Next Article in Journal
Modeling Reactive Oxygen Species-Induced Axonal Loss in Leber Hereditary Optic Neuropathy
Next Article in Special Issue
Focusing on the Role of Natural Products in Overcoming Cancer Drug Resistance: An Autophagy-Based Perspective
Previous Article in Journal
Alzheimer’s Disease: Challenges and a Therapeutic Opportunity to Treat It with a Neurotrophic Compound
Previous Article in Special Issue
Nrf2-Related Therapeutic Effects of Curcumin in Different Disorders
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Rosmarinic Acid and Related Dietary Supplements: Potential Applications in the Prevention and Treatment of Cancer

1
College of Integrated Traditional Chinese and Western Medicine, Changchun University of Chinese Medicine, Changchun 130117, China
2
Department of Respirology, First Affiliated Hospital to Changchun University of Chinese Medicine, Changchun 130021, China
3
College of Chinese Medicine, Changchun University of Chinese Medicine, Changchun 130117, China
4
School of pharmaceutical sciences, Changchun University of Chinese Medicine, Changchun 130117, China
5
Northeast Asia Research Institute of Traditional Chinese Medicine, Changchun University of Chinese Medicine, Changchun 130117, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Biomolecules 2022, 12(10), 1410; https://doi.org/10.3390/biom12101410
Submission received: 2 September 2022 / Revised: 24 September 2022 / Accepted: 28 September 2022 / Published: 2 October 2022

Abstract

:
Cancer constitutes a severe threat to human health and quality of life and is one of the most significant causes of morbidity and mortality worldwide. Natural dietary products have drawn substantial attention in cancer treatment and prevention due to their availability and absence of toxicity. Rosmarinic acid (RA) is known for its excellent antioxidant properties and is safe and effective in preventing and inhibiting tumors. This review summarizes recent publications on culture techniques, extraction processes, and anti-tumor applications of RA-enriched dietary supplements. We discuss techniques to improve RA bioavailability and provide a mechanistic discussion of RA regarding tumor prevention, treatment, and adjuvant therapy. RA exhibits anticancer activity by regulating oxidative stress, chronic inflammation, cell cycle, apoptosis, and metastasis. These data suggest that daily use of RA-enriched dietary supplements can contribute to tumor prevention and treatment. RA has the potential for application in anti-tumor drug development.

1. Introduction

Cancer is a significant public health problem worldwide. Diagnostic technologies and treatments, including surgery, targeted therapies, and immunotherapies, have made significant advancements in the past 30 years. The risk of cancer death has dropped by 32%; however, the progression of advanced tumors, post-treatment drug resistance, and recurrence remain the most critical aspects of clinical oncology [1]. Potential challenges, long-term treatment, and repeated hospitalizations severely impact the quality of life, substantial financial burden, and psychological stress [2]. Studies showed that some dietary supplements, ethnic herbs, and teas are used for tumor prevention and treatment [3,4,5]. Some natural products from diets or plants are potential anti-tumor drugs and chemotherapy sensitizers [6].
Rosmarinic acid (RA) is a flavonoid commonly found in plants in the Lamiaceae family. RA-rich plants such as Perilla frutescens (L.) Britton, Rosmarinus officinalis L., and Melissa officinalis L. are popular worldwide and used in tea, herbs, cooking condiments, spices, and fruits. RA is used to improve health because of its nutritional properties and has been noted to have potent antioxidant activity [7,8]. In the past ten years, it has been noted that these plants might prevent and treat tumors. Isolation of the anti-tumor components of the plant revealed that the active components include polyphenols. Studies found that RA can prevent tumorigenesis, inhibit tumor growth, and sensitize chemo-radiotherapy agents as adjuvant therapy [9,10,11]. The preparation of RA depends on the purification after biosynthesis of plants, and the recent research proposes synthesizing RA in vitro by engineering bacteria [12]. The bioavailability of RA is low; therefore, the improvements of the dosage form and the development of chemical delivery systems are necessary for anti-tumor applications [13]. This review summarizes the anti-tumor applications, extraction processes of RA-rich plants, and anti-tumor mechanisms to provide in-depth mechanistic insights. This review aims to provide the latest evidence on the biological properties and anti-tumor applications of RA and RA-enriched plants (Figure 1).

2. Methodology

A literature search was performed in PubMed and Google scholar from January 1998 to May 2022, and the last search date was 30 April 2022. The search term was “rosmarinic acid”. A secondary search was conducted by screening the list of articles that met the inclusion criteria. The keywords were “cancer” OR “tumor” OR “carcinoma” OR “malignancy”. The obtained 306 articles were screened, 31 review articles and 3 articles not published in English were removed. A further 175 relevant studies were excluded by reading the abstract, an additional 21 records identified as eligible articles. In total, 118 articles were sorted and classified. Finally, we organized the tables, wrote the text, and made figures to summarize the application of RA anticancer effects according to the SANRA and previously literature review [14,15].

3. Culture Techniques, Extraction Processes, and Anti-Tumor Applications of RA-Rich Plants

RA-rich plant extracts are functional ingredients and supplements that have become popular products in the health industry. Studies on the extraction processes and anti-tumor applications of RA-rich plants are summarized in Table 1.
Rosmarinus officinalis L. (rosemary) is a popular culinary herb worldwide and in European folk medicine. Aqueous extract of leaves inhibited the proliferation of cervical cancer, breast cancer, and T-cell leukemia cells [16]. Ethyl acetate extracts were enriched in RA, and they displayed antioxidant activity and promoted the apoptosis of colorectal cancer (CRC) cells [17]. The ethanol extract of rosemary dried leaves promoted apoptosis to enhance sensitivity to cisplatin (DDP) in ovarian carcinoma cells [18].
Perilla frutescens (L.) Britton is used as a medicinal plant in China, Japan, and Thailand. The aqueous extract of P. frutescens leaves are rich in RA and promote hepatocellular carcinoma (HCC) apoptosis by regulating apoptosis-related genes detected by cDNA microarrays [19]. Osakabe et al. optimized the extraction process of RA from leaves of P. frutescens with a concentration of 68% w/w. The extractive fraction and RA reduced inflammation and oxidative stress and reduced tumor size in skin cancers induced with 7,12-dimethylbenz[a]anthracene (DMBA) and 12-tetradecanoylphorbol 13-acetate (TPA) [20]. The seeds of P. frutescens are rich in omega-3 fatty acids and RA. After extracting oil from the seeds, RA can be enriched by 70% ethanol and ethyl acetate extraction. The RA-enriched fraction reduced reactive oxygen species (ROS) and inhibited invasion through the NF-κB pathway in A549 cells [21,22].
Melissa officinalis L. is a traditional herbal tea from the Mediterranean. Studies found that M. officinalis L. with ethanol extraction enriched RA [8,23]. These extracts had anti-tumor effects on HCT116 and H460 cells. The polyphenolic extract is a candidate for an antioxidant to protect human keratinocytes from UVB-induced skin damage [24]. Hydroxyphenylpyruvate reductase (HPPR) from M. officinalis L. was isolated and characterized as RA biosynthesis-related gene. Several terpenoid synthesis genes were identified and classified in this study [25].
Ocimum basilicum L. (i.e., basil) is consumed as a seasoning worldwide. Hosam et al. compared six cultivars of basils and found that basil leaf extracted from methanol had anti-tumor effects and was rich in RA [26]. Ethanol extract from basil leaves prevented metastasis in head and neck squamous cell carcinoma (HNSCC) [27]. In addition to the leaves, the callus of basil has high RA content. Saher et al. improved tissue culture technologies to increase RA production. They explored several plant growth regulators and found that 5 mg/L 6-benzylaminopurine (BAP) combined with 1 mg/L naphthalene acetic acid (NAA) yielded the best phenolic yield (346.08 mg/L), including 7.4 mg/g RA [28]. Subsequently, callus of basil grown on a medium supplemented with 10 mg/L CuO-NPs yielded the highest RA accumulation (11.4 mg/g) [29]. Light-emitting diode irradiation increased the RA content of callus 96.0 mg/g, 2.46-fold higher than control [30].
Extracts and processed products from Origanum vulgare L. are condiments in cooking, essential oils, and wine. It is a medicinal plant used to treat asthma, indigestion, headaches, and rheumatism in Turkey. Water-soluble ethyl acetate extract had antioxidant and anti-proliferative activities against C6 (rat glioma), and HeLa cells; RA, hesperetin, and hydroquinone were the active ingredients [31]. Juste et al. evaluated antioxidant and anticancer activities in various strains of O. vulgare and found that RA content was positively correlated with antioxidant activity [32].
Thyme is a perennial Lamiaceae herb native to temperate regions of Europe, North Africa, and Asia. It is used as a culinary seasoning and a medicinal plant in ethnomedicine. Thymus vulgaris L. callus crude extract (RA content 5.67 mg/g) inhibited human breast cancer cells [33]. Antioxidant and cytotoxic properties of Thymus longicaulis C. Presl were analyzed during various life cycle phases. Oct12 extract was rich in RA and showed a marked biological activity and cytotoxicity against several tumor cells [34].
The genus Salvia, also belonging to the Lamiaceae family, possesses anticancer medicinal properties. Salvia officinalis L. and Salvia fruticosa Mill. (Mediterranean medicinal plants) contain RA in aqueous extracts from 50 to 70 μg/mL. These extracts inhibited proliferation in breast cancer and colon cancer cell lines via the mitogen-activated protein kinase (MAPK)/ extracellular signal-regulated kinases 1 and 2 (ERK1/2) and the phosphatidylinositide-3-kinase (PI3K)/AKT pathways [35,36]. Research has shown that foliar spraying with NO and Si and under Cu stress in S. officinalis elevated total RA content by 2-fold above control leaves. The seedlings were irrigated with sodium silicate (1 mM Si), sodium nitroprusside (200 μM as a NO donor), and 200 μM CuSO4 [37]. Salvia miltiorrhiza Bunge is a popular Chinese medicinal herb. Methyl jasmonate (MeJA) enhanced the synthesis of RA in S. miltiorrhiza through regulation of the transcription factor gene SmMYB2 and secondary metabolism-related genes [tyrosine aminotransferase (TAT) and HPPR] [38,39]. Equal amounts of cellulase A and protamex mixture exhibited maximum effectiveness in extracting RA at 28.23 mg/g [40]. The callus culture of stem and leaf explants of S. miltiorrhiza promoted RA biosynthesis. RA and salvianolic acid B were cytotoxic primary phenolic compounds for acute lymphoblastic leukemia (ALL) cells [35]. Sage tea made from Salvia helps prevent colon cancer by inhibiting oxidation and DNA damage [41].
Prunella laciniata (L.) L. is a plant of the Labiaceae family that has been used as food and medicine in China for thousands of years. A tyrosine aminotransferase of Prunella vulgaris (PvTAT) is an RA biosynthesis enzyme applicable to engineering natural products [42]. Studies showed that 60% ethanol extract of P. laciniata showed high antioxidant activity in vitro and in vivo and inhibited tumor load in tumor-bearing C57BL/6 mice [43,44].
In addition, RA is the primary active ingredient in several plants. In Gastrocotyle hispida (Forssk.) Bunge, grown in Saudi Arabia, RA was a potent anti-breast cancer and anti-HCC active component [45]. The ethyl acetate fraction extract of Glechoma hederacea L. promoted mitochondrial membrane potential destruction and apoptosis in HCC cells. Substant polyphenols, including RA, caffeic acid, and ferulic acid, were separated using high-performance liquid chromatography [46]. RA is the principal polyphenol in Ehretia tinifolia L. and showed cytotoxicity and potent antioxidant activity against several cancer cell lines [47].
Plant culture techniques include the addition of nanoparticles, and the co-culture of plants and bacteria were used to increase RA yield. Young seedlings of Leonotis nepetifolia (L.) R.Br. were infected with Rhizobium rhizogenes strain A4. The dominant compounds in the extracts contained 2643 µg/g RA, which was 43% higher than in untransformed roots. The transformed roots extract showed better cytotoxic effects against breast cancer [HCC1937 cells the half-maximal inhibitory concentration (IC50) = 750 µg/mL] and leukemia (NALM-6 cells IC50 = 550 µg/mL), meanwhile, HUVEC normal cells had no change in cell viability at the same concentration [12].
Transformed roots of Dracocephalum kotschyi Boiss. were treated with 50 mg/L titanium dioxide nanoparticles (TiO2 NPs) for 24 h; this treatment raised RA levels to 530.5 μg/g by increasing the expression of PAL and RAS genes [48]. D. kotschyi was co-cultivated with Agrobacterium rhizogenes to mediate hairy root growth. Hairy roots were exposed to 75 mg/L Fe NP for 24 h, yielding RA content of 1194 μg/g [49].
The endangered plant species Satureja khuzistanica Jamzad (from Iran) yielded RA in methanol extracts ranging from 0.59% to 1.81%. Abbas et al. developed cell suspension cultures of S. khuzistanica supplemented with 100 μM MeJA as an elicitor to improve RA production to 3.9 g/L [50]. Subsequently, the authors found that suspension cultures treated with pre-optimized coronatine (1 µM) obtained 2.67 g/L RA production, and the crude extract induced apoptosis of MCF-7 cells [51]. These findings demonstrate the considerable potential of in vitro cell culture of plants to induce the biosynthesis of compounds for RA production.
Based on co-culture fermentation technology, RA is synthesized by fermenting plant pericarp and bacteria. Fresh grape skins were vacuum-cooled, powdered, and fermented by Lactobacillus plantarum KFY02 for 96 h. The fermentation broth was rich in RA, rutin, and resveratrol which have antioxidant and liver cancer inhibitory activities [52].
Synthetic biology has made rapid progress and shown broad application prospects in various fields. Engineering bacteria introduce genes into plants or other animals and uses known biochemical reactions in nature to produce small molecular compounds, primarily natural products. Several studies reported de novo synthesis of RA by engineered bacteria. Enzymes including rosmarinic acid synthase (RAS), 4-hydroxyphenlacetate 3-hydroxylase, D-lactate dehydrogenase, TAT, and tyrosine ammonia lyase catalyzed reactions for RA biosynthesis [53,54,55,56,57]. Yan et al. achieved 320.04 mg /L*h RA productivity from caffeic acid and 3,4-dihydroxyphenyllactic acid using an ATP and CoA cycle regeneration system [58]. Li et al. developed a three-strain co-culture synthetic route to produce RA from glucose; RA bioproduction reached 172 mg/L [59].
Table 1. The production process and anti-tumor effects of plants constitute a great source of RA.
Table 1. The production process and anti-tumor effects of plants constitute a great source of RA.
SourceBiotechnological Application for Production and Extraction ProcessRA Content Anti-tumor EffectRef
Rosmarinus officinalis L.Aqueous extract of leaves45.64 mg/gCervical cancer
Breast cancer
T-cell leukemia
[16]
Leaves removed the lipidic phase using hexane. Then, extracted in ethyl acetateApproximately 50.11% w/w RAColorectal cancer[17]
Dried leaves of Rosmarinus officinalis L. were extracted with 70% (v/v) ethyl alcohol overnight at 22 °C on a shaker. The stock solutions were collected from the supernatant-Ovarian carcinoma[18]
Perilla frutescens (L.) Britt.Fresh Perilla leaves were extracted with 1% w/v citric acid at 90 °C for 30 min, then mixed with n-butanol, dried, and dissolved in water. Elution with 0.1% w/v TFA containing 80% v/v methanol on Diaion HP2MG column68% w/w RA of freeze-dried powderSkin carcinogenesis[20]
The dried leaves were chopped, boiled in 1 L of distilled water for 1 h, and filtered. The supernatant was lyophilized. -HCC[19]
The seed meal was extracted in 70% ethanol and dried, then dissolved in ethyl acetate600.32–647.68 mg/gLung cancer[21,22]
Melissa officinalis L.50% ethanolic extracts of leavesN.A.Colorectal cancer[23]
Ethanolic extracts of dry leaves184.4 ± 0.3 mg/gLung cancer[8]
Ethanolic extractApproximately 18%Photoaging and skin cancer[24]
Ocimum tenuiflorum L.Leaves were soaked in 95% ethanol for two weeks, then filtered and driedApproximately 7.86 mg/gHNSCC[27]
Ocimum basilicum L.99% methanol extracts of dry leaves contained RA 3.01 mg/g3.01 mg/gCervical cancer
Breast cancer
T-cell leukemia
[26]
Callus of basil supplemented with 5 mg/L BAP and 1 mg/L NAA and extracted using 100% ethanol7.4 mg/g-[28]
Callus of basil grown on medium supplemented with 10 mg/L CuO-NPs, then extracted using 99.9% methanol11.4 mg/g-[29]
Callus of basil grown on with LED irradiation (24 h, 660 nm), then extracted using methanol96.0 mg/g-[30]
Origanum vulgare L.The aqueous part of the plant was chromatographed on silica gel and eluted with hexane0.15 mg/g RA/dry plantGlioma
Cervical cancer
[31]
Herb was ground and sieved using a 125-μm sieve. The powder was extracted with hot reflux in 90% (v/v) ethanol at 95 °C for 4 hApproximately 36 mg/gGlioma
Breast cancer
[32]
Thymus vulgaris L.Dried callus was extracted by Soxhlet continuous extraction device5.67 mg/gBreast cancer[33]
Thymus longicaulis C.PreslThe leaves were collected in October using 50% methanol for ultrasonic extraction3.03 mg/mLLeukemia
Glioma
Breast cancer
Colorectal cancer
[34]
Salvia officinalis L. and Salvia fruticosa Mill.Aqueous extracts52.0 and 71.5 μg/mL RA of water
extract
Colorectal cancer[35]
Salvia officinalis L.The seedlings were irrigated with 1 mM sodium silicate, 200 μM sodium nitroprusside, and 200 μM CuSO40.62 mg/g-[37]
Salvia miltiorrhiza BungeGround powder was enzymatically incubated and extracted with Cellulase A, Protamex (1:1), and distilled water at 30 °C for 2 h with stirring.28.23 mg/g-[40]
Prunella laciniata (L.) L.60% ethanol extract of leaves2.31 mg/gLung cancer[43,44]
Gastrocotyle hispida (Forssk.) Bunge80% methanol extracts from leaves-HCC
Breast cancer
[45]
Glechoma hederacea L.The whole plants were extracted in distilled water (3 hr at 100 °C) at a dilution of 1:50 (w/v), then extracted with ethyl acetate174.10 ± 5.80 mg/gHCC[46]
Ehretia tinifolia L.The juice in the fruit was applied onto an Amberlite XAD-7 column and eluted with methanol-Cervical cancer
Breast cancer
Colorectal cancer
[47]
Dracocephalum kotschyi Boiss.Transformed roots were influenced by 50 mg/L tTiO2 NPs for 24 h exposure time and incubated for one week. The transformed roots were harvested and extracted under 80% methanol ultrasound530.5 μg/g-[48]
In vitro grown leaves were co-cultivated with Agrobacterium rhizogenes strain to mediate hairy root. Hairy roots were exposed to 75 mg/L Fe NP for 24h, then harvested and extracted under 80% methanol ultrasound1194 μg/g-[49]
Leonotis nepetifolia (L.) R.Br.Young seedlings were infected with Rhizobium rhizogenes strain A4, then harvested and extracted under 80% methanol ultrasound2643 µg/gLung cancer
Breast cancer
T-cell leukemia
[12]
Satureja khuzistanica JamzadCell suspension cultures of plants supplemented with 100 μM MeJA for 21 days
Methanol extraction
3.9 g/L RA in cell suspension cultures-[50]
Cell suspension cultures of plants elicited with 1 µM coronatine2.67 g/L RA in cell suspension culturesBreast cancer[51]
Lactobacillus plantarumFresh grape skins were vacuum-cooled and powdered, fermented by Lactobacillus plantarum KFY02 for 96 h-HCC[52]

4. Improvement of Bioaccessibility and Bioavailability—Novel Technologies

The pharmacokinetic profile of RA was summarized by Nunes et al.; the benefits of RA as a supplement are limited due to formulation challenges, bioaccessibility, and bioavailability [7]. Therefore, it is essential to improve the bioavailability of RA, including the improvement of pharmaceutical technology and developments of drug delivery systems. For toxicology, a dose of 169.6 ± 32.4 mg/kg in Kunming mice (6 weeks old) was shown to be lethal, indicating that RA was slightly toxic [60]. Meanwhile, clinical studies should be considered for further investigation. There are several clinical studies using RA-enriched dietary supplements. Among them, there were no reports of adverse reactions [61,62]; however, these cannot explain the anti-tumor effects and potential toxicity of RA for humans.
A study evaluating the bioavailability and nutrient kinetics of Rosmarinus officinalis L. phenolic compounds in healthy humans found that phase II derivatives of RA were RA-glucuronide, methyl-RA-glucuronide, dimethyl-RA-glucuronide, and dimethyl-RA, suggesting absorption in the small intestine [13]. The absolute oral bioavailability of RA butyl ester was 10.52%, compared to only 1.57% in its original form [63]. The absolute bioavailability of RA was improved to 89.63 % after pulmonary administration [64].
Veras et al. tested excipients for RA. Microcrystalline cellulose and polyvinylpyrrolidone have compatibility against physical interactions, chemical incompatibilities, high temperatures, and water [65]. Encapsulation techniques involved oligosaccharides (e.g., cyclodextrins), increasing solubility in aqueous environments. Complexation of RA with cyclodextrin improved antioxidant activity [65]. Several delivery systems of nanoparticles, solid lipid nanoparticles (SLN), and phospholipid complexes have been applied to improve the bioavailability and absorption of RA in the gastrointestinal environment. RA-loaded silk fibroin nanoparticles had better bioavailability and induced apoptosis of breast and cervical cancer cells in vitro [66]. A study reported RA’s dose safety and toxicity loaded into SLN composed of Witepsol and Carnauba waxes [67]. RA is encapsulated in a hydrophobic bilayer that enhances bioavailability when exposed to the gastrointestinal tract. Xue et al. developed iron-crosslinked RA–lipid conjugates with high contents of RA and doxorubicin (DOX), which had better stability, bioavailability, and synergistic anti-breast cancer efficacy [68]. RA–phospholipid complexes increased oral bioavailability through enhanced intestinal permeability; an in vitro assessment determined that it had better permeation and antioxidant activity [69].

5. Biological Processes and Mechanism of Action of RA in Tumor Prevention and Treatment

5.1. Antioxidation and Anti-Inflammatory Effect

Oxidative stress is caused by the excessive accumulation of free radicals and involves the development of aging, cancer, heart failure, brain damage, and immune disorders. Therefore, the daily consumption of vitamin-rich foods as non-enzymatic antioxidant supplements, or superoxide dismutase (SOD), catalase (CAT), and other health products as the supplements of antioxidant enzymes can effectively remove free radicals. Studies have shown that phenolic antioxidant RA had the function of scavenging free radicals, including ROS and H2O2, and enhanced antioxidant enzymes and non-enzymic antioxidants [70,71]. The antioxidant effect of RA is mainly related to preventing tumorigenesis and chemosensitization.
Long-term exposure to ionizing radiation and chemical carcinogens induces tumorigenesis. Ultraviolet (UV) exposure and administration of chemical carcinogens including DMBA, TPA, 1,2-dimethylhydrazine (DMH), and azoxymethane (AOM) were used as models of tumorigenesis. Increased metabolic activity in cancerous tissues generates high concentrations of ROS leading to pro-tumorigenic events [72]. RA exhibited a potent scavenging effect on ABTS and DPPH radicals and prevented skin and oral carcinogenesis [70,73]. RA enhanced SOD, CAT, and glutathione peroxidase (GPx) activities and reduced lipid peroxidation and cytochrome P450, significantly reducing DMH-induced intestinal polyps in vivo [74,75,76,77]. The accumulation of ROS is often accompanied by inflammation, and skin cancer and CRC are usually associated with long-term chronic inflammation and oxidative stress. RA enhanced nuclear factor erythroid 2-related factor 2 (Nrf2)/heme oxygenase-1 (HO-1) antioxidant system to downregulate NOD-like receptor family pyrin domain containing 3 (NLRP3) and interleukin-1β (IL-1β) in a skin carcinogenesis model caused by UVB radiation [78]. In AOM and dextran sulfate sodium-induced colorectal carcinogenesis animal models, interleukin-6 (IL-6) levels and progression of colitis-associated colon cancer were decreased by RA. The mechanism involved the downregulation of Toll-like receptor 4 (TLR4)/nuclear factor-κB (NF-κB) and signal transducer and activator of transcription 3 (STAT3) [9,79]. RA prevents colorectal carcinogenesis due to antioxidation and anti-inflammatory effects. RA is metabolized and absorbed by the intestinal epithelium, suggesting the utility of daily dietary RA supplementation [13].
The application of the antioxidant effect of RA in tumor therapy is adjuvant therapy and improving tumor side effects. The combination of blue light and RA for HNSCC decreased H2O2 production and inhibited epithelial growth factor receptor (EGFR) activation in vitro [80]. Free radical scavenging increased RA synergism with cytarabine (Ara-C) against leukemia cells [81]. In addition, the antioxidant activity of RA can improve the toxicity of anti-tumor therapy. The antioxidation properties of RA protected ovaries without attenuating the anti-tumor effect of cisplatin [71]. RA improved the hepatorenal toxicity induced by methotrexate and cardiotoxicity induced by DOX based on antioxidant activity [82,83,84].
The anti-inflammatory targets of RA in tumor therapy are cyclooxygenase-2 (COX-2) and NF-κB. RA inhibited COX-2 activity and downregulated ERK1/2 to exert anti-inflammatory effects in lung, breast, and liver cancer cells [85,86]. The molecular simulation predicted that Arg120 in COX-2 was the active site of RA [86]. RA induced apoptosis of acute leukemia, liver cancer, and breast cancer by inhibiting NF-κB-mediated inflammation [87,88,89]. Wu et al. found that RA targeted I-kappaB kinase-β (IKK-β) to inhibit the NF-κB signaling pathway using molecular docking [90]. Inflammatory factors, including tumor necrosis factor-α (TNF-α), IL-1β, IL-6, and transforming growth factor-β (TGF-β), were reduced after anti-tumor therapy with RA [86,88].

5.2. Response to DNA Damage

ROS accumulation and chronic inflammation lead to DNA damage and carcinogenesis [91]. RA inhibited DNA damage due to potent antioxidant capacity, which plays an essential role in preventing tumorigenesis. DMH-induced rat colorectal polyp model and UV-irradiated mouse skin model have verified that it attenuated DNA damage and inhibited tumorigenesis [24,92]. In the anti-tumor process, RA acts as a chemosensitizer in a ROS-independent manner to inhibit DNA damage repair, thereby negatively responding to DNA damage [90]. RA is used as adjunctive therapy to destroy DNA structure (with alkylating agents) and inhibit RNA and DNA synthesis (DOX drugs). RA was combined with alkylating agents in the treatment of CRC resulting in the suppression of DNA repair proteins [41]. The efficacy of DOX in HCC was amplified by combining with RA, which induced mitochondrial dysfunction and DNA damage [93]. Zhang et al. showed that LncRNA MALAT-1 was regulated by RA and promoted DNA damage in ovarian cancer cells; however, the mechanism remains unclear [94].

5.3. Regulation of Cell Cycle and Tumor Proliferation

Sustained unplanned proliferation is one of the hallmarks of cancer, characterized by the potentially infinite proliferation of cancer cells due to the uninterrupted cell cycle and cell division. Cell cycle-related inhibitors (cyclin-dependent kinases 4/6 inhibitors) arrest tumors in the G1 phase, thereby preventing proliferation; this mechanism has been applied in the treatment of several tumors [95]. RA induced cell cycle arrest in treating multiple tumor cells, mainly through upregulation of p53 and p21 and downregulation of cyclins D1, E, and B1 [96]. RA induced G0/G1-phase arrest in breast and pancreatic cancer [97,98]. G2/M arrest occurred in treating kidney cancer and oral cancer [99,100]. Cell cycle arrest represents an opportunity for cancer cells to enter apoptosis. RA increased the expression of apoptosis-related proteins, including BCL-2 associated X (BAX), caspase-3, and caspase-8, and attenuated the expression of anti-apoptotic proteins B cell lymphoma-2 (BCL-2) and poly (ADP-ribose) polymerase (PARP) [51,98]. The upstream mechanism of RA inducing cell cycle arrest included histone deacetylases 2 (HDAC2) and glioma-associated oncogene homolog 1 (Gli1). RA mitigated the restriction of HDAC2 on p53, thereby triggering cell cycle arrest [96]. RA enhanced proteasome-mediated degradation of Gli1 and inhibited the expression of downstream cyclin D1 and snail1 [98].
In addition to inducing cell cycle arrest to inhibit proliferation, RA can also directly regulate cell proliferation-related targets. The anti-proliferative ability of RA was improved at lower concentrations in combination therapy [11,101]. RA behaved the excellent anti-proliferative activity against HeLa, HT29, A549 and MCF6 cancer cell lines with the IC50 values of 249.80, 277.85, 241.47, and 220.25 µM [102]. EGFR is a primary target of the anti-proliferation effects of RA [80,101]. Virtual drug screening analysis revealed that RA selectively inhibited EGFR and spleen tyrosine kinase (SYK). Kai-Cheng et al. synthesized three RA derivatives against drug-resistant EGFR [103]. Microtubule affinity regulating kinase 4 (MARK4) controlled the early step of cell division. Mini-chromosome maintenance complex component 7 (MCM7) initiated eukaryotic DNA replication. RA bound to the active pockets of cell proliferation-related proteins MARK4 and MCM7 with better potency and inhibited protein functions in silico [104,105].

5.4. Apoptosis-Inducing Effect

RA increased the ratio of BAX/BCL-2, activated caspase family proteins, and inhibited PARP, leading to apoptosis in several tumor cell lines [51,96,106]. RA promoted caspase family proteins activity observed in different types of tumor cells in vitro, including CRC, lung cancer, oral cancer, glioma, osteosarcoma, and ALL [10,60,90,100,106,107,108]. PI3K/AKT is the primary pathway of RA-mediated apoptosis. In treating HCC and glioma, RA acted as a Fyn inhibitor, promoting the expression of apoptosis-related proteins through the PI3K/AKT and NF-κB pathways [10,109]. RA downregulated the PI3K/AKT/the mechanistic target of rapamycin (mTOR) signaling pathway to induce apoptosis and inhibited epithelial-mesenchymal transition (EMT) and tumor growth in HCC and osteosarcoma [108,110]. A study applied RNA arrays to identify apoptosis genes regulated by RA in breast cancer cells. The TNF and TNF receptor superfamily were upregulated and were involved in several programmed cell death signaling pathways [97].
RA promoted apoptosis in combination therapy. Aslıhan et al. showed that RA enhanced caspase-3 activity and synergized with siRNA to inhibit heat shock protein 27 (HSP27), which directly induced apoptosis in human glioma cells [111]. Mucin 1 (MUC1) attenuated mitochondrial apoptotic factors and conferred resistance to cytarabine, gemcitabine, and cisplatin [112]. RA combined with a MUC1 inhibitor enhanced the inhibition of protein glycosylation-related enzymes. Combination therapy induced apoptosis-related proteins, including p53, BAX, BCL-2 associated agonist of cell death (BAD), and caspases-3, -8, and -9 [113]. Docking studies showed that RA possesses a good binding affinity to the p53 protein [114].

5.5. Suppression of Multidrug Resistance (MDR) Proteins

The ABCB1 gene and its functions encode MDR1/P-glycoprotein (P-gp) as an energy-dependent drug pump [115]. P-gp-mediated tumor resistance is combined with drugs, and the intracellular drugs are pumped extracellular through an ATP-dependent pathway, reducing intracellular drug concentration. RA has a remarkable sensitization to radiotherapy and chemotherapy [116] and has been observed as an inhibitory effect on P-gp in tumor therapy combined with DOX, DDP, and gemcitabine, leading to a sensitization effect on chemotherapy in gastric, breast, non-small-cell lung, and pancreatic cancers [11,117,118,119]. Studies indicated that MUC1 induces acquired chemoresistance by upregulating P-gp [120,121]. RA reduced MUC1 to sensitize chemotherapy in gastric cancer [113,122]; however, whether RA mediates P- gp through MUC1 requires further study.

5.6. Suppression of Glycolytic Pathway

Differentially expressed proteins after RA intervention in HCC were detected by proteomics analysis. Glycolysis and gluconeogenesis were significantly downregulated after RA intervention according to KEGG pathway enrichment. Inhibition of glycolysis reduced ATP production and inhibited the proliferation of HepG2 cells [123]. The Warburg effect and hypoxia-inducible factor 1 (HIF-1) strengthen energy metabolism, free radical accumulation, and chronic inflammation, promoting tumor angiogenesis and survival [124]. RA can mitigate IL-6/STAT3 and HIF-1α against the Warburg effect in gastric carcinoma and CRC [125,126].

5.7. EMT Inhibition

Invasive tumor cells exhibit characteristics associated with EMT, including mesenchymal cell morphology, loss of cell adhesion, upregulation of cell mobility, and expression of mesenchymal cell feature proteins [127]. RA regulated EMT-related proteins and inhibited tumor cell invasion [107]. RA promoted EMT through the upregulation of E-cadherin, inhibition of N-cadherin, and the concomitant inhibition of matrix metalloproteinases (MMPs), resulting in impaired invasive ability in osteosarcoma, pancreatic cancer, and CRC [108,128,129]. Studies suggested that RA suppressed the expression of Zinc finger E-box binding homeobox 1, snail1, and twist1, inhibited EMT, and increased chemosensitivity [117].

5.8. Anti-Angiogenesis and Metastasis

Both in vivo and in vitro studies have shown that RA has the ability to inhibit invasion and metastasis. First, RA can inhibit invasion ability through MMPs. The central role of MMPs in cancer metastasis is the degradation and remodeling of the extracellular matrix (ECM), which facilitates invasion and metastasis through peripheral cancer tissues. ECM-degrading proteolytic enzymes such as MMP-1, -2, -13, and -14 are involved [130]. Furthermore, the role of MMPs in promoting angiogenesis also promotes tumor progression and metastasis. MMP-9 regulates vascular endothelial growth factor (VEGF) translocation into cells to enable an angiogenic switch [131]. RA inhibited the expression of MMP-2 and MMP-9 and cell invasive ability in several tumor cell lines and attenuated lung metastasis of CRC in a mouse model [60,100,107]. The downregulation of MMP-2 and MMP-9 by RA as a Fyn inhibitor in treating HCC and glioma suppressed tumor invasion and migration [10,109]. Downregulation of AKT phosphorylation with repression by MMPs contributed to the suppression of tumor invasion ability [101,108]. Studies found that RA upregulated miRNAs; miR-506 and miR-1225-5p targeted the 3’ untranslated regions of MMPs to inhibit EMT and tumor metastasis [128,129].
Second, RA inhibits tumor metastasis through VEGF and IL-8 pathways. In vivo studies reported that RA inhibited lung metastasis and bone metastasis of breast cancer [60,132]. Downregulation of VEGF is also the most frequently reported target of RA treatment for metastasis [88,89,133]. Activation of EGFR and VEGF receptors promoted the expression of MMPs and VEGF through intracellular signaling cascades and inhibited the formation of metastatic lung nodule formation [60,132,134]. Huang et al. showed that ROS generation promoted VEGF expression and IL-8 release [134]. In addition, RA inhibited breast cancer metastasis by suppressing IL-8 through the NF-κB ligand/TNF receptor superfamily member 11a/osteoprotegerin pathway [132]. Nevertheless, the role of RA against metastasis requires further study.

6. Prevention of RA in Tumorigenesis

Antioxidation and free radical scavenging are the crucial functions of RA to prevent tumorigenesis. RA upregulated activity of SOD, CAT, glutathione (GSH), and GPx and downregulated thiobarbituric acid reactive substances and malondialdehyde (MDA) [73,74]. Sufficient evidence on the prevention of CRC is related to the potent antioxidant effect of RA on tissues during intestinal epithelial absorption to prevent polyps and tumorigenesis. Common inducers in animal models of colorectal cancer are DMH, AOM and dextran sodium sulfate (DSS). AOM and DMH can become carcinogenic through DNA alkylation, promoting the mispairing of bases. DSS is a synthetic sulfated polysaccharide, and its use alone was shown to cause colonic inflammation in mice, while a combination of AOM and DSS stably induced inflammatory colorectal cancer [135]. In colorectal carcinogenesis rat or mouse models, RA reduced the formation of aberrant crypt foci (ACF) and eliminated the progression of colitis-associated colon cancer [9,74,76,92]. The potential prevention of CRC by RA was mainly attributed to three aspects. First, the excellent antioxidant effect of RA could enhance antioxidant enzyme activity, including SOD, CAT, GSH, and GPx. In addition, RA attenuated DMH-induced upregulation of cytochrome P450 (CYP450) [74,77]. Second, RA inhibited the release of TNF-α, IL-6, and COX-2 pro-inflammatory factors [76,79]. The anti-inflammatory effect of RA has been related to the inhibition of TLR4/NF-κB and STAT3 [9]. Third, RA can reduce DNA damage against ACF formation [92]. A model of spontaneous CRC, C57BL/6J-ApcMin/+ mouse, was shown to be related to familial adenomatous polyposis-derived tumorigenesis. RA decreased the numbers of large adenomas (>3 mm) in C57BL/6J-ApcMin/+ mouse [136]. Skin carcinogenesis is primarily caused by exposure to UV in sunlight, which induces oxidative stress, and the formation of photoproducts and lesions in DNA. Impaired DNA repair may lead to mutagenesis and carcinogenesis [137,138]. RA ameliorated ROS generation, MDA content, and DNA damage in DMBA/TPA-induced skin papilloma mouse model and UV-irradiated keratinocytes [20,24,70,139]. RA downregulated NLRP3 and IL-1β production via the Nrf2/HO-1 antioxidant system [78]. Overall, the occurrence of skin cancer is related to UV exposure and oxidative stress, and the above effects of RA can effectively resist skin carcinogenesis. RA prevents oral cancer due to its antioxidant effects [73,140], as illustrated in Figure 2 and Table 2.

7. The Therapeutic Effect of RA on Cancer

RA inhibited several solid and hematologic tumors by inducing cell cycle arrest and apoptosis, and inhibiting EMT and tumor metastasis. Studies on the anti-tumor effects of RA through in vitro and in vivo models are summarized in Figure 3 and Table 3. In the case of glioma, RA was reported to promote apoptosis-related protein and exerted cytotoxicity in several glioma cell lines with an IC50 value ranging between 200 and 400 µM for 48 h [10,111]. RA induced cell apoptosis and inhibited the migration of oral cancer cells in vitro [100]. RA regulated apoptosis-related genes and changed the methylation pattern via DNA methyltransferases 1 (DNMT1) for breast cancer chemoprevention [97,141]. In addition, RA inhibited breast-derived bone metastases by suppressing IL-8 [132]. RA suppressed the viability of two gastric cancer cell lines at a lower IC50 concentration of 240 µM; meanwhile, RA suppressed tumor growth in gastric tumor-bearing mice by inhibiting of the Warburg effect [122,125]. RA was shown to possess a wide range of applications in the treatment of HCC, including induction of apoptosis, and inhibition of tumor growth and metastasis. Meanwhile, RA demonstrated little effects on the proliferation and morphology of normal human astrocytes cells [109]. Treatment of RA mediated the upregulation of caspase-3, -8, and -9 and inhibited BCL-2 expression to induce apoptosis in different HCC cell lines [19,106]. The downregulation of PI3K/AKT and glycolytic pathway by RA inhibited the cell proliferation and tumor growth of HCC [109,110,123]. Furthermore, RA analogue-11 is a synthesized RA analogue, which promoted apoptosis via the EGFR/AKT/NF-κB pathway in gastric cancer cells [142]. RA also inhibited VEGF expression and EMT to attenuate tumor invasion of HCC in vitro, but more in-depth evidence in in vivo studies is required [88,106,110]. Pancreatic cancer is a highly lethal disease and the fourth leading cause of cancer-related deaths worldwide [143]. RA was shown to induce apoptosis and inhibit pancreatic cancer invasion and proliferation in vitro and suppressed tumor growth in vivo [98,129]. CRC is the third most common cause of cancer-related deaths worldwide, with tumor metastasis occurring in approximately 45% of patients [144]. In CRC, RA demonstrated the potential to withstand CRC metastasis. A reduction in lung metastasis was observed in mice model after RA treatment [107]. Meanwhile, RA can downregulate EMT and MMPs to inhibit the invasion and migration ability of several CRC cell lines [60,107,128]. Inflammation affects cytokine receptor-mediated signaling pathways that mediate CRC tumor progression, including the TNF, IL-1, IL-6, and NF-κB pathways. Moreover, therapy-induced death of CRC cells can induce the production of TNF, IL-17, and IL-6 to save the remaining cells [145]. Therefore, avoiding the inflammatory response could help treat CRC. In this regard, RA was shown to suppress CRC inflammation by impairing the IL-6/STAT3 and NF-κB pathways [60,126]. In solid tumors, RA has also shown anti-tumor effects on ovarian cancer, cervical cancer, prostate cancer, and osteosarcoma, and the specific mechanisms are shown in Table 3. Through transcriptome sequencing analysis, RA methyl ester accelerated apoptosis in DDP resistant ovarian cancer cell line through inhibitory of Forkhead box M1 (FOXM1) [146]. And RA methyl ester also enhanced DDP sensitivity against cervical cancer by inhibiting mTOR/ribosomal protein S6 kinase β-1 (S6K1) pathway [147]. For hematological tumors, RA reported to induce cytotoxicity against multiple myeloma (MM) by inhibiting mitochondrial activity [148]. RA promoted apoptosis in leukemia cells by inhibiting NF-κB and ROS production [87,90], and the IC50 values of RA-treated normal lymphocytes were 1.7- to 5-fold higher than that of ALL cells [90].

8. Chemosensitivity Effect of RA on Tumor Therapy

Tumor resistance to chemotherapy is a significant cause of treatment failure, and has led to research on chemotherapeutic drug sensitizers. DDP is a platinum coordination complex, which can inhibit the DNA replication of cancer cells and damage the cell membrane structure. It is cytotoxic and more sensitive to fast proliferation cells, such as cancer cells. Therefore, it is widely used in the anti-tumor treatment of solid tumors [149]. Platinum-based drugs are used in the first-line treatment of lung, liver, and ovarian cancer. Four studies showed that RA increased the sensitivity of malignant tumor cells to DDP. RA downregulated MDR1 to increase the sensitivity of DDP in treating lung cancer [11]. The combination of RA and DDP induced G2/M phase arrest and apoptosis in renal cancer cells [99]. RA inhibited melanin synthesis and increased DDP sensitivity by inhibiting the ADAM17/EGFR/AKT/glycogen synthase kinase-3β (GSK3β) axis in melanoma [101]. In addition, RA showed synergistic anti-proliferation effect with DDP on ovarian cancer cells [18].
In breast cancer treatment, DOX and paclitaxel are used as sequential chemotherapy regimen [150]. RA mediated the sensitivity of DOX and paclitaxel by regulating p53 pathway and inducting apoptosis [89,118]. The first-line chemotherapy drug for gastric cancer treatment is 5-fluorouracil (5-FU). RA was shown to enhance chemosensitivity to 5-FU by increasing Forkhead box O4 (FOXO4) [151]. DOX is a chemotherapeutic drug used for gastric cancer treatment and RA was shown to reverse the resistance of SGC7901/Adr cells to DOX by inhibiting MDR1 [119]. In addition, RA cooperated with the anti-MUC1 antibody to promote apoptosis in human gastric carcinoma cells [113]. Although the recommended treatment of HCC is surgery, radiotherapy, and interventional therapy, chemotherapy and molecular-targeted therapy are still the main treatment options for advanced HCC. Combination therapy using RA and DOX can enhance DNA damage and BAX/BCL-2 ratio in HCC [93]. RA synergistically increased cytotoxicity and proteasome inhibition induced by MG132 in HCC [152]. RA also enhanced the efficacy of gemcitabine through the downregulation of multidrug resistance-associated protein 4 (MRP-4) and MRP-5 in Panc-1 pancreatic cancer cells [117]. Treatment with all-trans retinoic acid (ATRA) induced the differentiation of leukemia cells and increased the complete response rate of acute promyelocytic leukemia (APL) [153]. More than 80–90% of APL are expected to be cured with a therapeutic regimen based on ATRA and arsenic trioxide [154]. RA potentiated ATRA-induced macrophage differentiation in APL cells [155]. Then, RA synergistically inhibited DNA synthesis to potentiated the anti-proliferative effect of Ara-C [81]. In addition, RA can increase the sensitivity of physical therapy. Combining blue light and RA was shown to effectively decrease the cell proliferation of HNSCC [80]. RA specifically sensitized radiation to induce apoptosis in metastatic melanoma [156]. Using RA for synthesizing AuNPs plays an active role on the treatment of breast cancer [157].
Some RA-rich herbs have been used in adjuvant chemotherapy in ethnomedicine to sensitize cancer cells to conventional drugs and enhance their effects at minimal doses. Subsequent studies confirmed that RA indeed increased the sensitivity of commonly used chemotherapeutic drugs, including DDP, paclitaxel, 5-Fu, DOX, and Ara-C. The molecular targets involved in chemosensitization are displayed in Table 4.

9. Conclusions

Substantial evidence has been proven the potential benefits of RA and RA-enriched plants as drug candidates for the prevention and treatment of cancer. Among RA-rich plants, rosemary, basil, and Perilla frutescens (L.) Britt are potential anti-tumor plants as dietary supplements. In this review, antioxidative and anti-inflammatory effects of RA prevent tumorigenesis, and oral RA is a potential application to prevent CRC. RA exerts anti-tumor effects by inhibiting tumor cell proliferation and EMT, inducing cell cycle arrest and apoptosis, in which PI3K/AKT, NF-κB, IL-6/STAT3, p53, VEGF, and glycolysis pathways are involved. Inhibition of MDR protein by RA increases chemosensitivity in tumor therapy. In tumor therapy, RA is widely used in the treatment of digestive system tumors, including HCC and CRC. In addition, RA can increase the sensitivity of DDP and DOX drugs in the treatment of solid tumors. To improve the oral bioavailability of RA, modification of excipients, encapsulation using cyclodextrins, drug delivery systems, and derivatives of RA are promising candidates. This review provides a theoretical basis for the use of RA in the prevention and treatment of cancer. However, RA is worthy of further investigation based on high-throughput methods and clinical studies. It is expected to become one of the promising methods for preventing and treating cancer in the future.

Author Contributions

Conceptualization, Z.W. and J.W.; data curation, J.Z., L.X., D.J., Y.X. and L.T.; writing—original draft preparation, J.Z. and L.X.; writing—review and editing, T.W., D.Z., Z.W. and J.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Foundation of Jilin Province, grant number YDZJ202101ZYTS035 and the Education Department of Jilin Province, grant number JJKH20210958KJ.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

RA, Rosmarinic acid; CRC, colorectal cancer; DDP, cisplatin; HCC, hepatocellular carcinoma; DMBA, 7,12-dimethylbenz[a]anthracene; TPA, 12-tetradecanoylphorbol 13-acetate; ROS, reactive oxygen species; HPPR, Hydroxyphenylpyruvate reductase; HNSCC, head and neck squamous cell carcinoma; BAP, 6-benzylaminopurine; NAA, naphthalene acetic acid; MAPK, the mitogen-activated protein kinase; ERK1/2, extracellular signal-regulated kinases 1 and 2; PI3K, the phosphatidylinositide-3-kinase; MeJA, Methyl jasmonate; TAT, tyrosine aminotransferase; ALL, acute lymphoblastic leukemia; TiO2 NPs, titanium dioxide nanoparticles; IC50, half-maximal inhibitory concentration; RAS, rosmarinic acid synthase; SLN, solid lipid nanoparticles; DOX, doxorubicin; SOD, superoxide dismutase; CAT, catalase; DMH, 1,2-dimethylhydrazine; AOM, azoxymethane; GPx, glutathione peroxidase; EGFR, epithelial growth factor receptor; Ara-C, cytarabine; NLRP3, NOD-like receptor family pyrin domain containing 3; IL-1β, interleukin-1β; IL-6, interleukin-6; TLR4, Toll-like receptor 4; STAT3, signal transducer and activator of transcription 3; COX-2, cyclooxygenase-2; IKK-β, I-kappaB kinase-β; TNF-α, tumor necrosis factor-α; TGF-β, transforming growth factor-β; BAX, BCL-2 associated X; BCL-2, B cell lymphoma-2; PARP, poly (ADP-ribose) polymerase; HDAC2, histone deacetylases 2; Gli1, glioma-associated oncogene homolog 1; SYK, spleen tyrosine kinase; MARK4, microtubule affinity regulating kinase 4; MCM7, mini-chromosome maintenance complex component 7; BAX, BCL-2 associated X; mTOR, the mechanistic target of rapamycin; EMT, epithelial-mesenchymal transition; HSP27, heat shock protein 27; MUC1, Mucin 1; BAD, BCL-2 associated agonist of cell death; MDR, multidrug resistance; P-gp, P-glycoprotein; HIF-1, hypoxia-inducible factor 1; MMPs, matrix metalloproteinases; ECM, extracellular matrix; VEGF, vascular endothelial growth factor; GSH, glutathione; MDA, malondialdehyde; DSS, dextran sodium sulfate; ACF, aberrant crypt foci; CYP450, cytochrome P450; DNMT1, DNA methyltransferases 1; FOXM1, Forkhead box M1; S6K1, ribosomal protein S6 kinase β-1; MM, multiple myeloma; 5-FU, 5-fluorouracil; FOXO4, Forkhead box O4; MRP-4, multidrug resistance-associated protein 4; ATRA, all-trans retinoic acid.

References

  1. Siegel, R.L.; Miller, K.D.; Fuchs, H.E.; Jemal, A. Cancer statistics, 2022. CA A Cancer J. Clin. 2022, 72, 7–33. [Google Scholar] [CrossRef] [PubMed]
  2. Emery, J.; Butow, P.; Lai-Kwon, J.; Nekhlyudov, L.; Rynderman, M.; Jefford, M. Management of common clinical problems experienced by survivors of cancer. Lancet 2022, 399, 1537–1550. [Google Scholar] [CrossRef]
  3. Sardana, R.K.; Chhikara, N.; Tanwar, B.; Panghal, A. Dietary impact on esophageal cancer in humans: A review. Food Funct. 2018, 9, 1967–1977. [Google Scholar] [CrossRef] [PubMed]
  4. Ganesan, K.; Du, B.; Chen, J. Effects and mechanisms of dietary bioactive compounds on breast cancer prevention. Pharmacol. Res. 2022, 178, 105974. [Google Scholar] [CrossRef] [PubMed]
  5. Crowder, S.L.; Playdon, M.C.; Gudenkauf, L.M.; Ose, J.; Gigic, B.; Greathouse, L.; Peoples, A.R.; Sleight, A.G.; Jim, H.S.L.; Figueiredo, J.C. A Molecular Approach to Understanding the Role of Diet in Cancer-Related Fatigue: Challenges and Future Opportunities. Nutrients 2022, 14, 1496. [Google Scholar] [CrossRef]
  6. Khatoon, E.; Banik, K.; Harsha, C.; Sailo, B.L.; Thakur, K.K.; Khwairakpam, A.D.; Vikkurthi, R.; Devi, T.B.; Gupta, S.C.; Kunnumakkara, A.B. Phytochemicals in cancer cell chemosensitization: Current knowledge and future perspectives. Semin. Cancer Biol. 2022, 80, 306–339. [Google Scholar] [CrossRef]
  7. Nunes, S.; Madureira, A.R.; Campos, D.; Sarmento, B.; Gomes, A.M.; Pintado, M.; Reis, F. Therapeutic and nutraceutical potential of rosmarinic acid-Cytoprotective properties and pharmacokinetic profile. Crit. Rev. Food Sci. Nutr. 2017, 57, 1799–1806. [Google Scholar] [CrossRef]
  8. Magalhães, D.B.; Castro, I.; Lopes-Rodrigues, V.; Pereira, J.M.; Barros, L.; Ferreira, I.; Xavier, C.P.R.; Vasconcelos, M.H. Melissa officinalis L. ethanolic extract inhibits the growth of a lung cancer cell line by interfering with the cell cycle and inducing apoptosis. Food Funct. 2018, 9, 3134–3142. [Google Scholar] [CrossRef] [Green Version]
  9. Jin, B.R.; Chung, K.S.; Hwang, S.; Hwang, S.N.; Rhee, K.J.; Lee, M.; An, H.J. Rosmarinic acid represses colitis-associated colon cancer: A pivotal involvement of the TLR4-mediated NF-κB-STAT3 axis. Neoplasia 2021, 23, 561–573. [Google Scholar] [CrossRef]
  10. Liu, Y.; Xu, X.; Tang, H.; Pan, Y.; Hu, B.; Huang, G. Rosmarinic acid inhibits cell proliferation, migration, and invasion and induces apoptosis in human glioma cells. Int. J. Mol. Med. 2021, 47, 1–11. [Google Scholar] [CrossRef]
  11. Liao, X.Z.; Gao, Y.; Sun, L.L.; Liu, J.H.; Chen, H.R.; Yu, L.; Chen, Z.Z.; Chen, W.H.; Lin, L.Z. Rosmarinic acid reverses non-small cell lung cancer cisplatin resistance by activating the MAPK signaling pathway. Phytother. Res. 2020, 34, 1142–1153. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Kowalczyk, T.; Merecz-Sadowska, A.; Rijo, P.; Isca, V.; Picot, L.; Wielanek, M.; Śliwiński, T.; Sitarek, P. Preliminary Phytochemical Analysis and Evaluation of the Biological Activity of Leonotis nepetifolia (L.) R. Br Transformed Roots Extracts Obtained through Rhizobium rhizogenes-Mediated Transformation. Cells 2021, 10, 1242. [Google Scholar] [CrossRef] [PubMed]
  13. Achour, M.; Bravo, L.; Sarriá, B.; Ben Fredj, M.; Nouira, M.; Mtiraoui, A.; Saguem, S.; Mateos, R. Bioavailability and nutrikinetics of rosemary tea phenolic compounds in humans. Food Res. Int. 2021, 139, 109815. [Google Scholar] [CrossRef] [PubMed]
  14. Kamal, Z.; Bairage, J.J.; Moniruzzaman, P.R.D.; Tabibul, M.; Islam, M.O.F.; Islam, M.R.; Paul, P.K.; Mou, S.M.; Rahmatullah, M.J.W.J.o.P.; Sciences, P. Folk medicinal uses of some plants in Tangail district, Bangladesh. World J. Pharm. Pharm. Sci. 2014, 3, 52–63. [Google Scholar]
  15. Baethge, C.; Goldbeck-Wood, S.; Mertens, S. SANRA-a scale for the quality assessment of narrative review articles. Res. Integr. Peer Rev. 2019, 4, 5. [Google Scholar] [CrossRef] [Green Version]
  16. Elansary, H.O.; Mahmoud, E.A. Egyptian herbal tea infusions’ antioxidants and their antiproliferative and cytotoxic activities against cancer cells. Nat. Prod. Res. 2015, 29, 474–479. [Google Scholar] [CrossRef]
  17. Amar, Y.; Meddah, B.; Bonacorsi, I.; Costa, G.; Pezzino, G.; Saija, A.; Cristani, M.; Boussahel, S.; Ferlazzo, G.; Meddah, A.T. Phytochemicals, Antioxidant and Antiproliferative Properties of Rosmarinus officinalis L on U937 and CaCo-2 Cells. Iran. J. Pharm. Res. IJPR 2017, 16, 315–327. [Google Scholar]
  18. Tai, J.; Cheung, S.; Wu, M.; Hasman, D. Antiproliferation effect of Rosemary (Rosmarinus officinalis) on human ovarian cancer cells in vitro. Phytomedicine 2012, 19, 436–443. [Google Scholar] [CrossRef]
  19. Lin, C.S.; Kuo, C.L.; Wang, J.P.; Cheng, J.S.; Huang, Z.W.; Chen, C.F. Growth inhibitory and apoptosis inducing effect of Perilla frutescens extract on human hepatoma HepG2 cells. J. Ethnopharmacol. 2007, 112, 557–567. [Google Scholar] [CrossRef]
  20. Osakabe, N.; Yasuda, A.; Natsume, M.; Yoshikawa, T. Rosmarinic acid inhibits epidermal inflammatory responses: Anticarcinogenic effect of Perilla frutescens extract in the murine two-stage skin model. Carcinogenesis 2004, 25, 549–557. [Google Scholar] [CrossRef]
  21. Tantipaiboonwong, P.; Chaiwangyen, W.; Suttajit, M.; Kangwan, N.; Kaowinn, S.; Khanaree, C.; Punfa, W.; Pintha, K. Molecular Mechanism of Antioxidant and Anti-Inflammatory Effects of Omega-3 Fatty Acids in Perilla Seed Oil and Rosmarinic Acid Rich Fraction Extracted from Perilla Seed Meal on TNF-α Induced A549 Lung Adenocarcinoma Cells. Molecules 2021, 26, 6757. [Google Scholar] [CrossRef] [PubMed]
  22. Pintha, K.; Chaiwangyen, W.; Yodkeeree, S.; Suttajit, M.; Tantipaiboonwong, P. Suppressive Effects of Rosmarinic Acid Rich Fraction from Perilla on Oxidative Stress, Inflammation and Metastasis Ability in A549 Cells Exposed to PM via C-Jun, P-65-Nf-Κb and Akt Signaling Pathways. Biomolecules 2021, 11, 1090. [Google Scholar] [CrossRef] [PubMed]
  23. Encalada, M.A.; Hoyos, K.M.; Rehecho, S.; Berasategi, I.; de Ciriano, M.G.; Ansorena, D.; Astiasarán, I.; Navarro-Blasco, I.; Cavero, R.Y.; Calvo, M.I. Anti-proliferative effect of Melissa officinalis on human colon cancer cell line. Plant Foods Hum. Nutr. 2011, 66, 328–334. [Google Scholar] [CrossRef] [PubMed]
  24. Pérez-Sánchez, A.; Barrajón-Catalán, E.; Herranz-López, M.; Castillo, J.; Micol, V. Lemon balm extract (Melissa officinalis, L.) promotes melanogenesis and prevents UVB-induced oxidative stress and DNA damage in a skin cell model. J. Dermatol. Sci. 2016, 84, 169–177. [Google Scholar] [CrossRef]
  25. Mansouri, M.; Mohammadi, F. Transcriptome analysis to identify key genes involved in terpenoid and rosmarinic acid biosynthesis in lemon balm (Melissa officinalis). Gene 2021, 773, 145417. [Google Scholar] [CrossRef]
  26. Elansary, H.O.; Mahmoud, E.A. In vitro antioxidant and antiproliferative activities of six international basil cultivars. Nat. Prod. Res. 2015, 29, 2149–2154. [Google Scholar] [CrossRef]
  27. Utispan, K.; Niyomtham, N.; Yingyongnarongkul, B.E.; Koontongkaew, S. Ethanolic Extract of Ocimum sanctum Leaves Reduced Invasion and Matrix Metalloproteinase Activity of Head and Neck Cancer Cell Lines. Asian Pac. J. Cancer Prev. APJCP 2020, 21, 363–370. [Google Scholar] [CrossRef]
  28. Nazir, S.; Jan, H.; Tungmunnithum, D.; Drouet, S.; Zia, M.; Hano, C.; Abbasi, B.H. Callus Culture of Thai Basil Is an Effective Biological System for the Production of Antioxidants. Molecules 2020, 25, 4859. [Google Scholar] [CrossRef]
  29. Nazir, S.; Jan, H.; Zaman, G.; Khan, T.; Ashraf, H.; Meer, B.; Zia, M.; Drouet, S.; Hano, C.; Abbasi, B.H. Copper oxide (CuO) and manganese oxide (MnO) nanoparticles induced biomass accumulation, antioxidants biosynthesis and abiotic elicitation of bioactive compounds in callus cultures of Ocimum basilicum (Thai basil). Artif. Cells Nanomed. Biotechnol. 2021, 49, 626–634. [Google Scholar] [CrossRef]
  30. Nadeem, M.; Abbasi, B.H.; Younas, M.; Ahmad, W.; Zahir, A.; Hano, C. LED-enhanced biosynthesis of biologically active ingredients in callus cultures of Ocimum basilicum. J. Photochem. Photobiology. B Biol. 2019, 190, 172–178. [Google Scholar] [CrossRef]
  31. Erenler, R.; Sen, O.; Aksit, H.; Demirtas, I.; Yaglioglu, A.S.; Elmastas, M.; Telci, İ. Isolation and identification of chemical constituents from Origanum majorana and investigation of antiproliferative and antioxidant activities. J. Sci. Food Agric. 2016, 96, 822–836. [Google Scholar] [CrossRef] [PubMed]
  32. Baranauskaite, J.; Kubiliene, A.; Marksa, M.; Petrikaite, V.; Vitkevičius, K.; Baranauskas, A.; Bernatoniene, J. The Influence of Different Oregano Species on the Antioxidant Activity Determined Using HPLC Postcolumn DPPH Method and Anticancer Activity of Carvacrol and Rosmarinic Acid. BioMed Res. Int. 2017, 2017, 1681392. [Google Scholar] [CrossRef] [PubMed]
  33. Darw, H.Y.; Abdelmigid, H.; Albogami, S.; Alotaibi, S.; El-De, A.N.; Alnefaie, A. Induction of Biosynthetic Genes Related to Rosmarinic Acid in Plant Callus Culture and Antiproliferative Activity Against Breast Cancer Cell Line. Pak. J. Biol. Sci. 2020, 23, 1025–1036. [Google Scholar] [CrossRef] [PubMed]
  34. Galasso, S.; Pacifico, S.; Kretschmer, N.; Pan, S.P.; Marciano, S.; Piccolella, S.; Monaco, P.; Bauer, R. Influence of seasonal variation on Thymus longicaulis C. Presl chemical composition and its antioxidant and anti-inflammatory properties. Phytochemistry 2014, 107, 80–90. [Google Scholar] [CrossRef] [PubMed]
  35. Xavier, C.P.R.; Lima, C.F.; Fernandes-Ferreira, M.; Pereira-Wilson, C. Salvia Fruticosa, Salvia Officinalis, and Rosmarinic Acid Induce Apoptosis and Inhibit Proliferation of Human Colorectal Cell Lines: The Role in MAPK/ERK Pathway. Nutr. Cancer 2009, 61, 564–571. [Google Scholar] [CrossRef] [Green Version]
  36. Irtegun Kandemir, S.; Fidan, H.S.; Yener, I.; Mete, N.; Ertas, A.; Topcu, G.; Kolak, U. Investigation of cytotoxic and apoptotic effects of 63 compounds obtained from Salvia species: Promising anticancer agents. J. Food Biochem. 2022, 46, e14226. [Google Scholar] [CrossRef]
  37. Pirooz, P.; Amooaghaie, R.; Ahadi, A.; Sharififar, F.; Torkzadeh-Mahani, M. Silicon and nitric oxide synergistically modulate the production of essential oil and rosmarinic acid in Salvia officinalis under Cu stress. Protoplasma 2021, 259, 905–916. [Google Scholar] [CrossRef]
  38. Deng, C.; Wang, Y.; Huang, F.; Lu, S.; Zhao, L.; Ma, X.; Kai, G. SmMYB2 promotes salvianolic acid biosynthesis in the medicinal herb Salvia miltiorrhiza. J. Integr. Plant Biol. 2020, 62, 1688–1702. [Google Scholar] [CrossRef]
  39. Hou, Z.; Li, Y.; Su, F.; Chen, J.; Zhang, X.; Xu, L.; Yang, D.; Liang, Z. Application of (1)H-NMR combined with qRT-PCR technology in the exploration of rosmarinic acid biosynthesis in hair roots of Salvia miltiorrhiza Bunge and Salvia castanea f. tomentosa Stib. Planta 2020, 253, 2. [Google Scholar] [CrossRef]
  40. Su, C.H.; Pham, T.T.T.; Cheng, H.H. Aqueous enzymatic extraction of rosmarinic acid from Salvia officinalis: Optimisation using response surface methodology. Phytochem. Anal. PCA 2020, 31, 575–582. [Google Scholar] [CrossRef]
  41. Ramos, A.A.; Pedro, D.; Collins, A.R.; Pereira-Wilson, C. Protection by Salvia extracts against oxidative and alkylation damage to DNA in human HCT15 and CO115 cells. J. Toxicol. Environ. Health. Part A 2012, 75, 765–775. [Google Scholar] [CrossRef] [PubMed]
  42. Ru, M.; Wang, K.; Bai, Z.; Peng, L.; He, S.; Wang, Y.; Liang, Z. A tyrosine aminotransferase involved in rosmarinic acid biosynthesis in Prunella vulgaris L. Sci. Rep. 2017, 7, 4892. [Google Scholar] [CrossRef] [PubMed]
  43. Feng, L.; Jia, X.; Zhu, M.-M.; Chen, Y.; Shi, F. Antioxidant Activities of Total Phenols of Prunella vulgaris L. in Vitro and in Tumor-bearing Mice. Molecules 2010, 15, 9145–9156. [Google Scholar] [CrossRef] [PubMed]
  44. Feng, L.; Jia, X.-B.; Jiang, J.; Zhu, M.-M.; Chen, Y.; Tan, X.-B.; Shi, F. Combination of Active Components Enhances the Efficacy of Prunella in Prevention and Treatment of Lung Cancer. Molecules 2010, 15, 7893–7906. [Google Scholar] [CrossRef]
  45. Shahat, A.A.; Hidayathulla, S.; Khan, A.A.; Alanazi, A.M.; Al Meanazel, O.T.; Alqahtani, A.S.; Alsaid, M.S.; Hussein, A.A. Phytochemical profiling, antioxidant and anticancer activities of Gastrocotyle hispida growing in Saudi Arabia. Acta Trop. 2019, 191, 243–247. [Google Scholar] [CrossRef]
  46. Chao, W.W.; Liou, Y.J.; Ma, H.T.; Chen, Y.H.; Chou, S.T. Phytochemical composition and bioactive effects of ethyl acetate fraction extract (EAFE) of Glechoma hederacea L. J. Food Biochem. 2021, 45, e13815. [Google Scholar] [CrossRef]
  47. Monroy-García, I.N.; Carranza-Torres, I.E.; Carranza-Rosales, P.; Oyón-Ardoiz, M.; García-Estévez, I.; Ayala-Zavala, J.F.; Morán-Martínez, J.; Viveros-Valdez, E. Phenolic Profiles and Biological Activities of Extracts from Edible Wild Fruits Ehretia tinifolia and Sideroxylon lanuginosum. Foods 2021, 10, 2710. [Google Scholar] [CrossRef]
  48. Nourozi, E.; Hosseini, B.; Maleki, R.; Abdollahi Mandoulakani, B. Inductive effect of titanium dioxide nanoparticles on the anticancer compounds production and expression of rosmarinic acid biosynthesis genes in Dracocephalum kotschyi transformed roots. Plant Physiol. Biochem. PPB 2021, 167, 934–945. [Google Scholar] [CrossRef]
  49. Nourozi, E.; Hosseini, B.; Maleki, R.; Abdollahi Mandoulakani, B. Iron oxide nanoparticles: A novel elicitor to enhance anticancer flavonoid production and gene expression in Dracocephalum kotschyi hairy-root cultures. J. Sci. Food Agric. 2019, 99, 6418–6430. [Google Scholar] [CrossRef]
  50. Khojasteh, A.; Mirjalili, M.H.; Palazon, J.; Eibl, R.; Cusido, R.M. Methyl jasmonate enhanced production of rosmarinic acid in cell cultures of Satureja khuzistanica in a bioreactor. Eng. Life Sci. 2016, 16, 740–749. [Google Scholar] [CrossRef]
  51. Khojasteh, A.; Metón, I.; Camino, S.; Cusido, R.M.; Eibl, R.; Palazon, J. In Vitro Study of the Anticancer Effects of Biotechnological Extracts of the Endangered Plant Species Satureja Khuzistanica. Int. J. Mol. Sci. 2019, 20, 2400. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Liu, J.; Tan, F.; Liu, X.; Yi, R.; Zhao, X. Exploring the Antioxidant Effects and Periodic Regulation of Cancer Cells by Polyphenols Produced by the Fermentation of Grape Skin by Lactobacillus plantarum KFY02. Biomolecules 2019, 9, 575. [Google Scholar] [CrossRef] [PubMed]
  53. Jiang, J.; Bi, H.; Zhuang, Y.; Liu, S.; Liu, T.; Ma, Y. Engineered synthesis of rosmarinic acid in Escherichia coli resulting production of a new intermediate, caffeoyl-phenyllactate. Biotechnol. Lett. 2016, 38, 81–88. [Google Scholar] [CrossRef] [PubMed]
  54. Dewanjee, S.; Gangopadhyay, M.; Das, U.; Sahu, R.; Khanra, R. Enhanced rosmarinic acid biosynthesis in Solenostemon scutellarioides culture: A precursor-feeding strategy. Nat. Prod. Res. 2014, 28, 1691–1698. [Google Scholar] [CrossRef] [PubMed]
  55. Busch, T.; Petersen, M. Identification and biochemical characterisation of tyrosine aminotransferase from Anthoceros agrestis unveils the conceivable entry point into rosmarinic acid biosynthesis in hornworts. Planta 2021, 253, 98. [Google Scholar] [CrossRef]
  56. Levsh, O.; Pluskal, T.; Carballo, V.; Mitchell, A.J.; Weng, J.K. Independent evolution of rosmarinic acid biosynthesis in two sister families under the Lamiids clade of flowering plants. J. Biol. Chem. 2019, 294, 15193–15205. [Google Scholar] [CrossRef]
  57. Bloch, S.E.; Schmidt-Dannert, C. Construction of a chimeric biosynthetic pathway for the de novo biosynthesis of rosmarinic acid in Escherichia coli. Chembiochem A Eur. J. Chem. Biol. 2014, 15, 2393–2401. [Google Scholar] [CrossRef] [Green Version]
  58. Yan, Y.; Jia, P.; Bai, Y.; Fan, T.P.; Zheng, X.; Cai, Y. Production of rosmarinic acid with ATP and CoA double regenerating system. Enzym. Microb. Technol. 2019, 131, 109392. [Google Scholar] [CrossRef]
  59. Li, Z.; Wang, X.; Zhang, H. Balancing the non-linear rosmarinic acid biosynthetic pathway by modular co-culture engineering. Metab. Eng. 2019, 54, 1–11. [Google Scholar] [CrossRef]
  60. Xu, Y.; Xu, G.; Liu, L.; Xu, D.; Liu, J. Anti-invasion effect of rosmarinic acid via the extracellular signal-regulated kinase and oxidation-reduction pathway in Ls174-T cells. J. Cell Biochem. 2010, 111, 370–379. [Google Scholar] [CrossRef]
  61. Zhao, J.; Ji, D.; Zhai, X.; Zhang, L.; Luo, X.; Fu, X. Oral Administration of Prunella vulgaris L Improves the Effect of Taxane on Preventing the Progression of Breast Cancer and Reduces Its Side Effects. Front Pharm. 2018, 9, 806. [Google Scholar] [CrossRef] [PubMed]
  62. Baba, S.; Osakabe, N.; Natsume, M.; Yasuda, A.; Muto, Y.; Hiyoshi, K.; Takano, H.; Yoshikawa, T.; Terao, J. Absorption, metabolism, degradation and urinary excretion of rosmarinic acid after intake of Perilla frutescens extract in humans. Eur. J. Nutr. 2005, 44, 1–9. [Google Scholar] [CrossRef] [PubMed]
  63. Blažević, T.; Reznicek, G.; Ding, L.; Yang, G.; Haiss, P.; Heiss, E.H.; Dirsch, V.M.; Liu, R. Short Chain (≤C4) Esterification Increases Bioavailability of Rosmarinic Acid and Its Potency to Inhibit Vascular Smooth Muscle Cell Proliferation. Front Pharm. 2020, 11, 609756. [Google Scholar] [CrossRef] [PubMed]
  64. Lu, P.; Xing, Y.; Xue, Z.; Ma, Z.; Zhang, B.; Peng, H.; Zhou, Q.T.; Liu, H.; Liu, Z.; Li, J. Pharmacokinetics of salvianolic acid B, rosmarinic acid and Danshensu in rat after pulmonary administration of Salvia miltiorrhiza polyphenolic acid solution. Biomed. Chromatogr. BMC 2019, 33, e4561. [Google Scholar] [CrossRef]
  65. Veras, K.S.; Fachel, F.N.S.; Pittol, V.; Garcia, K.R.; Bassani, V.L.; Dos Santos, V.; Henriques, A.T.; Teixeira, H.F.; Koester, L.S. Compatibility study of rosmarinic acid with excipients used in pharmaceutical solid dosage forms using thermal and non-thermal techniques. Saudi Pharm. J. SPJ Off. Publ. Saudi Pharm. Soc. 2019, 27, 1138–1145. [Google Scholar] [CrossRef]
  66. Fuster, M.G.; Carissimi, G.; Montalbán, M.G.; Víllora, G. Antitumor Activity of Rosmarinic Acid-Loaded Silk Fibroin Nanoparticles on HeLa and MCF-7 Cells. Polymers 2021, 13, 3169. [Google Scholar] [CrossRef]
  67. Madureira, A.R.; Nunes, S.; Campos, D.A.; Fernandes, J.C.; Marques, C.; Zuzarte, M.; Gullón, B.; Rodríguez-Alcalá, L.M.; Calhau, C.; Sarmento, B.; et al. Safety profile of solid lipid nanoparticles loaded with rosmarinic acid for oral use: In vitro and animal approaches. Int. J. Nanomed. 2016, 11, 3621–3640. [Google Scholar] [CrossRef] [Green Version]
  68. Xue, X.; Ricci, M.; Qu, H.; Lindstrom, A.; Zhang, D.; Wu, H.; Lin, T.Y.; Li, Y. Iron-crosslinked Rososome with robust stability and high drug loading for synergistic cancer therapy. J. Control. Release 2021, 329, 794–804. [Google Scholar] [CrossRef]
  69. Huang, J.; Chen, P.X.; Rogers, M.A.; Wettig, S.D. Investigating the Phospholipid Effect on the Bioaccessibility of Rosmarinic Acid-Phospholipid Complex through a Dynamic Gastrointestinal in Vitro Model. Pharmaceutics 2019, 11, 156. [Google Scholar] [CrossRef] [Green Version]
  70. Sánchez-Campillo, M.; Gabaldon, J.A.; Castillo, J.; Benavente-García, O.; Del Baño, M.J.; Alcaraz, M.; Vicente, V.; Alvarez, N.; Lozano, J.A. Rosmarinic acid, a photo-protective agent against UV and other ionizing radiations. Food Chem. Toxicol. Int. J. Publ. Br. Ind. Biol. Res. Assoc. 2009, 47, 386–392. [Google Scholar] [CrossRef]
  71. Gui, H.; Jin, Y.; Lin, A.; Wang, P.; Wang, Y.; Zhu, H. Rosmarinic acid relieves cisplatin-induced ovary toxicity in female mice via suppression of oxidative stress and inflammation. J. Biochem. Mol. Toxicol. 2021, 35, e22839. [Google Scholar] [CrossRef] [PubMed]
  72. Helfinger, V.; Schröder, K. Redox control in cancer development and progression. Mol. Asp. Med. 2018, 63, 88–98. [Google Scholar] [CrossRef] [PubMed]
  73. Anusuya, C.; Manoharan, S. Antitumor initiating potential of rosmarinic acid in 7,12-dimethylbenz(a)anthracene-induced hamster buccal pouch carcinogenesis. J. Environ. Pathol. Toxicol. Oncol. Off. Organ Int. Soc. Environ. Toxicol. Cancer 2011, 30, 199–211. [Google Scholar] [CrossRef] [PubMed]
  74. Venkatachalam, K.; Gunasekaran, S.; Jesudoss, V.A.; Namasivayam, N. The effect of rosmarinic acid on 1,2-dimethylhydrazine induced colon carcinogenesis. Exp. Toxicol. Pathol. Off. J. Ges. Fur Toxikol. Pathol. 2013, 65, 409–418. [Google Scholar] [CrossRef]
  75. Karthikkumar, V.; Sivagami, G.; Vinothkumar, R.; Rajkumar, D.; Nalini, N. Modulatory efficacy of rosmarinic acid on premalignant lesions and antioxidant status in 1,2-dimethylhydrazine induced rat colon carcinogenesis. Env. Toxicol Pharm. 2012, 34, 949–958. [Google Scholar] [CrossRef]
  76. Karthikkumar, V.; Sivagami, G.; Viswanathan, P.; Nalini, N. Rosmarinic acid inhibits DMH-induced cell proliferation in experimental rats. J. Basic Clin. Physiol. Pharm. 2015, 26, 185–200. [Google Scholar] [CrossRef]
  77. Venkatachalam, K.; Gunasekaran, S.; Namasivayam, N. Biochemical and molecular mechanisms underlying the chemopreventive efficacy of rosmarinic acid in a rat colon cancer. Eur. J. Pharm. 2016, 791, 37–50. [Google Scholar] [CrossRef]
  78. Rodríguez-Luna, A.; Ávila-Román, J.; Oliveira, H.; Motilva, V.; Talero, E. Fucoxanthin and Rosmarinic Acid Combination Has Anti-Inflammatory Effects through Regulation of NLRP3 Inflammasome in UVB-Exposed HaCaT Keratinocytes. Mar. Drugs 2019, 17, 451. [Google Scholar] [CrossRef] [Green Version]
  79. Ilhan, N.; Bektas, I.; Susam, S.; Ozercan, I.H. Protective effects of rosmarinic acid against azoxymethane-induced colorectal cancer in rats. J. Biochem. Mol. Toxicol. 2021, 36, e22961. [Google Scholar] [CrossRef]
  80. Waer, C.N.; Kaur, P.; Tumur, Z.; Hui, D.D.; Le, B.; Guerra, C.; Henson, B.; Seleem, D.; Lewis, J. Rosmarinic Acid/ Blue Light Combination Treatment Inhibits Head and Neck Squamous Cell Carcinoma In Vitro. Anticancer Res. 2020, 40, 751–758. [Google Scholar] [CrossRef] [Green Version]
  81. Saiko, P.; Steinmann, M.T.; Schuster, H.; Graser, G.; Bressler, S.; Giessrigl, B.; Lackner, A.; Grusch, M.; Krupitza, G.; Bago-Horvath, Z.; et al. Epigallocatechin gallate, ellagic acid, and rosmarinic acid perturb dNTP pools and inhibit de novo DNA synthesis and proliferation of human HL-60 promyelocytic leukemia cells: Synergism with arabinofuranosylcytosine. Phytomedicine 2015, 22, 213–222. [Google Scholar] [CrossRef]
  82. Zhang, Q.; Li, J.; Peng, S.; Zhang, Y.; Qiao, Y. Rosmarinic Acid as a Candidate in a Phenotypic Profiling Cardio-/Cytotoxicity Cell Model Induced by Doxorubicin. Molecules 2020, 25, 836. [Google Scholar] [CrossRef] [PubMed]
  83. Jafaripour, L.; Naserzadeh, R.; Alizamani, E.; Javad Mashhadi, S.M.; Moghadam, E.R.; Nouryazdan, N.; Ahmadvand, H. Effects of Rosmarinic Acid on Methotrexate-induced Nephrotoxicity and Hepatotoxicity in Wistar Rats. Indian J. Nephrol. 2021, 31, 218–224. [Google Scholar] [CrossRef] [PubMed]
  84. Rahbardar, M.G.; Eisvand, F.; Rameshrad, M.; Razavi, B.M.; Hosseinzadeh, H. In Vivo and In Vitro Protective Effects of Rosmarinic Acid against Doxorubicin-Induced Cardiotoxicity. Nutr. Cancer 2021, 74, 747–760. [Google Scholar] [CrossRef] [PubMed]
  85. Scheckel, K.A.; Degner, S.C.; Romagnolo, D.F. Rosmarinic acid antagonizes activator protein-1-dependent activation of cyclooxygenase-2 expression in human cancer and nonmalignant cell lines. J. Nutr. 2008, 138, 2098–2105. [Google Scholar] [CrossRef] [Green Version]
  86. Tao, L.; Wang, S.; Zhao, Y.; Sheng, X.; Wang, A.; Zheng, S.; Lu, Y. Phenolcarboxylic acids from medicinal herbs exert anticancer effects through disruption of COX-2 activity. Phytomedicine 2014, 21, 1473–1482. [Google Scholar] [CrossRef]
  87. Moon, D.O.; Kim, M.O.; Lee, J.D.; Choi, Y.H.; Kim, G.Y. Rosmarinic acid sensitizes cell death through suppression of TNF-alpha-induced NF-kappaB activation and ROS generation in human leukemia U937 cells. Cancer Lett. 2010, 288, 183–191. [Google Scholar] [CrossRef]
  88. Cao, W.; Hu, C.; Wu, L.; Xu, L.; Jiang, W. Rosmarinic acid inhibits inflammation and angiogenesis of hepatocellular carcinoma by suppression of NF-κB signaling in H22 tumor-bearing mice. J. Pharm. Sci. 2016, 132, 131–137. [Google Scholar] [CrossRef] [Green Version]
  89. Mahmoud, M.A.; Okda, T.M.; Omran, G.A.; Abd-Alhaseeb, M.M. Rosmarinic acid suppresses inflammation, angiogenesis, and improves paclitaxel induced apoptosis in a breast cancer model via NF3 κB-p53-caspase-3 pathways modulation. J. Appl. Biomed. 2021, 19, 202–209. [Google Scholar] [CrossRef]
  90. Wu, C.F.; Hong, C.; Klauck, S.M.; Lin, Y.L.; Efferth, T. Molecular mechanisms of rosmarinic acid from Salvia miltiorrhiza in acute lymphoblastic leukemia cells. J. Ethnopharmacol. 2015, 176, 55–68. [Google Scholar] [CrossRef]
  91. Sabharwal, S.S.; Schumacker, P.T. Mitochondrial ROS in cancer: Initiators, amplifiers or an Achilles’ heel? Nat. Rev. Cancer 2014, 14, 709–721. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Furtado, R.A.; Oliveira, B.R.; Silva, L.R.; Cleto, S.S.; Munari, C.C.; Cunha, W.R.; Tavares, D.C. Chemopreventive effects of rosmarinic acid on rat colon carcinogenesis. Eur. J. Cancer Prev. 2015, 24, 106–112. [Google Scholar] [CrossRef]
  93. Huang, Y.; Cai, Y.; Huang, R.; Zheng, X. Rosmarinic Acid Combined with Adriamycin Induces Apoptosis by Triggering Mitochondria-Mediated Signaling Pathway in HepG2 and Bel-7402 Cells. Med. Sci. Monit. 2018, 24, 7898–7908. [Google Scholar] [CrossRef] [PubMed]
  94. Zhang, Y.; Hu, M.; Liu, L.; Cheng, X.L.; Cai, J.; Zhou, J.; Wang, T. Anticancer effects of Rosmarinic acid in OVCAR-3 ovarian cancer cells are mediated via induction of apoptosis, suppression of cell migration and modulation of lncRNA MALAT-1 expression. J. BU ON. Off. J. Balk. Union Oncol. 2018, 23, 763–768. [Google Scholar]
  95. Ingham, M.; Schwartz, G.K. Cell-Cycle Therapeutics Come of Age. J. Clin. Oncol. Off. J. Am. Soc. Clin. Oncol. 2017, 35, 2949–2959. [Google Scholar] [CrossRef] [PubMed]
  96. Jang, Y.G.; Hwang, K.A.; Choi, K.C. Rosmarinic Acid, a Component of Rosemary Tea, Induced the Cell Cycle Arrest and Apoptosis through Modulation of HDAC2 Expression in Prostate Cancer Cell Lines. Nutrients 2018, 10, 1784. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Messeha, S.S.; Zarmouh, N.O.; Asiri, A.; Soliman, K.F.A. Rosmarinic acid-induced apoptosis and cell cycle arrest in triple-negative breast cancer cells. Eur. J. Pharm. 2020, 885, 173419. [Google Scholar] [CrossRef] [PubMed]
  98. Zhou, X.; Wang, W.; Li, Z.; Chen, L.; Wen, C.; Ruan, Q.; Xu, Z.; Liu, R.; Xu, J.; Bai, Y.; et al. Rosmarinic Acid Decreases the Malignancy of Pancreatic Cancer Through Inhibiting Gli1 Signaling. Phytomedicine 2022, 95, 153861. [Google Scholar] [CrossRef] [PubMed]
  99. Chou, S.T.; Ho, B.Y.; Tai, Y.T.; Huang, C.J.; Chao, W.W. Bidirect effects from cisplatin combine with rosmarinic acid (RA) or hot water extracts of Glechoma hederacea (HWG) on renal cancer cells. Chin. Med. 2020, 15, 77. [Google Scholar] [CrossRef]
  100. Luo, Y.; Ma, Z.; Xu, X.; Qi, H.; Cheng, Z.; Chen, L. Anticancer effects of rosmarinic acid in human oral cancer cells is mediated via endoplasmic reticulum stress, apoptosis, G2/M cell cycle arrest and inhibition of cell migration. J. BU ON. Off. J. Balk. Union Oncol. 2020, 25, 1245–1250. [Google Scholar]
  101. Huang, L.; Chen, J.; Quan, J.; Xiang, D. Rosmarinic acid inhibits proliferation and migration, promotes apoptosis and enhances cisplatin sensitivity of melanoma cells through inhibiting ADAM17/EGFR/AKT/GSK3β axis. Bioengineered 2021, 12, 3065–3076. [Google Scholar] [CrossRef] [PubMed]
  102. Erenler, R.; Yildiz, I.; Aydin, A.; Genc, N. Antiproliferative and cytotoxic effects of bioactive compounds isolated from Onosma bourgaei. Med. Oncol. 2022, 39, 116. [Google Scholar] [CrossRef]
  103. Hsu, K.C.; Sung, T.Y.; Lin, C.T.; Chiu, Y.Y.; Hsu, J.T.; Hung, H.C.; Sun, C.M.; Barve, I.; Chen, W.L.; Huang, W.C.; et al. Anchor-based classification and type-C inhibitors for tyrosine kinases. Sci. Rep. 2015, 5, 10938. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Anwar, S.; Shamsi, A.; Shahbaaz, M.; Queen, A.; Khan, P.; Hasan, G.M.; Islam, A.; Alajmi, M.F.; Hussain, A.; Ahmad, F.; et al. Rosmarinic Acid Exhibits Anticancer Effects via MARK4 Inhibition. Sci. Rep. 2020, 10, 10300. [Google Scholar] [CrossRef] [PubMed]
  105. Parvizpour, S.; Masoudi-Sobhanzadeh, Y.; Pourseif, M.M.; Barzegari, A.; Razmara, J.; Omidi, Y. Pharmacoinformatics-based phytochemical screening for anticancer impacts of yellow sweet clover, Melilotus officinalis (Linn.) Pall. Comput. Biol. Med. 2021, 138, 104921. [Google Scholar] [CrossRef] [PubMed]
  106. Jin, B.; Liu, J.; Gao, D.; Xu, Y.; He, L.; Zang, Y.; Li, N.; Lin, D. Detailed studies on the anticancer action of rosmarinic acid in human Hep-G2 liver carcinoma cells: Evaluating its effects on cellular apoptosis, caspase activation and suppression of cell migration and invasion. J. BU ON. Off. J. Balk. Union Oncol. 2020, 25, 1383–1389. [Google Scholar]
  107. Han, Y.H.; Kee, J.Y.; Hong, S.H. Rosmarinic Acid Activates AMPK to Inhibit Metastasis of Colorectal Cancer. Front Pharm. 2018, 9, 68. [Google Scholar] [CrossRef] [PubMed]
  108. Ma, Z.; Yang, J.; Yang, Y.; Wang, X.; Chen, G.; Shi, A.; Lu, Y.; Jia, S.; Kang, X.; Lu, L. Rosmarinic acid exerts an anticancer effect on osteosarcoma cells by inhibiting DJ-1 via regulation of the PTEN-PI3K-Akt signaling pathway. Phytomedicine 2020, 68, 153186. [Google Scholar] [CrossRef]
  109. An, Y.; Zhao, J.; Zhang, Y.; Wu, W.; Hu, J.; Hao, H.; Qiao, Y.; Tao, Y.; An, L. Rosmarinic Acid Induces Proliferation Suppression of Hepatoma Cells Associated with NF-κB Signaling Pathway. Asian Pac. J. Cancer Prev. APJCP 2021, 22, 1623–1632. [Google Scholar] [CrossRef]
  110. Wang, L.; Yang, H.; Wang, C.; Shi, X.; Li, K. Rosmarinic acid inhibits proliferation and invasion of hepatocellular carcinoma cells SMMC 7721 via PI3K/AKT/mTOR signal pathway. Biomed Pharm. 2019, 120, 109443. [Google Scholar] [CrossRef]
  111. Şengelen, A.; Önay-Uçar, E. Rosmarinic acid and siRNA combined therapy represses Hsp27 (HSPB1) expression and induces apoptosis in human glioma cells. Cell Stress Chaperones 2018, 23, 885–896. [Google Scholar] [CrossRef] [PubMed]
  112. Lee, D.H.; Choi, S.; Park, Y.; Jin, H.S. Mucin1 and Mucin16: Therapeutic Targets for Cancer Therapy. Pharmaceuticals 2021, 14, 1053. [Google Scholar] [CrossRef] [PubMed]
  113. Radziejewska, I.; Supruniuk, K.; Bielawska, A. Anti-cancer effect of combined action of anti-MUC1 and rosmarinic acid in AGS gastric cancer cells. Eur. J. Pharm. 2021, 902, 174119. [Google Scholar] [CrossRef]
  114. Merecz-Sadowska, A.; Sitarek, P.; Śliwiński, T.; Zajdel, K.; Malinowska, K.; Zielińska-Bliźniewska, H.; Kucharska, E.; Zajdel, R. In Vitro and In Silico Studies on Leonotis nepetifolia (L.) R. Br. Root Extract against Cancer Cells. Curr. Pharm. Biotechnol. 2022, 23, 1383–1395. [Google Scholar] [CrossRef] [PubMed]
  115. Hall, M.D.; Handley, M.D.; Gottesman, M.M. Is resistance useless? Multidrug resistance and collateral sensitivity. Trends Pharmacol. Sci. 2009, 30, 546–556. [Google Scholar] [CrossRef] [PubMed]
  116. Robey, R.W.; Pluchino, K.M.; Hall, M.D.; Fojo, A.T.; Bates, S.E.; Gottesman, M.M. Revisiting the role of ABC transporters in multidrug-resistant cancer. Nat. Rev. Cancer 2018, 18, 452–464. [Google Scholar] [CrossRef] [PubMed]
  117. Yehya, A.H.S.; Asif, M.; Abdul Majid, A.M.S.; Oon, C.E. Complementary effects of Orthosiphon stamineus standardized ethanolic extract and rosmarinic acid in combination with gemcitabine on pancreatic cancer. Biomed. J. 2021, 44, 694–708. [Google Scholar] [CrossRef] [PubMed]
  118. Juskowiak, B.; Bogacz, A.; Wolek, M.; Kamiński, A.; Uzar, I.; Seremak-Mrozikiewicz, A.; Czerny, B. Expression profiling of genes modulated by rosmarinic acid (RA) in MCF-7 breast cancer cells. Ginekol. Pol. 2018, 89, 541–545. [Google Scholar] [CrossRef] [Green Version]
  119. Li, F.R.; Fu, Y.Y.; Jiang, D.H.; Wu, Z.; Zhou, Y.J.; Guo, L.; Dong, Z.M.; Wang, Z.Z. Reversal effect of rosmarinic acid on multidrug resistance in SGC7901/Adr cell. J. Asian Nat. Prod. Res. 2013, 15, 276–285. [Google Scholar] [CrossRef]
  120. Jin, W.; Liao, X.; Lv, Y.; Pang, Z.; Wang, Y.; Li, Q.; Liao, Y.; Ye, Q.; Chen, G.; Zhao, K.; et al. MUC1 induces acquired chemoresistance by upregulating ABCB1 in EGFR-dependent manner. Cell Death Dis. 2017, 8, e2980. [Google Scholar] [CrossRef] [Green Version]
  121. Nath, S.; Daneshvar, K.; Roy, L.D.; Grover, P.; Kidiyoor, A.; Mosley, L.; Sahraei, M.; Mukherjee, P. MUC1 induces drug resistance in pancreatic cancer cells via upregulation of multidrug resistance genes. Oncogenesis 2013, 2, e51. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Radziejewska, I.; Supruniuk, K.; Nazaruk, J.; Karna, E.; Popławska, B.; Bielawska, A.; Galicka, A. Rosmarinic acid influences collagen, MMPs, TIMPs, glycosylation and MUC1 in CRL-1739 gastric cancer cell line. Biomed. Pharm. 2018, 107, 397–407. [Google Scholar] [CrossRef] [PubMed]
  123. Ma, Z.J.; Yan, H.; Wang, Y.J.; Yang, Y.; Li, X.B.; Shi, A.C.; Jing-Wen, X.; Yu-Bao, L.; Li, L.; Wang, X.X. Proteomics analysis demonstrating rosmarinic acid suppresses cell growth by blocking the glycolytic pathway in human HepG2 cells. Biomed. Pharm. 2018, 105, 334–349. [Google Scholar] [CrossRef] [PubMed]
  124. Kelly, B.; O’Neill, L.A. Metabolic reprogramming in macrophages and dendritic cells in innate immunity. Cell Res. 2015, 25, 771–784. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Han, S.; Yang, S.; Cai, Z.; Pan, D.; Li, Z.; Huang, Z.; Zhang, P.; Zhu, H.; Lei, L.; Wang, W. Anti-Warburg effect of rosmarinic acid via miR-155 in gastric cancer cells. Drug Des. Dev. Ther. 2015, 9, 2695–2703. [Google Scholar] [CrossRef]
  126. Xu, Y.; Han, S.; Lei, K.; Chang, X.; Wang, K.; Li, Z.; Liu, J. Anti-Warburg effect of rosmarinic acid via miR-155 in colorectal carcinoma cells. Eur. J. Cancer Prev. 2016, 25, 481–489. [Google Scholar] [CrossRef]
  127. Aiello, N.M.; Kang, Y. Context-dependent EMT programs in cancer metastasis. J. Exp. Med. 2019, 216, 1016–1026. [Google Scholar] [CrossRef] [Green Version]
  128. Yang, K.; Shen, Z.; Zou, Y.; Gao, K. Rosmarinic acid inhibits migration, invasion, and p38/AP-1 signaling via miR-1225-5p in colorectal cancer cells. J. Recept. Signal Transduct. 2021, 41, 284–293. [Google Scholar] [CrossRef]
  129. Han, Y.; Ma, L.; Zhao, L.; Feng, W.; Zheng, X. Rosmarinic inhibits cell proliferation, invasion and migration via up-regulating miR-506 and suppressing MMP2/16 expression in pancreatic cancer. Biomed. Pharm. 2019, 115, 108878. [Google Scholar] [CrossRef]
  130. Kessenbrock, K.; Plaks, V.; Werb, Z. Matrix metalloproteinases: Regulators of the tumor microenvironment. Cell 2010, 141, 52–67. [Google Scholar] [CrossRef] [Green Version]
  131. Bergers, G.; Brekken, R.; McMahon, G.; Vu, T.H.; Itoh, T.; Tamaki, K.; Tanzawa, K.; Thorpe, P.; Itohara, S.; Werb, Z.; et al. Matrix metalloproteinase-9 triggers the angiogenic switch during carcinogenesis. Nat. Cell Biol. 2000, 2, 737–744. [Google Scholar] [CrossRef] [PubMed]
  132. Xu, Y.; Jiang, Z.; Ji, G.; Liu, J. Inhibition of bone metastasis from breast carcinoma by rosmarinic acid. Planta Med. 2010, 76, 956–962. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Ahamed, M.B.; Aisha, A.F.; Nassar, Z.D.; Siddiqui, J.M.; Ismail, Z.; Omari, S.M.; Parish, C.R.; Majid, A.M. Cat’s whiskers tea (Orthosiphon stamineus) extract inhibits growth of colon tumor in nude mice and angiogenesis in endothelial cells via suppressing VEGFR phosphorylation. Nutr. Cancer 2012, 64, 89–99. [Google Scholar] [CrossRef] [PubMed]
  134. Huang, S.S.; Zheng, R.L. Rosmarinic acid inhibits angiogenesis and its mechanism of action in vitro. Cancer Lett. 2006, 239, 271–280. [Google Scholar] [CrossRef]
  135. De Robertis, M.; Massi, E.; Poeta, M.L.; Carotti, S.; Morini, S.; Cecchetelli, L.; Signori, E.; Fazio, V.M. The AOM/DSS murine model for the study of colon carcinogenesis: From pathways to diagnosis and therapy studies. J. Carcinog. 2011, 10, 9. [Google Scholar] [CrossRef] [PubMed]
  136. Karmokar, A.; Marczylo, T.H.; Cai, H.; Steward, W.P.; Gescher, A.J.; Brown, K. Dietary intake of rosmarinic acid by Apc(Min) mice, a model of colorectal carcinogenesis: Levels of parent agent in the target tissue and effect on adenoma development. Mol. Nutr. Food Res. 2012, 56, 775–783. [Google Scholar] [CrossRef]
  137. Lubov, J.E.; Cvammen, W.; Kemp, M.G. The Impact of the Circadian Clock on Skin Physiology and Cancer Development. Int. J. Mol. Sci. 2021, 22, 6112. [Google Scholar] [CrossRef]
  138. Rudolf, J.; Raad, H.; Taieb, A.; Rezvani, H.R. NADPH Oxidases and Their Roles in Skin Homeostasis and Carcinogenesis. Antioxid. Redox Signal. 2018, 28, 1238–1261. [Google Scholar] [CrossRef]
  139. Gupta, D.; Archoo, S.; Naikoo, S.H.; Abdullah, S.T. Rosmarinic Acid: A Naturally Occurring Plant Based Agent Prevents Impaired Mitochondrial Dynamics and Apoptosis in Ultraviolet-B-Irradiated Human Skin Cells. Photochem. Photobiol. 2022, 98, 925–934. [Google Scholar] [CrossRef] [PubMed]
  140. Baldasquin-Caceres, B.; Gomez-Garcia, F.J.; López-Jornet, P.; Castillo-Sanchez, J.; Vicente-Ortega, V. Chemopreventive potential of phenolic compounds in oral carcinogenesis. Arch. Oral Biol. 2014, 59, 1101–1107. [Google Scholar] [CrossRef]
  141. Paluszczak, J.; Krajka-Kuźniak, V.; Baer-Dubowska, W. The effect of dietary polyphenols on the epigenetic regulation of gene expression in MCF7 breast cancer cells. Toxicol. Lett. 2010, 192, 119–125. [Google Scholar] [CrossRef]
  142. Li, W.; Li, Q.; Wei, L.; Pan, X.; Huang, D.; Gan, J.; Tang, S. Rosmarinic Acid Analogue-11 Induces Apoptosis of Human Gastric Cancer SGC-7901 Cells via the Epidermal Growth Factor Receptor (EGFR)/Akt/Nuclear Factor kappa B (NF-κB) Pathway. Med. Sci. Monit. Basic Res. 2019, 25, 63–75. [Google Scholar] [CrossRef]
  143. Kamisawa, T.; Wood, L.D.; Itoi, T.; Takaori, K. Pancreatic cancer. Lancet 2016, 388, 73–85. [Google Scholar] [CrossRef]
  144. Biller, L.H.; Schrag, D. Diagnosis and Treatment of Metastatic Colorectal Cancer: A Review. JAMA 2021, 325, 669–685. [Google Scholar] [CrossRef] [PubMed]
  145. Schmitt, M.; Greten, F.R. The inflammatory pathogenesis of colorectal cancer. Nat. Reviews. Immunol. 2021, 21, 653–667. [Google Scholar] [CrossRef] [PubMed]
  146. Lim, S.H.; Nam, K.H.; Kim, K.; Yi, S.A.; Lee, J.; Han, J.W. Rosmarinic Acid Methyl Ester Regulates Ovarian Cancer Cell Migration and Reverses Cisplatin Resistance by Inhibiting the Expression of Forkhead Box M1. Pharmaceuticals 2020, 13, 302. [Google Scholar] [CrossRef] [PubMed]
  147. Nam, K.H.; Yi, S.A.; Nam, G.; Noh, J.S.; Park, J.W.; Lee, M.G.; Park, J.H.; Oh, H.; Lee, J.; Lee, K.R.; et al. Identification of a novel S6K1 inhibitor, rosmarinic acid methyl ester, for treating cisplatin-resistant cervical cancer. BMC Cancer 2019, 19, 773. [Google Scholar] [CrossRef] [PubMed]
  148. Canturk, Z.; Dikmen, M.; Artagan, O.; Ozarda, M.G.; Ozturk, N. Cytotoxic Effects of Resveratrol, Rutin and Rosmarinic Acid on ARH-77 Human (Multiple Myeloma) Cell Line. Nat. Prod. Commun. 2016, 11, 1441–1444. [Google Scholar] [CrossRef] [Green Version]
  149. Galluzzi, L.; Senovilla, L.; Vitale, I.; Michels, J.; Martins, I.; Kepp, O.; Castedo, M.; Kroemer, G. Molecular mechanisms of cisplatin resistance. Oncogene 2012, 31, 1869–1883. [Google Scholar] [CrossRef] [Green Version]
  150. Citron, M.L.; Berry, D.A.; Cirrincione, C.; Hudis, C.; Winer, E.P.; Gradishar, W.J.; Davidson, N.E.; Martino, S.; Livingston, R.; Ingle, J.N.; et al. Randomized trial of dose-dense versus conventionally scheduled and sequential versus concurrent combination chemotherapy as postoperative adjuvant treatment of node-positive primary breast cancer: First report of Intergroup Trial C9741/Cancer and Leukemia Group B Trial 9741. J. Clin. Oncol. Off. J. Am. Soc. Clin. Oncol. 2003, 21, 1431–1439. [Google Scholar] [CrossRef]
  151. Yu, C.; Chen, D.Q.; Liu, H.X.; Li, W.B.; Lu, J.W.; Feng, J.F. Rosmarinic acid reduces the resistance of gastric carcinoma cells to 5-fluorouracil by downregulating FOXO4-targeting miR-6785-5p. Biomed. Pharm. 2019, 109, 2327–2334. [Google Scholar] [CrossRef] [PubMed]
  152. Ozgun, G.S.; Ozgun, E. The cytotoxic concentration of rosmarinic acid increases MG132-induced cytotoxicity, proteasome inhibition, autophagy, cellular stresses, and apoptosis in HepG2 cells. Hum. Exp. Toxicol. 2020, 39, 514–523. [Google Scholar] [CrossRef] [PubMed]
  153. Ni, X.; Hu, G.; Cai, X. The success and the challenge of all-trans retinoic acid in the treatment of cancer. Crit. Rev. Food Sci. Nutr. 2019, 59, S71–S80. [Google Scholar] [CrossRef] [PubMed]
  154. Stahl, M.; Tallman, M.S. Acute promyelocytic leukemia (APL): Remaining challenges towards a cure for all. Leuk. Lymphoma 2019, 60, 3107–3115. [Google Scholar] [CrossRef] [PubMed]
  155. Heo, S.K.; Noh, E.K.; Yoon, D.J.; Jo, J.C.; Koh, S.; Baek, J.H.; Park, J.H.; Min, Y.J.; Kim, H. Rosmarinic acid potentiates ATRA-induced macrophage differentiation in acute promyelocytic leukemia NB4 cells. Eur. J. Pharm. 2015, 747, 36–44. [Google Scholar] [CrossRef]
  156. Alcaraz, M.; Alcaraz-Saura, M.; Achel, D.G.; Olivares, A.; López-Morata, J.A.; Castillo, J. Radiosensitizing effect of rosmarinic acid in metastatic melanoma B16F10 cells. Anticancer Res. 2014, 34, 1913–1921. [Google Scholar]
  157. Ferreira-Gonçalves, T.; Gaspar, M.M.; Coelho, J.M.P.; Marques, V.; Viana, A.S.; Ascensão, L.; Carvalho, L.; Rodrigues, C.M.P.; Ferreira, H.A.; Ferreira, D.; et al. The Role of Rosmarinic Acid on the Bioproduction of Gold Nanoparticles as Part of a Photothermal Approach for Breast Cancer Treatment. Biomolecules 2022, 12, 71. [Google Scholar] [CrossRef]
Figure 1. Summary of anti-tumor effects of RA.
Figure 1. Summary of anti-tumor effects of RA.
Biomolecules 12 01410 g001
Figure 2. Mechanisms of tumorigenesis prevention effects of RA. When cells undergo oxidative stress or inflammatory factors (ILs), ROS and MDA are accumulated. The intracellular peroxidases SOD dismutate superoxide anion to H2O2, then GPx and CAT catalyze the decomposition of H2O2 into H2O to maintain cellular ROS homeostasis. When cells are continuously stimulated by ROS, accompanied by the activation of NF-κB and COX-2, the normal cells can abnormally proliferate, differentiate, and escape apoptosis, leading to tumorigenesis. RA inhibits the secretion of IL-1β and IL-6, the expression of NF-κB and COX-2, and downregulates the content of ROS and MDA.
Figure 2. Mechanisms of tumorigenesis prevention effects of RA. When cells undergo oxidative stress or inflammatory factors (ILs), ROS and MDA are accumulated. The intracellular peroxidases SOD dismutate superoxide anion to H2O2, then GPx and CAT catalyze the decomposition of H2O2 into H2O to maintain cellular ROS homeostasis. When cells are continuously stimulated by ROS, accompanied by the activation of NF-κB and COX-2, the normal cells can abnormally proliferate, differentiate, and escape apoptosis, leading to tumorigenesis. RA inhibits the secretion of IL-1β and IL-6, the expression of NF-κB and COX-2, and downregulates the content of ROS and MDA.
Biomolecules 12 01410 g002
Figure 3. The key mechanisms of anti-tumor effects of RA. Extrinsic apoptosis: RA induces extrinsic apoptosis by upregulating TNF-α and caspase-8. Endogenous apoptosis: RA inhibits BCL-2 and promotes the expression of BAX, caspase-3, -7, and -9 to mediate endogenous apoptosis. P53 pathway and cell cycle: RA inhibits MDM2 expression and promotes p53, p21 expression mediates apoptosis. P21 inhibits Cyclin D and Cyclin E to induce cell cycle arrest, and RA inhibits Cyclin B, Cyclin D, and Cyclin E to mediate cell cycle arrest and promote apoptosis. PI3K/AKT and NF-κB pathway: RA promotes the expression of tumor suppressor gene PTEN, inhibits PI3K expression, AKT phosphorylation, NF-κB expression, and p65 phosphorylation, which induces apoptosis through the inhibition of cell survival-related pathways.
Figure 3. The key mechanisms of anti-tumor effects of RA. Extrinsic apoptosis: RA induces extrinsic apoptosis by upregulating TNF-α and caspase-8. Endogenous apoptosis: RA inhibits BCL-2 and promotes the expression of BAX, caspase-3, -7, and -9 to mediate endogenous apoptosis. P53 pathway and cell cycle: RA inhibits MDM2 expression and promotes p53, p21 expression mediates apoptosis. P21 inhibits Cyclin D and Cyclin E to induce cell cycle arrest, and RA inhibits Cyclin B, Cyclin D, and Cyclin E to mediate cell cycle arrest and promote apoptosis. PI3K/AKT and NF-κB pathway: RA promotes the expression of tumor suppressor gene PTEN, inhibits PI3K expression, AKT phosphorylation, NF-κB expression, and p65 phosphorylation, which induces apoptosis through the inhibition of cell survival-related pathways.
Biomolecules 12 01410 g003
Table 2. Summary of tumorigenesis prevention effects of RA.
Table 2. Summary of tumorigenesis prevention effects of RA.
DiseaseModelTreatmentOutcomeRef
Colorectal carcinogenesisWistar male rats given DMH orally 20 mg/kg, once a dayRA 10mg/kg, once a dayInhibited the carcinogenic effect through circulatory antioxidant enzymes (SOD↑, CAT↑, GSH↑, and GPx↑)[75]
Colorectal carcinogenesisMale albino wistar rat given DMH 20 mg/kg subcutaneously for 4 weeksRA 2.5, 5, and 10 mg/kgReduced the polyp incidence through CYP450↓, lipid peroxidation↓, SOD↑, CAT↑, GPx↑, and GSH↑.[74]
Colorectal carcinogenesisWistar rats with subcutaneous injection of 40 mg/kg DMH for 2 weeksRA 4, 8 and 16 mg/kg body weightReduced DNA damage and frequency of the formation of ACF[92]
Colorectal carcinogenesisMale Wistar rats with subcutaneous injection of DMH 20 mg/kg.Oral RA 5mg/kg body weight 30 weeks in totalInhibited the tumor formation and reduced expressions of TNF-α, IL-6, and COX-2, and increased SOD, CAT, GPx, and TBARS[76]
Colorectal carcinogenesisMale Wistar rats with subcutaneous injection of DMH 20 mg/kg for 15 weeksDaily RA 5mg/kg orally Protected the activity of antioxidant enzymes (CYP450↓ and CYP4502E1↓) and reduced the formation of ACF[77]
Colorectal carcinogenesisMale Sprague-Dawley rats intraperitoneally injected with 15 mg/kg AOM once a week for 4 weeks RA 5 mg/kg orally per dayIncreased the total antioxidant status, and decreased the expression of IL-6 and total oxidant status[79]
Colorectal carcinogenesisMale BALB/c mice with oral administration of AOM5-ASA 75 mg/kg/day intraperitoneally for 7 days, then supplied drinking water containing 1–2% DSS for 49 daysRA 30 mg/kg/day orallyInhibited TLR4 mediated the activation of NF-κB and STAT3 and eliminated the progression of colitis-associated colon cancer[9]
Colorectal carcinogenesisAPC10.1 cells;
C57BL/6J-ApcMin/+ mouse model
RA 100 µM;
0.3% RA in the diet, 360 mg/kg per day
Decreased numbers of large adenomas (>3 mm)[136]
Skin carcinogenesisDMBA/TPA induced skin papilloma mouse modelTopical application RA 1.35 mg/mouseInhibited MDA, chemokines and arachidonic acid and prevented DNA from oxidative damage[20]
Skin carcinogenesisHaCaT cells exposed to UVARA 2.7–18 mg/mL Attenuated ROS generation and DNA damage in UVB-irradiated keratinocytes by LBE[24]
Skin carcinogenesisHaCaT cells exposed to UVBRA 2.5 or 5 µM Downregulated the inflammasome components (NLRP3 and IL-1β production) via Nrf2/HO-1 antioxidant system and prevented skin changes caused by UVB[78]
Skin carcinogenesisB16 melanoma cells;
Female albino Swiss mouses exposed to UVA light 3 times a week, total 100 times
2% RA in the diet to rats;
Cell administration RA at 1 mg/mL
RA increased the Tyr activity in vitro.
Oral RA inhibited skin changes caused by UVA exposure (skin photocarcinogenesis)
[70]
Oral carcinogenesis0.5% DMBA liquid paraffin treated on left buccal pouches of golden Syrian hamster model for 14 weeksRA orally 100 mg/kgSuppressed oral carcinogenesis through upregulation of SOD, CAT, GSH, GPx and downregulation of TBARS and BCL-2[73]
Oral carcinogenesisMale Syrian hamster intravenous injection of 0.5% DMBARA 1.3 mg/15mL Reduced the intensity and invasiveness of the tumor[140]
Tumor angiogenesisHuman umbilical vein endothelial cells (HUVECs)RA 50, 100 and 200 mMSuppression of ROS generation and downregulation the release of VEGF and IL-8[134]
Table 3. Summary of anti-tumor effects of RA.
Table 3. Summary of anti-tumor effects of RA.
DiseaseModel (IC50)TreatmentOutcomeRef
GliomaU251 and U343 glioma cellsRA 100, 200, and 400 µMInhibited BCL-2 and promoted the expression of BAX and cleaved caspase-3 protein, and downregulated PI3K/AKT/NF-κB signaling pathway through targeting Fyn.[10]
GliomaU-87 MG cells (IC50 for 48 h:373.48 μM)RA 80 and 215 µM Inhibited the expression of HSP27 and enhanced the activity of caspase-3[111]
Oral cancerSCC-15 cellsRA 10, 20, and 40 µMIncreased the expression of cleaved caspase-3 and BAX/BCL-2 ratio, induced G2/M cell cycle arrest, and inhibited migration through downregulation of MMP-2 and MMP-9[100]
Breast cancerMDA-MB-231 (IC50 for 48 h: 321.75 ± 9.75 uM) and MDA-MB-468 cells (IC50 for 48 h: 340.45 ± 7.57 uM)RA 125 and 250 µMInduced G0/G1 cell cycle arrest and apoptosis through regulation of apoptosis-related genes (HRK↑, TNFRSF25↑, BNIP3↑, TNF↑, GADD45A↑, BNIP3↑, TNFSF10↓, BIRC5↓ and TNFRSF11B↓)[97]
Breast cancerMCF7 cell line RA 20 and 40 µMRegulated the methylation pattern via DNMT1 for chemoprevention of cancer[141]
Breast-derived bone metastasesMDA-MB-231BO human bone-homing breast cancer cells (IC50: 118.04 µg/mL)RA 7.5, 15, 30, and 60 µg/mLInhibited the metastasis of breast cancer by suppression of IL-8 through NF-κB ligand/ TNF receptor superfamily member 11a /osteoprotegerin pathway[132]
Gastric cancerMKN45 cells (IC50 for 24 h: 240.2 μM);
MKN45 cells injected into BALB/c-nude mice
RA 60, 120.1, and 240.2 µM;
RA 2 mg/kg injected intraperitoneally for 14 days
Inhibited Warburg effect (glucose consumption, lactate generation, and HIF-1α) through downregulation of IL-6/STAT3 pathway[125]
Gastric cancerCRL-1739 cells (IC50 for 24 h: 240 μM)RA 100 and 200 μMInhibited the expression of MMP-9, TIMP-1, MUC1, Tn antigens and T antigens, increased the expression of collagen I[122]
Gastric cancerGES-1 (IC50 for 24 h: 289.425 ± 0.854 μmol/L) and SGC-7901 cells (IC50 for 24 h: 73.299 ± 2.011 μmol/L) RA analogue-11 10, 20, and 40μmol/LPromoted apoptosis via the EGFR/AKT/NF-κB pathway in gastric cancer cells.[142]
HCCHepG2 cellsRA 5 and 10 µg/mLInduced apoptosis through increasing the mRNA levels of Jun, Jun-B, Fos-B, BAX and caspase-8, and decreased BCL-2 mRNA expression[19]
HCCH22 tumor-bearing miceIntraperitoneal injection of RA 75, 150, and 300 mg/kgInhibited inflammatory cytokines (IL-1β, IL-6, TNF-α, TGF-β), angiogenic factors (VEGF) and phosphorylation of p65. The tumor inhibition rates in different concentrations of RA (39.03%, 42.98%, and 48.24%)[88]
HCCHepG2 cells (IC50 for 48 h: 33 ± 0.74 μg/mL)RA 6.25, 12.5, 25, 50, and 100 µg/mLInhibited the expression of GLUT-1 and HK-2 to suppress the glycolytic pathway.[123]
HCCHepG2 cellsRA 7, 14, and 28 µMInduced apoptosis (caspase-3↑, caspase-9↑ and BAX/BCL-2 ratio↑), inhibited migration, and invasion[106]
HCCHepG2 cellsRA 100, 200, and 400 µMReduced the expression of MMP-2, MMP-9, and BCL-2, promoted the expression of BAX and Caspase-3, and downregulated PI3K/AKT/NF-κB signaling pathway through targeting Fyn.[109]
HCCSMMC 7721 cells;
Tumor bearing model of nude mice
RA 20, 50, and 100 µmol/L;
RA 5, 10, and 20 mg/kg for 5 days
Downregulated PI3K/AKT/mTOR signaling pathway to induce apoptosis, inhibited EMT in vitro and tumor growth in vivo[110]
Pancreatic cancerPANC-1, PATU-8988, MIA PaCa-2 and BxPC-3 cells;
Tumor bearing model of nude mice (MIA PaCa-2 cells)
RA 100, 200, 300, 400, and 500 μM; Orally 50 mg/kg
RA 50 mg/kg orally for 30 days
Enhanced proteasome-mediated degradation of Gli1 and inhibited the expression of downstream VEGF, Cyclin D1 and snail1.
Induced apoptosis and inhibited invasion and proliferation in vitro;
Suppressed tumor growth in vivo
[98]
Pancreatic cancerPanc-1 (IC50 for 24 h: 104.2 ± 4.5 μM) and SW1990 cells (IC50 for 24 h: 118.9 ± 6.7 μM);
Nude mice injected subcutaneously into Panc-1 cells
RA 100 µM;
10 and 50 mg/kg orally for 30 days
Inhibited mRNA expression of MMP2 and MMP16 via miR-506;
Inhibited tumor growth in the xenograft mice model.
[129]
CRCHCT15 and CO115 cellsRA 10, 50, and 100 µMInhibited cell proliferation through inhibitory of phospho-ERK in HCT15[35]
CRCHCT8 (IC50: 298.1 μM), HCT116 (IC50: 319.8 μM), Ls174-T (IC50: 539.4 μM), and Lovo (IC50: 576.3 μM) cellsRA 0, 75, and 150 µMInhibited IL-1β, TNFα, IL-6, and STAT3 against Warburg effect [126]
CRCCT26 and HCT116;
BALB/c mice inoculated with CT26 via the lateral tail vein
RA 50, 100, and 200 µM;
oral injection of RA (100 mg/kg/day) for 14 days
Induced G0/G1 cell cycle arrest and apoptosis (caspases↑, Bcl-XL↓, and BCL-2↓), inhibited EMT and invasion via AMPK phosphorylation;
Reduced lung metastasis of CRC cells
[107]
Colon carcinoma
Lung cancer
Ls174-T human colon carcinoma cells.
Lewis lung carcinoma (LLC) cells injected into C57BL/6 mice
RA 37.5, 75, 150, and 300 µg/mL in vitro;
RA 1, 2, and 4 mg/kg intraperitoneal injection for 20 days
Inhibited the activities of EGFR and VEGFR, and then suppressed the nuclear translocation of NF-κB and activity of p-AKT and p-ERK resulting in downregulation of the mRNA and protein expression of MMP-2, MMP-9, and VEGF in vitro.
Inhibited the formation of metastasis nodules.
[60]
CRCHT-29 cellsRA 50, 100, and 200 µMInhibited EMT (E-cadherin, N-cadherin, MMP-1, -3, and -9) via the p38/AP-1 signaling[128]
Ovarian cancerOVCAR-3 cellsRA 10, 40, and 160 µMRegulated the expression of lncRNA MALAT-1, inhibited cell migration and induced apoptosis.[94]
Ovarian cancerSKOV-3, TOV-21G and DDP resistant daughter line TOV/CisRRA methyl ester 40 µM;
DDP 5µM;
combination therapy
Accelerated apoptosis in DDP resistant ovarian cancer cell line through inhibitory of FOXM1[146]
Cervical cancerHeLa and SiHa cellsRA methyl ester 80 µM;
DDP 5µM;
combination therapy
Exerted apoptosis effects against cervical cancer by inhibiting mTOR/S6K1 pathway[147]
Prostate cancerPC-3, DU145 cellsRA 200 µMInduced G0/G1 cell cycle arrest (Cyclin D1↓, Cyclin E↓, CyclinB1↓ and p21↑) and apoptosis, enhanced transcription of p53 by inhibition of HDAC2.[96]
OsteosarcomaU2OS (IC50 for 48h: 28 ± 1.14 μg/mL) and MG63 (IC50 for 48h: 25 ± 1.37 μg/mL) osteosarcoma cells.RA 12.5, 25, and 50 µg/mLInduced apoptosis (caspase-3, -8, and -9↑ and BAX/BCL-2 ratio↑), inhibited EMT and invasion (MMP-2↓, MMP-9↓) through DJ-1 mediated upregulation of PTEN and downregulation of PI3K/AKT [108]
MMARH-77 cellsRA 50, 100, and 200 µMExerted cytotoxic effects and decreased the mitochondrial activity[148]
LeukemiaU937 cells using TNF- α 10 ng/mL induced oxidative stressRA 60 µMReduced NF- κB and ROS production, promoted apoptosis[87]
Acute lymphoblastic leukemiaCCRF-CEM (IC50 for 48h: 14.6 ± 1.58 μM) and CEM/ADR5000 (IC50 for 48h: 44.5 ± 5.3 μM) cellsRA 15, 30, and 60 µMTargeted IKK-β to inhibit NF-κB signaling pathway, caused disruption of MMP and cell adhesion and promoted caspase-independent cell death[90]
Table 4. Summary of RA as chemosensitizers in tumor therapy.
Table 4. Summary of RA as chemosensitizers in tumor therapy.
DiseaseModelTreatmentOutcomeRef
Lung cancerA549 and A549/DDP (DDP resistance) cellsRA 10, 15, 20, and 40 µg/mL;
DDP 1 µg/mL;
combination therapy
Inhibited proliferation and invasion, and enhanced chemosensitivity to DDP based on downregulation of MDR1 mRNA expression [11]
Renal cancer786-O cellsRA 25, 50, and 100 µM;
DDP 5µM;
combination therapy
Induced G2/M phase arrest and apoptosis in renal cancer cells.[99]
Ovarian
cancer
A2780 and DDP resistant daughter line A2780CP70RA 2.5, 5, and 10 g/mLShowed synergistic anti-proliferation effect with DDP on A2780 cells[18]
Melanoma A375 cellsRA 50, 100, and 200 µg/mL;
DDP 8 µM;
combination therapy
Inhibited cell proliferation, invasion, and melanin synthesis, and increased apoptosis and DDP sensitivity via inhibitory of ADAM17/EGFR/AKT/GSK3β axis[101]
Breast cancerFemale Swiss albino mice with intradermal injection of 0.1 mL Ehrlich ascites carcinomaOral RA 50 mg/kg;
Paclitaxel 10 mg/kg/three times weekly intraperitoneally;
combination therapy
Exerted chemo-preventive in combination with paclitaxel, suppressed NF-κB, TNF-α, and VEGF, increased in apoptotic markers p53, caspase-3, and BAX/BCL-2 ratio [89]
Breast cancerMCF-7 cellsRA 1.5, 15, or 50 µM;
DOX 0.2 µM;
combination therapy
Decreased the MDM2 gene expression and potentiated the effect of DOX[118]
Gastric cancerAGS cellsRA 100 and 200 µM;
Anti-MUC1 antibody 5 µg/mL
combination therapy
Inhibited the expression of MUC1, BCL-2, Tn antigens and T antigens, increased the expression of caspase-9, BAX, and BAD[113]
Gastric cancerSGC7901/Adr cells (DOX resistance)RA 2.4 and 12 µMReversed the MDR of SGC7901/Adr cells, increased sensitivity to DOX and Rh123 through downregulating the expression of MDR1 transcript levels[119]
Gastric cancerSGC7901 and SGC7901/5-Fu (5-Fu resistance) cellsRA 15 µg/mL;
5-Fu 50 µg/mL;
combination therapy
Enhanced chemosensitivity to 5-Fu, increased FOXO4 by downregulating miR-6785-5p and miR-642a-3p[151]
HCCHepG2 and Bel-7402 CellsRA 25, 50, and 100 µg/mL;
DOX 0.4 µg/mL;
combination therapy
Enhanced DNA damage and apoptosis (BAX/BCL-2 ratio↑)[93]
HCCHepG2 cellsRA 10, 100, and 1000 mM;
MG132 1 µM;
combination therapy
Synergistically increased cytotoxicity, proteasome inhibition, autophagy, and apoptosis[152]
Pancreatic cancerPanc-1 cellsRA 10 and 20 µM;
Gemcitabine 12.5 nM;
combination therapy
Exerted anti-migration, pro-apoptosis effects and enhanced the efficacy of gemcitabine through downregulation of MRP-4, MRP-5, and Notch1 intracellular domain[117]
APLNB4 cellsRA 40 mM;
ATRA 10 nM;
combination therapy
RA potentiated ATRA-induced macrophage differentiation in APL cells and increased CCR-1, CCR-2, and ICAM-1 expression through activation of ERK and NF-κB[155]
APLHL-60 cellsRA 100, 125, and 150 µM;
Ara-C 5, 10, and 20 nM;
combination therapy
Synergistically inhibited DNA synthesis to potentiated the anti-proliferative effect of Ara-C[81]
HNSCCUM-SCC-1, UM-SCC-6, and OSC-2 cellsRA 80 µg/mL;
Blue light 400–500 nm; 60 J/cm2, 2 min;
combination therapy
Reduced EGFR activation and H2O2 production.[80]
Metastatic melanomaB16F10 cellsRA 20 and 40μM;
RA combination with X-rays
Specifically sensitized radiation induces apoptosis of tumor cells[156]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhao, J.; Xu, L.; Jin, D.; Xin, Y.; Tian, L.; Wang, T.; Zhao, D.; Wang, Z.; Wang, J. Rosmarinic Acid and Related Dietary Supplements: Potential Applications in the Prevention and Treatment of Cancer. Biomolecules 2022, 12, 1410. https://doi.org/10.3390/biom12101410

AMA Style

Zhao J, Xu L, Jin D, Xin Y, Tian L, Wang T, Zhao D, Wang Z, Wang J. Rosmarinic Acid and Related Dietary Supplements: Potential Applications in the Prevention and Treatment of Cancer. Biomolecules. 2022; 12(10):1410. https://doi.org/10.3390/biom12101410

Chicago/Turabian Style

Zhao, Jiachao, Liwei Xu, Di Jin, Yu Xin, Lin Tian, Tan Wang, Daqing Zhao, Zeyu Wang, and Jing Wang. 2022. "Rosmarinic Acid and Related Dietary Supplements: Potential Applications in the Prevention and Treatment of Cancer" Biomolecules 12, no. 10: 1410. https://doi.org/10.3390/biom12101410

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop