1,176
Views
0
CrossRef citations to date
0
Altmetric
Review Article

Anticancer perspectives of genistein: a comprehensive review

, , , , ORCID Icon, , , , , , , , & ORCID Icon show all
Pages 3305-3341 | Received 11 Aug 2023, Accepted 03 Nov 2023, Published online: 27 Nov 2023

ABSTRACT

There is a significant load of degenerative and chronic illnesses, especially cancer, which is one of the main reasons for morbidity and death globally. Polyphenolic phytochemicals found in many plant diets have been shown in epidemiologic and preclinical studies to have chemopreventive activities against numerous cancer types. As a result, there is growing interest in possible cancer chemopreventive medicines derived from natural compounds, such as polyphenols, which may provide a novel, cost-effective way to reduce the global cancer burden. Various epidemiologic researchers have found a link between a soy-rich diet and tumor avoidance, which has been linked to the existence of genistein, a phenolic component found in soy-based diets. Genistein controlled strong anti-inflammatory actions through the blocking of different signaling pathways such as (PGs) Prostaglandins, pro-inflammatory (ROS) reactive oxygen species and cytokines, (iNOS) inducible nitric oxide synthase and (NF-κB) nuclear factor kappa-B. Furthermore, therapeutic effects of genistein have been conveyed in various pathological situations through modifying intracellular paths such as Akt, mTOR, PI3K, PPARγ, NF-κB, Nrf2 and AMPK. Genistein works as a chemotherapeutic drug against several cancers, primarily through modifying the cell cycle, apoptosis, and angiogenesis, as well as limiting metastasis. Genistein also exhibits a synergistic attitude with renowned anticancer medicines including adriamycin, tamoxifen and docetaxel indicating a possible role in grouping treatment. The study presents the recent data available on genistein’s beneficial effects against various forms of cancers.

Introduction

Plants are a rich source of supplementary metabolites, which are physiologically active, have beneficial impacts on the human body, and have therapeutic characteristics for treating a variety of disorders. Multiple plants have medical characteristics; for example, coffee drinking has been shown to lower the possibility of diabetes mellitus. An investigative study discovered that consuming Allium cepa lessens the incidence of diabetes mellitus by reducing blood sugar levels. Likewise, Zingiber officinale offers diabetes-protective properties as well as enough glycemic action. Hesperetin, a flavonol produced from citrus trees, has considerable anti-cancer and anti-cardiovascular activity. Apigenin, a natural flavonoid present in a variety of vegetables and fruits, exhibits strong anti-cancer activity by regulating critical carcinogenic pathways and cell signaling. Various secondary metabolites of plants, particularly morin and resveratrol, have important therapeutic benefits. Genistein is one of the most significant phytoestrogens, with amazing biological impacts on the well-being of humans.[Citation1]

Genistein (4′, 5, 7-trihydroxyisoflavone) is a flavonoid that has been discovered naturally in a range of dietary items. Genistein is a flavonoid that was first extracted from the dyer’s broom Genista tinctoria in 1899. Prunitol and genistein are structurally similar. The synthetic chemistry of genistein was completed in 1928. Genistein has been identified in several types of glycosidic combinations in soy products. Genistein is a phytoestrogen because it has estrogenic actions in the human body. Genistein is structurally identical to 17-estradiol, may disrupt with hormone action through competing processes for receptor site, and has estrogenic impacts; yet, it has non-estrogen receptor-mediated cellular responses. Genistein also has powerful antioxidant effects. Genistein is an important regulator of several signaling pathways at the transcription and translation levels.[Citation2] Chemical structure of Genistein is presented in .

Figure 1. Structure of Genistein.

Figure 1. Structure of Genistein.

Soybeans and soybean derivatives such as soy milk, tofu, soy flour, miso, and natto are high in genistein. Though, a little amount of genistein also discovered in various pulses and fruits such as chickpeas, red clover, alfalfa, currants and resins etc. Genistein contains anti-cancer and anti-oxidant properties, and it has been discovered that genistein promotes glutathione peroxidase activation and limits DNA oxidation in human prostate carcinomas. Genistein is a chemoprotective agent and considered as a treatment therapy for several kinds of cancer including breast tumor, prostate cancer, colorectal cancer, liver tumor, and gastric cancer. Platelet activation and aggregation are critical in the pathophysiology and progression of vascular diseases. Flavonoids derived from plants have a possibility to treat cardiovascular disorders, and genistein has significant cardioprotective prospective. According to one study, genistein reduces inflammatory processes and provides cardioprotection in opposition to myocardial ischemia-reperfusion injury. Genistein contains potent biological properties that protect against breast cancer and cardiovascular diseases.[Citation3]

Genistein is also employed for managing effects of menopause such as night sweats, hot flashes, and insomnia, which are frequently caused by a lack of estrogen. Numerous investigations have also found that genistein has anti-diabetic properties because it modulates the breakdown of glucose. Genistein lowers the levels of visfatin in the blood, which aids to weight loss. According to the findings, dietary flavonoids like as genistein, quercetin, apigenin, and kaempferol are projected to have a greater degree of antioxidant activity than more traditional antioxidants such as vitamin E and vitamin C. This finding was made possible by the fact that flavonoids are found in plants. This forecast relies on the assumption that foods contain flavonoids in enough quantities. The flavonoid known as genistein may be found in a variety of plant-based meals, including fruits and vegetables. It is possible that some of the foods we eat include this molecule, which has qualities that make it effective against inflammation, oxidative damage, bacteria, and viruses.

Number of flavonoids are being studied for their possible anticancer effects. According to the results of a study, kaempferol was used to combat tumor. Kaempferol effectively suppressed the growth of triple-negative breast cancer (TNBC) MDA-MB-231 cells. Observations exhibited kaempferol inhibitory effects toward cellular growth and was resilient in MDA MB-231 cell lines compared to estrogen receptor-positive BT-474 cell lines.

As suggested by published research, genistein can boost the effectiveness of other therapies or work synergistically with other drugs to produce synergistic effects that are beneficial to the patient. Apoptosis can be caused by genistein because it causes an increase in the expression of caspase-3, Bax, and p21 while simultaneously causing a decrease in the expression of Akt, cyclin-B1, PLK-1, cyclin-A, CDK-2, CDC-2, and Bcl-2. In addition to this, it has been established that it raises the total amount of STAT3 protein that is degraded in liver cancer cells while simultaneously lowering the total amount of STAT3 activation.[Citation4] Genistein may be particularly effective to induce apoptosis and promote cancer cell death. In this article, in vivo and in vitro studies are discussed on health perspectives of genistein as an anti-cancer agent.

Perspective of genistein in reference to the prevention of cancers

Breast cancer

The effects of genistein, from soy, were investigated using BJ skin fibroblasts and MCF-7 tumor cells cultured in vitro. Different doses of genistein were used. According to the findings, genistein was able to suppress the growth of both types of cells. When it comes to Genistein’s culpability, there is no longer any room for debate. The actions of genistein can be triggered in a variety of cells, including normal dermal BJ fibroblasts and malignant MCF-7 cells. The amount of isoflavonoid that was found is directly related to the intensity of the impact that was seen. Regardless of the total amount of time that MCF-7 cells are exposed to the substance, a significant decrease in the number of cancer cells is observed after exposure to high concentrations of genistein for a period of 48 hours. There is a correlation between taking between ten and twenty milligrams of genistein and an increase in the percentage of slender fibroblasts. In spite of this, the genistein’s toxicity to fibroblasts is not noticed until 48 hours after the cells have been exposed to doses of 50 M or higher. The research shows that genistein does not come without any possible side effects, despite the fact that it has a strong potential to be used in the treatment of skin problems, incisions, and surgical scars in women who are undergoing or recovering from breast cancer therapy. Isoflavonoids have a potential therapeutic application, but more study has to be done on the cellular mechanisms that underlie this application. During this examination, it is extremely important to take into account both the genetic and the immunological elements of the illness. In addition, in vivo models need to be introduced into the research process in order to eliminate the possibility of the medication producing unintended side effects.[Citation5]

It has been found that the development of hormone-independent growth features in MCF-7 tumors is connected with the long-term treatment of modest dosages of the compound known as genistein. This was detected in cancers that had significantly high levels of HER2 expression. In spite of the fact that endocrine resistance has been related to persistent, low-level genistein-induced HER2 expression in human breast cancer, it is still unknown if this is the true cause of hormone resistance. Overexpression of HER2 has not been proven to be associated with resistance, contrary to the widespread perception that this is the case. It was feasible that different impacts on the expression of HER2 may be observed depending on the length of time that genistein therapy was received. A brief exposure to low concentrations of genistein mimics the actions of estrogen and suppresses the growth of HER2 in breast cancer cells. The sustained activation of EZH2 at Ser21 by ERK1/2 caused the quantity of lysine 27 trimethylation to decrease after extended exposure to low levels. This was the cause of the reduction. This activation was a direct consequence of the low-level exposure that had been received. A feedback loop between IL-8/HER2 and ERK1/2/EZH2/IL-6 was developed when H3K27me3 levels fell, IL-8 and IL-6 expression levels rose, and HER2 stages continued to expand. This loop was established due to the fact that H3K27me3 levels decreased. Tamoxifen-treated patients should make an effort to keep track of the amount of genistein that they take in during the course of their meals, since this is something that should be monitored. Breast cancer patients who use soy supplements in modest quantities for a lengthy period of time may be putting their health in jeopardy.[Citation6]

The metastatic breast tumors typically cause significant bone loss. As a direct consequence of this, the patient’s prognosis in general is becoming direr. The palliative character of the anti-resorptive drugs that are now available for the treatment of skeletal metastases means that they are often only offered after the appearance of secondary malignancies. As a consequence of this, researchers investigated the effects of genistein and coumestrol, both separately and in conjunction with one another, on the quantity of breast cancer cells, the expression of mediators of preferred skeletal metastasis, the attachment of bone matrix, and the formation of tumor-induced osteoclasts. An estrogen receptor-dependent experiment revealed that the combination of genistein and coumestrol resulted in a statistically significant decrease in the number of viable cells. The expression of the genes that are responsible for osteolysis (PTHrP and TNF-), the transition from epithelial to mesenchymal (snail), and bone attachment (CXCR4 and integrin V) was dramatically reduced when genistein and coumestrol were present. Because of the effects of genistein and coumestrol, tumors were unable to create osteoclasts and RANKL, which led to a significant decrease in breast cancer cell adherence to bone matrix. This was the outcome of the combination of the two compounds. In a study the researchers found that phytoestrogens disrupt important connections between tumors and bones that produce and sustain the formation of skeletal metastatic tissue. In addition to the discovery that breast cancer cells had a lower chance of surviving when exposed to phytoestrogens, another finding was published. Because of these connections, metastatic tissue can form and proliferate in the skeleton. It is possible that the effects of genistein on cell proliferation, angiogenesis, survival, and apoptosis are all factors that contribute to its capacity to suppress the growth of breast cancer. In specifically, the illness that is often referred to as breast cancer is used here as an example. According to a number of studies, the use of genistein lowers the chance of developing breast cancer; nevertheless, clinical trials have not consistently showed the same level of efficacy. It has also been demonstrated that higher amounts of genistein are associated with an increased likelihood of getting breast cancer.[Citation7]

It is of the utmost significance that genistein and E2 (17-estradiol) be able to engage in a two-way conversation with the PR and ER estrogen receptors. In comparison, the affinity that E2 possesses for both PR and ER was exactly same, but the affinity of genistein possesses for PR is considerably weaker than the affinity that it possesses for ER. After conducting an exhaustive analysis of the pertinent research that had been previously published, it was determined that genistein breast cancer activities were significantly influenced by a number of factors, including patient age, the presence or absence of menopause, transactivators and regulators of ER and PR expression and activities, normal intake mode, ER expression pattern and ratio, and genistein breast cancer activities.[Citation8]

Genistein appears to prevent triple-negative breast malignancies, which are cancers of the breast that do not have the human epidermal growth factor receptors HER2, ER, or PR. The lack of these receptors is what distinguishes this subtype of breast cancer from other types. The human epidermal growth factor receptors HER2, PR, and ER are not present in breast cancers that are referred to as triple-negative breast tumors (TNBCs). Genistein has shown promise as a therapy for cancer by demonstrating that it is feasible to inhibit the spread of cancer cells through a variety of different pathways. This is a crucial piece of new information. Because of these pathways, the Hedgehog-Gli1 signaling can be altered, tyrosine kinases can be blocked, the epigenetic activity can be changed, the MEK and Akt signaling pathways can be activated, and the cell cycle may be stopped. In addition, Genistein’s study has shown that it is possible to halt the spread of cancer cells to a variety of organs and other parts of the body, which is a very promising development.[Citation9]

The capacity of genistein to disrupt the cell cycle as well as its anti-metastatic qualities allow it to limit the growth of breast cancer cells. As a consequence of this, genistein is now in a position to slow or arrest the progression of the disease. On the other hand, the observed effects were depending on the dosage as well as the amount of exposure. In addition, there is some evidence that genistein is capable of preventing the growth and spread of breast cancer cells. There is some evidence to suggest that greater dosages of genistein may have the ability to suppress the growth and proliferation of cells. The levels of HIF-1, NF- κB, and VEGF, as well as an increase in the level of the tumor inhibitor p21, are among the signaling pathways that have been altered as a result of these adjustments. In addition, the levels of these signaling pathways have been affected. Among the other signaling pathways that have been disrupted are the ones that control the expansion of cancer cells. The researchers will have a deeper understanding of the molecular mechanisms that contribute to breast cancer, notably the roles that NF-κB and HIF-1 play as transcription factors.[Citation10]

As reported in another study, genistein was able to prevent the formation of cyclin B1 in BRCA1-mutated breast carcinomas. This resulted in an arrest of the G2/M phase of the cell cycle, which eventually resulted in the death of cancerous cells. In addition, breast cancer cells that have the BRCA1 mutation create more reactive oxygen species (ROS) per cell than normal breast cancer cells do. Normal breast cancer cells produce less ROS. Genistein therapy boosted Nrf2 expression, which led to reduced levels of reactive oxygen species (ROS).[Citation11] Researchers found that genistein is effective against triple-negative breast cancer by using two orthotopic preclinical patient-derived xenograft mice models, BCM-3204 and TM00091. Both of these models were used to conduct the research. Both of the PDX models were given the genistein medication, and as a result, the growth of the tumor was stopped. Through the use of transcriptomics, data demonstrated that the consumption of genistein affected the expression of a large number of genes known to have a role in the control of tumor growth in BCM-3204 PDX-derived TNBC tumors. An ovariectomized nude mouse breast cancer xenograft model was given in vitro administration of genistein and cisplatin at a dose of 5 mg kg-1d-1 of each substance. In xenograft tumor tissues, the average concentration of total genistein metabolite equivalent reached 0.2729 nmol g-1 of dry weight. When this threshold was reached, the effect that genistein had on cisplatin capacity to protect against cancer was significant.[Citation12]

Genistein was able to prevent cisplatin from suppressing cell growth and causing apoptosis by working on the basic processes.[Citation13] It was found that genistein has the ability to suppress the influence that hypoxia-inducible factor-1 (HIF-1) has on breast cancer. The efficacy of centchroman (CC) to treat breast cancer in a complementary way was enhanced by the food-based isoflavone genistein. This was achieved by preventing the passage of cells into the G2/M phase, which ultimately resulted in the death of the cells in a way that was dependent on ROS. Inducing mitochondrial malfunction, caspase-3, −7, and −9 activation, and PARP cleavage are all side effects of the combination of centchroman and genistein. Apoptosis is promoted as a result of the combination’s induction of a considerable drop in the phosphorylation of PI3K, Akt, and NF-κB. In addition, the combination was more effective than either centchroman or genistein alone in preventing the growth of tumors in the mouse 4T1 breast cancer model.[Citation14]

A similar research was conducted by employing the G0/G1MCF–7/ER-1 and MDA-MB-231/ER-1 cell sub-lines shown that greater concentrations (10–6 mol/l, 10–5 mol/l, and 10–4 mol/l) decreased cell proliferation and reduced their capacity to multiply after 48 hours.[Citation15]

Pancreatic cancer

The word “genistein” refers to an isoflavone that occurs naturally and has the capacity to prevent the growth of cancerous tumors. However, the fact that genistein is inefficient as a standalone treatment against cancer cells is a considerable barrier to advancement in this field. A research was conducted with the goals of first gaining an understanding of how pancreatic cancer (PC) cells developed genistein resistance and then using that information to build novel therapeutic approaches. In the beginning, it was explored that the major chemicals and signaling pathways linked with genistein resistance in PC cells by using bioinformatics tools. It was expected that genistein-resistant PC cells would exhibit low amounts of DEPTOR, which is an abbreviation for “the DEP domain-containing MTOR-interacting protein.” DEPTOR is a protein that normally disrupts the signaling pathway that is controlled by the mammalian target of rapamycin (mTOR). In later research and development, two unique kinds of genistein-resistant PC cells known as Panc-1 and PaCa were produced. Altering the expression of DEPTOR and treating the cells with the mTOR antagonist everolimus (ELM) were done in order to assess the relative contributions of DEPTOR and the mTOR antagonist to genistein resistance in the cells.[Citation16]

Research on cell apoptosis was carried out utilizing tumors that were produced from xenografts of mouse tissue in both laboratory settings (in vitro and in vivo). Both types of research made use of the cancerous conditions under study. Researchers were successful in identifying the upstream regulator of DEPTOR by making use of bioinformatics. In prostate cancer cells that were resistant to genistein treatment, the activation of the PI3K/AKT/mTOR signaling pathway was shown to be mediated by the downregulation of DEPTOR, according to the findings of a bioinformatics research. Because of the overexpression of DEPTOR, there was an increase in the synthesis of LDH as well as the activity of caspase-3, which ultimately led to the demise of PC cells. In addition, both the phosphorylation of mTOR and the amount of genistein necessary to inhibit it by 50% (known as the IC50) were shown to be lowered. Following treatment with ELM, there was a rise in both the capacity of genistein to rid rats of tumors and the sensitivity of PC cells to the effects of genistein when tested in vitro. Genistein inhibited the transcriptional activity of the ETS transcription factor ELK1 (ELK1), which resulted in a detrimental impact on the transcription of the DEPTOR gene. The overexpression of ELK1 prevented the DEPTOR gene from being translated, which led to a reduction in the cells’ sensitivity to genistein. In addition to this, it stopped ELM from working in PC cells in a way that would have made the cells more sensitive to genistein if it had been allowed to do so.[Citation17]

According to the findings of another research, ELK1 suppresses transcription of DEPTOR, which in turn phosphorylates mTOR and makes PC cells resistant to drugs. These findings were published in two separate studies. These findings were reported in two different investigators’ reports. These findings were independently discovered by two different researches and publicized. The administration of genistein to PC cells led to a reduction in the amount of the microRNA miR-223, as well as an increase in the amount of one of miR-223’s targets, Fbw7.[Citation18]

It was discovered that suppressing miR-223 led to cells developing more slowly and committing suicide. It was found that the substance genistein, which is now being explored as a potential therapy for pancreatic cancer, greatly lowered the production of the microRNA known as miR-27a in pancreatic cancer cells. This finding was made possible by the fact that pancreatic cancer cells were used to make the discovery. In pancreatic cancer cells, inhibiting miR-27a led to a reduction in cell proliferation as well as invasion.[Citation19]

Apoptosis, or the programmed death of cells, was also triggered as a result of this research. It was shown that genistein is able to decrease the expression of miR-223 in pancreatic cancer cells that have previously developed resistance to gemcitabine (GR). In this work, the reversibility of EMT as well as the synergistic effects of genistein and a miR-223 inhibitor were studied with regard to their impact on the proliferation, migration, and invasion of GR pancreatic cancer cells. In combination with genistein, the miR-223 inhibitor was provided to the subjects. Researchers made the discovery that genistein and a miR-223 inhibitor can lower the motility and invasion of cells while simultaneously boosting the sensitivity of GR cells to gemcitabine. In addition, genistein and a miR-223 inhibitor had the capacity to revert the EMT traits that are shown in GR cells. This was found to be the case. In human pancreatic cancer cells originating from the MIA PaCa-2 line and xenografts grown from these cells, the addition of genistein boosted the amount of apoptosis and autophagy that was triggered by 5-FU. Both of these cell types showed evidence of the influence that genistein had on them. The levels of the proteins beclin-1 and B-cell lymphoma 2 (Bcl2) altered as a consequence of the enhanced activity of autophagy. Research on animals contributed new pieces of evidence that helped bolster the reliability of these conclusions.[Citation20] Similarly, the volume of the final xenograft tumor was significantly decreased when treated with a combination of 5-fluorouracil and genistein. When compared to therapy with just 5-fluorouracil, this result was significantly better. Inducing autophagy and apoptosis at the same time turned out to be the most effective strategy for achieving success in this attempt.[Citation21]

Prostate cancer

The vast majority of prostate cancer (PCa) patients will, at some point, develop resistance to the medication they are receiving, leaving them with few potential curative choices. The most abundant source of bioactive molecules, natural products, might serve as a library for screening and finding viable candidates in the hunt for revolutionary prostate cancer treatment regimens. This is especially true for treatment regimens that are successful against prostate cancer that has developed resistant to previous treatments. This is especially true when taking into consideration the fact that natural goods are the most common source of bioactive chemicals. Flavones have been used in clinical and laboratory studies on prostate cancer for quite some time now. Examples of these flavones are apigenin and genistein. The research carried out over the course of the last 10 years that resulted in the discovery of apigenin and genistein as potential therapies for prostate cancer is the primary topic of this overview. Both of these isomers have been found to target the same signaling pathways, despite the fact that they function in PCa cells in quite different ways. This is true despite the fact that both of their structures are highly similar to one another. Taking into consideration the fact that research and testing are now being carried out on a variety of possible combinations, it would appear that genistein has the highest chance of getting authorized. It has been shown that these two flavones have anticancer action in both in vitro and animal experiments; however, clinical research is still necessary to prove their effectiveness. During the course of a mouse model of an in vitro xenograft tumor, different doses of genistein were given to the 22RV1, VCaP, and RWPE-1 cells for a period of 48 hours. The concentrations ranged anywhere from 0 to 100 mol/L in this experiment. The information that was obtained from this study has the potential to be extremely valuable for the creation of brand-new drugs and the treatment of prostate cancer using apigenin and genistein. Both of these treatments are now being investigated.[Citation22]

Apoptosis was brought on by genistein to an uptick in caspase-3 gene expression, as well as an increase in intracellular protein levels and enzyme activity. In addition, the cell growth inhibitor genistein was able to reduce the amount of p38MAPK protein in the body as well as the expression of the p38MAPK gene. The capacity of PC3 cells to participate in metastatic spread was dramatically reduced after the activity of MMP2 was inhibited.[Citation23]

Researchers have also reported that the treatment of PC3 cells with genistein at dosages of 120, 240, and 480 M lowered the synthesis of cellular NO, increased the production of catalase and glutathione, and triggered apoptosis in PC3 cells through a mechanism that was independent of mitochondria.[Citation24] The concentration of genistein that was used to treat PC3 and DU145 cells was 30 microM, whereas the concentration used to treat AG1024 cells was 10 microM. This combination had a synergistic impact on the cells, which revealed itself in a number of ways, including a major slowing of cell development and an acceleration of cell death. Synergistic effects are effects that occur when two or more factors work together to produce a greater effect than either one alone. While treatment with genistein inhibited both the HRR route and the NHEJ pathway by reducing the expression of Rad51 and Ku70, treatment with AG1024 solely inhibited the NHEJ pathway by inactivating Ku70. Genistein therapy was more effective than AG1024 treatment.[Citation25]

The effectiveness of the AG1024 therapy was inferior to that of the genistein therapy. In this sense, the results of the genistein therapy were superior to those of the AG1024 therapy. Additionally, the radiosensitization of prostate cancer cells with genistein and AG1024 was more successful than if each treatment had been utilized on its own. This was achieved by putting a stop to the generation of new cells, speeding up the process by which existing cells died, and blocking the HRR and NHEJ pathways. During the in vivo trials, animals were given genistein and AG1024, which resulted in a considerable reduction in the number of tumors identified in the treated animals in comparison to the number of tumors found in the untreated animals. The researchers found that the combination of genistein (30 mM) and AG1024 (10 mM) had a synergistic effect on the radiosensitivity of PCa cells by suppressing the NHEJ and HRR pathways.[Citation26]

A therapy with genistein-coupled polysaccharide (GCP) lowered the viability and proliferation of prostate cancer (CaP) cell lines LNCaP and PC-346C while concurrently improving their capability to block androgen receptor (AR) activity. Genistein-coupled polysaccharide (GCP) was administered. It is probable that it will suppress the expression of 3HSD, 17HSD, CYP17A, SRB1, and StAR, as well as other enzymes involved in the generation of intracrine androgens. Additionally, it may decrease the expression of additional enzymes. As a direct consequence of this, the amount of testosterone found within the body has the potential to increase by a factor of three. Three active GCP fractions were found after using a bioassay in combination with reverse-phase high-performance liquid chromatography fractionation. The following NMR and LC-MS studies of fraction 40, which had the highest degree of activity, revealed that genistein was the principal active component of GCP. This conclusion was reached based on the results of the fraction. The investigation of the portion that had the highest amount of activity brought up this realization. In addition, it was proven that genistein was the component of GCP that was responsible for its anti-proliferative, pro-apoptotic, and anti-AR characteristics. The effect of the vehicle on these biological processes was at least double that of the combination of GCP, fraction 40, and genistein. This is in compared to the control group.[Citation27]

Genistein did not have any impact on the expression of the other enzymes that are involved in the intracrine androgen synthesis pathway. The 3-HSD enzyme, the SRB1 enzyme, and the StAR enzyme are the other three. In comparison to GCP, genistein has the effect of significantly inhibiting the expression of CYP17A and 17-HSD. This decrease was a factor of 1.5 to 1, down from 1. The shift in expression that was seen was exactly this. The evolution and spread of prostate cancer has been connected to increased production of the proteins COX-2 and Glut-1, which has been linked, in turn, to modifications in signaling pathways that play a vital role in the course of the illness. The combination of plumbagin and genistein in the form of 100-nm nanoliposomes was able to limit the growth of prostate cancer cells in a synergistic manner. This course of action was adopted in order to accomplish the desired result. These chemicals inhibit Glut-1 receptors in addition to promoting the generation of reactive oxygen species (ROS), reducing the concentration of cellular GSH, inhibiting the synthesis of COX-2, and inhibiting the concentration of cellular GSH. It has been demonstrated that the combination of plumbagin and genistein can suppress the development of xenograft prostate cancers by roughly 80% without appreciably increasing the toxicity of either agent.[Citation28] Nanoliposomes that include plumbagins and genistein have the ability to block the PI3K/AKT3 signaling pathway and reduce the amount of Glut-1 transporters that are present in the body, hence preventing the development of cancer.[Citation29]

Liver cancer

Diethylnitrosamine, often known as DEN, was given orally to mice that were two weeks old in an effort to bring on the development of hepatocellular carcinoma. After then, injections of genistein were given to the mice once every two weeks for an additional five months, beginning when the mice were 40 weeks old and continuing until they were 62 weeks old. When genistein was eaten by food, there was a lower incidence of HCC, and the disease took longer to develop. When exposed to genistein, untreated Hep3B cells also exhibited apoptosis, which was evaluated by higher levels of phospho-AMPK in whole liver extracts.[Citation30]

After administering genistein, researchers found that there was an improvement in the rate of mitochondrial respiration as well as a reduction in the pro-inflammatory response. In addition, there was a reduction in liver damage, and it was demonstrated that phospho-AMPK was the factor responsible for the decrease in proinflammatory reactions. Researchers found that the presence of genistein led to an increase in microRNA that was produced during studies of hepatocellular carcinoma (HCC) conducted both in vitro and in vivo. These microRNAs have been shown to be capable of inhibiting not just the EIF5A2/PI3K/Akt pathway but also cell proliferation, migration, invasion, metastasis, and stemness.[Citation31]

After being treated with 5 M genistein, hepatocellular carcinoma (HCC) cells exhibit a significant increase in their radiosensitivity. An increase in DNA damage, chromosomal abnormalities, cell cycle arrest in the G2/M phase, and apoptosis are the effects of a therapy that was applied to HCC cells. This course of action is performed in order to accomplish this goal. The effects of X-rays on phospho-Bad (Ser136) are made worse by the fact that genistein raises levels of phospho-Chk2 (Thr68), phospho-ATM (Ser1981), and H2AX. In addition to this, genistein contributes to an increase in the concentration of phospho-ATM. In genistein-induced radiosensitivity, the expression of POU6F and CCNE2 is reduced, whilst the expression of FBXO32 and cyclin B1 is elevated.

According to the findings of a similar study, the presence of genistein in hepatocellular carcinoma PLC/PRF5 and HepG2 cell lines led to an increase in the level of gene expression for DNMT1, DNMT3a, and DNMTb. In addition, genistein encourages the process of apoptosis and blocks both the formation and multiplication of cells. In male Sprague-Dawley rats with hepatocellular cancer, versican, protein kinase C (PKC), and extracellular signal-regulated kinase 1 (ERK-1) were overexpressed. On the other hand, superoxide dismutase (SOD) and glutathione levels were reduced. These rats exhibited increased levels of oxidative stress, platelet-derived growth factor (PDGF), malondialdehyde (MDA), and platelet-derived growth factor (PDGF). These changes in the test individuals may be reverted to their original state by administering either 25 mg/kg or 75 mg/kg of genistein. HCC cells were shown to be more prone to mitochondrial apoptosis and aerobic glycolysis in an in vivo research that used subcutaneous xenograft mice as disease models. The study found that HCC cells were more vulnerable to these processes when the mice were given genistein at dosages of 25 or 75 mg/kg. Apoptosis was also triggered by genistein, which had the additional effect of inhibiting aerobic glycolysis in mitochondria.[Citation32]

In an interesting study, researchers observed that, the expression of E-cadherin and catenin was raised in HepG2, SMMC-7721, and Bel-7402 cells by genistein (dose dependent), whereas N-cadherin and vimentin were brought to lower levels. As a direct result of this alteration, there was an adjustment made to the amounts of both mRNA and protein. The epithelial-mesenchymal transition (EMT) that was promoted by TGF was stopped dead in its tracks by the contemporaneous injection of genistein. The expression of the mRNA and proteins Nuclear factor of activated T cells 1 (NFAT1), Abca3, Autotaxin, CD154, and Cox-2 was reduced by genistein in HepG2 cells. Ionomycin and phorbol 12-myristate 13-acetate (PMA) both boosted the activity of NFAT1, which brought about a reduction in the levels of E-cadherin and catenin and an increase in the levels of N-cadherin and vimentin. The presence of PMA and ionomycin together was able to counteract the effects that genistein had on the ability of HepG2 cells to migrate. The capacity of genistein to reverse the EMT, which was related with a drop in NFAT1 levels, resulted in a lower frequency of intrahepatic metastases when tested in trials conducted in living animals. Due to genistein’s capacity to suppress the EMT process, which was predominantly mediated by NFAT1, hepatocellular carcinoma cells were unable to migrate after being exposed to the compound.[Citation33]

The capability of genistein to revert EMT may give insight into the potential processes underpinning its success in the treatment of liver cancer. When human HepG2/C3A and HT29 colon adenocarcinoma cells were treated with 5 and 50 M genistein, the levels of the mRNA that encodes CYP1A1 and CYP1B1 were increased. On the other hand, it was found that the levels of mRNA that code for the enzymes CYP2D6, CYP26A1, and CYP26B1 were decreased. The amount of CYP P450 gene expression that was present in HT29 cells did not alter. Genistein acts as an anticancer agent in SNU-449 cells by preventing the growth of these cells, and its effectiveness is proportional to the concentration at which it is present. In addition, genistein caused an increase in the formation of intracellular ROS as well as the activation of c-Jun N-terminal kinases, p38, and apoptotic signal-regulating kinase 1, and it encouraged thioredoxin-1 levels to rise. All of these effects occurred within the cell.[Citation34]

Pretreatment with JNK and p38 inhibitors like SP600125 and SB203580 may reduce the amount of cell death that is initiated by genistein. Genistein was able to suppress the capacity of Hepa1–6 cells to proliferate while concurrently inducing apoptosis in a way that was dependent on concentration as well as the passage of time. After 24 hours of genistein administration, the greatest degree of cell growth inhibition reached was 52% (P 0.01), and the concentration at which 50% of the maximal inhibitory effect occurred (IC50) was 20 M. These results indicate that genistein inhibits cell growth. According to these studies, genistein has the potential to decrease the growth of cells. After 24, 48, and 72 hours, a concentration of 20 M genistein produced 35, 42, and 65% apoptotic cells, respectively.[Citation35]

When genistein was supplied to the PLC/PRF5 cell line of hepatocellular carcinoma (HCC) at various doses (1, 5, 10, 25, 50, 75, and 100 M/L) and durations (24, 48, and 72 h), there was a substantial reduction in the proliferation of liver cancer cells. It was demonstrated that the effectiveness of this suppression changed both with the passage of time and quantity. During the course of the experiment, there was a total of 53, 48, and 47% of viable cells found in the various treatment groups for the 25 M genistein concentration, respectively. It was discovered that a concentration of genistein equal to 25 mM was ideal for causing time-dependent apoptosis in a manner that was both powerful and effective. These findings were acquired using a technique called flow cytometry. During the course of the experiment, the percentage of test participants’ bodies that included apoptotic cells went from 44% to 56% to 60%, as reported by.[Citation36]

Genistein was shown to enhance ER levels while simultaneously decreasing DNMT1 gene expression in research conducted. A large rise in apoptosis was also caused by genistein, which led to a reduction in the viability of cells. Genistein significantly facilitated the process of apoptosis by both inducing apoptosis in HepG2 cell lines and reducing cell growth.[Citation37]

Gastric cancer

Establishment of cancer stem cells and the manifestation of the CD44 surface biomarker, which is distinctive to malignant stem cells, are both necessary steps in the process of oncogenesis. Gli1, the protein that is responsible for starting the Hedgehog signaling pathway, has been shown to play a role in all three of these processes. According to the evidence that is currently available, it has been demonstrated that genistein can reduce the expression of the genes Gli1 and CD44. Genistein has the potential to change a number of the biological properties of cancer stem cells due to its capacity to block Gli1-related signaling pathways. Throughout the study, the expression of CD44 provided the ability to single out specific cells that were generated from the human gastric cancer cell line MKN45. CD44+ cells that produced spherical colonies and had features associated with cancer stem cells expressed characteristics associated with stem cells. These cells exhibited properties that are typical of cancer progenitor cells. The CD44 (+) cells also had a larger number of transcripts implicated in the Sonic Hedgehog (Shh) signaling pathway than the CD44 (-) cells did. This was the case when compared to the CD44 (-) cells. In CD44+ cancer stem-like cells, the expression of CD44 mRNA, as well as Gli1 and protein, was decreased after treatment with genistein. Additionally, it has been proven that genistein is capable of inhibiting the expression of other stem cell markers. The ability of genistein to suppress Gli1 expression was tested using Gli1 siRNA, and the findings showed that the presence of genistein considerably improved the inhibition of CD44+ cell migration. Genistein was also shown to be efficient in reducing Gli2 expression. In conclusion, genistein reduces the expression of the Gli1 gene, which in turn decreases the ability of gastric cancer cells to behave as cancer stem cells. These cells are found in the intestinal tract.[Citation38]

Genistein inhibits the capacity of cells to penetrate, which is essential for the development of tumors and the spread of cancer. Genistein was found to be effective in preventing both of these outcomes. In the xenograft model, 15 M genistein was able to block stem cell-like features, diminish chemoresistance, decrease ABCG2 expression, and decrease ERK 1/2 activity in gastric cancer stem cells. These findings are the result of two distinct experiments, both of which were carried out independently.[Citation39,Citation40]

Colon cancer

The human cancer cell lines SW620 and SW480, as well as the nonmalignant carcinogenic cell line HaCaT, are all dose-dependently inhibited in vitro in terms of cell viability, proliferative activity, and the accelerated production of reactive oxygen species. This finding was made possible by the fact that all three cell lines were exposed to the same amount of the inhibitor. In addition to the noncancerous cell line HaCaT, this was shown in the human cancer cell lines SW480 and SW620.[Citation41]

The responses of human colon cancer cells (SW480 and SW620) and nonmalignant cell lines (CHO-K1 and HaCaT) to genistein were comparable in research study. It was discovered that genistein activated p53 and increased the levels of caspase 3/8 and tumor inhibitor protein (Tp53) in the colon cancer cell lines HT29 and SW620. These cell lines were used to study primary and metastatic forms of colon cancer. An increase in H2O2 levels and the development of filopodia were seen as a consequence of genistein’s ability to block the passage of SW620 cells through the G2/M phase of the cell cycle. In HT29 cells, genistein led to an increase in the number of stress fibers, as well as an increase in the generation of hydrogen peroxide, which ultimately resulted in the death of the cells. When compared to HT29 cells, SW620 cells demonstrated a reduced rate of mitochondrial production. Both malignant cell lines showed an increase in the expression of genes linked with inflammation.[Citation42]

There was a dose-dependent relationship between the inhibitory effects of genistein on the survival of HT29 colon cancer cells. This transpired as a direct consequence of an increase in NF-κB that was transferred into the nucleus of SW620 cells. After 72 hours of treatment with 10, 20, and 60 mol/l of genistein, there was a dose-dependent increase in the quantity of Wnt inhibitory factor 1 (WIF1) demethylation. In addition, treatment with genistein decreased the capacity of cells to migrate to new sites and invade nearby cells in a dose-dependent manner. This was true for both of these processes. During the research, reverse transcription, quantitative polymerase chain reaction, and western blotting were utilized in order to ascertain the amounts of mRNA and protein expression that were connected with invasion and migration.[Citation43]

After receiving an injection of genistein, there was a large rise in the levels of expression of metalloproteinase inhibitor 1 and E-cadherin, while there was a significant decrease in the levels of expression of matrix metalloproteinase MMP2 and MMP9. The administration of genistein resulted in a reduction in the expression levels of the cyclin D1 protein, as well as cMyc, which is a proto-oncogene protein, catenin, and pathway-associated proteins that are tied to the Wnt 1/catenin pathway. When compared to the control group, cells that had been transfected with WIF1 siRNA or cells that had been transfected with WIF1 siRNA and treated with genistein displayed a considerably higher potential for invasion and migration. This was the case regardless of whether or not the cells had also been treated with genistein. HT-29 cells that were treated with genistein and daidzein (DAI) underwent an increase in vimentin levels, activation of apoptosis, suppression of lipid droplet formation (LD), and downregulation of expression of Perilipin-1, ADRP, and Tip-47 family proteins. These events took place one after another as a result of this phenomenon. As a result of taking both drugs, the mRNA expression of PPAR, FAs, and FABP as well as glycerol-3-phosphate acyltransferase (GPAT3) and microsomal TG transfer protein (MTTP) all increased; however, the mRNA expression of UCP2 considerably reduced.[Citation44]

The expression of FOXO3a was enhanced by both drugs, while the expression of PI3K was reduced at the same time. In HT29 human colon cancer cells, the inclusion of genistein into PEGylated Silica Nanoparticles (Genistein-PEG-SiHNM) (PEG-SiNPs) increased the anti-proliferative and antioxidant capabilities of genistein.[Citation45]

The endogenous antioxidant enzymes were changed by genistein at a concentration of 200 mol/L, and it increased H2O2 generation, which concurrently activated two separate cell death pathways (apoptosis and autophagy). This had the impact of reducing the amount of cell migration that occurred. It has been demonstrated that genistein has the ability to block the EMT of colon cancer cells (HT-29 cells) by raising the activity of E-cadherin and lowering the activity of N-cadherin, as well as by inactivating EMT-related proteins such as Snail2/slug, ZEB2, ZEB1, FOXC2, FOXC1, and TWIST1. Genistein has the capacity to increase the expression of Bax/Bcl-2 and caspase-3 in HT-29 cells while simultaneously decreasing the expression of notch-1, NF-κB, and p-NF-κB. These findings were published in the journal Cell Reports. The capability of genistein to activate various stages of the caspase-3 pathway, including transcriptional, protein-based, and enzymatic processes, was responsible for the induction of apoptosis, which is also known as programmed cell death. In addition, genistein was able to stop the production of the p38 MAPK gene, which in turn stopped the expansion of the HT29 cells. This was done by suppressing the synthesis of the gene as well as reducing the quantity of phosphorylated active protein that it encodes.[Citation46]

Genistein significantly reduced the activity of MMP2 in addition to significantly inhibiting the capacity of HT29 colon cancer cells to grow. Numerous forms of cancer are responsible for the aberrant production of nuclear factor NF-κB, which is a signaling pathway that governs the transcriptional start of genes that are necessary for the tight regulation of numerous cellular activities. The control of a number of different cellular activities is the responsibility of this route. The ability of the cancer cells to spread to other parts of the body was directly hindered as a consequence of this. It has been shown that substances that inhibit the NF-κB pathway can reduce the risk of cancer development and progression. Concerning colon cancer, a great number of questions have still to be answered. In this work, it was established that genistein may trigger programmed cell death (apoptosis) in human colon cancer cells LoVo and HT-29 by blocking the NF-κB pathway, lowering Bcl-2 levels, and boosting Bax levels. This was done in order to show that genistein can kill cancer cells.[Citation47]

Genistein might be a useful therapeutic drug in the treatment of colon cancer. The human colon cancer cells SW-480 and HCT-116 were treated with genistein, daidzein, and biochanin, all of which are forms of A. These forms of A suppressed the growth of the cells and promoted apoptosis. These effects were seen on both of these cell lines. The impact of genistein was noticeably more effective than those of the other two drugs, and both the quantity taken and the length of time it was taken for were essential components in achieving the intended outcome. In addition, genistein prevents cells from moving on to the next phase of their life cycle when they are in the G2/M phase. The activation of p21waf1/cip1 and GADD45 was linked to this phenomenon. This was also linked to a decrease in the expression of CDC2 and CDC25A.[Citation48]

According to an interesting study, being exposed to genistein caused the cell cycle to be arrested in the G2/M phase in a manner that was reliant on p53. This finding is quite interesting in its own right. Researchers discovered that genistein and daidzein both have the ability to slow down the multiplication of Caco-2 cells. Both It was revealed that treatment with genistein alters the distribution of cells throughout the cell cycle. This is accomplished by lowering the expression of Cyclin B1 and Serine/Threonine-Protein Kinase 2 (Chk2) and raising the percentage of cells that are in the G2/M phase. Ionizing radiation and genistein derivatives worked together to produce a range of effects, including a reduction in EGFR activation and a synergistic or additive inhibition of the proliferation of HCT116 cancer cells. These were just two of the impacts. At doses of 25 and 50–100 M, respectively, the isoflavones genistein and daidzein, which are found in soybeans, were able to suppress the growth of the human colon adenocarcinoma grade II cell line (HT-29). After that, reverse transcription-polymerase chain reaction, or RT-PCR was applied, to investigate how genistein and daidzein changed factors that govern cell proliferation and contribute to the advancement of cancer. It has been proven that a concentration of genistein equal to fifty milligrams per liter is sufficient to suppress the expression of the CTNNBIP1 gene. Expression of either APC (adenomatous polyposis coli) or survivin (BIRC5).[Citation49]

An in-vivo experiment was conducted using a genistein medication, and the results suggested that it reduced the formation of tumors in mice that had undergone tumor transplants. The acronym “APC” refers to “adenomatous polyposis coli.” These cases of colorectal cancer in humans were shown to connect with higher levels of transforming growth factor beta-1 (TGF-1) and expression of the long non-coding RNA (lncRNA) TTTY18, according to a study that was conducted on humans. The quantities of proteins and messenger RNAs for myeloid leukemia cell differentiation protein (MCL1), beta amyloid A4 protein (APP), and vascular endothelial growth factor receptor 2 (KDR) all reduced in a dose-dependent manner. In the cells, the expression levels of the genes Akt-Ser473, Ki-67, and serum and glucocorticoid-regulated kinase 1 (SGK1) rose after going through these alterations. A drop in the expression levels of TTTY18, SGK1, AktSer473, and p38 MAPKTyr323 was seen throughout the course of an inquiry. The treatment with genistein caused colorectal cancer cells to display characteristics including a reduction in cell viability, an increase in cell mortality, a reduction in Ki-67-positive cells, a reduction in cell motility, and a reduction in cell viability. The following in vivo investigation demonstrated that tumor-bearing null mice treated with genistein exhibited considerable weight loss coupled with decreases in tumor-derived TGF-1 and TTTY18 levels.[Citation50]

According to research, the number of intracellularly positive cells for SGK1, AktSer473, and p38 MAPKTyr323 was significantly reduced. Additionally, the mice had lower levels of tumor-derived TTTY18. Apoptosis was induced in HCT-116 cells and cell growth was suppressed following the addition of genistein to the media used to culture the cells. The expression of the mRNAs encoding Akt, SGK1, and miR-95 was reduced in HCT-116 cells, and the phosphorylation of Akt was prevented by the presence of genistein. This was the consequence of a decrease in the activity of the miR-95 gene. In investigations conducted in vivo, the administration of genistein dramatically slowed down the progression of a xenograft tumor that had been transplanted in a mouse.[Citation51]

Oral cancer

By observing a number of different occurrences, the researchers were able to determine how genistein inhibited 7, 12-dimethylbenz[a]anthracene (DMBA) from causing oral tumors to develop in the buccal pouches of hamsters. These occurrences included a delay of six weeks in the beginning of clinicopathological abnormalities as well as minor dysplastic epithelial changes. There have been no dysplastic modifications throughout the course of the last month and a half, which indicates that this circumstance is now lot simpler to manage than it was previously. According to flow cytometry, DMBA induced a significant spike in S-phase fragment (SPF) values in 47.22% of the hamsters; however, these values saw a remarkable drop after being treated with genistein.[Citation52]

In a study conducted by researchers it was found that genistein was able to inhibit the progression of head and neck cancer (HNC-TICs) cells, as well as stop the epithelial-mesenchymal transition (EMT), increase miR-34a, which resulted in apoptosis, stop self-renewal, migration, and invasion, slow ALDH1 activity, and inhibit RTCB. This was connected to the rate of DNA replication that occurred within aneuploid cells in addition to the DNA pattern that those cells contained. Apoptosis was induced in the JHU011 oral squamous cell carcinoma cell line at the same time that epigenetic transcriptional repression was eliminated. In order to accomplish this goal, it was necessary to simultaneously downregulate the polycomb group (PcG) proteins BMI 1 and EZH2, as well as downregulate their target enzymes UbH2AK119 and H3K27me3. Additionally, the nanoformulation of synthetic genistein has the potential to alter the expression of EZH2 by inhibiting 3PK and causing its breakdown by proteasomes.[Citation53]

Researchers observed that genistein prevented the growth, migration, and creation of rhVEGF-A tubes when they used an animal model of mouth cancer sentinel lymph nodes. This inference may be made based on the findings of the study. During the process of lymphangiogenesis, it prevented the activation of proteins that are important in downstream signaling, such as FAK, PI3K, AKT, p38, and ERK. This was accomplished by inhibiting the expression of the p38 mitogen-activated protein kinase gene. In addition to this, it was found that it suppressed the synthesis of the hormone known as hypoxia-inducible factor-1 (HIF-1).[Citation54]

Genistein was administered to three separate cell lines generated from tongue cancer at doses of 20, 50, and 100 M during 24, 48, and 72 hours respectively. According to the findings, treatments that included 20 M after 24 hours, 20 or 50 M after 48 h, and 50 M after 72 hours resulted in a 50% reduction in the system and proliferation. According to clinical tests that proved the treatment’s effectiveness, the concentration of genistein dropped in a manner that was proportional to the amount of time that patients were given the medication. As a direct consequence of this experiment, even the migration of cells was dramatically slowed down.[Citation55]

According to the data obtained from another similar study, genistein was responsible for the reduction in the levels of vitronectin, OCT4, and survivin. Both the CNE2 and HONE1 human nasopharyngeal cancer cell lines were unable to produce tumor spheres as well as they would have under normal circumstances. As a direct result of the treatment with genistein, there was a decrease in the total number of EpCAM+ cells, as well as in the expression of NCSC markers, the rate of cell proliferation, and the proportion of NCSCs that went through the process of apoptosis.[Citation56]

Genistein is capable of inhibiting the signaling of the Sonic Hedgehog (SHH) protein. This protein is critical to the maintenance of the NCSC’s overall health. On the other hand, the stimulatory effects of purmorphamine on SHH signaling had the effect of lowering the inhibitive effects of genistein on NCSCs. This was the case because purmorphamine stimulated SHH signaling. It was demonstrated that the ability of genistein to limit the growth of esophageal cancer (EsC) cells was influenced both by the amount of the substance that was administered and the amount of time over which it was administered.[Citation57]

According to the findings of a research that were dependent on concentration, genistein was shown to produce a significant rise in the rate of apoptosis and to arrest the cell cycle in the G0/G1 phase. In addition, the normal epithelial cells that line the esophageal lining are not harmed by even high dosages of genistein, which is used to line the esophagus. This is the case even when the genistein is used to line the esophagus. While treating EsC cells with genistein led to a substantial increase in the expression of genes involved in the process of cell death, it also led to a significant decrease in the expression of genes involved in the cell cycle, which was an interesting finding. On a cellular and molecular level, this took happened.[Citation58]

In addition to this, it was shown that genistein drastically decreased the expression of the epidermal growth factor receptor (EGFR), as well as the phosphorylation of the epidermal growth factor receptor’s downstream signaling pathways, which included STAT3, MDM2, Akt, and JAK1/2. Within the context of the signaling pathway, this had an impact on the JAK1/2-STAT3 pathway as well as the AKT/MDM2/p53 pathway. This resulted in STAT3 and MDM2 being unable to enter the nucleus, which was a direct consequence of the problem. When given to nudized rats that were getting xenografts, the supplement genistein caused a significant and dose-dependent slowdown in the growth of tumors. In laryngeal cancer cells that already had high levels of miR-1469, the addition of genistein caused the production of Mcl1 to be inhibited, which resulted in an increase in the death rate of the cells. This was demonstrated by the fact that the cells died at a faster pace. The same molecular mechanism anomalies were also observed in vivo. It was found that therapy with genistein resulted in a considerable increase in the level of the tumor suppressor p53 in the body. The higher quantities of p53 encouraged the production of miR-1469, which resulted in elevated levels of miR-1469 and lower levels of Mcl1, respectively.[Citation18]

Genistein possesses the capability to obstruct the p53-miR-1469-Mcl1 pathway. This, in turn, lowers the proportion of laryngeal cancer cells that are able to survive. The presence of genistein led to an increase in the rate of apoptosis in response to a greater dosage as well as over the course of time. In addition to limiting the capacity of Hep-2 cells to divide and create more of themselves, genistein was also able to trigger cell death in these cells.[Citation59]

Cervical cancer

After twenty-four and forty-eight hours, the effects of genistein on the activation of HeLa cells were shown to be at their most apparent. After an initial adhesion period of two hours, DMEM was fortified with genistein solutions of the following concentrations: 0 M, 12.5 M, 25 M, 50 M, and 100 M. These were added to the culture medium. The CCK-8 test was employed to evaluate the pace at which cells were replicating themselves. After subjecting the cells to a treatment with 100 mM genistein, scratch test was carried out in order to establish whether or not they have the capability to migrate. The transwell experiment was carried out with the intention of discovering whether or not there is a chance of cell invasion and migration. After 24 and 48 hours of exposure to 100 M genistein, there was a considerable inhibition of the growth of HeLa cells. The techniques of Western blotting and quantitative real-time PCR were utilized in order to investigate the levels of expression of proteins and messenger RNAs (mRNAs) that are respectively associated with the FAK-paxillin and MAPK signaling pathways. The length of the treatment and the dose of genistein administered were closely related to the level of success achieved by the supplement. There was a statistically significant difference in the rates of scrape migration between the treatment group and the control group after being exposed to 100 M genistein for 24 and 48 hours. This difference was seen between the two groups. This distinction was seen in the scrape migration rates as well. In addition, malignant cells were unable to penetrate either the top chamber or the matrigel used in the experiment.[Citation60]

The expressions of the genes for FAK, paxillin, snail, and twist were significantly reduced by 100 M genistein in HeLa cells, which are human cervical cancer cells, as compared to the expressions of the genes in the control group. This was the case when 100 M genistein was used. Genistein had a considerable inhibitory effect on the development of HeLa cells, both in terms of the proliferation of the cells and their ability to expand metastatically. Furthermore, it was evident that the concentration of genistein had an impact on the activity that served as the primary focus of the investigation. As a result of the presence of genistein, both the expression and activation of the paxillin and FAK (Focal Adhesion Kinase) genes were stifled. Genistein also suppressed the activity of the FAK gene. The administration of genistein to HeLa human cervical cancer cells led to a significant reduction in the expression of DNMTs as well as HDACs, and this reduction was followed by a decline in the enzymatic activity of the cells. Genistein was able to suppress the expression of the Snail and Twist genes, in addition to the vimentin and catenin proteins. This decline was dependent on how much time had passed. According to molecular modeling, there is a chance that genistein will interact with members of the histone deacetylase (HDAC) and DNA methyltransferase (DNMT) families. This is the case even though there is no direct evidence to support this hypothesis. This lends credence to the idea that genistein may be capable of preventing the enzymes in issue from carrying out their functions. The expression of tumor suppressor genes (TSGs) such MGMT, RAR, p21, E-cadherin, and DAPK1 was restored after being treated with genistein over an extended period of time that was dependent on the length of the treatment.[Citation61]

Treatment to cisplatin at concentrations of 10 M or higher resulted in a decrease in the number of viable HeLa and CaSki cells. Additionally, this exposure caused the methylation of their promoter regions to be reversed. These findings can be found in their publication. When larger concentrations of cisplatin (8 M and 6 M) were coupled with genistein, the viability of HeLa and CaSki cells was significantly decreased. Both HeLa and CaSki cells had an increased sensitivity to cisplatin following treatment with genistein. This was the case for both cell lines. The CaSki cells in the group that got cisplatin in conjunction with genistein revealed a 37% decrease in p-ERK1/2, a 304% rise in p53 expression, and a 115% increase in cleaved caspase 3 levels when contrasted with the CaSki cells in the group that received cisplatin alone.[Citation62]

Genistein accelerated the proliferation of HeLa cells in human cervical cancer cells in a dose-dependent manner across a range of concentrations (0.001, 0.04, 0.1, and 1 molL-1) in human cervical carcinoma cells. The expression of Bcl2 decreased by 69% when comparing the group that got therapy with cisplatin alone to the group that received treatment with cisplatin and genistein combined. Cells obtained from human cervical cancer demonstrated this effect. After being treated with genistein a higher percentage of HeLa cells were found to be in the S phase of their cell cycle, while a lower number of cells were found to be in the G1 phase. In addition to this, the treatment with genistein resulted in a notable reduction in the quantity of apoptosis that happened in the cells when compared to the group that served as the control. This was the case when compared to the group of the placebo. In addition, it was demonstrated that the use of genistein increased the expression of estrogen receptors (ER), the Akt protein, and the nuclear NF-κB p65 protein. The significance of this discovery cannot be overstated. The nuclear NF-κB p65 protein, ER, Akt, and ER were the variables in question. ER was also a contributor to the problem. The results of this investigation suggest that there may be linkages between all three of these proteins.[Citation61]

Endometrial or uterine cancer

It has been shown that the three uterine sarcoma cell lines MES-SA-Dx5, MES-SA, and SK-UT-1 respond to genistein in a variety of different ways, all of which have to do with the rate at which the cells duplicate themselves. Specifically, this has been found to be the case in terms of the rate at which the cells replicate themselves. For the SK-UT-1 cell line, the relative half-maximal inhibitory values for genistein were 19.2 milligrams per milliliter, whereas for the MES-SA cell line; 13.1 milligrams per milliliter, and for the MES-SA-Dx5 cell line; 9.3 milligrams per milliliter. In all three of the cell lines that were studied, the activity of cell division was slowed down when genistein was added to the culture medium. Once the first delay of 48 hours had passed, the process of DNA fragmentation started once the inhibitory effects started to take effect. Western blot analyses revealed three primary expression patterns: the induction of p53 and Dickkopf-related protein 1 (DKK1), the suppression of histone deacetylase 4/5/7 (HDAC4/5/7), disheveled protein (DVL), BAX, survivin, and phosphorylated mitogen-activated protein kinase kinase (phospho-MEK), and the induction of p27 and catenin in the more resistant lines. The activity of TOPflash was shown to correlate with a reduction in the expression of catenin. Despite the presence of estrogen receptor alpha, the presence of genistein in endometrial cancer cells produces a persistent rise in the expression of PR-B and forkhead box protein O1. This takes place regardless of whether or not the cells possess an estrogen receptor alpha. Interfering with the G2 phase of the cell cycle and triggering apoptosis are two methods that may be utilized to inhibit the progression of cell growth. Genistein-induced PR expression, as opposed to changing the epigenetic status of the PR region, lowers the expression of CCAAT/enhancer-binding protein beta and initiates the c-Jun N-terminal kinase trial.[Citation63,Citation64]

In another study, after cultivating RL95–2 cells for 30 minutes in estrogen-depleted media with or without poly I:C, the cells were treated with genistein at doses of 10 (−7), 10 (−6), or 10 (−5) M or 17 (−9) M for a period of 48 hours. RL95–2 cells were grown in estrogen-depleted medium with or without poly I:C. There were three separate concentrations of genistein, and they were correspondingly labeled as 10 (−7), 10 (−6), and 10 (−5) M. Genistein has the ability to improve uterine immune function while simultaneously decreasing pathogen-induced endometrial inflammation. This is because genistein can raise the baseline level of TLR2 in human endometrial epithelial cells while at the same time decrease the level of TLR2 protein synthesis that is activated by viral components.[Citation65]

The effect of genistein on the growth of HeLa cells in vitro was dependent not only on the amount of the compound that was delivered but also on the period of time that had passed since the cells were exposed to the drug. The values for these variables ranged anywhere from 0.25 to 64 g/mL and ranging from 24 to 72 hours. Following a time period of 48 hours, the value for the IC (50) was determined to be 4.62 g/mL. After being subjected to DFMG for a period of time equal to that of the apoptotic process, the cells underwent the distinctive morphological changes that are associated with the death of cells. The cells showed distinct DNA ladders when they were run through an agarose gel electrophoresis, and the fraction of cells that belonged to the sub-G (1) population increased in a dose-dependent manner. At the same time as c-myc mRNA, c-myc protein, and the genes that are downstream of c-myc were pushed to proliferate at a faster pace, Bcl-2 protein was driven to a lower level of expression. These genes included the bax gene, as well as the cyto-c gene and the caspase-9 gene. When the c-myc gene was silenced using siRNA, the DFMG-induced increases in cell proliferation and death were significantly decreased. Increasing the level of c-myc expression through the utilization of cDNA transfection has the capacity to not only reduce cell proliferation but also to increase the effects of DFMG-induced apoptosis induction. It’s possible that one of these results will be helpful.[Citation66]

Ishikawa cells were subjected to genistein treatment for varied durations of time (eight, twenty-four, and forty-eight hours), as well as concentrations (10 pM, one nM, one million, and ten million). This was done in order to establish the Ishikawa cells as a trustworthy in vitro model for assessing the potential estrogenic activity of compounds of interest. The goal was to develop the Ishikawa cells as a dependable in vitro model. It was decided that a thorough microarray strategy would be used instead. It was shown that genistein had an influence on the expression of 5342 different genes. The dose and the amount of time that the effect was seen were both dependent on the effect. The 10 M concentration of genistein, which was the highest one that was investigated, was responsible for the majority of the changes that were observed in Ishikawa cells when it came to the expression of genes. The effects of 17-ethynyl estradiol on Ishikawa cells were explored (at equivalent doses, i.e., 10 M vs 1 M, respectively), and the estrogenic activity of genistein was assessed by assessing changes in the expression of 284 genes that were equally activated by both genistein and EE. 17-ethynyl estradiol was shown to have a progesterone-like action on Ishikawa cells. Data showed 66 genes that were similarly up- or down-regulated in vivo and in vitro by comparing the response of the Ishikawa cells to high doses of genistein and EE with that of the juvenile rat uterus to EE. The results of this comparison allowed us to identify genes that were similarly up- or down-regulated in both settings. In order to attain this goal, a comparison of the two responses to EE was carried out. In order to do this, a comparison was conducted between the reactions of the Ishikawa cells and the uterus of the adolescent rats when they were given high doses of both medications. Some of the molecular pathways that are influenced by genistein include regulation of translation, cell proliferation, and intracellular transport.[Citation67]

Lung cancer

The lung cancer cell line A549 LC cells were able to have their mortality rate increased by genistein in a manner that was dependent on both the passage of time and the quantity of the compound. In addition, genistein had less of an effect on LC cells as a result of the low quantities of inosine monophosphate dehydrogenase-2 (IMPDH2). The effect of genistein on LC cell viability and apoptosis was shown to be mitigated by the highly expressed protein kinase B (AKT1).[Citation68]

It has been demonstrated that genistein reduces the chance of developing lung cancer. The increased expression of AKT1 was the essential component that had a significant role in the accomplishment of this discovery. MnSOD is the factor that is responsible for the increase of the expression of FoxM1, which is also known as Forkhead box protein M1. When cancer stem-like cells (CSLCs) produced from non-small cell human lung cancer (NSCLCs) such as H460 or A549 cells were treated with 80 and 100 millimolar genistein, the cells’ viability was dramatically reduced. This occurred when the cells were exposed to genistein at a concentration of 80 and 100 mM. On the other hand, genistein was shown to have the opposite impact on IMR-90 cells, as was seen by the researchers. Genistein was able to inhibit the activity of sphere formation at sub-cytotoxic doses, which were defined as 20 and 40 M, in contrast to the group that served as the control. In addition, the levels of protein expression for CD133, CD44, Bmi1, and Nanog were all reduced as a result of the use of genistein. In addition, as compared to the group that served as a control, the effects of genistein on LCSLCs suggested a reduction in the frequency of their migratory and invasive tendencies. This was the case despite the fact that genistein was administered to both groups. The reduction of MnSOD and FoxM1 has the effect of exacerbating these side effects. The effects of genistein on the CSLC features of LCSLCs were either nullified or amplified, depending on whether or not MnSOD and FoxM1 were overexpressed. Genistein’s effects on the CSLC characteristics of LCSLCs were nullified. Cell lines H292 and A549 were taken from patients with non-small cell lung cancer (NSCLC). A dose-dependent drop in cell viability was generated by genistein, along with a reduction in cell proliferation, migration, and invasion, and an increase in apoptosis. Genistein also caused an increase in cell death. There was a drop in [cell proliferation, migration, and invasion], and there was decrease in cell viability.[Citation69]

It was discovered and reported that genistein influenced the expression of forkhead box M1 (FOXM1) and raised miR-873-5p. This information was obtained from the study that was carried out. The effects of genistein on the A549 human lung cancer cell were studied over the period of one, two, and three days using a range of concentrations of genistein dissolved in physiological saline. The concentrations of genistein used were 0 M, 10 M, 25 M, 50 M, 100 M, and 200 M. These concentrations were determined to be effective over a wide range of dosages after being subjected to testing.[Citation70]

Genistein greatly decreased cell growth, significantly increased cell apoptosis, significantly accelerated caspase-3/9 activation, and promoted the synthesis of microRNA-27a (miR-27a). Researchers were able to investigate the anticancer effects of genistein after it caused the SCLC cell line H446 to undergo apoptosis and arrest while simultaneously generating a significant reduction in the cell line’s potential for multiplication and migration. One of the many extra effects that genistein had on H446 cells was to enhance the anti-proliferative capabilities of cisplatin. Genistein also had a number of other impacts. Because of this, multiple FoxM1 target genes, including as Cdc25B, cyclin B1, and survivin, which control apoptosis and the cell cycle, were able to be downregulated as a result of the presence of genistein. This was possible because of the fact that these genes were able to respond to genistein. The fact that this leaves the FoxM1 protein unable to function properly is a significant development.[Citation71]

Increasing FoxM1 levels before to undergoing therapy with genistein may diminish the inhibitory effects of genistein on the proliferation of H446 cells. The anticancer activity of trichostatin A (TSA), which was studied in human lung cancer A549 cells, was shown to be boosted when genistein was present, according to the findings of the researchers. It is unknown if cancer cells coming from the various subtypes of lung cancer react in the same manner to the combination treatment as it is currently thought that they do. In this work, the effect of genistein on TSA was explored, as well as its influence on the anticancer activity of TSA in ABC-1, A549, and NCI-H460 (H460) cells. A comparison of the effects of genistein and TSA was also performed. After that, researchers looked into whether or not the actions of genistein are associated to an increase in the amount of acetylation of proteins that are neither histones nor non-histones. It was found that the impact of genistein on the inhibition of cell growth was significantly less in ABC-1 cells (p53 mutants) than it was in A549 and H460 cells. It was discovered that genistein could reduce the amount of apoptosis brought on by TSA in A549 and H460 cells, while it had no impact on ABC-1 cells. It was observed that the genistein-induced proliferation of A549 and H460 cells was at a lower level after the expression of p53 was suppressed. This was the case because the expression of p53 had been inhibited. In addition, the presence of genistein in A549 and H460 cells induced an increase in the total amount of histone H3/H4 acetylation that was brought on by TSA. Additionally, it was shown that genistein stimulated the acetylation of p53 in H460 cells.[Citation28]

In has been reported that anacardic acid was able to counteract the increase in TSA-induced cell death and acetylation of histones and p53 that was generated by genistein because it is an acetyltransferase inhibitor. The synthesis of the acetyltransferase protein p300 was able to be stimulated in A549 and NCI-H460 cells, respectively, by both genistein and TSA. In addition, it was established that the anticancer effectiveness of genistein was increased in rats that had the A549 tumor.[Citation72,Citation73]

Patients diagnosed with non-small-cell lung cancer (NSCLC) were treated with genistein. Anticancer effects were observed as a result of this treatment. Genistein induced a drop in the quantity of Bcl-xL that is located in the nucleus of cancer cells, while at the same time increased the amount that is present in the cytoplasm of those cells. Ionizing radiation (IR) is capable of causing DNA damage and death in non-small cell lung cancer (NSCLC) cell lines; however, low-cytoplasmic Bcl-xL NSCLC cell lines are far more vulnerable to this kind of damage and mortality than high-cytoplasmic Bcl-xL NSCLC cell lines. It was hypothesized that the amount of cytoplasmic Bcl-xL has an inverse association with the radiosensitivity of non-small-cell lung cancer; the hypothesis was confirmed by the findings of this clinical trial. Additionally, data showed that treatment with genistein made NSCLC cells more sensitive to the effects of radiation.[Citation74]

Genistein treatment upregulates the activities of Beclin-1 and Bcl-xL, this results in an increase in the amount of autophagy that was mediated by Beclin-1. As a consequence of this, the quantity of autophagy that was mediated by Beclin-1 increases. In a xenograft model of non-small cell lung cancer (NSCLC), both gensitein and cisplatin were proven to be effective against the illness through a variety of different routes. This was discovered through the use of the xenograft model. These activities included the activation of apoptosis and the prevention of tumor growth in addition to a significant reduction in the constitutive phosphorylation of AKT and phosphoinositide-3 kinase. According to research, lung cancer A549 cells treated with genistein had a significant increase in the levels of the pro-apoptotic factor Bax mRNA and protein 48 hours after the treatment, while the levels of the anti-apoptotic factor Bcl-2 mRNA and protein experienced a significant reduction.[Citation75]

Brain cancer

It has been demonstrated that radiotherapy is an effective treatment for glioblastoma multiforme (GBM), the most prevalent and fatal type of brain cancer. Additionally, it has been demonstrated that this malignancy responds favorably to the therapy. On the other hand, DNA-PKcs-positive GBM cells were able to migrate and invade more readily than DNA-PKcs-negative GBM cells, but the converse was not true; DNA-PKcs-negative GBM cells were not able to migrate and invade more easily than DNA-PKcs-positive GBM cells. In addition to this, it promoted activity in the DNA-PKcs/Akt2/Rac1 signaling pathway, which was a factor in the progression of malignant GBMs. An increase in the capacity of GBM cells to colonize new places was a factor that contributed to the accomplishment of this objective. In a study it was discovered that genistein has the ability to bind directly to DNA-PKcs and to disrupt the pathway that links DNA-PKcs to Akt2/Rac1. Because of this, it is feasible to stop the invasion and migration of GBM cells that are created by radiation, both in vitro and in vivo. This may be done both in the laboratory and in the patient. This is due to the fact that radiation is the root cause of these processes. Samples of high-grade and low-grade glioma cancer tissue were taken from a total of 16 different persons; the tissue was cleaned with PBS, diced into minute particles, absorbed with collagenase type I, and then cultured in DMEM with 10% fetal bovine serum. After the third cycle of cell division, the cells were treated with genistein, and quantitative real-time PCR (qRT-PCR) was performed to evaluate the quantities of transcripts produced by the MMP-2 and VEGF genes. A 580-fold decrease in MMP-2 expression was seen in low-grade glioma cells that were treated with genistein. This finding was determined by comparing treated cells to untreated cells. The MMP-2 imprint that was discovered to be present in high-grade lesions was proved to be two times lower when compared to the MMP-2 impression that was observed to be present in cells that originated from low-grade abrasions. Genistein induced a 4.7-fold decline in VEGF mRNA in high-grade glioma cells, while it had no effect on low-grade glioma cells. This finding was based on the fact that genistein was only tested on high-grade glioma cells.[Citation76]

During an experiment, 9 L gliosarcoma cells were injected intracranially into the brains of Fischer 344 male rats. These rats were employed as subjects. Genistein was constantly administered orally at a dose of either 100 or 200 milligrams per kilogram of body weight beginning on the third day after implantation and continuing until the 25th day of the study. The dosage was determined by the subject’s body weight. This method of administration was kept up for the entirety of the research. Propofol was given intramuscularly at a rate of 20 mg/kg once every five days up until the 25th day, two hours following the administration of genistein. Following treatment with genistein at doses ranging from 12.5–200 g over a period of 24 hours, human gliosarcoma cells (U251) exhibited a decrease in the number of viable cells present after being exposed to the compound. It was demonstrated that giving a therapy of genistein to rats with intracranial tumors, with or without the injection of propofol, either greatly reduced the size of the tumors or significantly increased the length of time that rats with tumors lived in a model of intracranial tumors. This was the case whether the treatment was administered with or without the injection of propofol. Because genistein is able to cleave caspase-3 and cause the production of pro-apoptotic proteins (Bax and Bad), the degree of apoptosis that occurs as a result of this capacity was significantly increased. In addition, it increased the synthesis of these proteins whether it was given by itself or in conjunction with propofol for the purpose of administration. This was true regardless of whether method of administration was used. Additionally, there was a decrease in the excessive production of TNF-, IL-1, and IL-6, which was associated to active NF-κB activation. This reduction was connected to the fact that there was a connection. This decrease has been seen. Both genistein and propofol caused gliosarcoma cells to commit suicide, which significantly slowed down the growth of the disease. It was demonstrated that the administration of propofol or genistein in combination with one another was more successful than the administration of each one on its own. This reveals that the anticancer effects of propofol may be enhanced by genistein and vice versa. Genistein shows promise as a radiosensitizer for improving the effectiveness of carbon ion radiotherapy against DNA-PKcs competent GBM in human glioblastoma (GBM) cell lines by suppressing DNA-PKcs phosphorylation and subsequently repressing the non-homologous end joining (NHEJ) repair of DSBs and delaying homologous recombination (HR) repair pathways.[Citation77]

Blood cancer

Genistein inhibits acute leukemia (AL) cells. The two AL cell lines both demonstrated genistein IC50 values that were lower than those reported by the bone marrow mesenchymal stem cell line. Wnt suppressor genes such as Wnt5a were positively influenced by genistein, which had the opposite effect on Wnt target genes such as c-myc and -catenin. Genistein also elevated genes that inhibit Wnt signaling, such as H4K20me1 and KMT5A. The amount of methylation and the degree to which H3K9ac was enriched in the Wnt5a promoter region remained relatively unchanged throughout the course of time. On the other hand, greater levels of H4K20me1 could be seen in the coding regions and the promoter of the Wnt5a gene. Additionally, the expression of Phospho-wee1 was suppressed as a result of genistein, but the expression of Phospho-cdc2, Myt1, Cyclin A, Cyclin E2, p21, and Phospho-histone H3 was elevated. At the point where the cell was in the G2/M phase, the advancement of the cell cycle was halted. Genistein is able to prevent the activation of the Wnt signaling pathway because it is capable of activating KMT5A and raising the amount of H4K20me1 that is present in both the promoter and the coding regions of the Wnt5a gene. As a consequence of this, there is an increased potential for the generation of Wnt5a. Genistein has the ability to halt the cell cycle in the G2/M phase.[Citation37]

VEGF or vascular endothelial growth factor, is a kind of angiogenic mediator that plays an essential role in both the invasion and metastasis of tumors. These factors’s capacity to be released can be found in leukemic cells. Any potential recurrence is made more difficult by the presence of the C-X-C chemokine receptor type-4 (CXCR-4), which has been linked to metastasis and is seen on malignant cells. It has been established that the isoflavonoid genistein, which comes from soy, has the capacity to prevent the process of angiogenesis. This ability was discovered via scientific research. (ALL) acute lymphoblastic leukemia cell cultures were utilized throughout the course of this study endeavor to investigate the effects of this medication on CXCR-4 and VEGF. For the purpose of determining the genistein’s level of cytotoxicity, a colorimetric experiment using MTS was carried out. Both the mRNA and protein levels of VEGF expression were examined in MOLT-4 and Jurkat cells that had been treated with genistein. The treatment had resulted in an increase in both of these levels. In addition to this, flow cytometry was performed in order to explore the influence that CXCR-4 had on the surfaces of the cells in order to see how they responded. In spite of the fact that genistein was more successful on MOLT-4 cells, it was discovered that it was less viable in two more cell types. After treatment with genistein in MOLT-4 and Jurkat cells, there was an increase in the levels of VEGF secretion and mRNA impression. On the other hand, there was a decrease in the levels of CXCR-4 surface expression.[Citation78]

Researchers suggest that genistein might not be an effective therapy for a number of different disorders. On the other hand, a different design has been identified, which can be beneficial for discovering CXCR-4 and VEGF modulators and, as a consequence, for creating novel therapeutics for blood cancer and other disorders associated to VEGF. The up-regulation of BTG3 as a tumor suppressor gene was improved when acute lymphoblastic leukemia (ALL) cell lines (MOLT4, MOLT17, and JURKAT) were treated with different doses of genistein (10, 25, 40, and 55 M) for 24, 48, and 72 hours. The BTG3 gene is responsible for inhibiting the proliferation of cancer cells. Elevation in BTG3 at the transcriptional level, it was revealed that the proliferation of cancer cells in MOLT4 and JUR was inhibited by 60–90%. This information was gleaned from their study.[Citation79]

Genistein caused morphological alterations in HL-60 cells, as well as a decrease in the number of viable cells, DNA damage, and fragmentation, which ultimately led to cell death. Genistein also caused DNA to become damaged and fragmented. Additionally, other outcomes came to pass. In HL-60 cells, genistein decreased the amount of m while concurrently generating a bigger creation of ROS and Ca2+. This occurred while genistein also increased the development of ROS. In HL-60 cells, genistein induced an increase in the expression of proteins that are associated with ER stress, apoptosis, and PARP cleavage; however, it caused a decrease in the expression of anti-apoptotic proteins such as Bcl-2 and Bid. These results suggest that genistein promotes cell death by increasing the expression of proteins that are involved with ER stress, apoptosis, and PARP cleavage. Some of the proteins that belong to this group are GRP78, IRE-1, Calpain 1, GADD153, caspase-4, caspase-7, and ATF-6. In HL-60 cells, there was a rise in the levels of the genes GADD153, Calpain 1, and GRP78 as a result of treatment with 40 M genistein. Within the context of the animal xenograft paradigm, genistein was administered intraperitoneally to mice at doses of 0.0, 0.2, and 0.4 mg/kg of the animal’s body weight over the course of 28 days. During this period, readings were obtained to determine the body weight as well as the volume of the tumor. It was discovered that when genistein was given to a group at a level of 0.4 mg/kg, it greatly decreased the weight of tumors while having no effect on the general body weight of the subjects. The presence of genistein in malignancies was observed to result in an increase in the expression of ATF-6, GRP78, Bax, Bad, and Bak.[Citation80]

In HL-60 cells, the combination of genistein and L-asparaginase (Asp) has been shown to be the factor that is responsible for causing apoptosis. When compared to the genistein or L-asparaginase (Asp) treatment on its own, this therapy led to a significantly lower number of total viable cells as well as a larger proportion of G2/M stage cell arrest, DNA damage, and cell death. Additionally, the number of total viable cells was much lower. In addition to this, the total number of viable cells was found to be drastically reduced. Additionally, the combined treatment of genistein and L-asparaginase (Asp) demonstrated a greater degree of m decrease than either the genistein therapy or the L-asparaginase (Asp) treatment alone could accomplish. The mitochondrial membrane potential was changed as a consequence of the interaction between genistein and L-asparaginase (Asp), which led to a malfunction in the mitochondria. Both combination therapies resulted in increased levels of Bak and Bax (pre-apoptotic proteins), an active form of caspase-3, and a larger fall in Bcl-2 (anti-apoptotic protein) when compared to either Genistein or L-asparaginase (Asp) therapy alone.[Citation81]

Acute myeloid leukemia, often known as AML, is a kind of cancer that attacks the progenitor cells of the hematological system. The end result of this onslaught is nearly perpetually the patient’s death. Research was conducted to determine whether or not genistein has the potential to be an effective alternative treatment for acute myeloid leukemia (AML). For the purpose of the model, two distinct cell lines were used: HL-60, which carries the FLT3 gene in its natural state, and MV4–11, which possesses the FLT3 gene with the ITD mutation. Both of these cell lines possessed the FLT3 gene in their respective natural states. The proliferation of both AML cell lines can be inhibited by genistein, as shown by a wide range of functional investigations as well as by this research that was based on an 8-plexed version of the iTRAQTM quantitative proteomics platform. The introduction of genistein into AML cells caused a blockage in the mTOR pathway, which led to a decrease in the amount of protein that was produced by the cells. Treatment with genistein produces apoptosis, which results in the death of cells in the long run.[Citation82]

Thyroid cancer

Both thymoquinone and genistein were shown to have a statistically significant negative influence on the viability of thyroid cancer (TC) cells, as well as on the mRNA expression levels of the hTERT, NF-κB, and VEGF-A genes. These results were found in a study that was conducted on mice. These findings came about as a consequence of an experiment that was carried out on mice. On the other hand, they were responsible for an important and favorable shift in the total quantity of PTEN and p21 mRNA expression that was present. There was a statistically significant increase in the total quantity of active CASP-3 protein as a result of the use of thymoquinone and genistein in the treatment of TCCs, which is the most common type of endocrine cancer.[Citation83]

TCCs are the most common form of endocrine cancer. This was the result of the therapy that was administered. Genistein was able to partially restore the epithelial-mesenchymal transition that was taking place, in addition to its ability to significantly limit the invasion of papillary thyroid cancer (PTC) cell lines. An investigation into the functional effects of genistein indicated that it considerably suppresses the protein-level effects of EMT by reducing the amount of catenin that is blocked by small interfering RNA (siRNA). In individuals with papillary thyroid cancer (PTC) whose thyroid cancer cells were between 2.5 and 80 ng/ml, the research found that genistein had an impact that was anticarcinogenic. The translocation of catenin into the cytoplasm of the cell is the most likely explanation for the impact that was seen. The presence of genistein resulted to an increase in the expression of GPR30 in the human thyroid squamous cell line SW579. This occurred despite the fact that ER was not present in these cells. Additionally, the expression of GPR30 was elevated as a result of exposure to genistein.[Citation84]

The presence of genistein in SW579 cells caused GPR30 to be inhibited, which led to a significant decrease in cell proliferation and an increase in cell mortality compared to when genistein was not present. The G2/M phase of the cell cycle is where the majority of the arrests in the cell cycle took place. The capacity of genistein to limit the ability of a thyroid cancer cell line (SW579) to invade and migrate may be connected, at least in part, to the fact that genistein suppresses MMP-2 expression. This interpretation is based on the fact that genistein inhibits the production of MMP-2.[Citation85]

Kidney cancer

When human bladder transitional cell carcinoma T24 cells were treated with genistein, the expression of CIP1/P21WAF1, which is a cyclin-dependent kinase (Cdk) suppressor complexed with Cdk2 and Cdc2, increased. CIP1/P21WAF1 inhibits the activity of cyclin-dependent kinases (Cdks). However, when cyclin A and cyclin B1 were introduced into the cells, its expression dropped significantly. There is a possibility that genistein is to blame for this upregulation. In addition, it has been shown that genistein is responsible for the activation of caspase-3, −8, and −9, which eventually leads to the cleavage of poly (ADP-ribose) polymerase. On the other hand, the employment of a pan-caspase inhibitor resulted in a large reduction in the quantity of apoptosis that was triggered by genistein. This was the case because the inhibition of many caspases was required. This provides support for the theory that the ability of genistein to trigger apoptosis was reliant on the activity of the caspase enzyme. Apoptosis is the programmed death of cells. In addition, genistein increased the ratio of Bax to Bcl-2 while concurrently decreasing the amount of cytochrome c that was allowed to be released into the cytoplasm. Because of this, the total degree of mitochondrial integrity was reduced as a consequence. In addition, LY294002 was able to preserve genistein’s capacity to block the PI3K/Akt signaling pathway while at the same time augmenting the apoptosis-inducing activities of both genistein and the PI3K/Akt inhibitor LY294002. When genistein was included in the experiment, the levels of reactive oxygen species (ROS) were significantly elevated; however, N-acetyl cysteine, which is more often referred to as NAC, was able to significantly lower the ROS levels. Akt/PI3K signaling, apoptosis, and arrest in the G2/M phase of the cell cycle were all induced by genistein; however, NAC was able to inhibit all of these effects.[Citation86]

Genistein is able to inhibit the formation of tumor spheres in 786-O and ACHN cells, as a result the level of renal CSC activity dropped by a substantial amount. This was done by simultaneously triggering apoptosis and reducing the growth of renal CSCs, as well as their signs and markers. These two strategies were applied in equal measure. In addition, the activity of the Sonic Hedgehog (Shh) pathway may be downregulated, which may have helped to the suppression of renal Cancer stem cells. Genistein is able to suppress the multiplication of progenitor cells that can result in kidney cancer. In order to achieve this objective, the amount of activation of the SH route has to be decreased.[Citation87]

In patients with renal cell carcinoma (RCC), genistein was able to slow the rate of cell proliferation as well as the rate of cell invasion, in addition to its ability to promote apoptosis. Additionally, the activity of the TCF reporter in RCC cells was able to be suppressed by the presence of genistein. It has been demonstrated that genistein significantly suppresses the expression of miR-1260b in RCC cells, which is a microRNA that is highly expressed in these cells. When the levels of miR-1260b expression in renal cancer tissues were compared to the levels of expression in normal tissues, the levels of expression in renal cancer tissues were significantly greater. This discovery has a substantial correlation to a decreased total survival time. In addition to this, the expression of miR-1260b in RCC cells led to an increase in the cancer cells’ capacity to penetrate healthy tissue and proliferate. When an inhibitor of miR-1260b was transfected into renal cancer cells, the 3’UTR luciferase activity of the target genes (sFRP1, Dkk2, and Smad4) was considerably decreased.[Citation88]

By inhibiting cell proliferation, which is essential for activities such as migration and dissemination of human renal cell carcinoma cell lines, genistein was able to thwart the progression of these malignant processes. Experiments using chromatin immunoprecipitation (ChIP) revealed that genistein increased ZO-1 expression by decreasing the quantity of PRC2 that was recruited to the ZO-1 promoter. This was demonstrated by the fact that the expression of ZO-1 was increased. Genistein prevents cancer from progressing by preventing HOTAIR from interacting with PRC2. This leads to a decrease in the total number of cancers that can develop as a result. The findings of immunoprecipitation testing indicate that genistein lowers levels of PRC2 EED. This illustrates that there is a dampening of the interaction between HOTAIR and PRC2 when there are low quantities of EED. Because EED overexpression in the presence of genistein restored PRC2 connection with HOTAIR and lowered ZO-1 transcription, it is feasible that genistein promotes ZO-1 via reducing the activity of HOTAIR/PRC2. This is because EED overexpression in the presence of genistein restored PRC2 connection with HOTAIR. This inference may be made because of the fact that genistein contributed to an increase in ZO-1. In addition, RIP investigations showed that HOTAIR interacts with SMARCB1 and ARID1A, which are both components of the human SWI/SNF chromatin remodeling complex. Both of these proteins are involved in the remodeling of chromatin. It has been demonstrated that genistein is a potent inhibitor of this interaction. In human cells, the alteration of chromatin is brought about by the coordinated actions of these two components.[Citation89]

Genistein suppresses SNAIL transcription through the use of a method that involves SMARCB1 and HOTAIR. These two proteins were involved in the process. This was found by combining the overexpression of HOTAIR in the cells with the suppression of SMARCB1 in the presence of genistein. This led to the discovery of the phenomenon. The cancer cells in the kidney were treated with genistein, which led to the death of the cancer cells and a decrease in their ability to replicate. In addition, it has been demonstrated that genistein raises the expression of CDKN2a while at the same time lowering the quantity of methylated CDKN2a.[Citation90,Citation91]

Eye cancer

Following the administration of genistein to CRL1872 malignant melanoma cells at concentrations of 25, 50, and 100 micrograms (mM), there was a discernible rise in the level of cyclin D1 expression. It was determined that a dosage of genistein consisting of 100 mg was sufficient to block the development of cyclin D1. Because genistein has the potential to activate apoptosis, the human retinoblastoma cell line Y79 was significantly hindered in its capacity to multiply and grow when it was not linked to an anchorage. Because genistein has the potential to destroy cells. The fact that genistein is capable of causing cell death was the most important contributor to the occurrence of this phenomenon. Screening with microRNA arrays revealed that genistein increases the amount of expression of the miR-145 gene, which led to the discovery of this relationship. The capacity of genistein to suppress the formation of retinoblastoma tumors was considerably helped by the presence of microRNA-145 as a key downstream effector. This was accomplished through the post-transcriptional regulation of ABCE1, which is an important downstream target. Genistein was successful in preventing the growth of retinoblastoma xenografts. According to the conclusions drawn from these investigations, genistein is able to accomplish this objective through increasing levels of miR-145.[Citation92] It is evident from the studies that genistein can interfere and downregulate the factors causing eye cancers. Activation of apoptosis via modulation at RNA level is remarkable feat exhibited by genistein.

Bone cancer

The peroxisome proliferator-activated receptor, or PPAR for short, possesses the features of an important regulator in a range of distinct metabolic pathways that are associated to cancer. Over the course of two days, MG-63 cells were subjected to a treatment in which they were given different amounts of genistein and/or GW9662, which is a selective PPAR competitor. The Cell Counting Kit-8 (CCK-8) was utilized so that researchers could ascertain how the effects of various chemicals altered the cellular viability of the cells they were studying. EdU, also known as 5-ethynyl-2”-deoxyuridine, was used in order to assess the rate at which new cells were being created. Another name for EdU is 5-ethynyl-2”-deoxyuridine. Researchers were able to uncover discrepancies in the course of the cell cycle and apoptosis by employing a method known as flow cytometry. The quantity of expression of the proteins that are involved in the PPAR pathway was assessed using western blot analysis. These proteins include BCL-2, PPAR, PTEN, Survivin, CIP1/P21WAF1, and Cyclin B1. In order to determine the amount of PPAR and PTEN mRNA expression, quantitative real-time RT-PCR was performed. Genistein is able to impede the growth and proliferation of OS cells, making it a potential treatment for this condition. Treatment with genistein in OS cells boosted PPAR expression. It has been postulated that the PPAR pathway is implicated in a significant portion of the genistein-induced modifications in cell proliferation, as stated by.[Citation93,Citation94]

A research was conducted to investigate the impact that genistein has on the nuclear factor B (NF-κB) and microRNA-29b (miR-29b) pathways that are responsible for the proliferation and death of multiple myeloma (MM) cells. A test using 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide was utilized in the investigation for the purpose of analyzing the process of cell proliferation. In addition, an Annexin V-fluorescein isothiocyanate/propidium iodide apoptosis assay and a caspase-3 activation test were carried out in order to establish whether or not apoptosis had taken place. In order to evaluate the amounts of miR-29b and NF-κB that were presumed to be present, reverse transcription, quantitative polymerase chain reaction, and western blotting were the procedures that were utilized, respectively. U266 cells were transfected with plasmids expressing for miR-29b and anti-miR-29b in order to gain a better knowledge of how genistein impacts MM. According to the findings, genistein was linked with a considerable slowing down of the pace at which cells developed, as well as an increase in the activity of caspase-3, and both of these effects were shown in U266 cells. Additionally, genistein was related with an increase in the activity of caspase-3. Both of these effects were observed in U266 cells. Genistein may raise the expression of miR-29b in U266 cells while concurrently lowering the levels of NF-κB protein in these cells. This was proved by the experiment described above. In addition, the findings indicated that miR-29b had the capability of altering the expression of NF-κB in U266. Genistein may inhibit the proliferation of human MM cells by concurrently increasing the amount of miR-29b and lowering the quantity of NF-κB.[Citation95]

Ovarian cancer

Ovarian cancer is one of the many types of cancer that can be inhibited by genistein. In the course of this investigation, we made use of a model that was based on the laying hen, which is well-known for the high rate of spontaneously occurring ovarian cancer that it experiences. After randomly assigning 100 laying hens to one of three groups, the following groupings were determined: standard (3.01 mg/hen), low genistein supplementation (52.48 mg/hen), and high genistein supplementation (106.26 mg/hen/day; each group). Ovarian tumors from hens that had been bred for a total of 78 weeks were utilized for the purposes of this study. These tumors were first collected, then examined, and finally discarded. It was found that taking genistein supplements led to a significant reduction in the overall risk of developing ovarian cancer, the number of tumors, and the size of tumors. Genistein was proven to reduce levels of oxidative stress indicators such as serum malondialdehyde and NF- κB and Bcl-2 expression in a research that investigated the molecular composition of ovarian tumors. On the other hand, it caused an increase in the quantity of expression of the proteins Nrf2, HO-1, and Bax in the ovarian tissues. The ingestion of genistein led to a decrease in the total amount of phosphorylation that occurred in mTOR, p70S6K1, and 4E-BP1, which showed that the mTOR pathway functioned at a more subdued level of activity. Genistein could play a part in the treatment of ovarian cancer with chemotherapy and emphasize the effect that it has on the molecular pathways that are connected to the course of the illness.[Citation96]

OVCAR-5 cells were treated with genistein in the absence of estrogen and monitored for any alterations that may have occurred. Cell enumeration and MTS testing both showed that the levels of cellular proliferation were unique from one another. In order to explore the levels of expression of important cell cycle regulators, the Western blot and real-time PCR methodologies were applied. According to the findings, genistein had a considerable favorable effect on both the proliferation and the viability of the OVCAR-5 cells. After receiving genistein therapy, the levels of cellular mRNA and protein expression for PCNA, Cyclin D1, and CDK4 increased, whilst the levels of p21 and p27 dropped. These changes occurred in response to the treatment. It was indicated that genistein has the ability to accelerate the proliferation and G1-S transition of OVCAR-5 ovarian cancer cells. This is in contrast to the findings of a large number of other investigations, which found that genistein does not have this capability. It is likely that the variance might be due to either distinct experimental circumstances or a wide diversity of ER expression patterns in the various cell lines.[Citation97]

Genistein lowers the chance of getting cancer, the mechanism by which genistein influences the glycogen synthase kinase-3 (GSK-3) pathway in ovarian cancer remains unknown. Genistein has been proven to inhibit the growth of ovarian cancer cells. Research was investigated to determine the effects of genistein on inflammatory biomarkers and GSK-3 signaling pathways in the ovaries of older laying hens that had been diagnosed with ovarian cancer. When the control group was given the same quantity of inflammatory proteins as the experimental group, the blood levels of inflammatory proteins such as tumor necrosis factor (TNF), interleukin-6 (IL-6), interleukin-8 (IL-8), and vascular endothelial growth factor (VEGF) were significantly higher. The results showed that when genistein was administered, these blood levels were significantly lower. Following therapy, it also resulted in a rise in protein kinase B (p-AKT) and insulin receptor substrate-1 (p-IRS-1), although it had the reverse impact on GSK-3. The dose that is administered has a direct bearing on the degree of success that may be achieved by the treatment. In the ovaries of older laying hens, an anticancer impact of genistein was revealed by a reduction in the levels of pro-inflammatory biomarkers as well as an inhibition of GSK-3 expression.[Citation96]

The key transcription factor known as c-Myc is one that has been implicated in the etiology of a number of different cancers. It was found that the levels of c-Myc mRNA and protein were much higher in samples collected from ovarian tissues with early stages of cancer compared to those taken from ovarian tissues with healthy stages of cancer. This was the case regardless of whether or not the samples were from cancerous or healthy ovarian tissue. It was discovered that patients whose blood contained elevated levels of c-Myc had a greater chance of being diagnosed with ovarian cancer at a clinical stage I (Ia+b/Ic) earlier in the disease’s progression. People who had considerably higher levels of total nuclear c-Myc expression lived noticeably longer than people who had significantly lower quantities of the protein. It was shown that the novel synthetic genistein analogue 7-difluromethoxyl-5,4’-di-n-octylgenistein (DFOG), which reduces PI3K/AKT signaling in vitro and in vivo, was more deadly in ovarian cancer cells when c-Myc was suppressed. Ovarian cancer cells are rendered more sensitive to the effects of DFOG as a direct result of Let-7d’s capacity to suppress c-Myc. This is because Let-7d is able to inhibit the activity of c-Myc. c-Myc is the primary therapeutic target of a given drug. Let-7d improves ovarian cancer cell sensitivity to DFOG by suppressing c-Myc and PI3K/AKT signaling. This is connected to early ovarian cancer, which is also associated with nuclear c-Myc expression.[Citation98]

In a typical scenario, cancer cells that have developed a resistance to taxol will have an increased level of TLR4 expression. It was able to discover 787 genes that are overexpressed in the ovarian cancer cell line SKOV3 and express TLR4 ectopically by doing a whole-genome transcriptome analysis. In the ovarian cancer cell line, these genes were able to be identified and located. After doing chromatin immunoprecipitation enrichment analysis, it was discovered that 27.8% of the genes that TLR4 elevated were also AR-regulated. As a direct result, the expression of AR was shown to be significantly elevated in taxol-resistant SKOV3 cells that exhibited an excessive level of TLR4 expression. On the other hand, when TLR4 was suppressed by the use of shRNA, there was a reduction in the expression of AR. Androgens, on the other hand, have been demonstrated to increase the activity of the AR, whilst small hairpin RNA (shRNA) has been shown to decrease the activity of the AR. It has been demonstrated that both of these mechanisms have an impact on the expression of genes that are connected with the AR. Cells that were resistant to taxol were more likely to have an overexpression of the AR-related genes DCDC2, ANKRD18B, ALDHA1, c14orf105, ITGBL1, and NEB. These genes were observed to be overexpressed more frequently. There is an adequate amount of data to support the hypothesis that these genes are involved in taxol resistance.[Citation99]

According to the findings of a published study, androgens have the ability to trigger the upregulation of a great number of genes. Many of these genes are involved in the pathways that lead to the production of steroids, thus this discovery is particularly relevant. Other effects, such as the suppression of AR-driven genes and a reduction in taxol resistance, can be detected in ovarian cancer cells as a result of the ability of genistein to lower levels of AR activation. This is possible because of the ability of genistein to inhibit the receptor for androgen receptor (AR). It has been established that genistein is the agent that causes this effect. Six TLR4- and AR-regulated genes were found to be involved in the overall process of building taxol resistance after it was established that these genes were implicated. In light of the findings presented here, which indicate that the TLR4/AR axis is a significant contributor to the emergence of taxol resistance.[Citation100]

TLR4/AR axis plays a significant part in the progression of taxol resistance. In order to create estimates on pharmacological interactions with targets, including secondary targets, the STRING database was utilized. The combination of genes that are functionally related to one another resulted in the creation of a module. After that, the analytical method known as Molecular Complex Detection (MCODE) was applied in order to conduct an evaluation of the PPI network modules considered to be the most significant. On the key genes, an enrichment analysis was performed by using the KEGG pathway database. The acronym for the Kyoto Encyclopedia of Genes and Genomes is KEGG, and its full name is the Kyoto Encyclopedia of Genes and Genomes. In addition to that, the Comparative Toxicogenomics Database, often known as CTD, was combed through for information about disease targets. As a consequence of this, it was found that genistein interacts with a total of 372, in addition to the initial 13 targets it was originally designed to interact with. After then, a method called MCODE analysis was used in order to look for important genes that were related with these targets. Additional study on the top 72 genes was carried out with the assistance of the KEGG database. CTD was able to acquire seven genes as a consequence of the identification of 123 genes that had a specific relationship with the marker “T” or “M.” This finding was made possible by the fact that 123 genes had a distinct interaction with the marker. The seven genes include CDKN1B, PTEN, EGFR, MAPK1, MAPK3, and PIK3C. The production of these genes required cooperation between three distinct pathways and 123 genes in all. Increased CDKN1B levels were revealed to have a strong link with overall survival log-rank. The Kaplan-Meier approach was applied.[Citation101]

It is becoming more and more obvious that microRNAs, which are also known as miRNAs, play an essential part in the modulation of a wide range of the pathobiological processes that are connected with cancer. This discovery comes on the heels of the discovery that soybeans have a wide range of other biological advantages. Although it has been shown that genistein has an effect on ovarian cancer cells, the particular biochemical mechanisms through which it exerts this effect are not yet fully known. Genistein has been shown to inhibit the growth of ovarian cancer cells. According to the findings of this study, the amount of miR-27a found in human ovarian cancer tissues is much higher than that found in normal ovarian cells. The miR-27a inhibitor had a strong inhibitory impact on the growth and migration of tumor cells when it was transfected into SKOV3 cells. Giving genistein to ovarian cancer cells resulted in a slower progression of the disease and a reduction in the migration of cancer cells across the body, as demonstrated by the findings of the research. Genistein also reduced the amount of time it took for cancer cells to spread throughout the body. Other cellular mechanistic analyses indicated that genistein significantly upregulated Sprouty2, a gene that is known to be a target of miR-27a, while concurrently downregulating miR-27a. This was demonstrated by the fact that genistein had the opposite effect on miR-27a. The investigation into this topic began when cells were treated with genistein. Ovarian cancer cells are dependent on the oncogenic miR-27a for growth and metastasis, and it also shows that genistein, a harmless miRNA inhibitor, may reduce the proliferation and migration of ovarian cancer cells to a higher extent. Findings suggest that ovarian cancer cells are dependent on the oncogenic miR-27a for growth and metastasis.[Citation102]

As suggested by scientific literature estrogen is a factor responsible for ovarian cancer. The word estrogen receptor type 2 is most commonly referred to by its acronym, which is ER, which stands for estrogen receptor. The amount of evidence that demonstrates ER’s ability to inhibit the development of cancer is consistently expanding. When compared to the phytoestrogens genistein and daidzein, which are both potentially present in soybeans, they have a larger tendency to bind ER. The ER receptor is highly responsive to ERB-041, which is a powerful and selective agonist. The goal of this study was to establish whether or whether there is a link between the chemicals genistein, daidzein, and ERB-041 and ovarian cancer. It has been demonstrated that the optimal dosages of genistein, daidzein, and ERB-041 for combating ovarian cancer cell lines are therapeutic levels of these three compounds. In order to evaluate migration, invasion, proliferation, cell cycle arrest, and apoptosis, respectively, the Transwell migration and invasion tests, XTT assays, focus formation, flow cytometry, and sphere formation assays were used. The downstream signaling pathways were traced using immunoblot analysis, which made this investigation possible.[Citation103]

According to the findings of a study, genistein, daidzein, and ERB-041 were able to suppress the capacity of ovarian cancer cells to significantly migrate, invade, and multiply. During the treatment process, these substances not only stopped the cycle of reproduction of the cell, but they also killed any cancer cells that they came into contact with. When the cells were treated with genistein, daidzein, or ERB-041, both the size of the spheres that were produced as well as the quantity of the spheres that were produced were significantly reduced. The combination of genistein, daidzein, and ERB-041 caused a reduction in the expression of p-FAK, p-PI3K, p-AKT, p-GSK3, p21, and cyclin D1 in ovarian cancer cells.[Citation104]

The connection between the estrogen receptor (ER) and the insulin-like growth factor-1 receptor (IGF-1 R) signaling pathway has a significant impact not only on the progression of estrogen-dependent cancers but also on their resistance to endocrine therapy. This interaction is responsible for both of these outcomes. The synthesis of components of the IGF-1 system is stimulated when estrogen (E2) is present in the body. Phosphorylation of the insulin receptor substrate-1 (IRS-1) and the initiation of mitogenic signaling cascades are also necessary steps in achieving this goal. Xenoestrogenic activity of bisphenol A (BPA) and the antiproliferative activity of genistein interacted with one another, as well as the impacts that these two activities had on one another. It was shown that BPA had an influence on this connection by elevating the expression of ER and IGF-1 R mRNA in BG-1 ovarian cancer cells, in addition to the phosphorylation of Akt and IRS-1 protein. The administration of BPA to the mouse model in which BG-1 cells were xenografted resulted in a significant increase in the volume of the tumor, as well as a noteworthy rise in the levels of ER, pIRS-1, and cyclin D1 expression inside the mass of the tumor. According to these data, it would appear that BPA promotes the ER and IGF-1 R signal crosstalk that, in turn, promotes the development of ovarian cancer. By inhibiting the production of ER, IGF-1 R, pIRS-1, and pAkt mRNA and protein expressions that were caused by BPA in a cell model, genistein was able to successfully combat the estrogenic effects of BPA. The purpose was to ascertain whether or not genistein was successful. This investigation set out to determine whether or not BPA interacts negatively with estrogen receptors in the body. In addition to this, it was demonstrated that the expressions of ER, pIRS-1, and pAkt were significantly reduced in vivo in a mouse model that made use of xenografts. In addition to this, it was proven that the capability of genistein to trigger apoptotic signaling cascades had an effect that was antiproliferative in nature. genistein inhibited the ability of the ER and IGF-1 R signaling pathways to communicate with one another after being triggered by either BPA or E2.[Citation105]

According to the results of a study, both E2 and BPA are capable of activating genes that are involved in epithelial-to-mesenchymal transition (EMT) and cell migration. They do this in a very methodical way by inhibiting the expression of E-cadherin while concurrently enhancing the protein synthesis of vimentin, cathepsin D, and MMP-2 via the ER signaling pathway. In a scratch experiment, the effect that E2, BPA, and NP had on the ability of BG-1 cells to migrate as a result of ER signaling was diminished when the cells were exposed to genistein. The expression of the proteins vimentin, cathepsin D, and MMP-2 had risen as a result of E2, BPA, or NP. However, the presence of genistein decreased the expression of these proteins. The proteins SnoN and Smad3 have different levels of expression in the cell after being exposed to E2, BPA, or NP. SnoN and pSmad3 are both engaged in the TGF signaling route; however, whereas SnoN is a negative regulator of the TGF signaling pathway, pSmad3 is a transcription factor in the downstream pathway. Both of these proteins are involved in the TGF signaling route. This reveals that in order for ER signaling to cause EMT and migration in BG-1 cells, simultaneous inhibition of TGF signaling by E2, BPA, and NP is required. NP is responsible for the suppression of TGF signaling. Radiation therapy is the most common cause of premature ovarian failure (POF); however, co-treatment of estrogenic medicines with genistein was able to counteract the estrogen-induced downregulation of TGF-signaling.[Citation106]

Impact of genistein was investigated, on the ovarian reserve of rats that had been radio-irradiated. After subjecting female Sprague-Dawley rats to 3.2 Gy of radiation to cause premature ovarian failure, the rats were then randomly randomized to receive either genistein (5 mg/kg, i.p.) or ethinyl estradiol (E2; 0.1 mg/kg, subcutaneously) once a day for 10 days. Both treatments were continued until the end of the experiment. The treatments were administered to the animals in a manner that was completely arbitrary. The patient’s levels of circulating estradiol and anti-mullerian hormone had dropped as a result of radiation therapy, but genistein was able to bring them back up. Glutathione peroxidase activity and endogenous glutathione levels are increased as a result of exposure to genistein. The number of primordial and developing follicles did not decrease. However, the number of atretic follicles was dramatically reduced. Genistein was responsible for the inhibition of the intrinsic apoptotic pathway, which in turn led to a decrease in the expression of both cytochrome c and caspase 3. Genistein was able to achieve this objective by suppressing the expression of Bax while concurrently boosting the expression of Bcl-2 in the cell. These encouraging findings from genistein are linked to an acceleration in the rate of granulosa cell proliferation. By increasing the expression of ER and FOXL-2 while concurrently decreasing the quantity of TGF, genistein was successful in preventing the transition of primordial follicles into more mature follicles at the molecular level.[Citation107] Recent studies of genistein against different cancers were shown in .

Table 1. Recent studies of genistein against different cancers.

Genistein may be an option worth considering as a kind of treatment for female cancer survivors who are concerned about maintaining ovarian function. Cancer stem cells, also known as CSCs and commonly shortened as CSCs, play an essential role in the development and progression of cancer. The aberrant expression of the AKT, ERK, and NF-κB signaling pathways has been associated with a variety of the distinct CSC subtypes. In the current investigation, it was looked into the potential effects that DFOG (synthetic genistein analogue: 7-difluromethoxyl-5,4’-di-n-octylgenistein) may have on a number of different signaling pathways in order to determine what sort of an effect it could have. It was found that OVCSLC (ovarian cancer stem-like cells) characteristics were present in spheroids generated from the SKOV3 cell line, and that the addition of DFOG resulted in a significant reduction in the stemness of the OVCSLCs. Additionally, the suppression of NF-κB activity as well as the phosphorylation of AKT and ERK1/2 proteins in OVCSLCs produced from SKOV3 cells were connected with the reduction of spheroid and colony formation that was brought about by DFOG. By activating FoxO3a and/or inactivating FoxM1 and targeting a spectrum of pro-survival (AKT and ERK1/2) and pro-inflammatory (NF-κB) pathways, DFOG was able to limit the oncogenicity of OVCSLCs.[Citation108] showed that genistein (GNST) as regulator of cellular pathways involved in cancers.

Figure 2. Genistein (GNST) as regulator of cellular pathways involved in cancers.

Figure 2. Genistein (GNST) as regulator of cellular pathways involved in cancers.

Conclusion

It has been shown that the natural isoflavone genistein, which can be found in legumes and soybeans, has the ability to suppress the growth of a wide range of cancers. This research is still ongoing. Genistein, according to the findings of several studies, has the ability to thwart the development of cancerous cells. Its several modes of action give it a great deal of promise as a possible strategy for the treatment as well as the prevention of cancer. This makes it an attractive candidate for research. It has been demonstrated that genistein can considerably reduce the chance of getting breast cancer. This is accomplished via genistein’s ability to prevent the development of cancer cells, induce programmed cell death also known as apoptosis, and lower the production of new blood vessels also known as angiogenesis. The ability of this compound to modify hormonal signaling pathways, in particular the estrogen receptor pathway, results in a reduction in the likelihood of the development of a hormone-dependent subtype of breast cancer. It has been established that the chemotherapy drugs that are used to treat breast cancer can benefit from an increase in genistein’s ability to improve both their effectiveness and their safety. It has the ability to limit the activity of essential enzymes in a way that leads to the development of prostate cancer, as well as anti-inflammatory qualities.

In addition to this, it blocks the signaling of androgen receptors, which halts the growth of cancer and keeps it from spreading. This is one of the ways that it combats the disease. When radiation therapy is combined with genistein treatment for prostate cancer, there is a chance that the overall effectiveness of the treatment for prostate cancer will improve. Inhibiting the growth of cancer cells, inducing programmed cell death (apoptosis), and reducing the production of new blood vessels are all characteristics of colorectal cancer, one of the most frequent kinds of the illness. Genistein does all three of these things. In addition, it limits the formation of new blood vessels. It is possible that the effect that it has on a wide variety of molecular targets, including as signaling cascades and growth factor receptors, is responsible for at least some of the anticancer qualities that it possesses. Furthermore, there is evidence to show that genistein can prevent the progression of metastatic colorectal cancer. This is in addition to the previously mentioned benefits. Because it prevents the production of new blood vessels (angiogenesis), produces programmed cell death (apoptosis), and slows the rate of cancer cell proliferation, genistein may be effective in the treatment of lung cancer.

As a result of its anti-inflammatory and antioxidant qualities, it contributes to the protection of lung tissue against the effects that carcinogens, which may or may not cause cancer, may have. Furthermore, genistein improves the efficiency of the chemotherapy drugs that are administered to patients who have lung cancer in order to treat the disease. In addition, there is evidence to suggest that genistein can limit the proliferation of malignant cells in the gastrointestinal tract, the cervix, and the ovaries. These are the organs that are most commonly affected by cancer. As a consequence of undergoing this treatment, there is an increase in the expression of tumor suppressor proteins, there is a reduction in the activity of oncogenic signaling pathways, and there is an alteration in the activity of cell cycle regulators. According to the findings of some preliminary research, genistein may also be effective in minimizing the bad effects of chemotherapy and overcoming drug resistance. This information comes from the research.

For the purposes of genistein therapy, further research is necessary in order to discover the most effective dose, treatment strategy, and patient selection. Taking into consideration all of the evidence at hand is the only way to successfully do this task. A number of factors, including individual variations in metabolic processes, the influence of genetic variables, and the possibility of drug-genistein interactions, are among those that need to be taken into consideration. The use of genistein as a natural anticancer drug for the treatment of a wide range of malignancies shows a lot of promise and might be very beneficial in the future. This specific field of research is particularly promising for a variety of reasons, some of which include its versatility in targeting a range of various pathways implicated in the growth of cancer, its compatibility with more conventional treatments, and its ability to overcome drug resistance. There is a need for more research in order to achieve a comprehensive understanding of the role that genistein plays in the treatment of cancer patients, the prevention of cancer, and the overall management of cancer. Genistein’s potential as a therapeutic drug will receive a boost as a result of this development.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Correction Statement

This article has been republished with minor changes. These changes do not impact the academic content of the article.

References

  • Rasheed, S.; Rehman, K.; Shahid, M.; Suhail, S.; Akash, M. S. Therapeutic Potentials of Genistein: New Insights and Perspectives. J. Food Biochem. 2022, 46(9), e14228. DOI: 10.1111/jfbc.14228.
  • Tuli, H. S.; Tuorkey, M. J.; Thakral, F.; Sak, K.; Kumar, M.; Sharma, A. K.; Sharma, U.; Jain, A.; Aggarwal, V.; Bishayee, A. Molecular Mechanisms of Action of Genistein in Cancer: Recent Advances. Front. Pharmacol. 2019, 10, 1336. DOI: 10.3389/fphar.2019.01336.
  • Yang, Z.; Kulkarni, K.; Zhu, W.; Hu, M. Bioavailability and Pharmacokinetics of Genistein: Mechanistic Studies on Its ADME. Anti-Cancer Agents Med. Chem. (Formerly Current Medicinal Chemistry-Anti-Cancer Agents). 2012, 12(10), 1264–1280. DOI: 10.2174/187152012803833107.
  • Sethi, G.; Rath, P.; Chauhan, A.; Ranjan, A.; Choudhary, R.; Ramniwas, S.; Sak, K.; Aggarwal, D.; Rani, I.; Tuli, H. S. Apoptotic Mechanisms of Quercetin in Liver Cancer: Recent Trends and Advancements. Pharmaceutics. 2023, 15(2), 712. DOI: 10.3390/pharmaceutics15020712.
  • Pawlicka, M. A.; Zmorzyński, S.; Popek-Marciniec, S.; Filip, A. A. The Effects of Genistein at Different Concentrations on MCF-7 Breast Cancer Cells and BJ Dermal Fibroblasts. Int. J. Mol. Sci. 2022, 23(20), 12360. DOI: 10.3390/ijms232012360.
  • Hu, C.; Wang, M.; Hu, M.; Ma, S.; Yang, B.; Xiao, W.; Zhou, Q.; Zhou, M.; Li, Z. Genistein Induces Endocrine Resistance in Human Breast Cancer by Suppressing H3K27 Trimethylation. Endocr. Relat. Cancer. 2023, 30(2), e220191. DOI: 10.1530/ERC-22-0191.
  • Mukund, V. G. Genistein: Its Role in Breast Cancer Growth and Metastasis. Curr. Drug Metab. 2020, 21(1), 6–10. DOI: 10.2174/1389200221666200120121919.
  • Chen, B.; Wu, L.; Tang, X.; Wang, T.; Wang, S.; Yu, H.; Wan, G.; Xie, M.; Zhang, R.; Xiao, H., et al. Quercetin Inhibits Tumorigenesis of Colorectal Cancer Through Downregulation of hsa_circ_0006990. Front Pharmacol. 2022, 13, 874696. DOI: 10.3389/fphar.2022.874696.
  • Malik, P.; Singh, R.; Kumar, M.; Malik, A.; Mukherjee, T. K. Understanding the Phytoestrogen Genistein Actions on Breast Cancer: Insights on Estrogen Receptor Equivalence, Pleiotropic Essence andEmerging Paradigms in Bioavailability Modulation. CTMC. 2023, 23(15), 1395–1413. DOI: 10.2174/1568026623666230103163023.
  • Bhat, S. S.; Prasad, S. K.; Shivamallu, C.; Prasad, K. S.; Syed, A.; Reddy, P.; Cull, C. A.; Amachawadi, R. G. Genistein: A Potent Anti-Breast Cancer Agent. Curr. Issues Mol. Biol. 2021, 43(3), 1502–1517. DOI: 10.3390/cimb43030106.
  • Kim, G. Y.; Suh, J.; Jang, J. H.; Kim, D. H.; Park, O. J.; Park, S. K.; Surh, Y. J. Genistein Inhibits Proliferation of BRCA1 Mutated Breast Cancer Cells: The GPR30-Akt Axis as a Potential Target. J. Cancer Prev. 2019, 24(4), 197–207. DOI: 10.15430/JCP.2019.24.4.197.
  • Ma, X.; Yu, X.; Min, J.; Chen, X.; Liu, R.; Cui, X.; Cheng, J.; Xie, M.; Diel, P.; Hu, X. Genistein Interferes with Antitumor Effects of Cisplatin in an Ovariectomized Breast Cancer Xenograft Tumor Model. Toxicol. Lett. 2022, 355, 106–115. DOI: 10.1016/j.toxlet.2021.11.013.
  • Mukund, V.; Saddala, M. S.; Farran, B.; Mannavarapu, M.; Alam, A.; Nagaraju, G. P. Molecular Docking Studies of Angiogenesis Target Protein HIF-1α and Genistein in Breast Cancer. Gene. 2019, 701, 169–172. DOI: 10.1016/j.gene.2019.03.062.
  • Kaushik, S.; Shyam, H.; Agarwal, S.; Sharma, R.; Nag, T. C.; Dwivedi, A. K.; Balapure, A. K. Genistein Potentiates Centchroman Induced Antineoplasticity in Breast Cancer via PI3K/Akt Deactivation and ROS Dependent Induction of Apoptosis. Life. sci. 2019, 239, 117073. DOI: 10.1016/j.lfs.2019.117073.
  • Jiang, H.; Fan, J.; Cheng, L.; Hu, P.; Liu, R. The anticancer activity of genistein is increased in estrogen receptor beta 1-positive breast cancer cells. Onco. Targets Ther. 2018, 11, 8153–8163. DOI: 10.2147/OTT.S182239.
  • Li, T.; Kuang, T.; Yang, Z.; Zhang, Q.; Zhang, W.; Fan, Y. Co-Treatment with Everolimus, an MTOR-Specific Antagonist, or Downregulation of ELK1 Enhances the Sensitivity of Pancreatic Cancer Cells to Genistein. Front Cell Dev Biol. 2021, 9, 633035. DOI: 10.3389/fcell.2021.633035.
  • Mesmar, F.; Dai, B.; Ibrahim, A.; Hases, L.; Jafferali, M. H.; Jose Augustine, J.; DiLorenzo, S.; Kang, Y.; Zhao, Y.; Wang, J.; et al. Clinical Candidate and Genistein Analogue AXP107-11 Has Chemoenhancing Functions in Pancreatic Adenocarcinoma Through G Protein-Coupled Estrogen Receptor Signaling. Cancer Med. 2019, 8(18), 7705–7719.
  • Ma, C. H.; Zhang, Y. X.; Tang, L. H.; Yang, X. J.; Cui, W. M.; Han, C. C.; Ji, W. Y. MicroRNA-1469, a p53-Responsive microRna Promotes Genistein Induced Apoptosis by Targeting Mcl1 in Human Laryngeal Cancer Cells. Biomed. Pharmacother. 2018, 106, 665–671. DOI: 10.1016/j.biopha.2018.07.005.
  • Xia, J.; Cheng, L.; Mei, C.; Ma, J.; Shi, Y.; Zeng, F.; Wang, Z.; Wang, Z. Genistein Inhibits Cell Growth and Invasion Through Regulation of MiR-27a in Pancreatic Cancer Cells. Curr. Pharm. Des. 2014, 20(33), 5348–5353. DOI: 10.2174/1381612820666140128215756.
  • Ma, J.; Zeng, F.; Ma, C.; Pang, H.; Fang, B.; Lian, C.; Yin, B.; Zhang, X.; Wang, Z.; Xia, J. Synergistic Reversal Effect of Epithelial-To-Mesenchymal Transition by MiR-223 Inhibitor and Genistein in Gemcitabine-Resistant Pancreatic Cancer Cells. Am. J. Cancer Res. 2016, 6(6), 1384–1395.
  • Ma, J.; Cheng, L.; Liu, H.; Zhang, J.; Shi, Y.; Zeng, F.; Miele, L.; Sarkar, F. H.; Xia, J.; Wang, Z. Genistein down-regulates miR-223 expression in pancreatic cancer cells. Curr. Drug Targets. 2013, 14(10), 1150–1156. DOI: 10.2174/13894501113149990187.
  • Ji, X.; Liu, K.; Li, Q.; Shen, Q.; Han, F.; Ye, Q.; Zheng, C. A Mini-Review of Flavone Isomers Apigenin and Genistein in Prostate Cancer Treatment. Front Pharmacol. 2022, 13, 851589. DOI: 10.3389/fphar.2022.851589.
  • Shafiee, G.; Saidijam, M.; Tayebinia, H.; Khodadadi, I. Beneficial Effects of Genistein in Suppression of Proliferation, Inhibition of Metastasis, and Induction of Apoptosis in PC3 Prostate Cancer Cells. Arch. Physiol. Biochem. 2022, 128(3), 694–702. DOI: 10.1080/13813455.2020.1717541.
  • Khamesi, S. M.; Salehi Barough, M.; Zargan, J.; Shayesteh, M.; Banaee, N.; Haji Noormohammadi, A.; Keshavarz Alikhani, H.; Mousavi, M. Evaluation of Anticancer and Cytotoxic Effects of Genistein on PC3 Prostate Cell Line Under Three-Dimensional Culture Medium. Iran. Biomed. J. 2022, 26(5), 380–388. DOI: 10.52547/ibj.3711.
  • Yu, X.; Yan, J.; Li, Y.; Cheng, J.; Zheng, L.; Fu, T.; Zhu, Y. Inhibition of Castration-Resistant Prostate Cancer Growth by Genistein Through Suppression of AKR1C3. Food & Nutrition Research. 2023, 67. DOI: 10.29219/fnr.v67.9024.
  • Tang, Q.; Ma, J.; Sun, J.; Yang, L.; Yang, F.; Zhang, W.; Li, R.; Wang, L.; Wang, Y.; Wang, H. Genistein and AG1024 Synergistically Increase the Radiosensitivity of Prostate Cancer Cells. Oncol. Rep. 2018, 40(2), 579–588. DOI: 10.3892/or.2018.6468.
  • Batra, N.; Sam, A.; Woldemariam, T.; Talbott, G.; de Vere White, R. W.; Ghosh, P. M.; Gaikwad, N. W.; Kotchoni, S. O.; Vinall, R. L. Genistein Combined Polysaccharide (GCP) Can Inhibit Intracrine Androgen Synthesis in Prostate Cancer Cells. Biomedicines. 2020, 8(8), 282. DOI: 10.3390/biomedicines8080282.
  • Tian, T.; Li, J.; Li, B.; Wang, Y.; Li, M.; Ma, D.; Wang, X. Genistein exhibits anti-cancer effects via down-regulating FoxM1 in H446 small-cell lung cancer cells. Tumour Biol. 2014, 35(5), 4137–4145. DOI: 10.1007/s13277-013-1542-0.
  • Song, Y. Y.; Yuan, Y.; Shi, X.; Che, Y. Y. Improved Drug Delivery and Anti-Tumor Efficacy of Combinatorial Liposomal Formulation of Genistein and Plumbagin by Targeting Glut1 and Akt3 Proteins in Mice Bearing Prostate Tumor. Colloids Surf. B Biointerfaces. 2020, 190, 110966. DOI: 10.1016/j.colsurfb.2020.110966.
  • Lee, S. R.; Kwon, S. W.; Lee, Y. H.; Kaya, P.; Kim, J. M.; Ahn, C.; Jung, E. M.; Lee, G. S.; An, B. S.; Jeung, E. B.; et al. Dietary Intake of Genistein Suppresses Hepatocellular Carcinoma Through AMPK-Mediated Apoptosis and Anti-Inflammation. BMC Cancer. 2019, 19(1), 6. DOI: 10.1186/s12885-018-5222-8.
  • Yang, X.; Jiang, W.; Kong, X.; Zhou, X.; Zhu, D.; Kong, L. Genistein Restricts the Epithelial Mesenchymal Transformation (EMT) and Stemness of Hepatocellular Carcinoma via Upregulating MiR-1275 to Inhibit the EIF5A2/PI3K/Akt Pathway. Biol. 2022, 11(10), 1383. DOI: 10.3390/biology11101383.
  • Sanaei, M.; Kavoosi, F.; Roustazadeh, A.; Golestan, F. Effect of Genistein in Comparison with Trichostatin a on Reactivation of DNMTs Genes in Hepatocellular Carcinoma. J. Clin. Transl. Hepatol. 2018, 6(2), 141–146. DOI: 10.14218/JCTH.2018.00002.
  • Dai, W.; Wang, F.; He, L.; Lin, C.; Wu, S.; Chen, P.; Zhang, Y.; Shen, M.; Wu, D.; Wang, C., et al. Genistein Inhibits Hepatocellular Carcinoma Cell Migration by Reversing the Epithelial-Mesenchymal Transition: Partial Mediation by the Transcription Factor NFAT1. Mol. Carcinog. 2015, 54(4), 301–311. DOI: 10.1002/mc.22100.
  • Lepri, S. R.; Sartori, D.; Semprebon, S. C.; Baranoski, A.; Coatti, G. C.; Mantovani, M. S. Genistein Affects Expression of Cytochrome P450 (CYP450) Genes in Hepatocellular Carcinoma (HEPG2/C3A) Cell Line. Drug Metab. Lett. 2018, 12(2), 138–144. DOI: 10.2174/1872312812666180709150440.
  • Sanaei, M.; Kavoosi, F.; Valiani, A.; Ghobadifar, M. A. Effect of Genistein on Apoptosis and Proliferation of Hepatocellular Carcinoma Hepa1-6 Cell Line. Int. J. Prev. Med. 2018, 9(1), 12. DOI: 10.4103/ijpvm.IJPVM_249_16.
  • Dastjerdi, M. N.; Kavoosi, F.; Valiani, A.; Esfandiari, E.; Sanaei, M.; Sobhanian, S.; Hakemi, M. G.; Mobarakian, M. Inhibitory Effect of Genistein on PLC/PRF5 Hepatocellular Carcinoma Cell Line. Int. J. Prev. Med. 2015, 6(1), 54. DOI: 10.4103/2008-7802.158914.
  • Roh, T.; Kim, S. W.; Moon, S. H.; Nam, M. J. Genistein induces apoptosis by down-regulating thioredoxin-1 in human hepatocellular carcinoma SNU-449 cells. Food. Chem. Toxicol. 2016, 97, 127–134. DOI: 10.1016/j.fct.2016.09.003.
  • Yu, D.; Shin, H. S.; Lee, Y. S.; Lee, D.; Kim, S.; Lee, Y. C. Genistein Attenuates Cancer Stem Cell Characteristics in Gastric Cancer Through the Downregulation of Gli1. Oncol. Rep. 2014, 31(2), 673–678. DOI: 10.3892/or.2013.2893.
  • Huang, W.; Wan, C.; Luo, Q.; Huang, Z.; Luo, Q. Genistein-Inhibited Cancer Stem Cell-Like Properties and Reduced Chemoresistance of Gastric Cancer. Int. J. Mol. Sci. 2014, 15(3), 3432–3443. DOI: 10.3390/ijms15033432.
  • Hou, S. G. Genistein: Therapeutic and Preventive Effects, Mechanisms, and Clinical Application in Digestive Tract Tumor. Evid. Based Complement. Alternat. Med. 2022, 2022, 1–10. DOI: 10.1155/2022/5957378.
  • Moreno-Quintero, G.; Castrillón-Lopez, W.; Herrera-Ramirez, A.; Yepes-Pérez, A. F.; Quintero-Saumeth, J.; Cardona-Galeano, W. Synthesis and Chemopreventive Potential of 5-Fu/genistein Hybrids on Colorectal Cancer Cells. Pharm. 2022, 15(10), 1299. DOI: 10.3390/ph15101299.
  • Alorda-Clara, M.; Torrens-Mas, M.; Morla-Barcelo, P. M.; Roca, P.; Sastre-Serra, J.; Pons, D. G.; Oliver, J. High Concentrations of Genistein Decrease Cell Viability Depending on Oxidative Stress and Inflammation in Colon Cancer Cell Lines. Int. J. Mol. Sci. 2022, 23(14), 7526. DOI: 10.3390/ijms23147526.
  • Castaño, M.; Martínez, E.; Osorio, M.; Castro, C. Development of Genistein Drug Delivery Systems Based on Bacterial Nanocellulose for Potential Colorectal Cancer Chemoprevention: Effect of Nanocellulose Surface Modification on Genistein Adsorption. Molecules. 2022, 27(21), 7201. DOI: 10.3390/molecules27217201.
  • Liang, Y. S.; Qi, W. T.; Guo, W.; Wang, C. L.; Hu, Z. B.; Li, A. K. Genistein and Daidzein Induce Apoptosis of Colon Cancer Cells by Inhibiting the Accumulation of Lipid Droplets. Food Nutr. Res. 2018, 62. DOI: 10.29219/fnr.v62.1384.
  • Pool, H.; Campos-Vega, R.; Herrera-Hernández, M. G.; García-Solis, P.; García-Gasca, T.; Sánchez, I. C.; Luna-Bárcenas, G.; Vergara-Castañeda, H. Development of Genistein-PEGylated Silica Hybrid Nanomaterials with Enhanced Antioxidant and Antiproliferative Properties on HT29 Human Colon Cancer Cells. Am. J. Transl. Res. 2018, 10(8), 2306–2323.
  • Zhou, P.; Wang, C.; Hu, Z.; Chen, W.; Qi, W.; Li, A. Genistein Induces Apoptosis of Colon Cancer Cells by Reversal of Epithelial-To-Mesenchymal via a Notch1/nf-Κb/slug/e-Cadherin Pathway. BMC Cancer. 2017, 17(1), 813. DOI: 10.1186/s12885-017-3829-9.
  • Shafiee, G.; Saidijam, M.; Tavilani, H.; Ghasemkhani, N.; Khodadadi, I. Genistein Induces Apoptosis and Inhibits Proliferation of HT29 Colon Cancer Cells. Int. J. Mol. Cell Med. 2016, 5(3), 178–191.
  • Zhang, Z.; Wang, C. Z.; Du, G. J.; Qi, L. W.; Calway, T.; He, T. C.; Du, W.; Yuan, C. S. Genistein Induces G2/M Cell Cycle Arrest and Apoptosis via ATM/p53-Dependent Pathway in Human Colon Cancer Cells. Int. J. Oncol. 2013, 43(1), 289–296. DOI: 10.3892/ijo.2013.1946.
  • Han, J.; Kurita, Y.; Isoda, H. Genistein-Induced G2/M Cell Cycle Arrest of Human Intestinal Colon Cancer Caco-2 Cells is Associated with Cyclin B1 and Chk2 Down-Regulation. Cytotechnology. 2013, 65(6), 973–978. DOI: 10.1007/s10616-013-9592-0.
  • Lepri, S. R.; Zanelatto, L. C.; da Silva, P. B.; Sartori, D.; Ribeiro, L. R.; Mantovani, M. S. Effects of Genistein and Daidzein on Cell Proliferation Kinetics in HT29 Colon Cancer Cells: The Expression of CTNNBIP1 (β-Catenin), APC (Adenomatous Polyposis Coli) and BIRC5 (Survivin). Hum. Cell. 2014, 27(2), 78–84. DOI: 10.1007/s13577-012-0051-6.
  • Qin, J.; Chen, J. X.; Zhu, Z.; Teng, J. A. Genistein Inhibits Human Colorectal Cancer Growth and Suppresses MiR-95, Akt and SGK1. Cell. Physiol. Biochem. 2015, 35(5), 2069–2077. DOI: 10.1159/000374013.
  • Hussein, A. M.; Attaai, A. H.; Zahran, A. M. Genistein anticancer efficacy during induced oral squamous cell carcinoma: an experimental study. J. Egypt. Natl. Canc. Inst. 2022, 34(1), 37. DOI: 10.1186/s43046-022-00140-5.
  • Dev, A.; Sardoiwala, M. N.; Kushwaha, A. C.; Karmakar, S.; Choudhury, S. R. Genistein Nanoformulation Promotes Selective Apoptosis in Oral Squamous Cell Carcinoma Through Repression of 3PK-EZH2 Signalling Pathway. Phytomedicine. 2021, 80, 153386. DOI: 10.1016/j.phymed.2020.153386.
  • Bae, M. G.; Hwang-Bo, J.; Lee, D. Y.; Lee, Y. H.; Chung, I. S. Effects of 6,8-Diprenylgenistein on VEGF-A-Induced Lymphangiogenesis and Lymph Node Metastasis in an Oral Cancer Sentinel Lymph Node Animal Model. Int. J. Mol. Sci. 2021, 22(2), 770. DOI: 10.3390/ijms22020770.
  • Hsieh, P. L.; Liao, Y. W.; Hsieh, C. W.; Chen, P. N.; Yu, C. C. Soy Isoflavone Genistein Impedes Cancer Stemness and Mesenchymal Transition in Head and Neck Cancer Through Activating MiR-34a/RTCB Axis. Nutrients. 2020, 12(7), 1924. DOI: 10.3390/nu12071924.
  • Ardito, F.; Di Gioia, G.; Pellegrino, M. R.; Muzio, L. L. Genistein as a Potential Anticancer Agent Against Head and Neck Squamous Cell Carcinoma. Curr. Top. Med. Chem. 2018, 18(3), 174–181. DOI: 10.2174/1568026618666180116122650.
  • Zhang, C.; Lv, B.; Yi, C.; Cui, X.; Sui, S.; Li, X.; Qi, M.; Hao, C.; Han, B.; Liu, Z. Genistein Inhibits Human Papillary Thyroid Cancer Cell Detachment, Invasion and Metastasis. J. Cancer. 2019, 10(3), 737–748. DOI: 10.7150/jca.28111.
  • Schroeter, A.; Aichinger, G.; Stornig, K.; Marko, D. Impact of Oxidative Metabolism on the Cytotoxic and Genotoxic Potential of Genistein in Human Colon Cancer Cells. Mol. Nutr. Food Res. 2019, 63(2), e1800635. DOI: 10.1002/mnfr.201800635.
  • Xi, P.; Zhang, S. The Research of Apoptosis and Proliferation Inhibition of Human Laryngeal Carcinoma Cell Line Hep-2 Induced by Genistein. Lin Chung Er Bi Yan Hou Tou Jing Wai Ke Za Zhi. 2015, 29(7), 638–641.
  • Chen, X.; Wu, Y.; Gu, J.; Liang, P.; Shen, M.; Xi, J.; Qin, J. Anti-invasive effect and pharmacological mechanism of genistein against colorectal cancer. BioFactors. 2020, 46(4), 620–628. DOI: 10.1002/biof.1627.
  • Sundaram, M. K.; Ansari, M. Z.; Al Mutery, A.; Ashraf, M.; Nasab, R.; Rai, S.; Rais, N.; Hussain, A. Genistein Induces Alterations of Epigenetic Modulatory Signatures in Human Cervical Cancer Cells. Anti Cancer Agents Med. Chem. 2018, 18(3), 412–421. DOI: 10.2174/1871520617666170918142114.
  • Liu, X.; Li, P.; Hirayama, R.; Niu, Y.; Liu, X.; Chen, W.; Jin, X.; Zhang, P.; Ye, F.; Zhao, T., et al. Genistein Sensitizes Glioblastoma Cells to Carbon Ions via Inhibiting DNA-PKcs Phosphorylation and Subsequently Repressing NHEJ and Delaying HR Repair Pathways. Radiother. Oncol. 2018, 129(1), 84–94. DOI: 10.1016/j.radonc.2018.04.005.
  • Yoriki, K.; Mori, T.; Aoyama, K.; Tarumi, Y.; Kataoka, H.; Kokabu, T.; Kitawaki, J. Genistein Induces Long-Term Expression of Progesterone Receptor Regardless of Estrogen Receptor Status and Improves the Prognosis of Endometrial Cancer Patients. Sci. Rep. 2022, 12(1), 10303. DOI: 10.1038/s41598-022-13842-6.
  • Carbonel, A. A.; Calió, M. L.; Santos, M. A.; Bertoncini, C. R.; Sasso Gda, S.; Simões, R. S.; Simões, M. J.; JM, S., Jr. Soybean Isoflavones Attenuate the Expression of Genes Related to Endometrial Cancer Risk. Climacteric. 2015, 18(3), 389–398. DOI: 10.3109/13697137.2014.964671.
  • Buathong, N.; Poonyachoti, S.; Deachapunya, C. Isoflavone Genistein Modulates the Protein Expression of Toll-like Receptors in Cancerous Human Endometrial Cells. J. Med. Assoc. Thai. 2015, 98(Suppl 9), S31–8.
  • Chen, Y.; Bai, J.; Xu, J.; Song, X. Effects of 7-Difluoromethy-5, 4’-Dimethoxygenistein on Proliferation and Apoptosis of Human Cervical Cancer Cells and Its Mechanism. Zhong Nan Da Xue Xue Bao Yi Xue Ban. 2016, 41(5), 463–470. DOI: 10.11817/j.issn.1672-7347.2016.05.003.
  • Naciff, J. M.; Khambatta, Z. S.; Carr, G. J.; Tiesman, J. P.; Singleton, D. W.; Khan, S. A.; Daston, G. P. Dose- and Time-Dependent Transcriptional Response of Ishikawa Cells Exposed to Genistein. Toxicol. Sci. 2016, 151(1), 71–87. DOI: 10.1093/toxsci/kfw024.
  • Xu, H.; Ma, H.; Zha, L.; Li, Q.; Pan, H.; Zhang, L. Genistein Promotes Apoptosis of Lung Cancer Cells Through the IMPDH2/AKT1 Pathway. Am. J. Transl. Res. 2022, 14(10), 7040–7051.
  • Fu, Z.; Cao, X.; Liu, L.; Cao, X.; Cui, Y.; Li, X.; Quan, M.; Ren, K.; Chen, A.; Xu, C., et al. Genistein Inhibits Lung Cancer Cell Stem-Like Characteristics by Modulating MnSod and FoxM1 Expression. Oncol. Lett. 2020, 20(3), 2506–2515. DOI: 10.3892/ol.2020.11802.
  • Yu, Y.; Xing, Y.; Zhang, Q.; Zhang, Q.; Huang, S.; Li, X.; Gao, C. Soy isoflavone genistein inhibits hsa_circ_0031250/miR-873-5p/FOXM1 axis to suppress non-small-cell lung cancer progression. IUBMB Life. 2021, 73(1), 92–107. DOI: 10.1002/iub.2404.
  • Yang, Y.; Zang, A.; Jia, Y.; Shang, Y.; Zhang, Z.; Ge, K.; Zhang, J.; Fan, W.; Wang, B. Genistein Inhibits A549 Human Lung Cancer Cell Proliferation via MiR-27a and MET Signaling. Oncol. Lett. 2016, 12(3), 2189–2193. DOI: 10.3892/ol.2016.4817.
  • Wu, T. C.; Lin, Y. C.; Chen, H. L.; Huang, P. R.; Liu, S. Y.; Yeh, S. L. The Enhancing Effect of Genistein on Apoptosis Induced by Trichostatin a in Lung Cancer Cells with Wild Type p53 Genes is Associated with Upregulation of Histone Acetyltransferase. Toxicol. Appl. Pharmacol. 2016, 292, 94–102. DOI: 10.1016/j.taap.2015.12.028.
  • Kamel, N. M.; Helmy, M. W.; Abdelfattah, E. Z.; Khattab, S. N.; Ragab, D.; Samaha, M. W.; Fang, J. Y.; Elzoghby, A. O. Inhalable Dual-Targeted Hybrid Lipid Nanocore-Protein Shell Composites for Combined Delivery of Genistein and All-Trans Retinoic Acid to Lung Cancer Cells. ACS Biomater. Sci. Eng. 2020, 6(1), 71–87. DOI: 10.1021/acsbiomaterials.8b01374.
  • Zhang, C.; Yang, F.; Ni, S.; Teng, W.; Ning, Y. Drug-Target-Disease Network Analysis of Gene-Phenotype Connectivity for Genistein in Ovarian Cancer. Onco. Targets Ther. 2018, 11, 8901–8908. DOI: 10.2147/OTT.S183302.
  • Liu, D.; Yan, L.; Wang, L.; Tai, W.; Wang, W.; Yang, C. Genistein Enhances the Effect of Cisplatin on the Inhibition of Non-Small Cell Lung Cancer A549 Cell Growth in vitro and in vivo. Oncol. Lett. 2014, 8(6), 2806–2810. DOI: 10.3892/ol.2014.2597.
  • Yazdani, Y.; Sharifi Rad, M. R.; Taghipour, M.; Chenari, N.; Ghaderi, A.; Razmkhah, M. Genistein Suppression of Matrix Metalloproteinase 2 (MMP-2) and Vascular Endothelial Growth Factor (VEGF) Expression in Mesenchymal Stem Cell Like Cells Isolated from High and Low Grade Gliomas. Asian Pac. J. Cancer Prev. 2016, 17(12), 5303–5307. DOI: 10.22034/APJCP.2016.17.12.5303.
  • Shahmoradi, M.; Banisharif-Dehkordi, F.; Kouhihabibidehkordi, M.; GhatrehSamani, M.; Jami, M. S.; Shirzad, H.; Pourgheysari, B. Molecular Effects of Genistein, as a Potential Anticancer Agent, on CXCR-4 and VEGF Pathway in Acute Lymphoblastic Leukemia. Mol. Biol. Rep. 2022, 49(6), 4161–4170. DOI: 10.1007/s11033-022-07163-0.
  • Ghasemi Goorbandi, R.; Mohammadi, M. R.; Malekzadeh, K. Synthesizing Efficacious Genistein in Conjugation with Superparamagnetic Fe3O4 Decorated with Bio-Compatible Carboxymethylated Chitosan Against Acute Leukemia Lymphoma. Biomater. Res. 2020, 24(1), 9. DOI: 10.1186/s40824-020-00187-2.
  • Abedi Nejad, M.; Nikbakht, M.; Afsa, M.; Malekzadeh, K. Restraining the Proliferation of Acute Lymphoblastic Leukemia Cells by Genistein Through Up-Regulation of B-Cell Translocation Gene-3 at Transcription Level. Galen. Med. J. 2019, 8, e1229. DOI: 10.31661/gmj.v8i0.1229.
  • Hsiao, Y. C.; Peng, S. F.; Lai, K. C.; Liao, C. L.; Huang, Y. P.; Lin, C. C.; Lin, M. L.; Liu, K. C.; Tsai, C. C.; Ma, Y. S., et al. Genistein Induces Apoptosis in vitro and Has Antitumor Activity Against Human Leukemia HL-60 Cancer Cell Xenograft Growth in vivo. Environ. Toxicol. 2019, 34(4), 443–456. DOI: 10.1002/tox.22698.
  • Hsiao, Y. C.; Chueh, F. S.; Ma, Y. S.; Lien, J. C.; Hsia, T. C.; Huang, W. W.; Chou, Y. C.; Chen, P. Y.; Chung, J. G.; Chen, H. Y., et al. Genistein Enhances the Effects of L-Asparaginase on Inducing Cell Apoptosis in Human Leukemia Cancer HL-60 Cells. Environ. Toxicol. 2021, 36(5), 764–772. DOI: 10.1002/tox.23078.
  • Narasimhan, K.; Lee, Y. M.; Lim, T. K.; Port, S. A.; Han, J. H.; Chen, C. S.; Lin, Q. Genistein Exerts Anti-Leukemic Effects on Genetically Different Acute Myeloid Leukemia Cell Lines by Inhibiting Protein Synthesis and Cell Proliferation While Inducing Apoptosis – Molecular Insights from an iTraq™ Quantitative Proteomics Study. Oncoscience. 2015, 2(2), 111–124. DOI: 10.18632/oncoscience.120.
  • Ozturk, S. A.; Alp, E.; Yar Saglam, A. S.; Konac, E.; Menevse, E. S. The Effects of Thymoquinone and Genistein Treatment on Telomerase Activity, Apoptosis, Angiogenesis, and Survival in Thyroid Cancer Cell Lines. J. Cancer Res. Ther. 2018, 14(2), 328–334. DOI: 10.4103/0973-1482.202886.
  • Ferrari, S. M.; Antonelli, A.; Guidi, P.; Bernardeschi, M.; Scarcelli, V.; Fallahi, P.; Frenzilli, G. Genotoxicity Evaluation of the Soybean Isoflavone Genistein in Human Papillary Thyroid Cancer Cells. Study of Its Potential Use in Thyroid Cancer Therapy. Nutr. Cancer. 2019, 71(8), 1335–1344. DOI: 10.1080/01635581.2019.1604004.
  • Zhao, D.; Yin, J.; Wu, S.; Wu, Z.; Pan, H. Effects of Genistein on Proliferation and Invasion of Human Thyroid Squamous Cell Carcinoma SW579 Cells. Wei Sheng Yan Jiu. 2018, 47(6), 984–992.
  • Park, C.; Cha, H. J.; Lee, H.; Hwang-Bo, H.; Ji, S. Y.; Kim, M. Y.; Hong, S. H.; Jeong, J. W.; Han, M. H.; Choi, S. H., et al. Induction of G2/M Cell Cycle Arrest and Apoptosis by Genistein in Human Bladder Cancer T24 Cells Through Inhibition of the ROS-Dependent PI3k/Akt Signal Transduction Pathway. Antioxidants (Basel). 2019, 8(9), 327. DOI: 10.3390/antiox8090327.
  • Li, E.; Zhang, T.; Sun, X.; Li, Y.; Geng, H.; Yu, D.; Zhong, C. Sonic Hedgehog Pathway Mediates Genistein Inhibition of Renal Cancer Stem Cells. Oncol. Lett. 2019, 18(3), 3081–3091. DOI: 10.3892/ol.2019.10657.
  • Hirata, H.; Ueno, K.; Nakajima, K.; Tabatabai, Z. L.; Hinoda, Y.; Ishii, N.; Dahiya, R. Genistein Downregulates Onco-MiR-1260b and Inhibits Wnt-Signalling in Renal Cancer Cells. Br. J. Cancer. 2013, 108(10), 2070–2078. DOI: 10.1038/bjc.2013.173.
  • Imai-Sumida, M.; Dasgupta, P.; Kulkarni, P.; Shiina, M.; Hashimoto, Y.; Shahryari, V.; Majid, S.; Tanaka, Y.; Dahiya, R.; Yamamura, S. Genistein Represses HOTAIR/Chromatin Remodeling Pathways to Suppress Kidney Cancer. Cell. Physiol. Biochem. 2020, 54(1), 53–70. DOI: 10.33594/000000205.
  • Ji, Z.; Huo, C.; Yang, P. Genistein Inhibited the Proliferation of Kidney Cancer Cells via CDKN2a Hypomethylation: Role of Abnormal Apoptosis. Int. Urol. nephrol. 2020, 52(6), 1049–1055. DOI: 10.1007/s11255-019-02372-2.
  • Guo, Y.; Zhang, A.; Ding, Y.; Wang, Y.; Yuan, W. Genistein Ameliorates Parathyroid Hormone-Induced Epithelial-To-Mesenchymal Transition and Inhibits Expression of Connective Tissue Growth Factor in Human Renal Proximal Tubular Cells. Arch. Med. Sci. 2013, 9(4), 724–730. DOI: 10.5114/aoms.2013.36929.
  • Wei, D.; Yang, L.; Lv, B.; Chen, L. Genistein Suppresses Retinoblastoma Cell Viability and Growth and Induces Apoptosis by Upregulating MiR-145 and Inhibiting Its Target ABCE1. Mol. Vis. 2017, 23, 385–394.
  • Song, M.; Tian, X.; Lu, M.; Zhang, X.; Ma, K.; Lv, Z.; Wang, Z.; Hu, Y.; Xun, C.; Zhang, Z.; et al. Genistein Exerts Growth Inhibition on Human Osteosarcoma MG-63 Cells via PPARγ Pathway. Int. J. Oncol. 2015, 46(3), 1131–1140. DOI: 10.3892/ijo.2015.2829.
  • Engel, N.; Adamus, A.; Schauer, N.; Kühn, J.; Nebe, B.; Seitz, G.; Kraft, K.; Nikitovic-Tzanakaki, D. Synergistic Action of Genistein and Calcitriol in Immature Osteosarcoma MG-63 Cells by SGPL1 Up-Regulation. PLoS. One. 2017, 12(1), e0169742. DOI: 10.1371/journal.pone.0169742.
  • Sharma, M.; Arora, I.; Chen, M.; Wu, H.; Crowley, M. R.; Tollefsbol, T. O.; Li, Y. Therapeutic Effects of Dietary Soybean Genistein on Triple-Negative Breast Cancer via Regulation of Epigenetic Mechanisms. Nutrients. 2021, 13(11), 3944. DOI: 10.3390/nu13113944.
  • Sahin, K.; Yenice, E.; Bilir, B.; Orhan, C.; Tuzcu, M.; Sahin, N.; Ozercan, I. H.; Kabil, N.; Ozpolat, B.; Kucuk, O. Genistein Prevents Development of Spontaneous Ovarian Cancer and Inhibits Tumor Growth in Hen Model. Cancer Prev. Res. (Phila). 2019, 12(3), 135–146. DOI: 10.1158/1940-6207.CAPR-17-0289.
  • Wang, Y.; Li, W.; Wang, Z.; Ren, H.; Li, Y.; Zhang, Y.; Yang, P.; Pan, S. Genistein Upregulates Cyclin D1 and CDK4 Expression and Promotes the Proliferation of Ovarian Cancer OVCAR-5 Cells. Clin. Chim. Acta. 2021, 512, 100–105. DOI: 10.1016/j.cca.2019.08.023.
  • Suraweera, T. L.; Merlin, J. J.; Dellaire, G.; Xu, Z.; Rupasinghe, H. V. Genistein and Procyanidin B2 Reduce Carcinogen-Induced Reactive Oxygen Species and DNA Damage Through the Activation of Nrf2/ARE Cell Signaling in Bronchial Epithelial Cells in vitro. Int. J. Mol. Sci. 2023, 24(4), 3676. DOI: 10.3390/ijms24043676.
  • Hwang, K. A.; Park, M. A.; Kang, N. H.; Yi, B. R.; Hyun, S. H.; Jeung, E. B.; Choi, K. C. Anticancer Effect of Genistein on BG-1 Ovarian Cancer Growth Induced by 17 β-Estradiol or Bisphenol a via the Suppression of the Crosstalk Between Estrogen Receptor α and Insulin-Like Growth Factor-1 Receptor Signaling Pathways. Toxicol. Appl. Pharmacol. 2013, 272(3), 637–646. DOI: 10.1016/j.taap.2013.07.027.
  • Huang, S. L.; Chang, T. C.; Chao, C. C. K.; Sun, N. K. Role of the TLR4-Androgen Receptor Axis and Genistein in Taxol-Resistant Ovarian Cancer Cells. Biochem Pharmacol. 2020, 177, 113965. DOI: 10.1016/j.bcp.2020.113965.
  • Zhang, L.; Ma, X.; Dong, Y. Effect of Genistein on Apoptosis of Lung Adenocarcinoma A549 Cells and Expression of Apoptosis Factors. J. Buon. 2018, 23(3), 641–646.
  • Xu, L.; Xiang, J.; Shen, J.; Zou, X.; Zhai, S.; Yin, Y.; Li, P.; Wang, X.; Sun, Q. Oncogenic MicroRNA-27a is a Target for Genistein in Ovarian Cancer Cells. Anti Cancer Agents Med. Chem. 2013, 13(7), 1126–1132. DOI: 10.2174/18715206113139990006.
  • Chan, K. K. L.; Siu, M. K. Y.; Jiang, Y. X.; Wang, J. J.; Leung, T. H. Y.; Ngan, H. Y. S. Estrogen Receptor Modulators Genistein, Daidzein and ERB-041 Inhibit Cell Migration, Invasion, Proliferation and Sphere Formation via Modulation of FAK and PI3K/AKT Signaling in Ovarian Cancer. Cancer Cell Int. 2018, 18(1), 65. DOI: 10.1186/s12935-018-0559-2.
  • Rendón, J. P.; Cañas, A. I.; Correa, E.; Bedoya-Betancur, V.; Osorio, M.; Castro, C.; Naranjo, T. W. Evaluation of the Effects of Genistein in vitro as a Chemopreventive Agent for Colorectal Cancer-Strategy to Improve Its Efficiency When Administered Orally. Molecules. 2022, 27(20), 7042. DOI: 10.3390/molecules27207042.
  • Hsiao, Y. C.; Chueh, F. S.; Ma, Y. S.; Lien, J. C.; Hsia, T. C.; Huang, W. W.; Chou, Y. C.; Chen, P. Y.; Chung, J. G.; Chen, H. Y., et al. Genistein Enhances the Effects of L‐Asparaginase on Inducing Cell Apoptosis in Human Leukemia Cancer HL‐60 Cells. Environ. Toxicol. 2021, 36(5), 764–772. DOI: 10.1002/tox.23078.
  • Kim, Y. S.; Choi, K. C.; Hwang, K. A. Genistein Suppressed Epithelial-Mesenchymal Transition and Migration Efficacies of BG-1 Ovarian Cancer Cells Activated by Estrogenic Chemicals via Estrogen Receptor Pathway and Downregulation of TGF-β Signaling Pathway. Phytomedicine. 2015, 22(11), 993–999. DOI: 10.1016/j.phymed.2015.08.003.
  • Haddad, Y. H.; Said, R. S.; Kamel, R.; Morsy, E. M. E.; El-Demerdash, E. Phytoestrogen Genistein Hinders Ovarian Oxidative Damage and Apoptotic Cell Death-Induced by Ionizing Radiation: Co-Operative Role of ER-β, TGF-β, and FOXL-2. Sci. Rep. 2020, 10(1), 13551. DOI: 10.1038/s41598-020-70309-2.
  • Ning, Y.; Xu, M.; Cao, X.; Chen, X.; Luo, X. Inactivation of AKT, ERK and NF-κB by genistein derivative, 7-difluoromethoxyl-5,4’-di-n-octylygenistein, reduces ovarian carcinoma oncogenicity. Oncol. Rep. 2017, 38(2), 949–958. DOI: 10.3892/or.2017.5709.