant and Anti-Inflammatory Activities of Selected Plant Compounds and Their Metal Ions Complexes. Molecules 2021, 26, 4886. [Google Scholar] [CrossRef]
  52. Qin, S.; Chen, Z.; Wen, Y.; Yi, Y.; Lv, C.; Zeng, C.; Chen, L.; Shi, M. Phytochemical Activators of Nrf2: A Review of Therapeutic Strategy in Diabetes. Acta Biochim. Biophys. Sin. 2022, 55, 11. [Google Scholar] [CrossRef]
  53. Ngo, V.; Duennwald, M.L. Nrf2 and Oxidative Stress: A General Overview of Mechanisms and Implications in Human Disease. Antioxidants 2022, 11, 2345. [Google Scholar] [CrossRef] [PubMed]
  54. Baird, L.; Yamamoto, M. The Molecular Mechanisms Regulating the KEAP1-NRF2 Pathway. Mol. Cell. Biol. 2020, 40, e00099-20. [Google Scholar] [CrossRef]
  55. He, F.; Ru, X.; Wen, T. NRF2, a Transcription Factor for Stress Response and Beyond. Int. J. Mol. Sci. 2020, 21, 4777. [Google Scholar] [CrossRef]
  56. Carlos-Reyes, Á.; López-González, J.S.; Meneses-Flores, M.; Gallardo-Rincón, D.; Ruíz-García, E.; Marchat, L.A.; Astudillo-de La Vega, H.; Hernández De La Cruz, O.N.; López-Camarillo, C. Dietary Compounds as Epigenetic Modulating Agents in Cancer. Front. Genet. 2019, 10, 79. [Google Scholar] [CrossRef] [PubMed]
  57. Chen, L.; Deng, H.; Cui, H.; Fang, J.; Zuo, Z.; Deng, J.; Li, Y.; Wang, X.; Zhao, L. Inflammatory Responses and Inflammation-Associated Diseases in Organs. Oncotarget 2018, 9, 7204–7218. [Google Scholar] [CrossRef]
  58. Yasmeen, F.; Pirzada, R.H.; Ahmad, B.; Choi, B.; Choi, S. Understanding Autoimmunity: Mechanisms, Predisposing Factors, and Cytokine Therapies. Int. J. Mol. Sci. 2024, 25, 7666. [Google Scholar] [CrossRef] [PubMed]
  59. Zhang, S.; Meng, Y.; Zhou, L.; Qiu, L.; Wang, H.; Su, D.; Zhang, B.; Chan, K.; Han, J. Targeting Epigenetic Regulators for Inflammation: Mechanisms and Intervention Therapy. MedComm 2022, 3, e173. [Google Scholar] [CrossRef]
  60. Rycyk-Bojarzyńska, A.; Kasztelan-Szczerbińska, B.; Cichoż-Lach, H.; Jargieło, A. Ganglioneuromatous Polyposis Associated with Type 2 B Multiple Endocrine Neoplasia (MEN 2B)—Case Report. Ann. Agric. Environ. Med. 2024, 31, 302–305. [Google Scholar] [CrossRef] [PubMed]
  61. Kaszycki, J.; Kim, M. Epigenetic Regulation of Transcription Factors Involved in NLRP3 Inflammasome and NF-kB Signaling Pathways. Front. Immunol. 2025, 16, 1529756. [Google Scholar] [CrossRef]
  62. Zhu, X.; Chen, Z.; Shen, W.; Huang, G.; Sedivy, J.M.; Wang, H.; Ju, Z. Inflammation, Epigenetics, and Metabolism Converge to Cell Senescence and Ageing: The Regulation and Intervention. Signal Transduct. Target. Ther. 2021, 6, 245. [Google Scholar] [CrossRef]
  63. Wu, D.; Shi, Y.; Zhang, H.; Miao, C. Epigenetic Mechanisms of Immune Remodeling in Sepsis: Targeting Histone Modification. Cell Death Dis. 2023, 14, 112. [Google Scholar] [CrossRef]
  64. Maiolino, G.; Fernández-Pascual, E.; Ochoa Arvizo, M.A.; Vishwakarma, R.; Martínez-Salamanca, J.I. Male Infertility and the Risk of Developing Testicular Cancer: A Critical Contemporary Literature Review. Medicina 2023, 59, 1305. [Google Scholar] [CrossRef]
  65. Jurkowska, R.Z. Role of Epigenetic Mechanisms in the Pathogenesis of Chronic Respiratory Diseases and Response to Inhaled Exposures: From Basic Concepts to Clinical Applications. Pharmacol. Ther. 2024, 264, 108732. [Google Scholar] [CrossRef]
  66. Tan, S.Y.X.; Zhang, J.; Tee, W.-W. Epigenetic Regulation of Inflammatory Signaling and Inflammation-Induced Cancer. Front. Cell Dev. Biol. 2022, 10, 931493. [Google Scholar] [CrossRef]
  67. Vezzani, B.; Carinci, M.; Previati, M.; Giacovazzi, S.; Della Sala, M.; Gafà, R.; Lanza, G.; Wieckowski, M.R.; Pinton, P.; Giorgi, C. Epigenetic Regulation: A Link between Inflammation and Carcinogenesis. Cancers 2022, 14, 1221. [Google Scholar] [CrossRef]
  68. Rickey, L.M.; Mueller, E.R.; Newman, D.K.; Markland, A.D.; Falke, C.; Rudser, K.; Smith, A.L.; Mueller, M.G.; Lowder, J.L.; Lukacz, E.S.; et al. Reliability of Uroflowmetry Pattern Interpretation in Adult Women. Neurourol. Urodyn. 2024, 43, 2084–2092. [Google Scholar] [CrossRef]
  69. Yang, Z.-H.; Dang, Y.-Q.; Ji, G. Role of Epigenetics in Transformation of Inflammation into Colorectal Cancer. World J. Gastroenterol. 2019, 25, 2863–2877. [Google Scholar] [CrossRef] [PubMed]
  70. Pandareesh, M.D.; Kameshwar, V.H.; Byrappa, K. Prostate Carcinogenesis: Insights in Relation to Epigenetics and Inflammation. Endocr. Metab. Immune Disord. Drug Targets 2021, 21, 253–267. [Google Scholar] [CrossRef]
  71. Deer, T.; Heros, R.; Tavel, E.; Wahezi, S.; Funk, R.; Buchanan, P.; Christopher, A.; Weisbein, J.; Gilligan, C.; Patterson, D.; et al. Comparing Conventional Medical Management to Spinal Cord Stimulation for the Treatment of Low Back Pain in a Cohort of DISTINCT RCT Patients. J. Pain Res. 2024, 17, 2741–2752. [Google Scholar] [CrossRef] [PubMed]
  72. Rostami, A.; White, K.; Rostami, K. Pro and Anti-Inflammatory Diets as Strong Epigenetic Factors in Inflammatory Bowel Disease. World J. Gastroenterol. 2024, 30, 3284–3289. [Google Scholar] [CrossRef]
  73. Liu, L.; Ni, S.; Zhang, L.; Chen, Y.; Xie, M.; Huang, X. Molecular Insights and Clinical Implications of DNA Methylation in Sepsis-Associated Acute Kidney Injury: A Narrative Review. BMC Nephrol. 2025, 26, 253. [Google Scholar] [CrossRef] [PubMed]
  74. Das, D.; Karthik, N.; Taneja, R. Crosstalk Between Inflammatory Signaling and Methylation in Cancer. Front. Cell Dev. Biol. 2021, 9, 756458. [Google Scholar] [CrossRef]
  75. Bagni, G.; Biancalana, E.; Chiara, E.; Costanzo, I.; Malandrino, D.; Lastraioli, E.; Palmerini, M.; Silvestri, E.; Urban, M.L.; Emmi, G. Epigenetics in Autoimmune Diseases: Unraveling the Hidden Regulators of Immune Dysregulation. Autoimmun. Rev. 2025, 24, 103784. [Google Scholar] [CrossRef]
  76. Li, C.; Gu, S.; Zhang, Y.; Zhang, Z.; Wang, J.; Gao, T.; Zhong, K.; Shan, K.; Ye, G.; Ke, Y.; et al. Histone Deacetylase in Inflammatory Bowel Disease: Novel Insights. Therap. Adv. Gastroenterol. 2025, 18, 17562848251318833. [Google Scholar] [CrossRef]
  77. Bacher, S.; Meier-Soelch, J.; Kracht, M.; Schmitz, M.L. Regulation of Transcription Factor NF-κB in Its Natural Habitat: The Nucleus. Cells 2021, 10, 753. [Google Scholar] [CrossRef]
  78. Grabiec, A.M.; Tak, P.P.; Reedquist, K.A. Targeting Histone Deacetylase Activity in Rheumatoid Arthritis and Asthma as Prototypes of Inflammatory Disease: Should We Keep Our HATs On? Arthritis Res. Ther. 2008, 10, 226. [Google Scholar] [CrossRef]
  79. Susini, P.; Marcaccini, G.; Cuomo, R.; Grimaldi, L.; Nisi, G. Thighs Lift in the Post-Bariatric Patient—A Systematic Review. J. Plast. Reconstr. Aesthet. Surg. 2024, 98, 357–372. [Google Scholar] [CrossRef]
  80. Alivernini, S.; Gremese, E.; McSharry, C.; Tolusso, B.; Ferraccioli, G.; McInnes, I.B.; Kurowska-Stolarska, M. MicroRNA-155—At the Critical Interface of Innate and Adaptive Immunity in Arthritis. Front. Immunol. 2018, 8, 1932. [Google Scholar] [CrossRef]
  81. Raisch, J. Role of microRNAs in the Immune System, Inflammation and Cancer. World J. Gastroenterol. 2013, 19, 2985. [Google Scholar] [CrossRef] [PubMed]
  82. Phimmachanh, M.; Han, J.Z.R.; O’Donnell, Y.E.I.; Latham, S.L.; Croucher, D.R. Histone Deacetylases and Histone Deacetylase Inhibitors in Neuroblastoma. Front. Cell Dev. Biol. 2020, 8, 578770. [Google Scholar] [CrossRef] [PubMed]
  83. Slaughter, M.J.; Shanle, E.K.; Khan, A.; Chua, K.F.; Hong, T.; Boxer, L.D.; Allis, C.D.; Josefowicz, S.Z.; Garcia, B.A.; Rothbart, S.B.; et al. HDAC Inhibition Results in Widespread Alteration of the Histone Acetylation Landscape and BRD4 Targeting to Gene Bodies. Cell Rep. 2021, 34, 108638. [Google Scholar] [CrossRef] [PubMed]
  84. Shalata, W.; Attal, Z.G.; Solomon, A.; Shalata, S.; Abu Saleh, O.; Tourkey, L.; Abu Salamah, F.; Alatawneh, I.; Yakobson, A. Melanoma Management: Exploring Staging, Prognosis, and Treatment Innovations. Int. J. Mol. Sci. 2024, 25, 5794. [Google Scholar] [CrossRef] [PubMed]
  85. Al Bitar, S.; Gali-Muhtasib, H. The Role of the Cyclin Dependent Kinase Inhibitor P21cip1/Waf1 in Targeting Cancer: Molecular Mechanisms and Novel Therapeutics. Cancers 2019, 11, 1475. [Google Scholar] [CrossRef] [PubMed]
  86. Sturmlechner, I.; Zhang, C.; Sine, C.C.; Van Deursen, E.-J.; Jeganathan, K.B.; Hamada, N.; Grasic, J.; Friedman, D.; Stutchman, J.T.; Can, I.; et al. P21 Produces a Bioactive Secretome That Places Stressed Cells under Immunosurveillance. Science 2021, 374, eabb3420. [Google Scholar] [CrossRef]
  87. Arora, I.; Sharma, M.; Sun, L.Y.; Tollefsbol, T.O. The Epigenetic Link between Polyphenols, Aging and Age-Related Diseases. Genes 2020, 11, 1094. [Google Scholar] [CrossRef]
  88. Manoukian, P.; Kuhnen, L.C.; van Laarhoven, H.W.M.; Bijlsma, M.F. Association of Epigenetic Landscapes with Heterogeneity and Plasticity in Pancreatic Cancer. Crit. Rev. Oncol. Hematol. 2025, 206, 104573. [Google Scholar] [CrossRef]
  89. Khan, M.I.; Rath, S.; Adhami, V.M.; Mukhtar, H. Targeting Epigenome with Dietary Nutrients in Cancer: Current Advances and Future Challenges. Pharmacol. Res. 2018, 129, 375–387. [Google Scholar] [CrossRef]
  90. Sood, U.; Garg, G.; Lal, R. Editorial: Thematic Issue on Modulating the Environment with Microbes. FEMS Microbes 2024, 5, xtae021. [Google Scholar] [CrossRef] [PubMed]
  91. He, W.-J.; Lv, C.-H.; Chen, Z.; Shi, M.; Zeng, C.-X.; Hou, D.-X.; Qin, S. The Regulatory Effect of Phytochemicals on Chronic Diseases by Targeting Nrf2-ARE Signaling Pathway. Antioxidants 2023, 12, 236. [Google Scholar] [CrossRef] [PubMed]
  92. Schiavoni, V.; Emanuelli, M.; Sartini, D.; Salvolini, E.; Pozzi, V.; Campagna, R. Curcumin and Its Analogues in Oral Squamous Cell Carcinoma: State-of-the-Art and Therapeutic Potential. Anticancer Agents Med. Chem. 2025, 25, 313–329. [Google Scholar] [CrossRef]
  93. Campagna, R.; Cecati, M.; Vignini, A. The Multifaceted Role of the Polyphenol Curcumin: A Focus on Type 2 Diabetes Mellitus. Curr. Diabetes Rev. 2025, 21, e15733998313402. [Google Scholar] [CrossRef] [PubMed]
  94. Schiavoni, V.; Emanuelli, M.; Milanese, G.; Galosi, A.B.; Pompei, V.; Salvolini, E.; Campagna, R. Nrf2 Signaling in Renal Cell Carcinoma: A Potential Candidate for the Development of Novel Therapeutic Strategies. Int. J. Mol. Sci. 2024, 25, 13239. [Google Scholar] [CrossRef] [PubMed]
  95. Dong, Z.; Hou, L.; Luo, W.; Pan, L.-H.; Li, X.; Tan, H.-P.; Wu, R.-D.; Lu, H.; Yao, K.; Mu, M.-D.; et al. Myocardial Infarction Drives Trained Immunity of Monocytes, Accelerating Atherosclerosis. Eur. Heart J. 2024, 45, 669–684. [Google Scholar] [CrossRef] [PubMed]
  96. Bergonzini, M.; Loreni, F.; Lio, A.; Russo, M.; Saitto, G.; Cammardella, A.; Irace, F.; Tramontin, C.; Chello, M.; Lusini, M.; et al. Panoramic on Epigenetics in Coronary Artery Disease and the Approach of Personalized Medicine. Biomedicines 2023, 11, 2864. [Google Scholar] [CrossRef] [PubMed]
  97. Ferrucci, L.; Fabbri, E. Inflammageing: Chronic Inflammation in Ageing, Cardiovascular Disease, and Frailty. Nat. Rev. Cardiol. 2018, 15, 505–522. [Google Scholar] [CrossRef]
  98. Marsit, C.J. Influence of Environmental Exposure on Human Epigenetic Regulation. J. Exp. Biol. 2015, 218, 71–79. [Google Scholar] [CrossRef]
  99. Zhang, Q.; Liu, Y.; Li, Y.; Bai, G.; Pang, J.; Wu, M.; Li, J.; Zhao, X.; Xia, Y. Implications of Gut Microbiota-Mediated Epigenetic Modifications in Intestinal Diseases. Gut Microbes 2025, 17, 2508426. [Google Scholar] [CrossRef]
  100. Begum, N.; Mandhare, A.; Tryphena, K.P.; Srivastava, S.; Shaikh, M.F.; Singh, S.B.; Khatri, D.K. Epigenetics in Depression and Gut-Brain Axis: A Molecular Crosstalk. Front. Aging Neurosci. 2022, 14, 1048333. [Google Scholar] [CrossRef]
  101. Liu, X.; Shao, J.; Liao, Y.-T.; Wang, L.-N.; Jia, Y.; Dong, P.; Liu, Z.; He, D.; Li, C.; Zhang, X. Regulation of Short-Chain Fatty Acids in the Immune System. Front. Immunol. 2023, 14, 1186892. [Google Scholar] [CrossRef]
  102. Rubio, K.; Hernández-Cruz, E.Y.; Rogel-Ayala, D.G.; Sarvari, P.; Isidoro, C.; Barreto, G.; Pedraza-Chaverri, J. Nutriepigenomics in Environmental-Associated Oxidative Stress. Antioxidants 2023, 12, 771. [Google Scholar] [CrossRef]
  103. Ji, H.; Khurana Hershey, G.K. Genetic and Epigenetic Influence on the Response to Environmental Particulate Matter. J. Allergy Clin. Immunol. 2012, 129, 33–41. [Google Scholar] [CrossRef]
  104. Breton, C.V.; Landon, R.; Kahn, L.G.; Enlow, M.B.; Peterson, A.K.; Bastain, T.; Braun, J.; Comstock, S.S.; Duarte, C.S.; Hipwell, A.; et al. Exploring the Evidence for Epigenetic Regulation of Environmental Influences on Child Health across Generations. Commun. Biol. 2021, 4, 769. [Google Scholar] [CrossRef] [PubMed]
  105. Ho, S.-M.; Johnson, A.; Tarapore, P.; Janakiram, V.; Zhang, X.; Leung, Y.-K. Environmental Epigenetics and Its Implication on Disease Risk and Health Outcomes. ILAR J. 2012, 53, 289–305, Erratum in ILAR J. 2017, 58, 413. [Google Scholar] [CrossRef]
  106. Alegría-Torres, J.A.; Baccarelli, A.; Bollati, V. Epigenetics and Lifestyle. Epigenomics 2011, 3, 267–277. [Google Scholar] [CrossRef] [PubMed]
  107. Barouki, R.; Melén, E.; Herceg, Z.; Beckers, J.; Chen, J.; Karagas, M.; Puga, A.; Xia, Y.; Chadwick, L.; Yan, W.; et al. Epigenetics as a Mechanism Linking Developmental Exposures to Long-Term Toxicity. Environ. Int. 2018, 114, 77–86. [Google Scholar] [CrossRef]
  108. Wang, G.; Su, H.; Guo, Z.; Li, H.; Jiang, Z.; Cao, Y.; Li, C. Rubus Occidentalis and Its Bioactive Compounds against Cancer: From Molecular Mechanisms to Translational Advances. Phytomedicine 2024, 126, 155029. [Google Scholar] [CrossRef]
  109. Wu, X.; Wu, H.; Zhong, M.; Chen, Y.; Su, W.; Li, P. Epigenetic Regulation by Naringenin and Naringin: A Literature Review Focused on the Mechanisms Underlying Its Pharmacological Effects. Fitoterapia 2025, 181, 106353. [Google Scholar] [CrossRef]
  110. Babar, Q.; Saeed, A.; Tabish, T.A.; Pricl, S.; Townley, H.; Thorat, N. Novel Epigenetic Therapeutic Strategies and Targets in Cancer. Biochim. Biophys. Acta (BBA)—Mol. Basis Dis. 2022, 1868, 166552. [Google Scholar] [CrossRef] [PubMed]
  111. Lagosz-Cwik, K.B.; Melnykova, M.; Nieboga, E.; Schuster, A.; Bysiek, A.; Dudek, S.; Lipska, W.; Kantorowicz, M.; Tyrakowski, M.; Darczuk, D.; et al. Mapping of DNA Methylation-Sensitive Cellular Processes in Gingival and Periodontal Ligament Fibroblasts in the Context of Periodontal Tissue Homeostasis. Front. Immunol. 2023, 14, 1078031. [Google Scholar] [CrossRef]
  112. Nisar, A.; Jagtap, S.; Vyavahare, S.; Deshpande, M.; Harsulkar, A.; Ranjekar, P.; Prakash, O. Phytochemicals in the Treatment of Inflammation-Associated Diseases: The Journey from Preclinical Trials to Clinical Practice. Front. Pharmacol. 2023, 14, 1177050. [Google Scholar] [CrossRef]
  113. Hossain, M.S.; Wazed, M.A.; Asha, S.; Amin, M.R.; Shimul, I.M. Dietary Phytochemicals in Health and Disease: Mechanisms, Clinical Evidence, and Applications—A Comprehensive Review. Food Sci. Amp. Nutr. 2025, 13, e70101. [Google Scholar] [CrossRef]
  114. Hosseini, B.; Berthon, B.S.; Saedisomeolia, A.; Starkey, M.R.; Collison, A.; Wark, P.A.B.; Wood, L.G. Effects of Fruit and Vegetable Consumption on Inflammatory Biomarkers and Immune Cell Populations: A Systematic Literature Review and Meta-Analysis. Am. J. Clin. Nutr. 2018, 108, 136–155. [Google Scholar] [CrossRef]
  115. Caban, M.; Lewandowska, U. Polyphenols and the Potential Mechanisms of Their Therapeutic Benefits against Inflammatory Bowel Diseases. J. Funct. Foods 2022, 95, 105181. [Google Scholar] [CrossRef]
  116. Hasnat, H.; Shompa, S.A.; Islam, M.M.; Alam, S.; Richi, F.T.; Emon, N.U.; Ashrafi, S.; Ahmed, N.U.; Chowdhury, M.N.R.; Fatema, N.; et al. Flavonoids: A Treasure House of Prospective Pharmacological Potentials. Heliyon 2024, 10, e27533. [Google Scholar] [CrossRef]
  117. Fahmy, M.I.; Sadek, M.A.; Abdou, K.; El-Dessouki, A.M.; El-Shiekh, R.A.; Khalaf, S.S. Orientin: A Comprehensive Review of a Promising Bioactive Flavonoid. Inflammopharmacology 2025, 33, 1713–1728. [Google Scholar] [CrossRef] [PubMed]
  118. Liga, S.; Paul, C.; Péter, F. Flavonoids: Overview of Biosynthesis, Biological Activity, and Current Extraction Techniques. Plants 2023, 12, 2732. [Google Scholar] [CrossRef] [PubMed]
  119. Masenga, S.K.; Kabwe, L.S.; Chakulya, M.; Kirabo, A. Mechanisms of Oxidative Stress in Metabolic Syndrome. Int. J. Mol. Sci. 2023, 24, 7898. [Google Scholar] [CrossRef] [PubMed]
  120. Hinz, M.; Scheidereit, C. The IκB Kinase Complex in NF-κB Regulation and Beyond. EMBO Rep. 2014, 15, 46–61. [Google Scholar] [CrossRef]
  121. Roberti, A.; Chaffey, L.E.; Greaves, D.R. NF-κB Signaling and Inflammation-Drug Repurposing to Treat Inflammatory Disorders? Biology 2022, 11, 372. [Google Scholar] [CrossRef]
  122. Chauhan, A.; Islam, A.U.; Prakash, H.; Singh, S. Phytochemicals Targeting NF-κB Signaling: Potential Anti-Cancer Interventions. J. Pharm. Anal. 2022, 12, 394–405. [Google Scholar] [CrossRef]
  123. Saha, S.; Buttari, B.; Panieri, E.; Profumo, E.; Saso, L. An Overview of Nrf2 Signaling Pathway and Its Role in Inflammation. Molecules 2020, 25, 5474. [Google Scholar] [CrossRef]
  124. Romero-Durán, M.A.; Silva-García, O.; Perez-Aguilar, J.M.; Baizabal-Aguirre, V.M. Mechanisms of Keap1/Nrf2 Modulation in Bacterial Infections: Implications in Persistence and Clearance. Front. Immunol. 2024, 15, 1508787. [Google Scholar] [CrossRef] [PubMed]
  125. Dewanjee, S.; Bhattacharya, H.; Bhattacharyya, C.; Chakraborty, P.; Fleishman, J.; Alexiou, A.; Papadakis, M.; Jha, S.K. Nrf2/Keap1/ARE Regulation by Plant Secondary Metabolites: A New Horizon in Brain Tumor Management. Cell Commun. Signal. 2024, 22, 497. [Google Scholar] [CrossRef] [PubMed]
  126. Pouremamali, F.; Pouremamali, A.; Dadashpour, M.; Soozangar, N.; Jeddi, F. An Update of Nrf2 Activators and Inhibitors in Cancer Prevention/Promotion. Cell Commun. Signal. 2022, 20, 100. [Google Scholar] [CrossRef]
  127. Zhang, L.; Xu, L.-Y.; Tang, F.; Liu, D.; Zhao, X.-L.; Zhang, J.-N.; Xia, J.; Wu, J.-J.; Yang, Y.; Peng, C.; et al. New Perspectives on the Therapeutic Potential of Quercetin in Non-Communicable Diseases: Targeting Nrf2 to Counteract Oxidative Stress and Inflammation. J. Pharm. Anal. 2024, 14, 100930. [Google Scholar] [CrossRef]
  128. Sharbafshaaer, M.; Pepe, R.; Notariale, R.; Canale, F.; Tedeschi, G.; Tessitore, A.; Bergamo, P.; Trojsi, F. Beyond Antioxidants: The Emerging Role of Nrf2 Activation in Amyotrophic Lateral Sclerosis (ALS). Int. J. Mol. Sci. 2025, 26, 9872. [Google Scholar] [CrossRef] [PubMed]
  129. Li, C.; Powell, K.; Giliberto, L.; LeDoux, C.; d’Abramo, C.; Sciubba, D.; Al Abed, Y. Non-Electrophilic Activation of NRF2 in Neurological Disorders: Therapeutic Promise of Non-Pharmacological Strategies. Antioxidants 2025, 14, 1047. [Google Scholar] [CrossRef]
  130. Schreiber, S.P.; Villalba, J.; Meyer-Ficca, M.L. Potential Epigenetic Impacts of Phytochemicals on Ruminant Health and Production: Connecting Lines of Evidence. Animals 2025, 15, 1787. [Google Scholar] [CrossRef]
  131. Thakur, V.S.; Deb, G.; Babcook, M.A.; Gupta, S. Plant Phytochemicals as Epigenetic Modulators: Role in Cancer Chemoprevention. AAPS J. 2014, 16, 151–163. [Google Scholar] [CrossRef]
  132. Pang, Y.; Zhang, H.; Li, H.; Wang, X.; Wei, P.; Yi, L.; Lin, S. Epigenetic Insights into Aging: Emerging Roles of Natural Products in Therapeutic Interventions. Phytother. Res. 2025, 39, 3300–3322. [Google Scholar] [CrossRef]
  133. Zhang, Z.; Wang, G.; Li, Y.; Lei, D.; Xiang, J.; Ouyang, L.; Wang, Y.; Yang, J. Recent Progress in DNA Methyltransferase Inhibitors as Anticancer Agents. Front. Pharmacol. 2022, 13, 1072651. [Google Scholar] [CrossRef]
  134. Akone, S.H.; Ntie-Kang, F.; Stuhldreier, F.; Ewonkem, M.B.; Noah, A.M.; Mouelle, S.E.M.; Müller, R. Natural Products Impacting DNA Methyltransferases and Histone Deacetylases. Front. Pharmacol. 2020, 11, 992. [Google Scholar] [CrossRef] [PubMed]
  135. Daskalos, A.; Oleksiewicz, U.; Filia, A.; Nikolaidis, G.; Xinarianos, G.; Gosney, J.R.; Malliri, A.; Field, J.K.; Liloglou, T. UHRF1-mediated Tumor Suppressor Gene Inactivation in Nonsmall Cell Lung Cancer. Cancer 2011, 117, 1027–1037. [Google Scholar] [CrossRef] [PubMed]
  136. Cacan, E.; Ali, M.W.; Boyd, N.H.; Hooks, S.B.; Greer, S.F. Inhibition of HDAC1 and DNMT1 Modulate RGS10 Expression and Decrease Ovarian Cancer Chemoresistance. PLoS ONE 2014, 9, e87455. [Google Scholar] [CrossRef] [PubMed]
  137. Číž, M.; Dvořáková, A.; Skočková, V.; Kubala, L. The Role of Dietary Phenolic Compounds in Epigenetic Modulation Involved in Inflammatory Processes. Antioxidants 2020, 9, 691. [Google Scholar] [CrossRef]
  138. Boaru, D.L.; Fraile-Martinez, O.; De Leon-Oliva, D.; Garcia-Montero, C.; De Castro-Martinez, P.; Miranda-Gonzalez, A.; Saez, M.A.; Muñon-Zamarron, L.; Castillo-Ruiz, E.; Barrena-Blázquez, S.; et al. Harnessing the Anti-Inflammatory Properties of Polyphenols in the Treatment of Inflammatory Bowel Disease. Int. J. Biol. Sci. 2024, 20, 5608–5672. [Google Scholar] [CrossRef]
  139. Eckschlager, T.; Plch, J.; Stiborova, M.; Hrabeta, J. Histone Deacetylase Inhibitors as Anticancer Drugs. Int. J. Mol. Sci. 2017, 18, 1414. [Google Scholar] [CrossRef] [PubMed]
  140. Rebollo-Hernanz, M.; Zhang, Q.; Aguilera, Y.; Martín-Cabrejas, M.A.; Gonzalez De Mejia, E. Phenolic Compounds from Coffee By-Products Modulate Adipogenesis-Related Inflammation, Mitochondrial Dysfunction, and Insulin Resistance in Adipocytes, via Insulin/PI3K/AKT Signaling Pathways. Food Chem. Toxicol. 2019, 132, 110672. [Google Scholar] [CrossRef]
  141. Dierckx, T.; Haidar, M.; Grajchen, E.; Wouters, E.; Vanherle, S.; Loix, M.; Boeykens, A.; Bylemans, D.; Hardonnière, K.; Kerdine-Römer, S.; et al. Phloretin suppresses neuroinflammation by autophagy-mediated Nrf2 activation in macrophages. J. Neuroinflammation 2021, 18, 148. [Google Scholar] [CrossRef]
  142. Liu, Z.; Huang, M.; Hong, Y.; Wang, S.; Xu, Y.; Zhong, C.; Zhang, J.; Zhuang, Z.; Shan, S.; Ren, T. Isovalerylspiramycin I Suppresses Non-Small Cell Lung Carcinoma Growth through ROS-Mediated Inhibition of PI3K/AKT Signaling Pathway. Int. J. Biol. Sci. 2022, 18, 3714–3730. [Google Scholar] [CrossRef]
  143. Tuli, H.S.; Tuorkey, M.J.; Thakral, F.; Sak, K.; Kumar, M.; Sharma, A.K.; Sharma, U.; Jain, A.; Aggarwal, V.; Bishayee, A. Molecular Mechanisms of Action of Genistein in Cancer: Recent Advances. Front. Pharmacol. 2019, 10, 1336. [Google Scholar] [CrossRef]
  144. Record, I.R.; Dreosti, I.E.; McInerney, J.K. The Antioxidant Activity of Genistein in Vitro. J. Nutr. Biochem. 1995, 6, 481–485. [Google Scholar] [CrossRef]
  145. Kim, D.H.; Jung, W.-S.; Kim, M.E.; Lee, H.-W.; Youn, H.-Y.; Seon, J.K.; Lee, H.-N.; Lee, J.S. Genistein Inhibits Pro-Inflammatory Cytokines in Human Mast Cell Activation through the Inhibition of the ERK Pathway. Int. J. Mol. Med. 2014, 34, 1669–1674. [Google Scholar] [CrossRef]
  146. Liu, T.; Zhang, L.; Joo, D.; Sun, S.-C. NF-κB Signaling in Inflammation. Signal Transduct. Target. Ther. 2017, 2, 17023. [Google Scholar] [CrossRef] [PubMed]
  147. Xia, L.; Zhu, G.; Peng, Q.; Li, X.; Zou, X.; Zhang, W.; Zhao, L.; Li, X.; Wu, P.; Luo, A.; et al. Natural Products Combating EGFR-TKIs Resistance in Cancer. Eur. J. Med. Chem. Rep. 2025, 13, 100251. [Google Scholar] [CrossRef]
  148. Zhao, Z.; Shin, H.S.; Satsu, H.; Totsuka, M.; Shimizu, M. 5-Caffeoylquinic Acid and Caffeic Acid Down-Regulate the Oxidative Stress- and TNF-α-Induced Secretion of Interleukin-8 from Caco-2 Cells. J. Agric. Food Chem. 2008, 56, 3863–3868. [Google Scholar] [CrossRef]
  149. Tyszka-Czochara, M.; Bukowska-Strakova, K.; Kocemba-Pilarczyk, K.A.; Majka, M. Caffeic Acid Targets AMPK Signaling and Regulates Tricarboxylic Acid Cycle Anaplerosis While Metformin Downregulates HIF-1α-Induced Glycolytic Enzymes in Human Cervical Squamous Cell Carcinoma Lines. Nutrients 2018, 10, 841. [Google Scholar] [CrossRef] [PubMed]
  150. Zhu, Y.; Yue, P.; Dickinson, C.F.; Yang, J.K.; Datanagan, K.; Zhai, N.; Zhang, Y.; Miklossy, G.; Lopez-Tapia, F.; Tius, M.A.; et al. Natural Product Preferentially Targets Redox and Metabolic Adaptations and Aberrantly Active STAT3 to Inhibit Breast Tumor Growth in Vivo. Cell Death Dis. 2022, 13, 1022. [Google Scholar] [CrossRef] [PubMed]
  151. Zhao, Y.; Liu, J. Anti-Inflammatory Effects of p-Coumaric Acid in LPS-Stimulated RAW264.7 Cells: Involvement of NF-κB and MAPKs Pathways. Med. Chem. 2016, 6, 327–330. [Google Scholar] [CrossRef]
  152. Wu, C.; Sun, C.; Han, X.; Ye, Y.; Qin, Y.; Liu, S. Sanyin Formula Enhances the Therapeutic Efficacy of Paclitaxel in Triple-Negative Breast Cancer Metastases through the JAK/STAT3 Pathway in Mice. Pharmaceuticals 2022, 16, 9. [Google Scholar] [CrossRef]
  153. Casari, G.; Romaldi, B.; Scirè, A.; Minnelli, C.; Marzioni, D.; Ferretti, G.; Armeni, T. Epigenetic Properties of Compounds Contained in Functional Foods Against Cancer. Biomolecules 2024, 15, 15. [Google Scholar] [CrossRef]
  154. Surai, P. Silymarin as a Natural Antioxidant: An Overview of the Current Evidence and Perspectives. Antioxidants 2015, 4, 204–247. [Google Scholar] [CrossRef]
  155. Bouyahya, A.; El Hachlafi, N.; Aanniz, T.; Bourais, I.; Mechchate, H.; Benali, T.; Shariati, M.A.; Burkov, P.; Lorenzo, J.M.; Wilairatana, P.; et al. Natural Bioactive Compounds Targeting Histone Deacetylases in Human Cancers: Recent Updates. Molecules 2022, 27, 2568. [Google Scholar] [CrossRef]
  156. Neelab; Zeb, A.; Jamil, M. Milk Thistle Protects against Non-Alcoholic Fatty Liver Disease Induced by Dietary Thermally Oxidized Tallow. Heliyon 2024, 10, e31445. [Google Scholar] [CrossRef] [PubMed]
  157. Sugihara, A.; Nguyen, L.C.; Shamim, H.M.; Iida, T.; Nakase, M.; Takegawa, K.; Senda, M.; Jida, S.; Ueno, M. Mutation in Fission Yeast Phosphatidylinositol 4-Kinase Pik1 Is Synthetically Lethal with Defect in Telomere Protection Protein Pot1. Biochem. Biophys. Res. Commun. 2018, 496, 1284–1290. [Google Scholar] [CrossRef] [PubMed]
  158. Baş, H.; Kalender, Y.; Pandir, D.; Kalender, S. Effects of Lead Nitrate and Sodium Selenite on DNA Damage and Oxidative Stress in Diabetic and Non-Diabetic Rat Erythrocytes and Leucocytes. Environ. Toxicol. Pharmacol. 2015, 39, 1019–1026. [Google Scholar] [CrossRef]
  159. Astrain-Redin, N.; Sanmartin, C.; Sharma, A.K.; Plano, D. From Natural Sources to Synthetic Derivatives: The Allyl Motif as a Powerful Tool for Fragment-Based Design in Cancer Treatment. J. Med. Chem. 2023, 66, 3703–3731. [Google Scholar] [CrossRef]
  160. Zhang, M.; Pan, H.; Xu, Y.; Wang, X.; Qiu, Z.; Jiang, L. Allicin Decreases Lipopolysaccharide-Induced Oxidative Stress and Inflammation in Human Umbilical Vein Endothelial Cells through Suppression of Mitochondrial Dysfunction and Activation of Nrf2. Cell Physiol. Biochem. 2017, 41, 2255–2267. [Google Scholar] [CrossRef] [PubMed]
  161. Chen, M.; Gu, H.; Ye, Y.; Lin, B.; Sun, L.; Deng, W.; Zhang, J.; Liu, J. Protective Effects of Hesperidin against Oxidative Stress of Tert-Butyl Hydroperoxide in Human Hepatocytes. Food Chem. Toxicol. 2010, 48, 2980–2987. [Google Scholar] [CrossRef]
  162. Addi, M.; Elbouzidi, A.; Abid, M.; Tungmunnithum, D.; Elamrani, A.; Hano, C. An Overview of Bioactive Flavonoids from Citrus Fruits. Appl. Sci. 2021, 12, 29. [Google Scholar] [CrossRef]
  163. Li, C.; Schluesener, H. Health-Promoting Effects of the Citrus Flavanone Hesperidin. Crit. Rev. Food Sci. Nutr. 2017, 57, 613–631. [Google Scholar] [CrossRef]
  164. Tejada, S.; Pinya, S.; Martorell, M.; Capó, X.; Tur, J.A.; Pons, A.; Sureda, A. Potential Anti-Inflammatory Effects of Hesperidin from the Genus Citrus. Curr. Med. Chem. 2019, 25, 4929–4945. [Google Scholar] [CrossRef]
  165. Fernández-Bedmar, Z.; Anter, J.; Alonso-Moraga, A.; Martín De Las Mulas, J.; Millán-Ruiz, Y.; Guil-Luna, S. Demethylating and Anti-hepatocarcinogenic Potential of Hesperidin, a Natural Polyphenol of Citrus Juices. Mol. Carcinog. 2017, 56, 1653–1662. [Google Scholar] [CrossRef]
  166. Jiang, W.; Xia, T.; Liu, C.; Li, J.; Zhang, W.; Sun, C. Remodeling the Epigenetic Landscape of Cancer—Application Potential of Flavonoids in the Prevention and Treatment of Cancer. Front. Oncol. 2021, 11, 705903. [Google Scholar] [CrossRef]
  167. Dixon, R. Genistein. Phytochemistry 2002, 60, 205–211. [Google Scholar] [CrossRef]
  168. Goh, Y.X.; Jalil, J.; Lam, K.W.; Husain, K.; Premakumar, C.M. Genistein: A Review on Its Anti-Inflammatory Properties. Front. Pharmacol. 2022, 13, 820969. [Google Scholar] [CrossRef]
  169. Setchell, K.D.R.; Cassidy, A. Dietary Isoflavones: Biological Effects and Relevance to Human Health. J. Nutr. 1999, 129, 758S–767S. [Google Scholar] [CrossRef]
  170. Valsecchi, A.E.; Franchi, S.; Panerai, A.E.; Rossi, A.; Sacerdote, P.; Colleoni, M. The Soy Isoflavone Genistein Reverses Oxidative and Inflammatory State, Neuropathic Pain, Neurotrophic and Vasculature Deficits in Diabetes Mouse Model. Eur. J. Pharmacol. 2011, 650, 694–702. [Google Scholar] [CrossRef] [PubMed]
  171. Zhou, X.; Yuan, L.; Zhao, X.; Hou, C.; Ma, W.; Yu, H.; Xiao, R. Genistein Antagonizes Inflammatory Damage Induced by β-Amyloid Peptide in Microglia through TLR4 and NF-κB. Nutrition 2014, 30, 90–95. [Google Scholar] [CrossRef]
  172. Smolińska, E.; Moskot, M.; Jakóbkiewicz-Banecka, J.; Węgrzyn, G.; Banecki, B.; Szczerkowska-Dobosz, A.; Purzycka-Bohdan, D.; Gabig-Cimińska, M. Molecular Action of Isoflavone Genistein in the Human Epithelial Cell Line HaCaT. PLoS ONE 2018, 13, e0192297. [Google Scholar] [CrossRef] [PubMed]
  173. Hämäläinen, M.; Nieminen, R.; Asmawi, M.; Vuorela, P.; Vapaatalo, H.; Moilanen, E. Effects of Flavonoids on Prostaglandin E2 Production and on COX-2 and mPGES-1 Expressions in Activated Macrophages. Planta Med. 2011, 77, 1504–1511. [Google Scholar] [CrossRef] [PubMed]
  174. Ko, K.-P.; Kim, S.-W.; Ma, S.H.; Park, B.; Ahn, Y.; Lee, J.W.; Lee, M.H.; Kang, E.; Kim, L.S.; Jung, Y.; et al. Dietary Intake and Breast Cancer among Carriers and Noncarriers of BRCA Mutations in the Korean Hereditary Breast Cancer Study. Am. J. Clin. Nutr. 2013, 98, 1493–1501. [Google Scholar] [CrossRef]
  175. Bilir, B.; Sharma, N.V.; Lee, J.; Hammarstrom, B.; Svindland, A.; Kucuk, O.; Moreno, C.S. Effects of Genistein Supplementation on Genome-Wide DNA Methylation and Gene Expression in Patients with Localized Prostate Cancer. Int. J. Oncol. 2017, 51, 223–234. [Google Scholar] [CrossRef]
  176. Lu, H.; Shi, J.-X.; Zhang, D.-M.; Wang, H.-D.; Hang, C.-H.; Chen, H.-L.; Yin, H.-X. Inhibition of Hemolysate-Induced iNOS and COX-2 Expression by Genistein through Suppression of NF-κB Activation in Primary Astrocytes. J. Neurol. Sci. 2009, 278, 91–95. [Google Scholar] [CrossRef]
  177. Habtemariam, S. The Molecular Pharmacology of Phloretin: Anti-Inflammatory Mechanisms of Action. Biomedicines 2023, 11, 143. [Google Scholar] [CrossRef]
  178. Hao, X.; Bai, Y.; Li, W.; Zhang, M.X. Phloretin Attenuates Inflammation Induced by Subarachnoid Hemorrhage through Regulation of the TLR2/MyD88/NF-kB Pathway. Sci. Rep. 2024, 14, 26583. [Google Scholar] [CrossRef]
  179. Lee, S.-W.; Lee, G.; Jo, J.-H.; Yang, Y.; Ahn, J.-H. Biosynthesis of Phloretin and Its C-Glycosides through Stepwise Culture of Escherichia Coli. Appl. Biol. Chem. 2024, 67, 99. [Google Scholar] [CrossRef]
  180. Mansuri, M.L.; Parihar, P.; Solanki, I.; Parihar, M.S. Flavonoids in Modulation of Cell Survival Signalling Pathways. Genes Nutr. 2014, 9, 400. [Google Scholar] [CrossRef] [PubMed]
  181. Chang, W.-T.; Huang, W.-C.; Liou, C.-J. Evaluation of the Anti-Inflammatory Effects of Phloretin and Phlorizin in Lipopolysaccharide-Stimulated Mouse Macrophages. Food Chem. 2012, 134, 972–979. [Google Scholar] [CrossRef] [PubMed]
  182. Vargas-Mendoza, N. Hepatoprotective Effect of Silymarin. World J. Hepatol. 2014, 6, 144. [Google Scholar] [CrossRef]
  183. Wadhwa, K.; Pahwa, R.; Kumar, M.; Kumar, S.; Sharma, P.C.; Singh, G.; Verma, R.; Mittal, V.; Singh, I.; Kaushik, D.; et al. Mechanistic Insights into the Pharmacological Significance of Silymarin. Molecules 2022, 27, 5327. [Google Scholar] [CrossRef]
  184. Lovelace, E.S.; Wagoner, J.; MacDonald, J.; Bammler, T.; Bruckner, J.; Brownell, J.; Beyer, R.P.; Zink, E.M.; Kim, Y.-M.; Kyle, J.E.; et al. Silymarin Suppresses Cellular Inflammation by Inducing Reparative Stress Signaling. J. Nat. Prod. 2015, 78, 1990–2000. [Google Scholar] [CrossRef]
  185. Chittezhath, M.; Deep, G.; Singh, R.P.; Agarwal, C.; Agarwal, R. Silibinin Inhibits Cytokine-Induced Signaling Cascades and down-Regulates Inducible Nitric Oxide Synthase in Human Lung Carcinoma A549 Cells. Mol. Cancer Ther. 2008, 7, 1817–1826. [Google Scholar] [CrossRef]
  186. Zhao, Y.; Zhou, Y.; Gong, T.; Liu, Z.; Yang, W.; Xiong, Y.; Xiao, D.; Cifuentes, A.; Ibáñez, E.; Lu, W. The Clinical Anti-Inflammatory Effects and Underlying Mechanisms of Silymarin. iScience 2024, 27, 111109. [Google Scholar] [CrossRef]
  187. Bin-Jumah, M.N.; Nadeem, M.S.; Gilani, S.J.; Mubeen, B.; Ullah, I.; Alzarea, S.I.; Ghoneim, M.M.; Alshehri, S.; Al-Abbasi, F.A.; Kazmi, I. Lycopene: A Natural Arsenal in the War against Oxidative Stress and Cardiovascular Diseases. Antioxidants 2022, 11, 232. [Google Scholar] [CrossRef] [PubMed]
  188. Imran, M.; Ghorat, F.; Ul-Haq, I.; Ur-Rehman, H.; Aslam, F.; Heydari, M.; Shariati, M.A.; Okuskhanova, E.; Yessimbekov, Z.; Thiruvengadam, M.; et al. Lycopene as a Natural Antioxidant Used to Prevent Human Health Disorders. Antioxidants 2020, 9, 706. [Google Scholar] [CrossRef]
  189. Kulawik, A.; Cielecka-Piontek, J.; Czerny, B.; Kamiński, A.; Zalewski, P. The Relationship Between Lycopene and Metabolic Diseases. Nutrients 2024, 16, 3708. [Google Scholar] [CrossRef]
  190. Birková, A. Caffeic Acid: A Brief Overview of Its Presence, Metabolism, and Bioactivity. Bioact. Compd. Health Dis. 2020, 3, 74. [Google Scholar] [CrossRef]
  191. Alam, M.; Ahmed, S.; Elasbali, A.M.; Adnan, M.; Alam, S.; Hassan, M.I.; Pasupuleti, V.R. Therapeutic Implications of Caffeic Acid in Cancer and Neurological Diseases. Front. Oncol. 2022, 12, 860508. [Google Scholar] [CrossRef] [PubMed]
  192. Ku, H.-C.; Lee, S.-Y.; Yang, K.-C.; Kuo, Y.-H.; Su, M.-J. Modification of Caffeic Acid with Pyrrolidine Enhances Antioxidant Ability by Activating AKT/HO-1 Pathway in Heart. PLoS ONE 2016, 11, e0148545. [Google Scholar] [CrossRef] [PubMed]
  193. Pavlíková, N. Caffeic Acid and Diseases—Mechanisms of Action. Int. J. Mol. Sci. 2022, 24, 588. [Google Scholar] [CrossRef]
  194. Zielińska, D.; Zieliński, H.; Laparra-Llopis, J.M.; Szawara-Nowak, D.; Honke, J.; Giménez-Bastida, J.A. Caffeic Acid Modulates Processes Associated with Intestinal Inflammation. Nutrients 2021, 13, 554. [Google Scholar] [CrossRef]
  195. Yao, J.; Liu, Y.; Lin, H.; Shao, C.; Jin, X.; Peng, T.; Liu, Y. Caffeic Acid Activates Nrf2 Enzymes, Providing Protection against Oxidative Damage Induced by Ionizing Radiation. Brain Res. Bull. 2025, 224, 111325. [Google Scholar] [CrossRef]
  196. Yin, Y.; Wang, Z.; Hu, Y.; Wang, J.; Wang, Y.; Lu, Q. Caffeic Acid Hinders the Proliferation and Migration through Inhibition of IL-6 Mediated JAK-STAT-3 Signaling Axis in Human Prostate Cancer. Oncol. Res. 2024, 32, 1881–1890. [Google Scholar] [CrossRef]
  197. Shin, K.-M.; Kim, I.-T.; Park, Y.-M.; Ha, J.; Choi, J.-W.; Park, H.-J.; Lee, Y.S.; Lee, K.-T. Anti-Inflammatory Effect of Caffeic Acid Methyl Ester and Its Mode of Action through the Inhibition of Prostaglandin E2, Nitric Oxide and Tumor Necrosis Factor-α Production. Biochem. Pharmacol. 2004, 68, 2327–2336. [Google Scholar] [CrossRef]
  198. Cortez, N.; Villegas, C.; Burgos, V.; Cabrera-Pardo, J.R.; Ortiz, L.; González-Chavarría, I.; Nchiozem-Ngnitedem, V.-A.; Paz, C. Adjuvant Properties of Caffeic Acid in Cancer Treatment. Int. J. Mol. Sci. 2024, 25, 7631. [Google Scholar] [CrossRef]
  199. Park, J.-Y.; Yasir, M.; Lee, H.; Han, E.-T.; Han, J.-H.; Park, W.; Kwon, Y.-S.; Chun, W. Caffeic Acid Methyl Ester Inhibits LPS-induced Inflammatory Response through Nrf2 Activation and NF-κB Inhibition in Human Umbilical Vein Endothelial Cells. Exp. Ther. Med. 2023, 26, 559. [Google Scholar] [CrossRef] [PubMed]
  200. Wan, F.; Zhong, R.; Wang, M.; Zhou, Y.; Chen, Y.; Yi, B.; Hou, F.; Liu, L.; Zhao, Y.; Chen, L.; et al. Caffeic Acid Supplement Alleviates Colonic Inflammation and Oxidative Stress Potentially Through Improved Gut Microbiota Community in Mice. Front. Microbiol. 2021, 12, 784211. [Google Scholar] [CrossRef]
  201. Ding, Q.; Xu, Y.-M.; Lau, A.T.Y. The Epigenetic Effects of Coffee. Molecules 2023, 28, 1770. [Google Scholar] [CrossRef] [PubMed]
  202. Contardi, M.; Lenzuni, M.; Fiorentini, F.; Summa, M.; Bertorelli, R.; Suarato, G.; Athanassiou, A. Hydroxycinnamic Acids and Derivatives Formulations for Skin Damages and Disorders: A Review. Pharmaceutics 2021, 13, 999. [Google Scholar] [CrossRef]
  203. Xia, T.; Zhang, B.; Duan, W.; Zhang, J.; Wang, M. Nutrients and Bioactive Components from Vinegar: A Fermented and Functional Food. J. Funct. Foods 2020, 64, 103681. [Google Scholar] [CrossRef]
  204. Jeong, Y.H.; Yang, H.J.; Li, W.; Oh, Y.-C.; Choi, J.-G. Immune-Enhancing Effects of Gwakhyangjeonggi-San in RAW 264.7 Macrophage Cells through the MAPK/NF-κB Signaling Pathways. Int. J. Mol. Sci. 2024, 25, 9246. [Google Scholar] [CrossRef]
  205. Moustardas, P.; Aberdam, D.; Lagali, N. MAPK Pathways in Ocular Pathophysiology: Potential Therapeutic Drugs and Challenges. Cells 2023, 12, 617. [Google Scholar] [CrossRef]
  206. Moon, H.-R.; Yun, J.-M. P-Coumaric Acid Modulates Cholesterol Efflux and Lipid Accumulation and Inflammation in Foam Cells. Nutr. Res. Pract. 2024, 18, 774. [Google Scholar] [CrossRef] [PubMed]
  207. Medeiros-Linard, C.F.B.; Andrade-da-Costa, B.L.D.S.; Augusto, R.L.; Sereniki, A.; Trevisan, M.T.S.; Perreira, R.D.C.R.; De Souza, F.T.C.; Braz, G.R.F.; Lagranha, C.J.; De Souza, I.A.; et al. Anacardic Acids from Cashew Nuts Prevent Behavioral Changes and Oxidative Stress Induced by Rotenone in a Rat Model of Parkinson’s Disease. Neurotox. Res. 2018, 34, 250–262. [Google Scholar] [CrossRef] [PubMed]
  208. Augusto, R.L.; Mendonça, I.P.; De Albuquerque Rego, G.N.; Pereira, D.D.; Da Penha Gonçalves, L.V.; Dos Santos, M.L.; De Souza, R.F.; Moreno, G.M.M.; Cardoso, P.R.G.; De Souza Andrade, D.; et al. Purified Anacardic Acids Exert Multiple Neuroprotective Effects in Pesticide Model of Parkinson’s Disease: In Vivo and in Silico Analysis. IUBMB Life 2020, 72, 1765–1779. [Google Scholar] [CrossRef]
  209. Tomasiak, P.; Janisiak, J.; Rogińska, D.; Perużyńska, M.; Machaliński, B.; Tarnowski, M. Garcinol and Anacardic Acid, Natural Inhibitors of Histone Acetyltransferases, Inhibit Rhabdomyosarcoma Growth and Proliferation. Molecules 2023, 28, 5292. [Google Scholar] [CrossRef]
  210. Rais, N.; Ved, A.; Ahmad, R.; Kumar, M.; Deepak Barbhai, M.; Radha; Chandran, D.; Dey, A.; Dhumal, S.; Senapathy, M.; et al. S-Allyl-L-Cysteine—A Garlic Bioactive: Physicochemical Nature, Mechanism, Pharmacokinetics, and Health Promoting Activities. J. Funct. Foods 2023, 107, 105657. [Google Scholar] [CrossRef]
  211. Arreola, R.; Quintero-Fabián, S.; López-Roa, R.I.; Flores-Gutiérrez, E.O.; Reyes-Grajeda, J.P.; Carrera-Quintanar, L.; Ortuño-Sahagún, D. Immunomodulation and Anti-Inflammatory Effects of Garlic Compounds. J. Immunol. Res. 2015, 2015, 401630. [Google Scholar] [CrossRef] [PubMed]
  212. Sasi, M.; Kumar, S.; Kumar, M.; Thapa, S.; Prajapati, U.; Tak, Y.; Changan, S.; Saurabh, V.; Kumari, S.; Kumar, A.; et al. Garlic (Allium sativum L.) Bioactives and Its Role in Alleviating Oral Pathologies. Antioxidants 2021, 10, 1847. [Google Scholar] [CrossRef]
  213. Ansary, J.; Forbes-Hernández, T.Y.; Gil, E.; Cianciosi, D.; Zhang, J.; Elexpuru-Zabaleta, M.; Simal-Gandara, J.; Giampieri, F.; Battino, M. Potential Health Benefit of Garlic Based on Human Intervention Studies: A Brief Overview. Antioxidants 2020, 9, 619. [Google Scholar] [CrossRef]
  214. Nian, H.; Delage, B.; Pinto, J.T.; Dashwood, R.H. Allyl Mercaptan, a Garlic-Derived Organosulfur Compound, Inhibits Histone Deacetylase and Enhances Sp3 Binding on the P21WAF1 Promoter. Carcinogenesis 2008, 29, 1816–1824. [Google Scholar] [CrossRef]
  215. Taleghani, A.; Emami, S.A.; Tayarani-Najaran, Z. Artemisia: A Promising Plant for the Treatment of Cancer. Bioorganic Med. Chem. 2020, 28, 115180. [Google Scholar] [CrossRef]
  216. Shoaib, M.; Shah, I.; Ali, N.; Adhikari, A.; Tahir, M.N.; Shah, S.W.A.; Ishtiaq, S.; Khan, J.; Khan, S.; Umer, M.N. Sesquiterpene Lactone! A Promising Antioxidant, Anticancer and Moderate Antinociceptive Agent from Artemisia Macrocephala Jacquem. BMC Complement. Altern. Med. 2017, 17, 27. [Google Scholar] [CrossRef]
  217. Zhu, M.; Wang, Y.; Han, J.; Sun, Y.; Wang, S.; Yang, B.; Wang, Q.; Kuang, H. Artesunate Exerts Organ- and Tissue-Protective Effects by Regulating Oxidative Stress, Inflammation, Autophagy, Apoptosis, and Fibrosis: A Review of Evidence and Mechanisms. Antioxidants 2024, 13, 686. [Google Scholar] [CrossRef]
  218. Zhang, Q.; Yi, H.; Yao, H.; Lu, L.; He, G.; Wu, M.; Zheng, C.; Li, Y.; Chen, S.; Li, L.; et al. Artemisinin Derivatives Inhibit Non-Small Cell Lung Cancer Cells Through Induction of ROS-Dependent Apoptosis/Ferroptosis. J. Cancer 2021, 12, 4075–4085. [Google Scholar] [CrossRef]
  219. Efferth, T.; Oesch, F. The Immunosuppressive Activity of Artemisinin-type Drugs towards Inflammatory and Autoimmune Diseases. Med. Res. Rev. 2021, 41, 3023–3061. [Google Scholar] [CrossRef]
  220. Laurindo, L.F.; Santos, A.R.D.O.D.; Carvalho, A.C.A.D.; Bechara, M.D.; Guiguer, E.L.; Goulart, R.D.A.; Vargas Sinatora, R.; Araújo, A.C.; Barbalho, S.M. Phytochemicals and Regulation of NF-kB in Inflammatory Bowel Diseases: An Overview of In Vitro and In Vivo Effects. Metabolites 2023, 13, 96. [Google Scholar] [CrossRef]
  221. Ma, Z.; Woon, C.Y.-N.; Liu, C.-G.; Cheng, J.-T.; You, M.; Sethi, G.; Wong, A.L.-A.; Ho, P.C.-L.; Zhang, D.; Ong, P.; et al. Repurposing Artemisinin and Its Derivatives as Anticancer Drugs: A Chance or Challenge? Front. Pharmacol. 2021, 12, 828856. [Google Scholar] [CrossRef] [PubMed]
  222. Sagar, S.M.; Yance, D.; Wong, R.K. Natural Health Products That Inhibit Angiogenesis: A Potential Source for Investigational New Agents to Treat Cancer—Part 1. Curr. Oncol. 2006, 13, 14–26. [Google Scholar] [CrossRef] [PubMed]
  223. Kazmi, S.T.B.; Fatima, H.; Naz, I.; Kanwal, N.; Haq, I. Pre-Clinical Studies Comparing the Anti-Inflammatory Potential of Artemisinic Compounds by Targeting NFκB/TNF-α/NLRP3 and Nrf2/TRX Pathways in Balb/C Mice. Front. Pharmacol. 2024, 15, 1352827. [Google Scholar] [CrossRef] [PubMed]
  224. Silva, G.D.S.E.; Marques, J.N.D.J.; Linhares, E.P.M.; Bonora, C.M.; Costa, É.T.; Saraiva, M.F. Review of Anticancer Activity of Monoterpenoids: Geraniol, Nerol, Geranial and Neral. Chem.-Biol. Interact. 2022, 362, 109994. [Google Scholar] [CrossRef]
  225. Ben Ammar, R.; Mohamed, M.E.; Alfwuaires, M.; Abdulaziz Alamer, S.; Bani Ismail, M.; Veeraraghavan, V.P.; Sekar, A.K.; Ksouri, R.; Rajendran, P. Anti-Inflammatory Activity of Geraniol Isolated from Lemon Grass on Ox-LDL-Stimulated Endothelial Cells by Upregulation of Heme Oxygenase-1 via PI3K/Akt and Nrf-2 Signaling Pathways. Nutrients 2022, 14, 4817. [Google Scholar] [CrossRef]
  226. Kim, H.; Li, S.; Nilkhet, S.; Baek, S.J. Anti-Cancer Activity of Rose-Geranium Essential Oil and Its Bioactive Compound Geraniol in Colorectal Cancer Cells. Appl. Biol. Chem. 2025, 68, 30. [Google Scholar] [CrossRef]
  227. Bhattacharjee, S.; Dashwood, R.H. Epigenetic Regulation of NRF2/KEAP1 by Phytochemicals. Antioxidants 2020, 9, 865. [Google Scholar] [CrossRef]
  228. Nurkolis, F.; Taslim, N.A.; Syahputra, R.A.; d’Arqom, A.; Tjandrawinata, R.R.; Purba, A.K.R.; Mustika, A. Food Phytochemicals as Epigenetic Modulators in Diabetes: A Systematic Review. J. Agric. Food Res. 2025, 21, 101873. [Google Scholar] [CrossRef]
  229. Gowd, V.; Kanika; Jori, C.; Chaudhary, A.A.; Rudayni, H.A.; Rashid, S.; Khan, R. Resveratrol and Resveratrol Nano-Delivery Systems in the Treatment of Inflammatory Bowel Disease. J. Nutr. Biochem. 2022, 109, 109101. [Google Scholar] [CrossRef]
  230. Song, X.; Wang, Y.; Gao, L. Mechanism of Antioxidant Properties of Quercetin and Quercetin-DNA Complex. J. Mol. Model. 2020, 26, 133. [Google Scholar] [CrossRef]
  231. Olasehinde, T.A.; Olaokun, O.O. Apigenin and Inflammation in the Brain: Can Apigenin Inhibit Neuroinflammation in Preclinical Models? Inflammopharmacology 2024, 32, 3099–3108. [Google Scholar] [CrossRef]
  232. Pinto, C.; Cidade, H.; Pinto, M.; Tiritan, M.E. Chiral Flavonoids as Antitumor Agents. Pharmaceuticals 2021, 14, 1267. [Google Scholar] [CrossRef]
  233. Iordache, M.P.; Buliman, A.; Costea-Firan, C.; Gligore, T.C.I.; Cazacu, I.S.; Stoian, M.; Teoibaș-Şerban, D.; Blendea, C.-D.; Protosevici, M.G.-I.; Tanase, C.; et al. Immunological and Inflammatory Biomarkers in the Prognosis, Prevention, and Treatment of Ischemic Stroke: A Review of a Decade of Advancement. Int. J. Mol. Sci. 2025, 26, 7928. [Google Scholar] [CrossRef]
  234. Ding, K.; Jiang, W.; Jia, H.; Lei, M. Synergistically Anti-Multiple Myeloma Effects: Flavonoid, Non-Flavonoid Polyphenols, and Bortezomib. Biomolecules 2022, 12, 1647. [Google Scholar] [CrossRef]
  235. Fatima, Z.; Itrat, N.; Israr, B.; Ahmad, A.M.R. Therapeutic Efficacy of Soy-Derived Bioactives: A Systematic Review of Nutritional Potency, Bioactive Therapeutics, and Clinical Biomarker Modulation. Foods 2025, 14, 3447. [Google Scholar] [CrossRef] [PubMed]
  236. Xie, Q.; Bai, Q.; Zou, L.; Zhang, Q.; Zhou, Y.; Chang, H.; Yi, L.; Zhu, J.; Mi, M. Genistein Inhibits DNA Methylation and Increases Expression of Tumor Suppressor Genes in Human Breast Cancer Cells. Genes Chromosomes Cancer 2014, 53, 422–431. [Google Scholar] [CrossRef]
  237. Montgomery, M.; Srinivasan, A. Epigenetic Gene Regulation by Dietary Compounds in Cancer Prevention. Adv. Nutr. 2019, 10, 1012–1028. [Google Scholar] [CrossRef]
  238. Zahari, C.N.M.C.; Mohamad, N.V.; Akinsanya, M.A.; Gengatharan, A. The Crimson Gem: Unveiling the Vibrant Potential of Lycopene as a Functional Food Ingredient. Food Chem. Adv. 2023, 3, 100510. [Google Scholar] [CrossRef]
  239. Tufail, T.; Bader Ul Ain, H.; Noreen, S.; Ikram, A.; Arshad, M.T.; Abdullahi, M.A. Nutritional Benefits of Lycopene and Beta-Carotene: A Comprehensive Overview. Food Sci. Nutr. 2024, 12, 8715–8741. [Google Scholar] [CrossRef]
  240. Maaz, M.; Sultan, M.T.; Khalid, M.U.; Raza, H.; Imran, M.; Hussain, M.; Al Abdulmonem, W.; Alsagaby, S.A.; Abdelgawad, M.A.; Ghoneim, M.M.; et al. A Comprehensive Review on the Molecular Mechanism of Lycopene in Cancer Therapy. Food Sci. Amp. Nutr. 2025, 13, e70608. [Google Scholar] [CrossRef]
  241. Mannino, F.; D’Angelo, T.; Pallio, G.; Ieni, A.; Pirrotta, I.; Giorgi, D.A.; Scarfone, A.; Mazziotti, S.; Booz, C.; Bitto, A.; et al. The Nutraceutical Genistein-Lycopene Combination Improves Bone Damage Induced by Glucocorticoids by Stimulating the Osteoblast Formation Process. Nutrients 2022, 14, 4296. [Google Scholar] [CrossRef]
  242. Zhang, K.; Wang, J.; Xu, B. Critical Review on Molecular Mechanisms for Genistein’s Beneficial Effects on Health Through Oxidative Stress Reduction. Antioxidants 2025, 14, 904. [Google Scholar] [CrossRef]
  243. Gao, Y.; Zhang, S.; Zhang, X.; Du, Y.; Ni, T.; Hao, S. Crosstalk between Metabolic and Epigenetic Modifications during Cell Carcinogenesis. iScience 2024, 27, 111359. [Google Scholar] [CrossRef]
  244. Böhm, V.; Bitsch, R. Intestinal Absorption of Lycopene from Different Matrices and Interactions to Other Carotenoids, the Lipid Status, and the Antioxidant Capacity of Human Plasma. Eur. J. Nutr. 1999, 38, 118–125. [Google Scholar] [CrossRef]
  245. Naveen Kumar, P.; Elango, P.; Asmathulla, S.; Kavimani, S. A Systematic Review on Lycopene and Its Beneficial Effects. Biomed. Pharmacol. J. 2017, 10, 2113–2120. [Google Scholar] [CrossRef]
  246. Narra, F.; Piragine, E.; Benedetti, G.; Ceccanti, C.; Florio, M.; Spezzini, J.; Troisi, F.; Giovannoni, R.; Martelli, A.; Guidi, L. Impact of Thermal Processing on Polyphenols, Carotenoids, Glucosinolates, and Ascorbic Acid in Fruit and Vegetables and Their Cardiovascular Benefits. Comp. Rev. Food Sci. Food Safe 2024, 23, e13426. [Google Scholar] [CrossRef] [PubMed]
  247. Shanaida, M.; Mykhailenko, O.; Lysiuk, R.; Hudz, N.; Balwierz, R.; Shulhai, A.; Shapovalova, N.; Shanaida, V.; Bjørklund, G. Carotenoids for Antiaging: Nutraceutical, Pharmaceutical, and Cosmeceutical Applications. Pharmaceuticals 2025, 18, 403. [Google Scholar] [CrossRef] [PubMed]
  248. Yin, S.; Xu, X.; Li, Y.; Fang, H.; Ren, J. Lycopene as a Potential Anticancer Agent: Current Evidence on Synergism, Drug Delivery Systems and Epidemiology (Review). Oncol. Lett. 2025, 30, 462. [Google Scholar] [CrossRef] [PubMed]
  249. Dong, R.; Wang, J.; Guan, R.; Sun, J.; Jin, P.; Shen, J. Role of Oxidative Stress in the Occurrence, Development, and Treatment of Breast Cancer. Antioxidants 2025, 14, 104. [Google Scholar] [CrossRef]
  250. Paschos, A.; Pandya, R.; Duivenvoorden, W.C.M.; Pinthus, J.H. Oxidative Stress in Prostate Cancer: Changing Research Concepts towards a Novel Paradigm for Prevention and Therapeutics. Prostate Cancer Prostatic Dis. 2013, 16, 217–225. [Google Scholar] [CrossRef]
  251. Selvaraj, N.R.; Nandan, D.; Nair, B.G.; Nair, V.A.; Venugopal, P.; Aradhya, R. Oxidative Stress and Redox Imbalance: Common Mechanisms in Cancer Stem Cells and Neurodegenerative Diseases. Cells 2025, 14, 511. [Google Scholar] [CrossRef]
  252. Remigante, A.; Morabito, R. Cellular and Molecular Mechanisms in Oxidative Stress-Related Diseases. Int. J. Mol. Sci. 2022, 23, 8017. [Google Scholar] [CrossRef]
  253. López-Contreras, A.K.; Martínez-Ruiz, M.G.; Olvera-Montaño, C.; Robles-Rivera, R.R.; Arévalo-Simental, D.E.; Castellanos-González, J.A.; Hernández-Chávez, A.; Huerta-Olvera, S.G.; Cardona-Muñoz, E.G.; Rodríguez-Carrizalez, A.D. Importance of the Use of Oxidative Stress Biomarkers and Inflammatory Profile in Aqueous and Vitreous Humor in Diabetic Retinopathy. Antioxidants 2020, 9, 891. [Google Scholar] [CrossRef]
  254. Ávila-Gálvez, M.Á.; Giménez-Bastida, J.A.; González-Sarrías, A.; Espín, J.C. New Insights into the Metabolism of the Flavanones Eriocitrin and Hesperidin: A Comparative Human Pharmacokinetic Study. Antioxidants 2021, 10, 435. [Google Scholar] [CrossRef]
  255. Hu, L.; Luo, Y.; Yang, J.; Cheng, C. Botanical Flavonoids: Efficacy, Absorption, Metabolism and Advanced Pharmaceutical Technology for Improving Bioavailability. Molecules 2025, 30, 1184. [Google Scholar] [CrossRef]
  256. Cord, D.; Rimbu, M.C.; Tanase, C.; Tablet, C.; Duca, G. Molecular Docking Study of Some Active Principles from Silybum Marianum, Chelidonium Majus, Ginkgo Biloba, Gelsemium Sempervirens, Artemisia Annua, and Taraxacum Officinale. Chem. J. Mold. 2025, 20, 100–105. [Google Scholar] [CrossRef]
  257. Rao, A.; Kumar, A.H.S. Computational Pharmacology Analysis of Lycopene to Identify Its Targets and Biological Effects in Humans. Appl. Sci. 2025, 15, 7815. [Google Scholar] [CrossRef]
  258. Liang, X.; Ma, C.; Yan, X.; Liu, X.; Liu, F. Advances in Research on Bioactivity, Metabolism, Stability and Delivery Systems of Lycopene. Trends Food Sci. Technol. 2019, 93, 185–196. [Google Scholar] [CrossRef]
  259. Yen, G.-C.; Cheng, H.-L.; Lin, L.-Y.; Chen, S.-C.; Hsu, C.L. The Potential Role of Phenolic Compounds on Modulating Gut Microbiota in Obesity. J. Food Drug Anal. 2020, 28, 195–205. [Google Scholar] [CrossRef]
  260. El-Saadony, M.T.; Saad, A.M.; Korma, S.A.; Salem, H.M.; Abd El-Mageed, T.A.; Alkafaas, S.S.; Elsalahaty, M.I.; Elkafas, S.S.; Mosa, W.F.A.; Ahmed, A.E.; et al. Garlic Bioactive Substances and Their Therapeutic Applications for Improving Human Health: A Comprehensive Review. Front. Immunol. 2024, 15, 1277074. [Google Scholar] [CrossRef]
  261. Rîmbu, M.C.; Cord, D.; Savin, M.; Grigoroiu, A.; Mihăilă, M.A.; Gălățanu, M.L.; Ordeanu, V.; Panțuroiu, M.; Țucureanu, V.; Mihalache, I.; et al. Harnessing Plant-Based Nanoparticles for Targeted Therapy: A Green Approach to Cancer and Bacterial Infections. Int. J. Mol. Sci. 2025, 26, 7022. [Google Scholar] [CrossRef]
  262. Xia, Y.; Shi, C.; Lu, J.; Zhu, Z.; Li, M.; Pan, Y.; Huang, X.; Zhang, L.; Liu, A. Artemisinin and Its Derivatives from Molecular Mechanisms to Clinical Applications: New Horizons Beyond Antimalarials. Int. J. Mol. Sci. 2025, 26, 8409. [Google Scholar] [CrossRef]
  263. Pavan, B.; Dalpiaz, A.; Marani, L.; Beggiato, S.; Ferraro, L.; Canistro, D.; Paolini, M.; Vivarelli, F.; Valerii, M.C.; Comparone, A.; et al. Geraniol Pharmacokinetics, Bioavailability and Its Multiple Effects on the Liver Antioxidant and Xenobiotic-Metabolizing Enzymes. Front. Pharmacol. 2018, 9, 18. [Google Scholar] [CrossRef] [PubMed]
  264. Di Pietro, N.; Baldassarre, M.P.A.; Cichelli, A.; Pandolfi, A.; Formoso, G.; Pipino, C. Role of Polyphenols and Carotenoids in Endothelial Dysfunction: An Overview from Classic to Innovative Biomarkers. Oxid. Med. Cell. Longev. 2020, 2020, 6381380. [Google Scholar] [CrossRef] [PubMed]
  265. Rîmbu, M.C.; Popescu, L.; Mihăilă, M.; Sandulovici, R.C.; Cord, D.; Mihăilescu, C.-M.; Gălățanu, M.L.; Panțuroiu, M.; Manea, C.-E.; Boldeiu, A.; et al. Synergistic Effects of Green Nanoparticles on Antitumor Drug Efficacy in Hepatocellular Cancer. Biomedicines 2025, 13, 641. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Phytochemical-mediated regulation of redox balance, inflammation, and epigenetic mechanisms: Phytochemicals play multifaceted roles in preserving cellular homeostasis by coordinating redox balance, inflammatory signaling, and epigenetic control. By scavenging reactive oxygen species (ROS) and activating antioxidant defense systems such as the Nrf2/ARE pathway, phytochemicals enhance the expression of enzymes including superoxide dismutase (SOD), catalase (CAT), and glutathione peroxidase (GPx). Simultaneously, they inhibit pro-inflammatory mediators—NF-κB, COX-2, iNOS, IL-1β, and TNF-α—attenuating cytokine overproduction and tissue damage. At the epigenetic level, phytochemicals modulate DNA methyltransferases (DNMTs), histone deacetylases (HDACs), and histone acetyltransferases (HATs), restoring physiological gene expression patterns by reversing hypermethylation of tumor suppressor genes and promoting histone acetylation. Through these convergent molecular actions, phytochemicals link oxidative and inflammatory responses to chromatin remodeling and transcriptional reprogramming, supporting chemopreventive and therapeutic effects in chronic diseases.
Figure 1. Phytochemical-mediated regulation of redox balance, inflammation, and epigenetic mechanisms: Phytochemicals play multifaceted roles in preserving cellular homeostasis by coordinating redox balance, inflammatory signaling, and epigenetic control. By scavenging reactive oxygen species (ROS) and activating antioxidant defense systems such as the Nrf2/ARE pathway, phytochemicals enhance the expression of enzymes including superoxide dismutase (SOD), catalase (CAT), and glutathione peroxidase (GPx). Simultaneously, they inhibit pro-inflammatory mediators—NF-κB, COX-2, iNOS, IL-1β, and TNF-α—attenuating cytokine overproduction and tissue damage. At the epigenetic level, phytochemicals modulate DNA methyltransferases (DNMTs), histone deacetylases (HDACs), and histone acetyltransferases (HATs), restoring physiological gene expression patterns by reversing hypermethylation of tumor suppressor genes and promoting histone acetylation. Through these convergent molecular actions, phytochemicals link oxidative and inflammatory responses to chromatin remodeling and transcriptional reprogramming, supporting chemopreventive and therapeutic effects in chronic diseases.
Molecules 30 04317 g001
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Cord, D.; Rîmbu, M.C.; Iordache, M.P.; Albulescu, R.; Pop, S.; Tanase, C.; Popa, M.-L. Phytochemicals as Epigenetic Modulators in Chronic Diseases: Molecular Mechanisms. Molecules 2025, 30, 4317. https://doi.org/10.3390/molecules30214317

AMA Style

Cord D, Rîmbu MC, Iordache MP, Albulescu R, Pop S, Tanase C, Popa M-L. Phytochemicals as Epigenetic Modulators in Chronic Diseases: Molecular Mechanisms. Molecules. 2025; 30(21):4317. https://doi.org/10.3390/molecules30214317

Chicago/Turabian Style

Cord, Daniel, Mirela Claudia Rîmbu, Marius P. Iordache, Radu Albulescu, Sevinci Pop, Cristiana Tanase, and Maria-Linda Popa. 2025. "Phytochemicals as Epigenetic Modulators in Chronic Diseases: Molecular Mechanisms" Molecules 30, no. 21: 4317. https://doi.org/10.3390/molecules30214317

APA Style

Cord, D., Rîmbu, M. C., Iordache, M. P., Albulescu, R., Pop, S., Tanase, C., & Popa, M.-L. (2025). Phytochemicals as Epigenetic Modulators in Chronic Diseases: Molecular Mechanisms. Molecules, 30(21), 4317. https://doi.org/10.3390/molecules30214317

Article Metrics

Back to TopTop