Skip to main content

Human metabolic effects of BPA and the application of a hybrid photocatalytic membrane for BPA contaminated water

Abstract

Bisphenol A (BPA) is widely used in manufacturing various consumer products and detected in various water sources. Epidemiological studies reveal a correlation between BPA exposure and metabolic system disorders, including obesity, diabetes, and cardiovascular issues. Experimental studies further support these findings by demonstrating the adverse impact of BPA on physiological processes, contributing to the onset of metabolic disorders. Despite its detrimental health effects, removal of BPA poses a formidable challenge due to its intricate molecular structure, resistant to conventional water treatment methods. To address this, our review comprehensively summarizes human BPA exposure data and in vivo/in vitro mammalian studies, offering a comparative analysis of treatment technologies with a focus on documented health impacts. Biological treatment removes BPA efficiently, however, maintaining ideal bacterial populations and controlling biomass concentration provide difficulties that affect operational stability and scalability. In the meantime, despite the high removal rate of physiochemical treatment such as absorption and membrane technology, they consume significant amounts of energy and generate chemical residues that could retain toxicity. In this regard, a hybrid photocatalytic membrane emerges as a promising solution, forming the basis for our comparative evaluation in wastewater treatment and water purification. By effectively degrading BPA and mitigating the BPA toxicity, the photocatalytic membrane helps reduce human exposure to this harmful compound. This technology presents a viable approach to tackle BPA-related environmental challenges while shedding light on its intricate metabolic effects on human health.

1 Introduction

Bisphenol A (BPA; 4,4’-dihydroxy-2,2-diphenylpropane) is widely used as a non-polymer additive in plastics manufacturing, in the production of polycarbonate and epoxy resins. BPA is a white, crystalline solid that is soluble in fats and almost insoluble in water with molecular weight of 228.29 g cm−3. The melting and boiling point are 156 °C and 220 °C at a pressure of 5 MPa, respectively [1]. Figure 1 shows the molecular structures of BPA and of the similar 17β-estradiol based on the presence of OH groups and fused ring structures [2].

Fig. 1
figure 1

Molecular structure of bisphenol A (BPA) appears to be similar to 17β-estradiol, allowing the chemical to bind to estrogen receptor Molecular structures were retrieved from United States Environmental Protection Agency, www.comptox.epa.gov)

Worldwide, BPA is commonly used as synthetic chemical in manufacturing various consumer products including infant feeding and drinking water bottles, plastic food containers and utensils, food packaging, toys, eyeglass lenses, dental and medical products, electronic appliances, beverage containers, water pipes [3], compact discs, impact-resistant safety equipment, detergents, flame retardants, food, cosmetics, and pesticides. Thus, the risks of BPA exposure among the majority of world population would be high. Dodds and Lawson [4] noted that BPA was first reported as a synthetic estrogen in 1890. Krishnan et al. [5] first recognized BPA as an endocrine disruptor chemical. Vandenberg et al. [6] have reviewed the level of BPA in human tissues and fluid and suggested that BPA levels were detected exceeding the safe dose in human due to use of products such as plastic products and containers, which result in increased in diseases and cancers. Studies have demonstrated that BPA has detrimental effects on the metabolism of adults, and a number of studies cited hereafter established the link between BPA and the development of metabolic diseases. The effects on fetuses, newborns, and young children are even more detrimental [7]. Specifically, the changes in endogenous estrogen gene expression after BPA exposure showed that the presence of BPA altered estrogen signaling [8]. Furthermore, changes in developing tissues and organs at an early age increase the risks of developing adult diseases in childhood and later in life [9]. Studies have linked BPA with the development of metabolic diseases that include obesity (ICD-10: E66.9) [10, 11], diabetes (ICD-10: R73.03) [12, 13], hypertension (ICD-10: I10) [14,15,16], hyperthyroidism (ICD-10: E05.90), hypothyroidism (ICD-10: E03.9) [17,18,19], and cardiovascular disease (ICD-10: I25.10) [20,21,22]. However, the mechanism of BPA effect towards human health is unique. Some tissues/organs may have been affected by very low concentration, and others may have been affected at higher concentration depending on the time of exposure, and duration. Therefore, there is no standard safe concentration besides the one that has been documented for low and no adverse effect level.

Chemical replacements such as bisphenol S (BPS), bisphenol F (BPF), and bisphenol AF (BPAF) have been used in some applications to replace BPA. The creation of bio-based polymers made from renewable resources is an additional approach. Although various options are available, each has a unique set of factors to take into account, and their effects on the environment and safety may differ. Furthermore, they might have comparable health risks as BPA. Harnett et al. [23] emphasize the harmful effects of BPA and its analogs, namely tetramethyl BPF (TMBPF), BPAF, and BPS, shedding light on their cytotoxic impact. Despite the fact that some of these compounds were made with the idea of "safety by design," current research and this study show that many of these compounds are, in fact, "regrettable substitutions," since they are more poisonous and potent on human cells than their parent compounds. Because of the unique qualities of BPA that make it advantageous in some production processes, finding a replacement for BPA that works in all applications is still difficult.

BPA is a pervasive environmental material due to its mass manufacturing and consumption. Many ecological media, including water, soil, biomass, and air, have been shown to contain BPA. The traces of BPA found in surface water, ground water, sea water, and other sources of water are difficult to detect and remove [24].The intricate and demanding removal process of BPA relies predominantly on current treatment methods, which involve biological and advanced oxidation process. Biological treatments are up to 90% effective, and some oxidation and photocatalytic treatments can remove BPA to less than the detection limits [25, 26]. For biological treatments, the pH must be regulated to control the rate of removal [27] and enzyme cultures are difficult to acquire, maintain, and apply or alternatively, the seeding with well-acclimated microorganisms [28] is difficult to maintain as indigenous consortia typically have competitive advantages.

The two important advanced oxidation treatments of BPA apply chlorination or photocatalysis. The chlorination of BPA usually increases human toxicity by estrogenic disruption, accumulation in body fat, and maternal transfer to new-born compared to unchlorinated BPA [29]. Often investigated and applied to treat many water-borne contaminants, photocatalysis effectively transforms recalcitrant organic pollutants into less toxic metabolites in comparison with any other conventional treatment. The typical photocatalytic treatment involves a suspension of either nano or microscale photocatalytic particles in a reactor. Despite the efficiency of photocatalysis, the separation of the photocatalytic particles is a major problem. Beside reusability of the recovered photocatalytic particles, the separation is important to avoid the adverse biological effects of the photocatalyst even in the absence of light [30]. Therefore, the development of membranes that bind photocatalytic particles and permit light penetration is very promising in removing BPA from water.

This paper reviews epidemiological and laboratory investigations on the effect of BPA exposure to humans causing metabolic disorders. This BPA exposure assessment is necessary to evaluate treatments of BPA contaminated water from surface water, groundwater, seawater, and other sources. Other investigators have separately reviewed (1) the investigation of the occurrence and impact of BPA on human health and (2) the treatment technologies for the removal of BPA and other pollutants in wastewater treatment, particularly in engineering literature. In the meantime, some investigators have also reviewed on BPA occurrences, health effects and remediation mechanisms including membrane technology. However, those reviews did not synthesize the human health effects of BPA well enough to fully guide the selection of the better treatment methods to remove harmful amounts of BPA from water. This review emphasizes on the significance of UV and visible light radiation in the application of photocatalytic hollow-fiber membranes as these membranes demonstrate the capability to reduce BPA levels in contaminated water to below detection limits. Given that drinking water stands out as the primary source of BPA exposure to humans, the review extensively delves into sources from surface fresh waters and wells. Additionally, it examines variations in BPA levels associated with the household use of portable water. This approach ensures a comprehensive exploration of BPA exposure, particularly in the context of drinking water and its various sources.

2 Effect of bisphenol A exposure on human health

2.1 Epidemiological studies

Upon ingestion, BPA is swiftly absorbed by the stomach and intestines. The liver metabolizes it into the primary glucuronic acid-bound form, which is then excreted through urine with a short half-life. Urine is commonly used as a bio-monitoring matrix and reflects different BPA concentrations across various ages, occupations, and genders globally [31]. BPA affects the endocrine and reproductive systems by interacting with various receptors and enzyme pathways involved in steroid metabolism [31]. It disrupts normal body functions by binding to these receptors, either mimicking or blocking the effects of natural hormones [32]. The endocrine system plays a crucial role in regulating various physiological processes, including metabolism. Hormones, which are regulated by the endocrine system, influence metabolism and play a role in the development of metabolic syndrome. Many epidemiological studies link BPA exposure to impairments of human (1) metabolic [33, 34], (2) reproductive [35], and (3) neurodevelopmental health [36,37,38]. Figure 2 shows the association of BPA exposure with development of metabolic diseases and neurodevelopmental impairments.

Fig. 2
figure 2

Effects of bisphenol A (BPA) exposure on the metabolic and neurodevelopmental disorders (designated as rectangles): (ICD-06) impairment of neurodevelopment, (ICD-05) obesity, (ICD-05) diabetes, and (ICD-11) cardiovascular disease (CVD).  Dashed lines define the indicative linkage between the major disorders and possible causes of the impairment.  Thin arrows define causation of the major disorders: obesity, diabetes, CVD, and neurodevelopmental impairment by the four health impairments known to be caused by BPA or intermediate conditions: glucose intolerance and insulin resistance caused by BPA accumulation in fat that causes obesity and elevated lipids and triglycerides, respectively.  Broad arrows define the major health effects caused by BPA

Exposure to BPA in utero causes miscarriage and low birth weights [39,40,41], and increases the risk of obesity and cardiovascular disease during adulthood [42]. Study has demonstrated that BPA exposure in utero alters the plasma lipid profile by increasing triglycerides, which eventually leads to the accumulation of lipid droplets in fat tissues [11]. Hence, adult obesity prevails because of juvenile exposure to BPA. Altered glucose metabolism also increases the risk of obesity but further causes type 2 diabetes [43]. Obesity and diabetes both increase the risk of cardiovascular disease [42]. The accumulation of BPA in the body may affect thyroid gland functions, and eventually leads to impaired brain development in young children, thus, increases the risk of neurodevelopmental impairment [38]. The onset of any or all the following conditions associated with obesity leads to insulin resistance, glucose intolerance, dyslipidemia and cardiovascular disease [42, 43].

Among the most common sources of BPA exposure is occupational exposure. Study by Chu et al. [44], showed that direct contact with BPA daily induced skin allergy symptoms and the development of dermatitis for 22 workers at construction sites. Apart from physical skin symptoms, BPA exposure has been shown to elicit negative effects on reproductive system. Study by Li et al. [45] reported a significant risk of sexual dysfunction and low libido among BPA-exposed workers in a factory producing and packaging epoxy resins. The increased BPA concentration in worker serum was associated with the decrease in the mean testosterone concentration in serum.

Study has also reported that 28 workers at a semiautonomous epoxy resin factory who had inhaled BPA, had inadequate thyrotropin, the pituitary hormone that stimulates thyroid production of thyroxine (T4) and triiodothyronine (T3), which in turn stimulates the metabolism in most tissues of the human body [46]. To assess endocrine activity and hormone production of the thyroid, the authors have measured total and free T3, and free T4. Abnormal levels of the total and free T3, and free T4 were measured in the workers exposed to BPA. These thyroid hormones exceeded the normal range for adults [46]: total T3 (1200 to 2700 pM), free T4 (10.2 to 24.5 pM), and free T3 (2.76 to 6.45 pM). The function of the thyroid gland is of concern when free T3 exceeds 6.46 pM (diagnosis of hyperthyroidism is being established). In workers with the greatest BPA exposure, they also reported a positive association between free T4 and lactate dehydrogenase (Enzyme Commission or EC 1.1.1.27), and between insulin and thyrotropin (thyroid stimulating hormone). Thus, they also demonstrate that exposure to BPA disrupts the homeostasis of pituitary and thyroid hormones and leads to the development of type 2 diabetes as noted in Fig. 2.

Meanwhile, in children, studies conducted by [47, 48] estimated BPA exposure among children in day care facilities versus households. Surprisingly, BPA was detected in both liquid and solid foods, and the estimated BPA dose was 43 ng kg−1 d−1 for children of age 2 to 5 years [47]. The BPA samples were collected from environmental media: air, dust, and soil and personal media: liquid food, solid food, hand wipes, and urine [47]. The food samples were collected from both the day care facilities and homes where children live. In day care center, BPA contamination was typically higher in liquid food (average 384 ng kg−1) compared to solid food contamination detected over a range of 196 to 932 ng kg−1. In household samples, BPA contamination was less. However, the study did not adequately explain why BPA contamination varied between samples collected in day care centers versus homes.

Wilson et al. [48] found that 99% of the BPA exposure originated from the diet of children at an estimated dose of 52 to 74 ng kg−1 d−1, and an estimated inhalation dose of 0.24 to 0.41 ng kg−1 d−1 at day care facilities. The children of ages 1.5 to 5 years were randomly chosen from eligible day care centers and households [48]. The 2007 study detected BPA in more than half of the samples of indoor air and samples of hardwood floors and food preparation surfaces. In addition, nearly all the wipes used to sample the hands of the children had detectable BPA. More than 68% of the samples from the liquid food given to the children had detectable BPA, whereas 83% of the solid food samples contained detectable BPA. Interestingly, both studies by [47, 48] found that the exposure and doses of BPA were higher for children kept in day care centres compared to children stayed at home. Presumably day care centers serve more pre-processed, packaged food whereas households prepare meals from more basic foodstuff with less packaging in plastics.

Different investigators reviewed the U.S. National Health and Nutrition Examination Survey data of 2003 to 2006. The study reported there is no association between urinary BPA and diabetes [49]. Using the same data from the U.S. National Survey, however, another study found a significant relationship between the BPA in urine and diabetes [50]. In the reproductive organs, research discovered that the serum level of BPA increases with the level of androgen in women with polycystic ovary syndrome. Men normally have more androgen than women [51]. Perera and Herbstman [36], reported a negative correlation between androgen level and BPA in men who were exposed to epoxy resins. As of 2021, the authors of this review have not found epidemiological evidence to contradict Vandenberg et al. [6] who noted that whether the BPA effect on the androgen level being the pathway for BPA impact on metabolic diseases was unknown.

From chromosomal abnormalities research, [39,40,41] reported that higher BPA exposure is associated with complications during pregnancy. All three studies found that women carrying their pregnancy with chromosomal abnormalities had higher maternal serum levels of BPA, and higher BPA levels were associated with a history of recurring miscarriage [41]. These data are based on small sample sizes with limited information about each subject and about other causative factors.

2.2 BPA and metabolic disease

Disruption of normal metabolism is an important effect on the conversion of ingested food into energy. Metabolism involves thousands of enzymes involved in numerous pathways. When affected by metabolic diseases, the essential cellular enzymes involved in expressing, transporting, and processing proteins, carbohydrates, and lipids are disrupted. Investigation of the effect of BPA exposure on metabolism has been explored in vivo using mice, rats, and other rodents. Fetal and perinatal exposure of mice and other rodents to relevant doses of BPA affected organs and led to weight increases of the offspring in adulthood. The incremental weight increases to attain rodent adulthood depends on gender, doses of BPA, and duration of BPA exposure [52, 53], which is associated with the 55 [54].

In addition, perinatal BPA exposure in pregnant mice led to an increase in insulin resistance during the period of pregnancy. An individual who is diagnosed with insulin resistance usually develops type 2 diabetes or has a higher risk of cardiovascular diseases. Furthermore, insulin resistance is associated with an increase in triglyceride and leptin concentrations [55]. Surprisingly, BPA has a higher impact on human pancreatic β-cells compared to mice; Soriano et al. [56] also reported the adverse effect of BPA on glucose metabolism, which diminished the hormone adiponectin release from human adipocytes (fat cells). These findings suggested that BPA could be considered as a risk factor for metabolic disorder development in humans.

Both in vitro and in vivo investigations [11, 52, 53, 57] have reported the effect of BPA exposure on metabolism in children and adults. Several studies [21, 56, 58] conclude that limited exposure to BPA may have more impact on health compared to greater doses. Numerous studies have confirmed that alterations in the mechanism of BPA resulted in disruptions within the body, even at low levels of harm. Figure 3 shows the adverse effect of BPA exposure on vital organs. The study that was conducted on fetal lungs showed a similar hindrance in lung development between moderate and high amounts of BPA exposure [59]. Furthermore, researchers have also observed that BPA has an immediate effect on cardiac myocytes with a non-linear dose response [57] which supports the impact to human health from limited exposure of BPA as compared to high dosage. Women tend to have more estrogenic effects compared to men. Several cross-sectional studies [19, 60, 61] observed negative effects of BPA exposure, but these studies do not indicate acute effects that require immediate water or wastewater treatment. Further long-term investigations of molecular or genetic modifications are necessary to put the briefly documented cross-sectional effects into long-term chronic effects that require treatment standards. In particular, the programming of diseases that occur in adulthood due to BPA exposure prenatal, perinatal, or during younger childhood must be investigated.

Fig. 3
figure 3

Bisphenol A adverse effects on vital organs: brain (due to thyroid hormone disruptions), heart (heart failure and several heart diseases), liver (disruption of glucose and lipid metabolism, increased risk of type 2 diabetes and obesity), and kidney (interruption in leptin). Notes: CVD is cardiovascular diseases, body mass index is calculated by dividing the weight (kg) by height (m), chronic hyperinsulinaemia is a condition of excess level of insulin in the blood and usually promoted by obesity and associated with adipose dysfunction, and leptin is a hormone secreted by white adipose tissue that has an important role in the regulation of body weight; a larger amount of plasma leptin is associated with chronic kidney disease, and is thus, indirectly related to the kidney and the prevalence of obesity

The effect of BPA is particularly noticeable in the vital organs essential for survival. The vital organs affected by BPA include the brain [62], heart [63], kidneys [64], liver [65], and lungs [59]. Excessive levels of BPA have been correlated with obesity [58], diabetes [55, 59], cardiovascular diseases [18, 57], polycystic ovarian syndrome [66], and low sperm count [67]. Depending on dose, BPA stimulates in vitro differentiation of murine fibroblasts into adipocytes and increases lipid storage [68]. Small doses of BPA inhibit adiponectin secretion by human adipocyte cell cultures in vitro and stimulate the secretion of interleukin-6 and a necrosis factor, which are two inflammatory adipokines, suggesting possible involvement in insulin resistance, obesity, and metabolic syndrome [54, 69, 70].

2.2.1 Cardiovascular disease

The studies of the relationship between human BPA exposure and cardiovascular disease are limited, but coronary artery disease in men and women exposed to BPA has been reported. Rodent exposure to BPA worsened cardiovascular disease morbidity and mortality [71]. Larger urinary BPA concentrations in humans has been associated with hypertension [15], heart attack [72], coronary heart diseases [73], and peripheral arterial disease [74]. Chronic exposure to BPA has shown to cause maladaptive changes in heart size and shape (cardiac remodelling), atherosclerosis, and altered blood pressure in female mice [72]. Patel et al. [72] also reported that BPA exposure resulted in alteration of cardiac calcium handling, which involved the expression of the protein sarco-endoplasmic reticulum calcium-ATPase where disruption of the protein would lead to electrophysiological dysfunction of the heart. The study demonstrated the negative impact of BPA exposure on heart was not limited to cardiac remodelling but caused cardiac conduction dysregulation occurring at protein level by alternating their expression level negatively. In addition, study by Jiang et al. [75] had evaluated (1) the structure of general heart tissue, (2) the morphology of cardiomyocytes (striated self-beating and cylindrical rod-shaped muscle cells), and (3) the mitochondria in Wistar rats (R. norvegicus) exposed to BPA at a dosage of 50 µg kg−1 d−1. They observed that prolonged BPA exposure induced cardiac impairments by disrupting mitochondrial function in cardiomyocytes. The study has shown that disruption of cardiomyocytes mitochondrial function induced cardiac impairments by promoting myocardium hypertrophy and cardiomyocyte enlargement.

In addition, Gonzalex et al. [76] investigated whether BPA leads to myocardial interstitial fibrosis by inducing proliferation of cardiac fibroblast cells that increase collagen production. The study has demonstrated that BPA was able to induce myocardial interstitial fibrosis which contributes to chronic dysfunction of the left ventricular, eventually leads to heart failure. Cardiac impairment induced by BPA is not only limited to mitochondrial dysfunction and myocardial interstitial fibrosis. Hu et al. [77] reported a decrease in cultured cardiac cells (isolated from healthy male Sprague–Dawley) upon exposure to BPA. Furthermore, BPA promotes the proliferation of cardiac fibroblasts via extracellular signal-regulated protein kinases 1 and 2 (ERK1/2; EC 2.7.11.1), which increases collagen production. Patel et al. [78] found that chronic oral BPA exposure at the beginning of gestation 11.5 days led to reduced survival of the mice after an experimental myocardial infarction was induced by isoproterenol when the mice reached adulthood, 16 weeks after birth. A substantial reduction was observed in survival among the chronically BPA-exposed C5bl/6n mice (Mus musculus) immediately after 3 weeks after the experimental myocardial infarction. The data presented by this group of researchers suggested that after the experimental myocardial infarction, cardiac remodeling and thus heart recovery was reduced by chronic exposure to BPA [78].

Yan et al. [79] investigated the effect of acute BPA exposure to the heart of rats subjected to global ischemia (inadequate blood supply), followed by reperfusion of the blood supply. They observed a period of sustained ventricular arrhythmias of the female Sprague–Dawley rats (R. norvegicus) heart exposed to 1 nM BPA compared to the male heart. The ventricular arrhythmias developed secondary to BPA exposure in female rats was due to a more sensitive female physiology response towards estrogen-like compounds compared to males [79]. They further demonstrated that acute exposure to BPA worsened the ventricular arrhythmia following ischemic-reperfusion injury or reoxygenation injury of the heart of female rats (R. norvegicus). This worsening of ventricular arrhythmia is consistent with Posnack et al. [71] who found slowed electrical conduction in the hearts of BPA-exposed, female rats (R. norvegicus).

Moreover, higher total urinary BPA was associated with an increased incidence of hypertension [74]. A study of elderly Korean adults without a prior history of hypertension revealed similar findings in which BPA occurrence in urine significantly increased the onset of hypertension. Bae et al. [80] also found that BPA exposure is associated with reduced heart rate variability. Patel et al. [78] observed no differences in blood pressure of the male humans exposed to BPA. However, female suffered a small increase in systolic blood pressure associated with a BPA exposure at 5.0 µg kg−1 d−1. Earlier Patel et al. [72] determined that BPA had a sex-specific impact on the blood pressure when they compared the systolic, diastolic, and arterial pressures between females and males. Kilic et al. [81] showed that more than a dozen genes and genetic loci were affected in BPA-exposed human subject. One of which is the sirtuin protein (SIRT1), a NAD (nicotinamide adenine dinucleotide)-dependent histone deacetylase (EC 3.5.1.98) located in the nucleus. This protein is expressed in many tissues. The protein SIRT1 is activated in distinct metabolic and stress response pathways. The expression of this protein is targeting many cellular proteins, such as PPAR-γ, PGC-1α, FoxO, AMPK, NF-kB, PTP, eNOS, and p53.

In summary, BPA exposure was proven to have an association with cardiovascular disease and hypertension development. The exposure impacts the expression of a number of cardiac proteins and seems to be more evident in exposed females; thus, increasing the possibility of women to experience cardiac crisis compared to men. Moreover, a longer duration of BPA exposure could modify and induce cardiac conduction. Acute exposure showed an impact on immediate observation such as arrhythmia compared to chronic. The dosage of BPA is also an important factor in the development of cardiac injuries. Most of the studies suggested that a small dose of BPA has more impact on cardiovascular disease development than a large dose [80, 82].

2.2.2 Obesity

Obese people usually have a 20% greater body weight than the person with a normal body mass index. In adults, a body mass index more than 30 kg m−2 is defined as obese. However, the calculated body mass index might reflect different physiology and metabolic health [82]. The energy expenditure should be greater than the energy intake to maintain or reduce weight gain. Obesity occurs due to overconsumption of fatty foods and having a sedentary lifestyle. Studies conducted have agree that exposure to endocrine disruptors (mainly from food and beverages) influence lipid metabolism thus causing obesity.

An epidemiological cross-sectional analysis carried out by Hong et al. [60] comprised of 296 healthy, reproductive-aged Korean women above 18 years old demonstrated that urinary BPA positively associated with the body mass index, waist circumference, and fasting serum insulin. In contrast, Do et al. [83] analyzed the Canadian Health Measures Survey (2007–2011) that monitored 4733 Canadian adults aged above 18 years old and reported that urinary BPA is not associated with increased waist circumference but was positively associated with obesity defined by the body mass index. However, Hong et al. [60] in the study of 11,293 Navajo (American Indian) and Alaska Native adults reported an association of waist circumference and of obesity with the urinary BPA concentration [84].

Another interesting study has described a positive correlation between BPA and an excessive body fat, and increased serum leptin. Zhao et al. [85] also examined the relationship between BPA and both osteocalcin level and bone mineral density. Their cohort study of 246 healthy premenopausal women from Shanghai, China age 20 years and older with regular menstrual cycles showed that the exposure to BPA has an undetectable impact on bone mineral density. However, BPA in urine was positively associated with body fat mass and leptin. Bone mineral density was essentially unaffected by BPA, perhaps because the healthy participants were not overweight. Still, further investigations need to be conducted to observe the long-term effect of BPA exposure on bone mineral density due to reports of inconsistency in the effects of leptin signalling on neuroendocrine regulation and bone metabolism [83].

Vafeiadi et al. [10] reported that early life BPA exposure increased the risk of obesity and cardiometabolic disorders, based on a study involving 500 mother–child pairs from a pregnancy cohort in Greece. Larger BPA concentrations in the urine of children were associated with increased body mass index, waist circumference, and the sum of skinfold thickness at the age of 4 years. However, a larger cohort should be assembled to determine whether obesity later in childhood indeed resulted from prenatal or early childhood BPA exposure as they observed a negative association between body mass index and adiposity in girls whose mothers were exposed to BPA during early pregnancy.

In a cross-sectional study of a nationally representative subsample of 2838 randomly selected children and adolescents aged 6 to 19 years in the United States National Health and Nutrition Examination Surveys during 2003 to 2008, urinary BPA was significantly associated with obesity for Caucasians, but not for blacks or Hispanics [61]. An interesting finding from a study of 177 (120 normal and 57 obese) Saudi children aged 13 to 16 years revealed a positive association between serum BPA and obesity. Obese children had a larger serum BPA level compared to normal weight children. Al-Daghri et al. [86] also found a significant association of elevated low-density lipoprotein cholesterol with serum BPA level, but only in girls, thus, suggesting that effect of BPA is gender specific.

Xie et al. [11] found that microRNA-21a-5p regulated adipogenic differentiation. The results showed that overexpression of the biological regulator microRNA-21a-5p significantly decreased the red lipid droplets and triglycerides in BPA exposed 3T3-L1 cells (derived from mouse embryo), a cell line that differentiates into adipocytes. They also concluded that microRNA-21a-5p overexpression attenuated BPA-induced obesity in vivo by targeting map2k3 gene through map2k3 p38 mitogen-activated protein kinase (MKK3-p38 MAPK) in 3T3-L1 cells. These findings suggested a potential therapeutic strategy for BPA-induced obesity. Although BPA in urine was positively associated with body mass index, fat mass, leptin, and serum insulin (during fasting), further investigation is needed, especially regarding the range of ages of the people tested. Whether obesity is affected by prenatal or early life BPA exposure is still inconclusive as of 2021.

2.2.3 Diabetes

BPA has an intense effect on the mouse endocrine pancreas, an essential function involved in glucose metabolism. BPA disrupts pancreatic β-cells, which are crucial in producing insulin and that induce hyperinsulinemia and insulin resistance in male rats (R. norvegicus) [87]. Insulin is a peptide hormone, essential in maintaining the blood glucose level [88]. Excess glucose induces secretion of insulin from pancreatic β-cells. The cardiac muscles are one of the important tissues that exerted response within the presence of insulin, which in turn, responsible for transporting glucose to all living tissues within the blood circulation. Apart from cardiac muscles, adipocytes and skeletal muscles also exerted response towards presence of insulin [89]. Sivashanmugam et al. [13] assessed the effect of BPA on insulin signaling molecules in the cardiac muscles of adult-male, Wistar rats (R. norvegicus). The selected adult-male, Wistar rats (200 to 250 g body weight) were divided into four groups of which receiving four different 30-d doses orally of 0, 10, 100, and 400 mg BPA kg−1 body weight d−1. They showed that the group of rats that received the largest dose (400 mg BPA kg−1 body weight−1 d−1) suffered significant protein decreases in insulin receptors and pIRTyr1162 proteins, with no significant changes in IRS1 (insulin receptor substrate-1) and kinase B or Akt proteins [13]. However, the authors observed a decrease in the three proteins that affect the cardiovascular function: (1) pIRS1Tyr632 (at BPA 30-d doses of 100 and 400 mg BPA kg−1 body weight d−1), (2) pAkt Ser473 (400 mg BPA kg−1 body weight d−1) and (3) glucose transporter 4 (GLUT4) at 100 and 400 mg BPA, which cycles between cell cytosol and the plasma membrane. They concluded that BPA exposure has an adverse effect on cardiac insulin signal transduction that may affect cardiac function.

Sakurai et al. [12] reveal that in adipocytes (fat cells), BPA exposure increases the basal and insulin-activated glucose transport by stimulating glucose transporter 4 expressions. Adult male mice (M. musculus) exposed to BPA developed both hyperinsulinemia and mild insulin resistance [90]. More prolonged exposure to BPA in rats (R. norvegicus), up to 4 days, resulted in chronic hyperinsulinemia and insulin resistance.

In addition, Soriano et al. [56] found that BPA exposure to humans and adult female mice (M. musculus) interfered with glucose metabolism and may increase the possibility of diabetes type 2, cardiovascular diseases, and insulin resistance. They found that in the human pancreatic β-cells, glucose-induced insulin secretion is stimulated at an environmentally relevant dose of BPA (1 nM), while for mice, exposure to increasing doses of BPA increased insulin secretion. Soriano et al. [56] also found reduced adiponectin (protein hormone and adipokine) in human adipocytes (fat storage cells), which confirmed a disruption of the glucose metabolism. Guan et al. [91] evaluated the effect of BPA exposure on lipid metabolism of the fish, Gobiocypris rarus. The authors found that BPA induced triglyceride increases in male fish suggesting that the fatty acid β-oxidation pathway, after BPA exposure, might be gender-dependent.

2.2.4 Other metabolic diseases

Several studies investigated the development of other metabolic diseases following BPA exposure, which include hyper- and hypothyroidism, as well as cancer. In addition, BPA binds to estrogen receptors and interferes with endogenous estrogen (demonstrated in vitro with cell lines). BPA also binds to androgen receptors, thus acting as an androgen antagonists or testosterone blockers in vitro. Furthermore, BPA binds to thyroid receptors, to disrupt the function of the two thyroid hormones. In a cross-sectional study from February to August 2012 involving 39 obese and overweight children aged 3 to 8 years from Dayton, Ohio in the United States, Khalil et al. [19] found that obese male children had significantly less free thyroxine as compared to female children. The thyrotropin was inversely associated with BPA. Earlier Braun et al. [17] suggested that BPA disrupts the thyroid function.

Chapalamadugu et al. [18] reported that BPA is a thyroid hormone receptors antagonist and able to inhibit the induction of gene Myh6, thus, interfering with the development and growth of the fetal heart into the postnatal heart. As an endocrine disruptor, BPA affects the nuclear hormone receptors, which bind steroids that act as hormones and regulate target gene transcription. Moriyama et al. [92] found that BPA reduced the thyroid hormone triiodothyronine (T3) binding to the nuclear thyroid receptors and inhibited the transcription process. They determined that BPA suppressed the activity of the transcription factor Gal4-TRα1 up to half of the respective control level. This finding also demonstrated that BPA impaired thyroid function by inhibiting the hormone triiodothyronine or T3 binding to nuclear thyroid receptors, thus suppressing transcriptional activity. Study by Heindel et al. [22] reported that BPA exposure disrupted thyroid hormone signaling during preclinical development and validation of an animal model that mimics human responses. The end point of the research program reported that two years of BPA exposure on Sprague–Dawley female rats (R. norvegicus) did not exhibit concerning health risks. However, the study investigated neoplastic changes rather than alteration on thyroid function secondary to BPA exposure suggesting that BPA exposure posed concerning threat on thyroid functions and metabolism [22].

BPA is a potential risk factor in breast cancer [93]. Two studies found in MMTV-erbB2 mice (M. musculus) that even small BPA doses could accelerate mammary tumorigenesis [94] and metastasis [95]. Small BPA doses induce aggressive tumorigenesis in breast cancer patients [96]. There is an association between BPA exposure and complex endometrial hyperplasia and endometrial cancer [97]. At very small doses (ng BPA kg−1 body weight d−1), the level found in the blood of general populations, induced chemo-resistance in several malignant human breast cell lines, regardless of whether these cells are estrogen receptor positive or negative while it increases the expression of several anti-apoptotic proteins [98].

Another study on prostate cancer found that mice (M. musculus) fetal or perinatal exposure to BPA or diethylstilboestrol is associated with prostate hyperplasia and pre-malignant lesions in adults [99]. The perinatal exposure to BPA or estradiol (Fig. 1) increases the susceptibility to prostate cancer, either spontaneously or after a second estrogenic injection [100]. Heindel et al. [22] noted the possibility of whether BPA exposure would reprogram genes could be determined by comparing the methylome and transcriptome from rat (R. norvegicus) prostate cells with uterus cells. Interestingly, they found that BPA exposure also increases the chances of having testicular cancer. At low doses (nM or pM), BPA acted only through an intracellular seven-transmembrane G protein-coupled estrogen receptor (GPR30) by a promotive effect, while at greater doses (mM), BPA also acted through an estrogen receptor, which counteracted the promotive GPR30 mediated effect [101].

3 Distribution of BPA in the environment and occurrences of BPA in houseshold water

Widely used as a liner in food and beverage packaging [102], as well as in other plastics, toys, and electrical equipment [103] BPA exposure ranges from a variety of sources, including direct contact with contaminated plastic materials, utensils, and equipment, swallowing contaminated food and water, and inhaling polluted air. Since meeting daily nutritional demands is necessary for survival, food exposure is the most significant factor for BPA contamination. BPA contamination through food exposure arises from the use of BPA in the production of various plastic food serving containers made of polycarbonate (PC) and polyvinyl chloride (PVC), where BPA seeps into food and drinks. Food cans are also made from epoxy resins for the interior coatings, hence contribute significantly to food adulteration [104]. Furthermore, the bioaccumulation of BPA in the fish from the domestic pond occurs resulting from BPA leaching of the plastic storage tanks and other paraphernalia used in fish farming [105]. Interestingly, rivers, reservoirs, lakes, and estuaries are regarded as the main BPA sinks; thus explaining the uptake of BPA in the aquatic species. The level of BPA in seafood such as prawn, crab, blood cockle, white clam, squid, and ocean fish from the supermarket in Singapore was reported between 13.3 and 213.1 ng g−1 wet weight. BPA levels were undetectable (< 0.2 µg L−1) in river water but ranged from 2 to 8.8 µg kg−1 and 0.3 to 12 µg kg−1 wet weight in periphyton and benthic organisms, respectively. As important food sources for invertebrates and some fish, benthos are aquatic species that live on the bottom of lakes, rivers and creeks or on the sea floor.

Apart from food, there are a few other potential sources that might influence the levels of BPA in humans, such as drinking water, air and dust, which receive lesser attention. Dermal exposure is the second most common way that BPA is absorbed which rises from direct contact with toys, medical equipment, and thermal paper. Thirdly, BPA exposure can occur via inhalation of BPA-containing vapors, mists, dust, and gasses [104]. Table 1 provides some concentrations of BPA based on the few published exposure assessments involving food wrapped in plastics, other food containers, air and dust, dental sealants, and paper currencies.

Table 1 Bisphenol A concentrations in different media reported for 2010 to 2021

3.1 Occurrences in household water

Numerous water sources, including drinking water supplies, groundwater, and surface water, have been shown to contain BPA. The reason for this prevalent phenomenon is the pervasive usage of BPA, which is released into the environment during the production of plastics, epoxy resins, and other consumer products. BPA can enter surface water bodies like rivers and lakes by leaching from plastic garbage, runoff from cities, and industrial discharges. Contamination of groundwater can occur when BPA-containing materials seep into the soil. Furthermore, BPA may seep into sources of drinking water due to the deterioration of plastic pipes or the leaching of products containing BPA that are utilized in water treatment procedures. Household water is potable domestic water used indoor and outdoor for drinking, food preparation, and bathing, among others. Although BPA reacts rapidly in chlorinated water, this contaminant can persist to household taps [1] and beyond. BPA diglycidyl ether is used as a monomer in the synthesis of epoxy resins to manufacture polycarbonate products [112], and as a stabiliser in PVC products [113, 114]. In addition, epoxy coatings are used on the internal surfaces of water mains, and other connecting pipes, and household connectors. As a result, BPA can be released into drinking water and surface water from household water pipes and polyethylene tanks.

3.1.1 Tap water

Epoxy resins containing BPA are widely used to reline the interior surfaces of the water mains and water pipes [115] due to transparency, large impact strength, resistance to acids as well as stability in oil and intense temperatures (up to 145 °C) [113]. The widespread use of BPA in PVC pipe coatings has led to the emerging trace of BPA in household water. The contamination of BPA into tap water might happen due to the BPA leaching from the PVC hose as the water passing through [116]. The assessment of the BPA occurrence in Malaysia has reported BPA from 3.5 to 59.8 ng L−1 in 30 water samples collected directly from tap connected to the main distribution pipes [117]. Shockingly, large BPA concentrations of 56.4 and 59.8 ng L−1, respectively, was detected in two samples collected from source fitted with filter devices. Meanwhile, samples collected from source without filter devices have been detected with BPA with a mean concentration of 9.6 ± 4.3 ng L−1.

Previously, BPA was reported present in water samples from leaching test for epoxy resins [118] and materials and equipment used for water supplies [119]. Continuous test of tap water from pipes coated with epoxy resins A and B were performed in the presence of chlorine [120]. After 24 months, BPA was detected at all times in the retained tap water of both pipes A and B with the concentration varies from 3.9 to 130 ng L−1, and 2.6 to 12 ng L−1, respectively. A most recent investigation [116] to assess the occurrence of BPA in tap water reported that BPA was ubiquitous in the tested samples. For this assessment, the release of BPA from water supplied through pipes made of different materials such as PVC, stainless steel and galvanized pipes. Water samples were obtained during fall, winter, and spring for seasonal comparative purposes. The study findings indicated that BPA was found in two samples from Taipei and one from Kaohsiung, which were obtained from buildings using PVC pipes at concentrations of 12, 10 and 3 ng L−1, respectively. As expected, the concentrations increase with contact times and ambient temperatures. BPA concentrations were also affected by ambient temperatures, which decreased in winter and increased again in spring.

A further investigation of BPA presence in tap water was conducted earlier in Guangzhou [121]. For this analysis, six samples of tap water were obtained from six separate drinking water sources and the findings revealed that five of the tap water sample plants were found to contain BPA concentrations ranging from 2.3 to 317 ng L−1. The result also suggested that BPA levels were close to BPA levels in the drinking water survey results [122]. Like the previous study, this result also suggested that the concentration of BPA in tap water may vary in different seasons in which the concentration was lower in winter than in summer. In summary, different levels of BPA were found in most of the samples of tap water with the highest levels found in the samples obtained from taps connected to PVC pipes and water filter systems. To conclude, the reliability of the water treatment process does not guarantee the full elimination of BPA from drinking water, as pollutants can also be introduced by contamination of supply materials during contact with water.

3.1.2 Drinking water

Conventional drinking water treatment plants (DWTPs) typically provide coagulation, sedimentation, filtration, and chlorination. Although BPA quickly degrades in chlorinated water, DWTPs are not specifically designed to remove endocrine disrupting chemicals; thus, BPA in ambient surface or ground waters can persist through the water purification processes. BPA in drinking water has been recorded mainly because of lacquer coatings of concrete tanks and the lining of many fittings in contact with drinking water [102]. Other than water supplies, BPA may also leach from plasticisers and by-products during the manufacturing process, in addition to the potential for BPA presence in drinking water. The occurrence of BPA was previously investigated in a study in which they measured the water treatment efficiency in Taiwan [113] from 11 water treatment plants in various regions of Taiwan. Water samples were obtained three times a year from 2009 to 2011. The findings showed that BPA was found in most of the samples with BPA levels in treated water varying from below the quantification mark to 38 ng L−1.

Fan et al. [123]. confirmed the identification of BPA and the chlorinated intermediates in Chinese drinking water from 62 DWTPs in 31 major cities across China using pre-chlorination, flocculation, sedimentation, filtration, and disinfection treatments. 61 DWTPS used chlorine or sodium hypochlorite as a disinfectant, one used chlorine dioxide. Of the 54 water samples from DWTPs, 31 were from reservoirs and 23 were from river water, while eight were from groundwater. BPA was detected in all 62 water source samples with concentrations ranging from 4.7 to 512 ng L−1. Concentrations were generally higher compared to drinking water samples reported in the U.S. and Canada [122, 124]. BPA in drinking water samples from 62 DWTPs in 31 cities in China were detected at 60 of 62 plants [123]. The average concentrations of BPA collected from river were the highest followed by reservoir and groundwater with concentrations at 106, 79.8 and 62.8 ng L−1 respectively. As the water source for DWTPs in China is well protected, the concentrations of BPA recorded were lower (range between 2.2 and 1030 ng L−1). BPA with concentrations less than 5 to 14 ng L−1 were also reported in 19 DWTPs used for the production of drinking water in the U.S. [122] as well as in surface water of south Germany [125], Canada [124], and France [126]. In France, the quantification of BPA from DWTPs located in the French Poitou–Charentes area was conducted [126]. Samples were collected from the surface and treated water from eight municipal DWTPs. The result shows that BPA was detected at concentrations ranging from 6.7 to 29.7 ng L−1 and from 2.0 to 16.9 ng L−1, respectively.

Ironically, BPA level has decreased from 37 to 79% in most sampling sites which have been correlated with treatment processes. In the Poitou–Charentes area, different combinations of water treatment processes have been applied. Most BPA was extracted during the chlorination process of due to the intense chlorine reactivity. Next, the filtering phase was shown to eliminate BPA in most water treatment plant [127]. Ozonation [128] and activated carbon have been proposed as an effective way of reducing BPA [129]. Different results were published with complete removal ranging from 70–100% [127].

Levels of BPA in China drinking water have been reported to be between less than 2 and 128 ng L−1. Meanwhile, chlorinated BPA in drinking waters were reported to be an average concentration of 7.1 ng L−1 as compared to the sample collected from the reservoirs with an average concentration of 79.8 ng L−1 [127]. The reported levels of BPA in the drinking water were greater than those reported in southern Germany [125], the French Poitou–Charentes area [126], Malaysia [117], Canada [124], and United States [122]. Interestingly, significantly large concentrations of BPA ranging from 0.07 to 4.21 g dm−3 was detected in water stored in PC bottles with the greater BPA levels determined in newly manufactured bottles [130]. BPA with concentrations ranging from 17.6 to 324 ng L−1 was also detected in bottled mineral and drinking water from 21 brands at the local supermarket in Guangzhou. In Japan, bottled mineral water stored at 50 °C was detected with high concentrations of BPA in comparison with samples stored at 25 °C [131]. The minimal and maximal concentrations of BPA reported were 3.3 ± 2.6 and 11.3 ± 5.3 ng L−1, respectively [117]. Similarly, BPA level at median 4.6 ng L−1 was detected in bottled water in Greece [132].

Table 2 shows that BPA has been detected in various surface waters ranging from less than the quantitation limit to 21 ng L−1 Greater levels of BPA found in some waters may be due to insufficient treatment and management of wastewaters from homes, villages, cities, and industries where typical wastewater treatment plants (WWTPs) are ineffective [117]. The highest BPA level in drinking water was found in samples obtained from taps connected to PVC pipes and water filters devices. In this regard, BPA are widespread as coatings for many fittings in contact with drinking water [102]. Trace levels of BPA have been found in most potable drinking water samples. Filtered mineral water had lower BPA levels, while samples processed in poor storage conditions had slightly high BPA levels. However, BPA consumption from drinking water is very low and is less than 0.01% of the tolerable daily intake. The study shows that BPA is a pervasive pollutant in soil, tap and bottled mineral water.

Table 2 Occurrences of Bisphenol A in different water sources worldwide

4 Removal of BPA from water

Experimental and epidemiological investigations (Sect. 2.0) have identified correlations between BPA exposure and diverse health issues, with a specific emphasis on the metabolic system, while the occurrence of BPA in the environment (Sect. 3.0), particularly in drinking water, has emphasize the significance of drinking water as a major route of BPA exposure for both humans and animals. The discussion underscores that the presence of BPA in drinking water directly aligns with the potential health impacts elucidated in the metabolic syndrome section. By highlighting the link between environmental exposure and health outcomes, the paper provides insight into how BPA, as an endocrine-disrupting compound, may contribute to adverse metabolic effects. Given the essential nature of water consumption, drinking water becomes a primary source through which individuals come into contact with BPA in their daily lives, therefore methods to efficiently remove BPA from the freshwater, wastewater, and soils are urgently needed [142, 143]. The aquatic assimilation of BPA is largely dependent on biological transformation due to the presence of microorganisms, bacteria, and other biological entities [144].

Through various enzymatic reactions and metabolic pathways, these microorganisms can initiate the breakdown and transformation of BPA into different compounds. The benefits of biodegradation of BPA may include the ability to operate at low concentrations of enzymatic reaction over a wide pH and temperature range, as well as the ease with which the process can be controlled. Therefore, in this section, the degradation of BPA by enzymatic treatment and bacteria transformation was reviewed to highlight the advantages and potential applications of biodegradation. These methods, inherently connected to the biological processes discussed in the exposure section, present promising avenues for environmentally friendly and sustainable BPA remediation. However, the costs of implementing such technologies at full-scale facilities, and data from studies on the treatment of solutions containing BPA at concentrations two to four orders of magnitude higher than the environmental occurrence is difficult to extrapolate to the real-world situation [145]. In the meantime, the complete elimination of BPA from drinking water cannot be guaranteed by the effectiveness of the treatment process because BPA is used so extensively to reline water mains and as coatings for numerous fittings that come into contact with drinking water, which can allow the contaminants to leak into the water supply [102]. Therefore, the inclusion of photocatalysis underscores the diversity of approaches, as it harnesses advanced oxidation processes to degrade BPA. The synergistic combination of photocatalysis and membrane filtration enhances the overall efficiency, ensuring that even trace amounts of BPA are targeted and eliminated. In this regard, the photocatalytic membrane system holds the potential to serve as the ultimate step in drinking water treatment, ensuring the thorough removal of BPA. Furthermore, photocatalytic membrane introduces an alternative and complementary strategy, broadening the toolkit for addressing BPA contamination in various environmental contexts. The comparison of biodegradation and photocatalysis underscores the necessity for a multimodal approach to properly handle BPA pollution, including both environmental sustainability and practical viability.

4.1 Biodegradation treatment

4.1.1 Enzymatic treatment

Enzymatic treatment has garnered significant attention from environmentalists as promising treatment for the removal of BPA as it offers a targeted and environmentally friendly approach to remediation. The breaking of chemical bonds within the BPA structure is mostly initiated by enzymes such as esterases and hydrolases [146]. The intermediate molecules produced by this enzymatic process are metabolized further into less complicated and maybe less hazardous chemicals [147]. The subsequent stepwise breakdown of these intermediates, facilitated by additional enzymes and metabolic pathways, ultimately transforms BPA into simpler and less toxic compounds. Xing et al. [148] proposed degradation pathways of BPA by dialysis membrane enclosed laccase catalysis (DMELC) as shown in the Fig. 4. The process initiates with laccase oxidization, leading to the breakage of the C–-C bond of isopropyl connecting two phenols. This results in the formation of intermediates 4-(2-hydroxypropan-2-yl) phenol through electron transfer. Subsequently, this intermediate is susceptible to further transformation into 4-(prop-1-en-2-yl) phenol (product 1). Concurrently, the loss of electrons and alkylation of 4-(2-hydroxypropan-2-yl) phenol generates either 3-methyl-2,3-dithtdrobenzofuran (product 2) or 2-methyl-2,3-dithtdrobenzofuran (product 3). Additionally, 4-(prop-1-en-2-yl) phenol has the potential to undergo methylation, leading to the formation of 1-methyl-4-(prop-1-en-2-yl)cyclohex-2-en-1-ol (product 4). With regard to the formation of the intermediates, product 1 was identified as Class I with low toxicity, products 2 and 3 were classified as persistent substances [148]. The possible pathway of degradation BPA will lead to complete mineralization to water and carbon dioxide. While these proposed pathways provide insights into the degradation process, the precise mechanism requires further elucidation in subsequent research endeavours. Furthermore, the exact degradation mechanism can vary among different enzymes involved in the breakdown of BPA. Enzymes are highly specific biological catalysts, and different enzymes recognize and interact with substrates in distinct ways. In the context of BPA degradation, various enzymes, such as laccases and hydrolases, may be involved, each with its unique mode of action.

Fig. 4
figure 4

Proposed degradation pathways of BPA by DMELC [148] (License No: 5790651247422)

Several oxidoreductases (e.g., peroxidases and polyphenol oxidases) expressed by microorganisms and plants transform aromatic compounds, such as BPA, nonylphenol, and other phenolic compounds. Previously, the degradation of BPA by manganese peroxidase (MnP) from white-red mycelia Basidiomycete (Pleurotus ostreatus) showed that approximately 80% of BPA was degraded within 12 days of incubation [149]. In-vitro study suggested that single-electron oxidation by MnP in the presence of H2O2 initiated the degradation of BPA followed by the metabolization into phenol, 4-isopropenylphenol, 4-isopropylphenol, and hexestrol [150].

Tsutsumi et al. [151] conducted a similar study on the elimination of BPA by MnP from lignin-degrading basidiomycetes and found that BPA was completely eliminated after a 1-h treatment. However, after 1-h, 40% of BPA's estrogen activity remained in the reaction mixtures. Extension of the treatment time to 12-h completed the removal of the estrogenic activity of BPA. Phanerochaete chrysosporium, the best-studied white-rot fungus, secretes two heme peroxidases under ligninolytic conditions, namely lignin peroxidase (LiP) and MnP [150]. 50% BPA removal was achieved after 1-h reaction of LiP and micellar solution, and the reaction was reactivated by adding H2O2 as the reversed micellar solution. However, the available literature on LiP is limited. Previously, the oxidation of BPA by horseradish (Armoracia rusticana) peroxidase (HRP) has also been studied [152]. HRP in the presence of a hydrogen donor catalyses the reduction of H2O2 to H2O; followed the oxidation of BPA. In the study, the incubation of BPA with different concentration of HRP (0.3, 5.0, and 66.7 U mL−1) show that the removal of BPA increased according to the enzyme level of activity with 99% achieved by 66.7 U mL−1. Interestingly, the oxidation of BPA also removed the estrogen-like activity.

Table 3 shows various enzymatic treatments (peroxidise) for the removal of BPA. Peroxidases specifically HRP, bitter gourd peroxidises (BGP), soybean peroxidase (SBP), and polyphenol oxidase from various enzyme sources are useful in the removal of BPA under different conditions. The optimal pH can be varied from pH 5 to pH 10 depending on the enzyme types. Similarly, the optimal temperature was between 25 to 40 °C with complete removal of BPA estrogenic activity [12, 153]. Interestingly, BGP was quite effective in the degradation of BPA. About 90% BPA was removed by BGP at pH 7.0 and 40 °C [154]. The greatest enzyme activity was observed in the presence of polyphone oxidase from potato with more than 95% degradation for 60 min [155].

Table 3 Degradation of BPA by enzymatic treatment

4.1.2 Bacterial transformation

BPA removal by bacteria is another prominent method. Using microbial communities with certain enzymes and metabolic pathways, BPA can be broken down and transformed through the process of bacterial remediation. The ability of this mechanism, called biodegradation, to lower BPA levels in a variety of environmental matrices, such as polluted soils and water systems, makes it especially noteworthy. Generally, BPA is recognized by bacteria which then carry it into their cells. Here, certain enzymes start the breakdown process which results in the development of intermediate chemicals. These intermediates are further metabolized through a variety of metabolic pathways and then gradually changing into less complex and possibly hazardous chemicals. Certain breakdown products might be incorporated into the bacterial cell, supporting other cellular functions or energy production. In the end, the breakdown products are either integrated into the bacterial biomass or released into the surrounding environment [162].

Various BPA-degrading bacteria have been identified and isolated to treat BPA in WWTPs including Sphingomonas sp., Pseudomonas sp., Streptomyces sp., Bacillus sp. and several more [163, 164]. Bacteria with greater BPA biodegradability are limited because the efficiency of the degradation process is very different from strain specificity in addition to the biomass amount, biological compounds, pH, and the presence of oxygen. Sphingomonas strains were identified from the sludge of WWTPs at a plastics manufacturing facility. This novel bacterium designated strain MV1 was Gram-negative, aerobic bacillus [165] and currently was the most frequently isolated BPA-degrading bacteria which use BPA as the sole source of carbon and energy. Strain MV1 consumed 10 g L−1 BPA within 4 d with the formation of 4-hydroxybenzoic acid, and 4-hydroxyacetophenone (4-HAP) as intermediate products. Greater BPA degradation was achieved by Sphingomonas sp. strains, which 115 mg L−1 BPA was consumed within 15-h [166]. Interestingly, Sphingomonas bisphenolicum strain AO1 from the soil of a vegetable growing field in Japan transformed 115 mg L−1 BPA after 6-h exposure [167].

In addition, Pseudomonas putida strains KA4 and KA5 were the two bacterial strains that were frequently isolated from the environment and showed greater BPA biodegradability using BPA as the sole carbon and energy source [168]. The degradation of BPA can reach up to 90% of BPA when the initial BPA concentration was 1 mg L−1. Another BPA-degrading bacterial Pseudomonas monteilii strain N-502 was isolated [169]. 500 mg L−1 BPA was completely degraded within 10 d accelerated by folic acid, Mg2+ and Ca2+. Ten more Pseudomonas strains could completely transformed 115 mg/L BPA [166]. Also, Bacillus pumilus strains isolated from Korean food, kimchi, were effective in removing BPA [170]. BPA with concentrations of 25 and 50 mg L−1 were successfully transformed without the detectible formation of 4-HAP in a medium supplemented with nutrients in the form of peptone, a beef extract, and a yeast extract in the presence of 10% NaCl [170].

Table 4 demonstrates different removal ratios for bacterial degradation of BPA. Recent studies found removal to nondetectable levels to 90% removal of BPA by bacterial strains tolerating high concentrations of this pollutant [171, 172]. Kang et al. [173] showed that Streptomyces sp. strain had greater BPA degradation potential with more than 90% in 10 d. Meanwhile several white-rot fungi removed BPA to undetectable levels in 12 h when the initial concentration of BPA was 50 mg mL−1 [174]. Interestingly, the transformation rate was higher in nutrient-rich YMG medium as compared to the minimal salts medium., Sphingomonas sp. strain BP-7 and Sphingomonas yanoikuyae BP-11R efficiently degraded 300 mg L−1 BPA without the formation of detectable BPA intermediates in the presence of activated carbon [175]. Furthermore, the degradation of BPA in the rhizosphere was enhanced in the presence of BPA-degrading bacterium, Novosphingobium sp. strain TYA-1 and a BPF-degrading bacterium, Sphingobium yanoikuyae strain TYF-1, isolated from the rhizosphere of Phragmites australis. BPA persisted in the soil in the absence of P. australis; hence, the interactions of P. australis in the rhizosphere efficiently removed BPA [176].

Table 4 Transformation of BPA by different microorganisms

While bacteria and enzymes can facilitate the degradation of BPA, recycling or direct disposal of the resultant by-products, especially in landfills or sludge, may pose potential environmental risks. The process of BPA degradation may not always yield complete breakdown, leading to the formation of intermediates that could retain toxicity. Without a thorough analysis of these by-products, their introduction into recycling processes or direct disposal in landfills could result in unintended environmental contamination. Additionally, the potential for leaching of these compounds from landfills into the surrounding environment raises concerns about soil and water pollution. It is crucial to consider the long-term environmental impact and ensure compliance with environmental regulations to avoid adverse consequences and safeguard ecosystems. Advocating for a comprehensive approach that includes rigorous assessment, adherence to regulations, and sustainable disposal practices is essential to address the potential safety risks associated with recycling or direct disposal of BPA or its degradation by-products. Further research is warranted to better understand the fate and environmental implications of these by-products.

4.1.3 Limitation of biodegradation treatment of BPA

Although effective, biodegradation of BPA has certain drawback. In biodegradation, the breakdown of BPA is microbe dependent. These microbes can differ in their BPA specificity which could restrict the efficacy of the biological treatment since certain bacteria lack the enzyme machinery needed for successful BPA breakdown [178]. Previously, Acinetobacter sp. K1MN and Pseudomonas sp. BG12 have shown the capacity for BPA removal with reduction of its toxic effect [179]. In contrast, some strains of Lactococcus showed BPA adsorption but not degradation. In this regard, the hydrophobicity of BPA may be an important factor for its adsorption with lactococci where lactococci might adsorb BPA by hydrophobic-binding effect of proteins on cell surface [180]. Consequently, it is critical to find novel bacterial strains and look into how well they degrade BPA for effective biodegradation treatment. BPA may also have adverse effects on bacteria which may limit their capacity when attempting to process BPA. The toxicity of BPA might interfere with or disrupt the normal metabolic processes of the bacteria responsible for degradation which further hinder their ability to carry out efficient degradation processes [162]. Environmental factors such temperature, pH, and nutrient availability have a significant impact on how well biodegradation treatments work [181]. Microorganisms cannot function as well in suboptimal settings, which can lessen the efficacy of treatment [182]. Maintaining ideal conditions for microbial activity can be challenging when implementing biological treatment systems, and it may call for careful control. This complexity may increase the overall cost of the treatment procedure and create operational challenges. Besides that, extended periods of time may be necessary for biodegradation treatment to effectively remove BPA in certain situations, particularly when the concentration of BPA is high. When dealing with high BPA amounts, biological treatment may be less successful as microorganisms have the potential to degrade BPA at a comparatively sluggish rate. From the reported literatures, Sphingomonas sp. strain AO1 completely degrade 115 mg L−1 of BPA within 6 h, however Sphingomonas sp. strain MV1 completely degrade 10 g L−1 BPA within 4 days [183]. With regard to this, post treatment or combination of other techniques might be required to remove BPA efficiently. Another limitation of biodegradation treatment that is worth highlighting is the formation of intermediate compound during the BPA degradation process. These intermediates may possess distinct chemical properties and could potentially introduce new challenges, as their environmental fate and toxicity may differ from that of the original BPA [184]. Finally, the application of biodegradation treatment to BPA contamination in non-watery matrices or solid substrates is limited because it is generally intended for aqueous conditions. When applied to solid substrates, the fundamental design of biodegradation treatment which is optimized for aqueous environments may result in diminished efficacy. Effective BPA removal or degradation is restricted by the unique physicochemical properties of solid materials, which calls for a sophisticated strategy catered to the unique features of these matrices. The complexity of implementing and managing biological treatment systems, coupled with their limited applicability to aqueous environments, adds operational challenges and considerations for optimizing BPA remediation strategies. Understanding all biodegradation treatment limitations is crucial for optimizing the use of biological treatment methods and for considering alternative or complementary approaches when addressing BPA contamination in different environmental settings.

4.2 Advanced oxidation processes

4.2.1 Photocatalysis

The development of advanced treatment techniques is now a subject of significant interest. Advanced oxidation processes have been commonly used for the treatment of various wastewaters because strong oxidants can easily transform a broad variety of organic and inorganic contaminants such as heavy metals, metalloids, and certain ions present in water [185]. Advanced oxidation processes use intensively reactive species such as hydroxyl radicals, superoxide anion radicals, and H2O2 as the main oxidizing agent. Generally, photocatalysis is a physical process of oxidation on the surface of catalyst excited by the source of light (Fig. 5) [186]. Photocatalytic degradation oxidizes complex organic compounds into smaller molecules, such as carbon dioxide and water, under light irradiation. In addition, photocatalysis may mineralize organic compounds to the harmless materials under sensible conditions due to several advantages include complete mineralization and up to 90% transformation efficiency in treating organic compounds at trace concentrations [187].

Fig. 5
figure 5

principle action of photocatalytic degradation [186] *(Disclaimer: Copyright © 2013 Dipak Nipane et al. This is an open access article distributed under the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.)

Among advanced oxidation processes, photocatalysis with titanium oxide (TiO2), zinc oxide (ZnO), ferric oxide (Fe2O3), cadmium sulfide (CdS), and zinc sulfide (ZnS) has been reported with efficient degradation [188]. Table 5 demonstrates the degradation of BPA by various photocatalysts at various optimum conditions. For instance, in a study conducted by Ohko et al. [189], 99% of BPA at an initial concentration of 40 ppm was degraded by TiO2 powder in the presence of Hg-Xe lamp after 15 h reaction. Similarly, 90% BPA degradation (initial concentration was 0.1 mM) was achieved in the presence of 0.8 g L−1 TiO2 particles after 4 h of reaction under the same light irradiations [190]. Interestingly, increased volume of photocatalyst would efficiently enhance the degradation of BPA where 98% degradation of BPA was reported in a 2 h of reaction in the presence of 120 mg L−1 TiO2 99% anatase in the presence of UV light irradiations. Zn–O photocatalyst is another semiconductor oxide that has been widely studied due to the favorable photocatalytic properties. 98% of BPA at an initial concentration of 25 ppm was degraded by 0.1% w/w zinc powder after 5 h of reaction [190]. Bechambi et al. [191] found undetectable BPA after 24 h of reaction with 1 g L−1 Ce-ZnO powder, in the presence of 5.1 M H2O2 under UV irradiation. Recently, Hunge et al. [192] synthesised TiO2-3@nanodiamond (ND) composite photocatalysts which revealed that neutral and acidic conditions enhanced BPA degradation and complete removal was achieved with 8 mg of TiO2-3@ND composite at pH 5.1 under UV light after 100 min of experimental works. With regard to Table 5, although photocatalytic activity is efficient, ZnO based photocatalyst has a wide band gap and easily photocorrosive [193]. Silver-based photocatalyst such as oxoacid salt (Ag3PO4) is promising as an efficient photocatalyst [194]. There are materials less explored with pollutant management studies with literature on those materials scarce. All detectable BPA at an initial concentration of 10 ppm was transformed by 1.67 mg mL−1 Ag3PO4 and 150 mg Ag3PO4 graphene hydrogel in the presence of Xe lamp after 10 and 12 min of reaction, respectively. Similar to Zn–O based photocatalyst, Ag-based photocatalyst tends to undergo photo-corrosion when exposed to a light source. This instability is the major limitation as a potential photocatalyst for pollutants removal. Anyway the photocatalytic reaction using TiO2 has been extensively studied due to inexpensive, photostability, and long-term chemical stability against chemical corrosion [195]. Additionally, TiO2 has become the focus of intense interest due to the capability to utilise UV light to efficiently catalyse reactions for the removal of a large variety of organic and inorganic contaminants such as heavy metals, metalloids, and certain ions [185, 196].

Table 5 Degradation of bisphenol A by different photocatalysis

Based on an extensive evaluation of the literatures by Reddy et al. [205], substantial progress has been made with visible light-active photocatalysts especially in the case of titania-based materials. Black titania catalysts have shown excellent light absorption capacities and efficiencies in the treatment of aqueous pollutants. It is worth mentioning that titania, silver, and bismuth photocatalysts have excellent performance for the assessment of photocatalytic performance. Similarly, Nath et al. [184] describe application of several heterogeneous catalysts for the breakdown of BPA via Fenton-like processes. New materials for catalyst synthesis such as g-C3N4 and MXenes provide an interesting improvement in degradation efficiency [184]. While photocatalyst materials work excellently in BPA degradation, it is crucial to ensure that they can be readily separated from the reaction mixture in order to maximize their recycling and reusability. The overall sustainability and economy of photocatalytic processes are enhanced by the use of readily separable photocatalyst materials.

4.2.2 Hybrid process: photocatalytic membrane

Usually, a photocatalytic reaction is conducted in a suspension of catalyst particles; however, the separation of the TiO2 particles is a major problem despite the efficiency of the photocatalytic activity. Fabrication of nanostructured TiO2 hollow fibre photocatalytic membrane [206] demonstrated organic removal of 90% by filtration and photocatalytic degradation of the membrane in the presence of UV irradiation. Recent fabrication of hollow fibre membranes from polyethersulfone (PES) and TiO2 for water treatment was demonstrated to be photoactive in the degradation of Methylene Blue [207]. However, the selective layer was damaged during the UV-A irradiation of only 2 h at 0.6 mW cm−2, which indicated that PES-TiO2 ultrafiltration membranes are not resistant to UV radiation. Polyvinylidene fluoride (PVDF) hollow fibre membranes with nano-TiO2 additives were fabricated previously [208] for dense algae cultivation and wastewater polishing. The nano TiO2-embedded membranes improved the surface hydrophilicity with the total resistance of about 50% lower than that of the control. This study demonstrated that TiO2 nanocomposite membranes had a better antifouling property for dense algae cultivation and wastewater polishing. Only a few were reported on the photodegradation of BPA by TiO2 membrane as listed in Table 6. However, the degradation of BPA by photocatalytic TiO2 hollow fibre membrane is still less explored, hence, available information is limited.

Table 6 Degradation of BPA by photocatalysis immobilised in a membrane matrix

The incorporation of TiO2 particles in the membrane matrix would eliminate the post-treatment process; however, it will decrease the photocatalyst active surface area [212]. This leads to the emerging in the exploration and fabrication of photocatalytic DLHF membrane for the excellent performance of immobilised photocatalyst. The outer surface layer of the membrane has embedded immobilised TiO2 particles that catalytically react with contaminants in the presence of visible light and UV radiation, while the inner membrane layer will act as a separation layer. Interestingly, large loading of more than 50% by weight of photocatalyst immobilised at the outer layer membrane surface will ensure the fast transformation of the pollutant in the membrane photoreactor (MPR, Fig. 6). Briefly, MPR consists of submerged photocatalytic membrane modules fitted with light (LED and UV) source. The membrane modules will be immersed, and suction will be applied by the peristaltic pump for permeate collection. The compact design of photocatalytic membrane simultaneously transforms BPA and filter particulates, which allowed the membrane to be directly used for water treatment without leaving any photocatalytic particles in the water.

Fig. 6
figure 6

Photocatalytic dual layer hollow fibre membrane in the MPR system

Dzinun et al. [213] has successfully fabricated a photocatalytic DLHF membrane with surface TiO2 immobilised for the removal of nonylphenol. The DLHF membrane has a sandwich-like structure where both layers were filled with sponge like structure with the presence of finger-like void in the inner and outer space. The distribution of TiO2 on the outer layer membrane surface has improved the membrane hydrophilicity, pore size, and permeability performance with almost 85% of nonylphenol removal was achieved under UVA irradiation. Previously, the fabrications of TiO2/ZnO DLHF membranes were proposed as a superior sensor material in comparison to regular single layer hollow fibres (SLHFs) for the detection of reducing gases, such as CO [214]. The study reported that TiO2/ZnO DLHF membranes demonstrated a better behaviour compared to SLFHs when detecting reducing gases but showed a lower response to oxidising gases such as NO2. In this regard, the inner TiO2 layer is responsible for electron absorption, enabling the use of a thicker outer layer than the one that would be required to achieve the same sensing properties in the case of SLHFs. Hence, TiO2/ZnO DLHF membranes were proven to be advantageous over regular SLHFs for the detection of reducing gases.

Recently, Fu et al. [215] summarized the progress in photocatalytic degradation of BPA. According to the literature, a few issues still need to be addressed to improve the efficiency of photocatalytic degradation include: (1) low efficiency of photogenerated electron–hole pair recombination and visible light utilization; (2) large mass transfer resistance; and (3) poor selectivity of photocatalysts. To enhance photocatalytic efficiency, expanding the visible spectrum response range and minimizing the recombination of photogenerated electron–hole pairs are crucial. Different approaches, like using conductive materials as photocatalysts, have been suggested to shift the spectrum bands from UV to visible light, enhancing electron–hole separation and charge transfer speed. Interestingly, Kamaludin et al. [142] successfully fabricated visible light driven photocatalytic DLHF) membrane for the removal of BPA from water. Removal efficiencies of 88 and 87% of BPA were achieved after 360 min of visible and UV light irradiation, respectively. Interestingly, N-doped TiO2 DLHF membrane effectively mitigates the effect of BPA. This finding shows that N-doped TiO2 DLHF membrane is a promising option for the removal of BPA from contaminated water sources. Recently, Mohamad Noor et al. [216] successfully developed a Cu2O composited PVDF DLHF membrane. Under visible light irradiation, about 75% of the 10 ppm BPA was successfully removed in 360 min, without the copper element being leached. Similarly, Sukaini et al. [217] reported that Cu2O/WO3/PVDF DLHF membrane efficiently remove 98% of 2-ppm BPA without copper leaching into the water sample. Interestingly, under dark conditions, the DLHF sample showed the capability of energy storage performance and could drive certain degradation after lighting off up to 71% of 2-ppm BPA. This finding showed that the development of photocatalysts with energy storage capability can provide a significant contribution toward more practical photocatalytic membranes, so photodegradation can occur even in dark conditions. In the meantime, Zakria et al. [218] successfully photodegraded about 74% of 2 mg/L BPA in synthetic wastewater while outstanding 91% removal of BPA in real treated sewage wastewater achieved using Cu2O/PVDF thin film hollow fiber membrane after 360 min light irradiation. Despite the excellent photodegradation performance, the fabrication of DLHF membranes for photocatalytic degradation of BPA is still less explored. Similarly, the mitigation of BPA health effect on the treated water was not assessed. Therefore, relevant literature is limited.

To further emphasize the selection of photocatalytic membranes for BPA removal, comparison between biodegradation treatment and photocatalytic membrane is summarized into Table 7. This comparison will encompass factors such as efficiency, cost-effectiveness, environmental impact, scalability, and practical feasibility to provide valuable insights into the unique attributes of each method. This comparison will also provide a comprehensive view to support the preference for photocatalytic membranes in BPA degradation.

Table 7 Comparison between biodegradation treatment and photocatalytic membrane

Photocatalytic membranes, while offering notable advantages in the degradation of BPA, are not without significant drawbacks. One well-known limitation is their reliance on energy-intensive UV light for optimal performance, introducing potential operational cost concerns and limiting energy efficiency. However, photocatalytic membranes benefits, such the capacity to effectively break down persistent pollutants like BPA and act as a physical barrier for membrane-based separation underscore their potential as a multifaceted solution for water treatment. The ability to simultaneously address chemical degradation and physical separation enhances their overall efficacy. For instance, while photocatalysis may face challenges like energy consumption and potential catalyst fouling, the integration of photocatalytic membranes combines the benefits of membrane filtration and advanced oxidation processes. This integration enhances the overall efficiency of BPA removal by addressing both physical separation and degradation aspects in a single step. Comparing photocatalytic membranes with other treatment options, such as biodegradation and enzymatic treatments, it reveals a nuanced landscape. Biodegradation and enzymatic treatments offer eco-friendly solutions leveraging natural processes but may face limitations in terms of scalability and the need for optimal environmental conditions [219]. Photocatalytic membranes, on the other hand, provide a continuous and scalable treatment method that operates under various environmental conditions [220]. However, the energy requirements and potential catalyst fouling must be carefully considered. In summary, the advantages of photocatalytic membranes, particularly their dual functionality, can offset certain drawbacks. When compared with other treatment options, the choice between them should be contextual, considering factors such as energy efficiency, scalability, and the specific environmental conditions in which the treatment will be applied. The holistic evaluation of these factors will guide the selection of the most suitable and effective treatment approach for addressing BPA contamination.

5 Conclusions

This paper provided an outline on the presence of BPA in surface water, household water, and drinking water and adverse health effect associated with BPA exposure. Numerous studies on metabolic diseases have proven that metabolic diseases in adults can originate in utero and during infancy. Furthermore, experimental research animals support the concepts of a causal link between adverse events in utero and early life with the development of cardiovascular and metabolic disease in later life. In addition, BPA levels detected in the umbilical cord, foetal cord serum and amniotic fluid indicate that BPA can cross trans-placentally and potentially impact the developing foetus. Multiple studied rodents and primate demonstrated the long-term detrimental effects of BPA exposure on foetal organs. Therefore, BPA exposure during pregnancy should be given a better emphasis to lower the burden on the development of metabolic diseases in the future.

BPAs could be successfully removed by enzymatic and microbial treatment in the presence of suitable mediators. AOPs, including photocatalysis, ozonation, and Fenton reactions, have shown effectiveness in breaking down BPA and its byproducts through the generation of highly reactive species. Integrating AOPs with photocatalytic membranes could further enhance the removal efficiency and ensure complete mineralization of BPA, addressing both chemical degradation and physical separation in one step. The appeal of photocatalytic degradation stands out due to its unique properties, including anti-fouling, superhydrophilicity, and concurrent photocatalytic oxidation and separation. Photocatalytic membranes emerge as a superior solution for degrading BPA in drinking water and its sources when compared to biodegradation treatments. The continuous and scalable nature of photocatalytic membranes ensures uninterrupted BPA degradation, making them suitable for both small-scale applications and large-scale municipal water treatment. Their robustness and ability to operate under diverse environmental conditions further enhance their reliability. Importantly, photocatalytic membranes provide a dual functionality by simultaneously physically separating contaminants and chemically degrading BPA. This distinguishes them from biodegradation methods, which may face challenges in scalability and robustness, relying on microbial activity that can be sensitive to fluctuations in environmental parameters. Photocatalytic membranes, with reduced dependence on specific microbial conditions, offer a versatile and effective approach to address BPA contamination in drinking water, presenting a holistic solution that combines chemical degradation with efficient physical separation. However, some technical issues still need to be resolved such as increase the contact area between photocatalyst, and UV for more effective photocatalytic activity. Furthermore, solar irradiated photocatalytic membranes need to be explored, since most of photocatalytic membranes in the literature are excited only by UV. Finally, there is still a lack of durability data of these photocatalytic membranes. It is inevitable that some photocatalyst may leach off the membranes, hence shortening their lifespan. As the use of BPA is expected to increase in the future, more deliberate efforts to develop diverse yet effective treatment options are imperative to manage water quality. Further research is urgently needed to investigate the transformation pathways of BPA and the intermediates as well as to determine the effect and fate of the intermediates in the environment.

Availability of data and materials

Not applicable.

References

  1. Michalowicz J. Bisphenol A – Sources, toxicity and biotransformation. Environ Toxicol Phar. 2014;37:738–58.

    Article  Google Scholar 

  2. Bolli A, Galluzzo P, Ascenzi P, Del Pozzo G, Manco I, Vietri MT, et al. Laccase treatment impairs bisphenol A-induced cancer cell proliferation affecting estrogen receptor alpha-dependent rapid signals. Iubmb Life. 2008;60:843–52.

    Article  Google Scholar 

  3. Bondesson M, Jonsson J, Pongratz I, Olea N, Cravedi JP, Zalko D, et al. A CASCADE of effects of bisphenol A. Reprod Toxicol. 2009;28:563–7.

    Article  Google Scholar 

  4. Dodds EC, Lawson W. Synthetic strogenic agents without the phenanthrene nucleus. Nature. 1936;137:996.

    Article  Google Scholar 

  5. Krishnan AV, Stathis P, Permuth SF, Tokes L, Feldman D. Bisphenol-A: an estrogenic substance is released from polycarbonate flasks during autoclaving. Endocrinology. 1993;132:2279–86.

    Article  Google Scholar 

  6. Vandenberg LN, Hauser R, Marcus M, Olea N, Welshons WV. Human exposure to bisphenol A (BPA). Reprod Toxicol. 2007;24:139–77.

    Article  Google Scholar 

  7. Jiang Y, Li J, Xu S, Zhou Y, Zhao H, Li Y, et al. Prenatal exposure to bisphenol A and its alternatives and child neurodevelopment at 2 years. J Hazard Mater. 2020;388:121774.

    Article  Google Scholar 

  8. Gertz J, Reddy TE, Varley KE, Garabedian MJ, Myers RM. Genistein and bisphenol A exposure cause estrogen receptor 1 to bind thousands of sites in a cell type-specific manner. Genome Res. 2012;22:2153–62.

    Article  Google Scholar 

  9. Santangeli S, Consales C, Pacchierotti F, Habibi HR, Carnevali O. Transgenerational effects of BPA on female reproduction. Sci Total Environ. 2019;685:1294–305.

    Article  Google Scholar 

  10. Vafeiadi M, Roumeliotaki T, Myridakis A, Chalkiadaki G, Fthenou E, Dermitzaki E, et al. Association of early life exposure to bisphenol A with obesity and cardiometabolic traits in childhood. Environ Res. 2016;146:379–87.

    Article  Google Scholar 

  11. Xie X, Song J, Li G. MiR-21a-5p suppresses bisphenol A-induced pre-adipocyte differentiation by targeting map2k3 through MKK3/p38/MAPK. Biochem Bioph Res Co. 2016;473:140–6.

    Article  Google Scholar 

  12. Sakurai K, Kawazuma M, Adachi T, Harigaya T, Saito Y, Hashimoto N, et al. Bisphenol A affects glucose transport in mouse 3T3-F442A adipocytes. Brit J Pharmacol. 2004;141:209–14.

    Article  Google Scholar 

  13. Sivashanmugam P, Mullainadhan V, Karundevi B. Dose-dependent effect of Bisphenol-A on insulin signaling molecules in cardiac muscle of adult male rat. Chem-Biol Interact. 2017;266:10–6.

    Article  Google Scholar 

  14. Shiue I, Shiue YY. The role of housing characteristics in biomarkers: US NHANES, 2003-2006. Int J Cardiol. 2013;168:5069–71.

    Article  Google Scholar 

  15. Shankar A, Teppala S. Relationship between serum cystatin C and hypertension among US adults without clinically recognized chronic kidney disease. J Am Soc Hypertens. 2011;5:378–84.

    Article  Google Scholar 

  16. Ranciere F, Lyons JG, Loh VHY, Botton J, Galloway T, Wang T, et al. Bisphenol A and the risk of cardiometabolic disorders: a systematic review with meta-analysis of the epidemiological evidence. Environ Health-Glob. 2015;14:46.

    Google Scholar 

  17. Braun JM, Kalkbrenner AE, Calafat AM, Yolton K, Ye X, Dietrich KN, et al. Impact of early-life bisphenol A exposure on behavior and executive function in children. Pediatrics. 2011;128:873–82.

    Article  Google Scholar 

  18. Chapalamadugu KC, Vandevoort CA, Settles ML, Robison BD, Murdoch GK. Maternal bisphenol A exposure impacts the fetal heart transcriptome. PloS ONE. 2014;9:e89096.

    Article  Google Scholar 

  19. Khalil N, Ebert JR, Wang L, Belcher S, Lee M, Czerwinski SA, et al. Bisphenol A and cardiometabolic risk factors in obese children. Sci Total Environ. 2014;470–471:726–32.

    Article  Google Scholar 

  20. Gao XQ, Wang HS. "Impact of bisphenol A on the cardiovascular system — Epidemiological and experimental evidence and molecular mechanisms. Int J Env Res Pub He. 2014;11:8399–413.

    Article  Google Scholar 

  21. Ge LC, Chen ZJ, Liu HY, Zhang KS, Liu H, Huang HB, et al. Involvement of activating ERK1/2 through G protein coupled receptor 30 and estrogen receptor α/β in low doses of bisphenol A promoting growth of Sertoli TM4 cells. Toxicol Lett. 2014;226:81–9.

    Article  Google Scholar 

  22. Heindel JJ, Newbold RR, Bucher JR, Camacho L, Delclos KB, Lewis SM, et al. NIEHS/FDA CLARITY-BPA research program update. Reprod Toxicol. 2015;58:33–44.

    Article  Google Scholar 

  23. Harnett KG, Chin A, Schuh SM. BPA and BPA alternatives BPS, BPAF, and TMBPF, induce cytotoxicity and apoptosis in rat and human stem cells. Ecotoxicol Environ Saf. 2021;216:112210.

    Article  Google Scholar 

  24. Pironti C, Ricciardi M, Proto A, Bianco PM, Montano L, Motta O. Endocrine-disrupting compounds: an overview on their occurrence in the aquatic environment and human exposure. Water. 2021;13:1347.

    Article  Google Scholar 

  25. Shi JH, Suzuki Y, Lee BD, Nakai S, Hosomi M. Isolation and characterization of the ethynylestradiol-biodegrading microorganism Fusarium proliferatum strain HNS-1. Water Sci Technol. 2002;45:175–9.

    Article  Google Scholar 

  26. Cao TT, Zhou DD, Wang XD, Cui C. Photocatalytic degradation of bisphenol A over Co-BiOCl/biochar hybrid catalysts: Properties, efficiency and mechanism. J Mol Liq. 2022;362:119622.

    Article  Google Scholar 

  27. Auriol M, Filali-Meknassi Y, Tyagi RD, Adams CD, Surampalli RY. Endocrine disrupting compounds removal from wastewater, a new challenge. Process Biochem. 2006;41:525–39.

    Article  Google Scholar 

  28. Balest L, Mascolo G, Di Iaconi C, Lopez A. Removal of endocrine disrupter compounds from municipal wastewater by an innovative biological technology. Water Sci Technol. 2008;58:953–6.

    Article  Google Scholar 

  29. Pahigian JM, Zuo Y. Occurrence, endocrine-related bioeffects and fate of bisphenol A chemical degradation intermediates and impurities: A review. Chemosphere. 2018;207:469–80.

    Article  Google Scholar 

  30. Ma WF, Zhang Y, Li LL, You LJ, Zhang P, Zhang YT, et al. Tailor-made magnetic Fe3O4@mTiO2 microspheres with a tunable mesoporous anatase shell for highly selective and effective enrichment of phosphopeptides. ACS Nano. 2012;6:3179–88.

    Article  Google Scholar 

  31. Xing J, Zhang S, Zhang M, Hou J. A critical review of presence, removal and potential impacts of endocrine disruptors bisphenol A. Comp Biochem Phys C. 2022;254:109275.

    Google Scholar 

  32. Kawa IA, Akbar M, Fatima Q, Mir SA, Jeelani H, Manzoor S, et al. Endocrine disrupting chemical Bisphenol A and its potential effects on female health. Diabetes Metab Syndr Clin Res Rev. 2021;15:803–11.

    Article  Google Scholar 

  33. Singla P, Bardoloi A, Parkash AA. Metabolic effects of obesity: A review. World J Diabetes. 2010;1:76–88.

    Article  Google Scholar 

  34. Han C, Hong YC. Bisphenol A, hypertension, and cardiovascular diseases: Epidemiological, laboratory, and clinical trial evidence. Curr Hypertens Rep. 2016;18:11.

    Article  Google Scholar 

  35. Giulivo M, Lopez de Alda M, Capri E, Barcelo D. Human exposure to endocrine disrupting compounds Their role in reproductive systems metabolic syndrome and breast cancer A review. Environ Res. 2016;151:251–64.

    Article  Google Scholar 

  36. Perera F, Herbstman J. Prenatal environmental exposures, epigenetics, and disease. Reprod Toxicol. 2011;31:363–73.

    Article  Google Scholar 

  37. Casas M, Forns J, Martinez D, Avella-Garcia C, Valvi D, Ballesteros-Gomez A, et al. Exposure to bisphenol A during pregnancy and child neuropsychological development in the INMA-Sabadell cohort. Environ Res. 2015;142:671–9.

    Article  Google Scholar 

  38. Miki Y, Hata S, Nagasaki S, Suzuki T, Ito K, Kumamoto H, et al. Steroid and xenobiotic receptor-mediated effects of bisphenol A on human osteoblasts. Life Sci. 2016;155:29–35.

    Article  Google Scholar 

  39. Hanaoka T, Kawamura N, Hara K, Tsugane S. Urinary bisphenol A and plasma hormone concentrations in male workers exposed to bisphenol A diglycidyl ether and mixed organic solvents. Occup Environ Med. 2002;59:625–8.

    Article  Google Scholar 

  40. Yamada H, Furuta I, Kato EH, Kataoka S, Usuki Y, Kobashi G, et al. Maternal serum and amniotic fluid bisphenol A concentrations in the early second trimester. Reprod Toxicol. 2002;16:735–9.

    Article  Google Scholar 

  41. Sugiura-Ogasawara M, Ozaki Y, Sonta S, Makino T, Suzumori K. Exposure to bisphenol A is associated with recurrent miscarriage. Hum Reprod. 2005;20:2325–9.

    Article  Google Scholar 

  42. Aghakhanian F, Wong C, Tan JSY, Yeo LF, Ramadas A, Edo J, et al. Metabolic syndrome and cardiometabolic risk factors among indigenous Malaysians. Public Health. 2019;176:106–13.

    Article  Google Scholar 

  43. Pillon NJ, Loos RJF, Marshall SM, Zierath JR. Metabolic consequences of obesity and type 2 diabetes: Balancing genes and environment for personalized care. Cell. 2021;184:1530–44.

    Article  Google Scholar 

  44. Chu CY, Ponten A, Sun CC, Jee SH. Concomitant contact allergy to the resins, reactive diluents and hardener of a bisphenol A/F-based epoxy resin in subway construction workers. Contact Dermatitis. 2006;54:131–9.

    Article  Google Scholar 

  45. Li D, Zhou Z, Qing D, He Y, Wu T, Miao M, et al. Occupational exposure to bisphenol-A (BPA) and the risk of self-reported male sexual dysfunction. Hum Reprod. 2010;25:519–27.

    Article  Google Scholar 

  46. Wang L, Ying GG, Chen F, Zhang LJ, Zhao JL, Lai HJ, et al. Monitoring of selected estrogenic compounds and estrogenic activity in surface water and sediment of the Yellow River in China using combined chemical and biological tools. Environ Pollut. 2012;165:241–9.

    Article  Google Scholar 

  47. Wilson NK, Chuang JC, Lyu C, Menton R, Morgan MK. Aggregate exposures of nine preschool children to persistent organic pollutants at day care and at home. J Expo Sci Env Epid. 2003;13:187–202.

    Article  Google Scholar 

  48. Wilson NK, Chuang JC, Morgan MK, Lordo RA, Sheldon LS. An observational study of the potential exposures of preschool children to pentachlorophenol, bisphenol-A, and nonylphenol at home and daycare. Environ Res. 2007;103:9–20.

    Article  Google Scholar 

  49. LaKind JS, Goodman M, Naiman DQ. Use of NHANES data to link chemical exposures to chronic diseases: a cautionary tale. PLoS One. 2012;7:e51086.

    Article  Google Scholar 

  50. Mirmira P, Evans-Molina C. Bisphenol A, obesity, and type 2 diabetes mellitus: genuine concern or unnecessary preoccupation? Transl Res. 2014;164:13–21.

    Article  Google Scholar 

  51. Takeuchi T, Tsutsumi O, Ikezuki Y, Takai Y, Taketani Y. Positive relationship between androgen and the endocrine disruptor, bisphenol A, in normal women and women with ovarian dysfunction. Endocr J. 2004;51:165–9.

    Article  Google Scholar 

  52. Vom Saal FS, Cooke PS, Buchanan DL, Palanza P, Thayer KA, Nagel SC, et al. A physiologically based approach to the study of bisphenol A and other estrogenic chemicals on the size of reproductive organs, daily sperm production, and behavior. Toxicol Ind Health. 1998;14:239–60.

    Article  Google Scholar 

  53. Somm E, Schwitzgebel VM, Toulotte A, Cederroth CR, Combescure C, Nef S, et al. Perinatal exposure to bisphenol A alters early adipogenesis in the rat. Environ Health Persp. 2009;117:1549–55.

    Article  Google Scholar 

  54. Alonso-Magdalena P, Quesada I, Nadal A. Endocrine disruptors in the etiology of type 2 diabetes mellitus. Nat Rev Endocrinol. 2011;7:346–53.

    Article  Google Scholar 

  55. Alonso-Magdalena P, Vieira E, Soriano S, Menes L, Burks D, Quesada I, et al. Bisphenol A exposure during pregnancy disrupts glucose homeostasis in mothers and adult male offspring. Environ Health Persp. 2010;118:1243–50.

    Article  Google Scholar 

  56. Soriano S, Alonso-Magdalena P, Garcia-Arevalo M, Novials A, Muhammed SJ, Salehi A, et al. Rapid insulinotropic action of low doses of bisphenol-A on mouse and human islets of Langerhans: role of estrogen receptor β. PLoS One. 2012;7:e31109.

    Article  Google Scholar 

  57. Liang Q, Gao X, Chen Y, Hong K, Wang HS. Cellular mechanism of the nonmonotonic dose response of bisphenol A in rat cardiac myocytes. Environ Health Perspect. 2014;122:601–08.

    Article  Google Scholar 

  58. Rubin BS, Murray MK, Damassa DA, King JC, Soto A. M. Perinatal exposure to low doses of bisphenol A affects body weight, patterns of estrous cyclicity, and plasma LH levels. Environ Health Perspect. 2001;109:675–80.

    Article  Google Scholar 

  59. Hijazi A, Guan H, Cernea M, Yang K. Prenatal exposure to bisphenol A disrupts mouse fetal lung development. FASEB J. 2015;29:4968–77.

    Article  Google Scholar 

  60. Hong S H et al. Urinary bisphenol A is associated with insulin resistance and obesity in reproductive‐aged women. Clinical Endocrinol. 2017;86:506–12.

    Article  Google Scholar 

  61. Trasande L, Attina TM, Blustein J. Association between urinary bisphenol A concentration and obesity prevalence in children and adolescents. Jama. 2012;308:1113–21.

    Article  Google Scholar 

  62. Nakamura D, Yanagiba Y, Duan Z, Ito Y, Okamura A, Asaeda N, et al. Bisphenol A may cause testosterone reduction by adversely affecting both testis and pituitary systems similar to estradiol. Toxicol Lett. 2010;194:16–25.

    Article  Google Scholar 

  63. Lombo M, Fernandez-Diez C, Gonzalez-Rojo S, Navarro C, Robles V, Herraez MP. Transgenerational inheritance of heart disorders caused by paternal bisphenol A exposure. Environ Pollut. 2015;206:667–78.

    Article  Google Scholar 

  64. Moreno-Gomez-Toledano R, Arenas MI, Velez-Velez E, Coll E, Quiroga B, Bover J, et al. Bisphenol A exposure and kidney diseases: systematic review, meta-analysis, and NHANES 03–16 study. Biomolecules. 2021;11:1046.

    Article  Google Scholar 

  65. DeBenedictis B, Guan H, Yang K. Prenatal exposure to bisphenol A disrupts mouse fetal liver maturation in a sex-specific manner. J Cell Biochem. 2016;117:344–50.

    Article  Google Scholar 

  66. Fenichel P, Chevalier N, Brucker-Davis F. Bisphenol A: an endocrine and metabolic disruptor. Annales d'endocrinologie. 2013;74:211–20.

    Article  Google Scholar 

  67. Hugo ER, Brandebourg TD, Woo JG, Loftus J, Alexander JW, Ben-Jonathan N. Bisphenol A at environmentally relevant doses inhibits adiponectin release from human adipose tissue explants and adipocytes. Environ Health Perspect. 2008;116:1642–7.

    Article  Google Scholar 

  68. N'Tumba-Byn T, Moison D, Lacroix M, Lecureuil C, Lesage L, Prud'homme SM, et al. Differential effects of bisphenol A and diethylstilbestrol on human, rat and mouse fetal leydig cell function. PLoS One. 2012;7:e51579.

    Article  Google Scholar 

  69. Wada K, Sakamoto H, Nishikawa K, Sakuma S, Nakajima A, Fujimoto Y, et al. Life style-related diseases of the digestive system: endocrine disruptors stimulate lipid accumulation in target cells related to metabolic syndrome. J Pharmacol Sci. 2007;105:133–7.

    Article  Google Scholar 

  70. Ben-Jonathan N, LaPensee CR, LaPensee EW. What can we learn from rodents about prolactin in humans? Endocr Rev. 2008;29:1–41.

    Article  Google Scholar 

  71. Posnack NG, Jaimes R, 3rd, Asfour H, Swift LM, Wengrowski AM, Sarvazyan N, et al. Bisphenol A exposure and cardiac electrical conduction in excised rat hearts. Environ Health Perspect. 2014;122:384–90.

    Article  Google Scholar 

  72. Patel BB, Raad M, Sebag IA, Chalifour LE. Lifelong exposure to bisphenol a alters cardiac structure/function, protein expression, and DNA methylation in adult mice. Toxicol Sci. 2013;133:174–85.

    Article  Google Scholar 

  73. Malik R, Mushtaque RS, Siddiqui UA, Younus A, Aziz MA, Humayun C, et al. Association between coronary artery disease and microRNA: literature review and clinical perspective. Cureus. 2017;9:e1188.

    Google Scholar 

  74. Shankar A, Teppala S, Sabanayagam C. Bisphenol A and peripheral arterial disease: results from the NHANES. Environ Health Perspect. 2012;120:1297–300.

    Article  Google Scholar 

  75. Jiang Y, Xia W, Yang J, Zhu Y, Chang H, Liu J, et al. BPA-induced DNA hypermethylation of the master mitochondrial gene PGC-1α contributes to cardiomyopathy in male rats. Toxicology. 2015;329:21–31.

    Article  Google Scholar 

  76. Gonzalez A, Schelbert EB, Díez J, Butler J. Myocardial interstitial fibrosis in heart failure: biological and translational perspectives. J Am Coll Cardiol. 2018;71:1696–706.

    Article  Google Scholar 

  77. Hu Y, Zhang L, Wu X, Hou L, Li Z, Ju J, et al. Bisphenol A, an environmental estrogen-like toxic chemical, induces cardiac fibrosis by activating the ERK1/2 pathway. Toxicol Lett. 2016;250–251:1–9.

    Article  Google Scholar 

  78. Patel BB, Raad M, Sebag IA, Chalifour LE. Sex-specific cardiovascular responses to control or high fat diet feeding in C57bl/6 mice chronically exposed to bisphenol A. Toxicol Rep. 2015;2:1310–8.

    Article  Google Scholar 

  79. Yan S, Song W, Chen Y, Hong K, Rubinstein J, Wang HS. Low-dose bisphenol A and estrogen increase ventricular arrhythmias following ischemia–reperfusion in female rat hearts. Food Chem Toxicol. 2013;56:75–80.

    Article  Google Scholar 

  80. Bae S, Kim JH, Lim YH, Park HY, Hong YC. Associations of bisphenol A exposure with heart rate variability and blood pressure. Hypertension. 2012;60:786–93.

    Article  Google Scholar 

  81. Kilic U, Gok O, Bacaksiz A, Izmirli M, Elibol-Can B, Uysal O. SIRT1 gene polymorphisms affect the protein expression in cardiovascular diseases. PLoS One. 2014;9:e90428.

    Article  Google Scholar 

  82. Peters U, Dixon AE, Forno E. Obesity and asthma. J Allergy Clin Immun. 2018;141:1169–79.

    Article  Google Scholar 

  83. Do MT, Chang VC, Mendez MA, de Groh M. Urinary bisphenol A and obesity in adults: results from the Canadian Health Measures Survey. Health Promot Chronic Dis Prev Can. 2017;37:403–12.

    Article  Google Scholar 

  84. Slattery ML, Ferucci ED, Murtaugh MA, Edwards S, Ma KN, Etzel RA, et al. Associations among body mass index, waist circumference, and health indicators in American Indian and Alaska Native adults. Am J Health Promot. 2010;24:246–54.

    Article  Google Scholar 

  85. Zhao HY, Bi YF, Ma LY, Zhao L, Wang TG, Zhang LZ, et al. The effects of bisphenol A (BPA) exposure on fat mass and serum leptin concentrations have no impact on bone mineral densities in non-obese premenopausal women. Clin Biochem. 2012;45:1602–6.

    Article  Google Scholar 

  86. Al-Daghri NM, Abd-Alrahman SH, Bukhari I, Panigrahy A, Khattak MN, Alokail M. BPA exposure is related to metabolic changes in obese Saudi children. Int J Clin Exp Patho. 2017;10:9910–6.

    Google Scholar 

  87. Bansal A, Rashid C, Xin F, Li C, Polyak E, Duemler A, et al. Sex- and dose-specific effects of maternal bisphenol A exposure on pancreatic islets of first- and second-generation adult mice offspring. Environ Health Perspect. 2017;125:097022.

    Article  Google Scholar 

  88. Wilcox G. Insulin and insulin resistance. Clin Biochem Rev. 2005;26:19-39.

    Google Scholar 

  89. Shao D, Tian R. Glucose transporters in cardiac metabolism and hypertrophy. Compr Physiol. 2015;6:331–51.

    Article  Google Scholar 

  90. Alonso-Magdalena P, Laribi O, Ropero AB, Fuentes E, Ripoll C, Soria B, et al. Low doses of bisphenol A and diethylstilbestrol impair Ca2+ signals in pancreatic α-cells through a nonclassical membrane estrogen receptor within intact islets of Langerhans. Environ Health Perspect. 2005;113:969–77.

    Article  Google Scholar 

  91. Guan Y, Gao J, Zhang Y, Chen S, Yuan C, Wang Z. Effects of bisphenol A on lipid metabolism in rare minnow Gobiocypris rarus. Comp Biochem Phys C. 2016;179:144–9.

    Google Scholar 

  92. Moriyama K, Tagami T, Akamizu T, Usui T, Saijo M, Kanamoto N, et al. Thyroid hormone action is disrupted by bisphenol A as an antagonist. J Clin Endocr Metab. 2002;87:5185–90.

    Article  Google Scholar 

  93. Smith-Bindman R. Environmental causes of breast cancer and radiation from medical imaging: findings from the Institute of Medicine report. Arch Intern Med. 2012;172:1023–7.

    Article  Google Scholar 

  94. Paulose T, Speroni L, Sonnenschein C, Soto AM. Estrogens in the wrong place at the wrong time: Fetal BPA exposure and mammary cancer. Reprod Toxicol. 2015;54:58–65.

    Article  Google Scholar 

  95. Jenkins S, Wang J, Eltoum I, Desmond R, Lamartiniere CA. Chronic oral exposure to bisphenol A results in a nonmonotonic dose response in mammary carcinogenesis and metastasis in MMTV-erbB2 mice. Environ Health Perspect. 2011;119:1604–9.

    Article  Google Scholar 

  96. Dairkee SH, Seok J, Champion S, Sayeed A, Mindrinos M, Xiao W, et al. Bisphenol A induces a profile of tumor aggressiveness in high-risk cells from breast cancer patients. Cancer Res. 2008;68:2076–80.

    Article  Google Scholar 

  97. Hiroi H, Tsutsumi O, Takeuchi T, Momoeda M, Ikezuki Y, Okamura A, et al. Differences in serum bisphenol a concentrations in premenopausal normal women and women with endometrial hyperplasia. Endocr J. 2004;51:595–600.

    Article  Google Scholar 

  98. Lapensee EW, Tuttle TR, Fox SR, Ben-Jonathan N. Bisphenol A at low nanomolar doses confers chemoresistance in estrogen receptor-α-positive and -negative breast cancer cells. Environ Health Perspect. 2009;117:175–80.

    Article  Google Scholar 

  99. Prins GS, Korach KS. The role of estrogens and estrogen receptors in normal prostate growth and disease. Steroids. 2008;73:233–44.

    Article  Google Scholar 

  100. Prins GS, Tang WY, Belmonte J, Ho SM. Perinatal exposure to oestradiol and bisphenol A alters the prostate epigenome and increases susceptibility to carcinogenesis. Basic Clin Pharmacol Toxicol. 2008;102:134–8.

    Article  Google Scholar 

  101. Roger C, Lambard S, Bouskine A, Mograbi B, Chevallier D, Nebout M, et al. Estrogen-induced growth inhibition of human seminoma cells expressing estrogen receptor β and aromatase. J Mol Endocrinol. 2005;35:191–9.

    Article  Google Scholar 

  102. Grasselli F, Baratta L, Baioni L, Bussolati S, Ramoni R, Grolli S, et al. Bisphenol A disrupts granulosa cell function. Domest Anim Endocrin. 2010;39:34–9.

    Article  Google Scholar 

  103. Flint S, Markle T, Thompson S, Wallace E. Bisphenol A exposure, effects, and policy: A wildlife perspective. J Environ Manage. 2012;104:19–34.

    Article  Google Scholar 

  104. Manzoor MF, Tariq T, Fatima B, Sahar A, Tariq F, Munir S, et al. An insight into bisphenol A, food exposure and its adverse effects on health: A review. Front Nutr. 2022;9:1047827.

    Article  Google Scholar 

  105. Cunha SC, Cunha C, Ferreira AR, Fernandes JO. Determination of bisphenol A and bisphenol B in canned seafood combining QuEChERS extraction with dispersive liquid-liquid microextraction followed by gas chromatography-mass spectrometry. Anal Bioanal Chem. 2012;404:2453–63.

    Article  Google Scholar 

  106. Giussani DA, Camm EJ, Niu Y, Richter HG, Blanco CE, Gottschalk R, et al. Developmental programming of cardiovascular dysfunction by prenatal hypoxia and oxidative stress. PLoS One. 2012;7:e31017.

    Article  Google Scholar 

  107. O’Mahony J, Moloney M, McCormack M, Nicholls IA, Mizaikoff B, Danaher M. Design and implementation of an imprinted material for the extraction of the endocrine disruptor bisphenol A from milk. J Chromatogr B. 2013;931:164–9.

    Article  Google Scholar 

  108. Osman MA, Mahmoud GI, Elgammal MH, Hasan RS. Studying of bisphenol A levels in some canned food, feed and baby bottles in Egyptian markets. Fresen Environ Bull. 2018;27:9374–81.

    Google Scholar 

  109. Fu P, Kawamura K. Ubiquity of bisphenol A in the atmosphere. Environ Pollut. 2010;158:3138–43.

    Article  Google Scholar 

  110. Adeyi AA, Babalola BA. Bisphenol-A (BPA) in Foods commonly consumed in Southwest Nigeria and its Human Health Risk. Sci Rep. 2019;9:17458.

    Article  Google Scholar 

  111. Liao C, Kannan K. High levels of bisphenol A in paper currencies from several countries, and implications for dermal exposure. Environ Sci Technol. 2011;45:6761–8.

    Article  Google Scholar 

  112. Corrales J, Kristofco LA, Steele WB, Yates BS, Breed CS, Williams ES, et al. Global Assessment of Bisphenol A in the Environment: Review and Analysis of Its Occurrence and Bioaccumulation. Dose-response. 2015;13:1559325815598308.

    Article  Google Scholar 

  113. Chen HW, Liang CH, Wu ZM, Chang EE, Lin TF, Chiang PC, et al. Occurrence and assessment of treatment efficiency of nonylphenol, octylphenol and bisphenol-A in drinking water in Taiwan. Sci Total Environ. 2013;449:20–8.

    Article  Google Scholar 

  114. Biles JE, McNeal TP, Begley TH. Determination of bisphenol A Migrating from epoxy can coatings to infant formula liquid concentrates. J Agr Food Chem. 1997;45:4697–700.

    Article  Google Scholar 

  115. Yamamoto T, Yasuhara A. Determination of bisphenol A migrated from polyvinyl chloride hoses by GC/MS. Bunseki Kagaku. 2000;49:443–7.

    Article  Google Scholar 

  116. Cheng YC, Chen HW, Chen WL, Chen CY, Wang GS. Occurrence of nonylphenol and bisphenol A in household water pipes made of different materials. Environ Monit Assess. 2016;188:562.

    Article  Google Scholar 

  117. Santhi VA, Sakai N, Ahmad ED, Mustafa AM. Occurrence of bisphenol A in surface water, drinking water and plasma from Malaysia with exposure assessment from consumption of drinking water. Sci Total Environ. 2012;427-428:332–8.

    Article  Google Scholar 

  118. Bae B, Jeong JH, Lee SJ. The quantification and characterization of endocrine disruptor bisphenol-A leaching from epoxy resin. Water Sci Technol. 2002;46:381–7.

    Article  Google Scholar 

  119. Marques dos Santos M, Li C, Jia S, Thomas M, Gallard H, Croue JP, et al. Formation of halogenated forms of bisphenol A (BPA) in water: Resolving isomers with ion mobility – mass spectrometry and the role of halogenation position in cellular toxicity. J Hazard Mater. 2024;465:133229.

    Article  Google Scholar 

  120. Kosaka K, Hayashida T, Terasaki M, Asami M, Yamada T, Itoh M, et al. Elution of bisphenol A and its chlorination by-products from lined pipes in water supply process. Water Supply. 2012;12:791–8.

    Article  Google Scholar 

  121. Li X, Ying GG, Su HC, Yang XB, Wang L. Simultaneous determination and assessment of 4-nonylphenol, bisphenol A and triclosan in tap water, bottled water and baby bottles. Environ Int. 2010;36:557–62.

    Article  Google Scholar 

  122. Benotti MJ, Trenholm RA, Vanderford BJ, Holady JC, Stanford BD, Snyder SA. Pharmaceuticals and endocrine disrupting compounds in U.S. drinking water. Environ Sci Technol. 2009;43:597–603.

    Article  Google Scholar 

  123. Fan Z, Hu J, An W, Yang M. Detection and occurrence of chlorinated byproducts of bisphenol a, nonylphenol, and estrogens in drinking water of china: comparison to the parent compounds. Environ Sci Technol. 2013;47:10841–50.

    Article  Google Scholar 

  124. Kleywegt S, Pileggi V, Yang P, Hao C, Zhao X, Rocks C, et al. Pharmaceuticals, hormones and bisphenol A in untreated source and finished drinking water in Ontario, Canada — Occurrence and treatment efficiency. Sci Total Environ. 2011;409:1481–8.

    Article  Google Scholar 

  125. Kuch HM, Ballschmiter K. Determination of endocrine-disrupting phenolic compounds and estrogens in surface and drinking water by HRGC−(NCI)−MS in the picogram per liter range. Environ Sci Technol. 2001;35:3201–6.

    Article  Google Scholar 

  126. Dupuis A, Migeot V, Cariot A, Albouy-Llaty M, Legube B, Rabouan S. Quantification of bisphenol A, 353-nonylphenol and their chlorinated derivatives in drinking water treatment plants. Environ Sci Pollut R. 2012;19:4193–205.

    Article  Google Scholar 

  127. Gallart-Ayala H, Moyano E, Galceran MT. On-line solid phase extraction fast liquid chromatography–tandem mass spectrometry for the analysis of bisphenol A and its chlorinated derivatives in water samples. J Chromatogr A. 2010;1217:3511–8.

    Article  Google Scholar 

  128. Deborde M, von Gunten U. Reactions of chlorine with inorganic and organic compounds during water treatment—Kinetics and mechanisms: A critical review. Water Res. 2008;42:13–51.

    Article  Google Scholar 

  129. Stackelberg PE, Gibs J, Furlong ET, Meyer MT, Zaugg SD, Lippincott RL. Efficiency of conventional drinking-water-treatment processes in removal of pharmaceuticals and other organic compounds. Sci Total Environ. 2007;377:255–72.

    Article  Google Scholar 

  130. Colin A, Bach C, Rosin C, Munoz JF, Dauchy X. Is drinking water a major route of human exposure to alkylphenol and bisphenol contaminants in France? Arch Environ Contam Toxicol. 2014;66:86–99.

    Article  Google Scholar 

  131. Toyo’oka T, Oshige Y. Determination of alkylphenols in mineral water contained in PET Bottles by liquid chromatography with coulometric detection. Anal Sci. 2000;16:1071–6.

    Article  Google Scholar 

  132. Amiridou D, Voutsa D. Alkylphenols and phthalates in bottled waters. J Hazard Mater. 2011;185:281–6.

    Article  Google Scholar 

  133. Stachel B, Ehrhorn U, Heemken O-P, Lepom P, Reincke H, Sawal G, et al. Xenoestrogens in the River Elbe and its tributaries. Environ Pollut. 2003;124:497–507.

    Article  Google Scholar 

  134. Belfroid A, van Velzen M, van der Horst B, Vethaak D. Occurrence of bisphenol A in surface water and uptake in fish: evaluation of field measurements. Chemosphere. 2002;49:97–103.

    Article  Google Scholar 

  135. Dorn PB, Chou C-S, Gentempo JJ. Degradation of bisphenol A in natural waters. Chemosphere. 1987;16:1501–7.

    Article  Google Scholar 

  136. Wan Y, Xia W, Yang S, Pan X, He Z, Kannan K. Spatial distribution of bisphenol S in surface water and human serum from Yangtze River watershed, China: Implications for exposure through drinking water. Chemosphere. 2018;199:595–602.

    Article  Google Scholar 

  137. Lee CC, Jiang LY, Kuo YL, Hsieh CY, Chen CS, Tien CJ. The potential role of water quality parameters on occurrence of nonylphenol and bisphenol A and identification of their discharge sources in the river ecosystems. Chemosphere. 2013;91:904–11.

    Article  Google Scholar 

  138. Jin X, Jiang G, Huang G, Liu J, Zhou Q. Determination of 4-tert-octylphenol, 4-nonylphenol and bisphenol A in surface waters from the Haihe River in Tianjin by gas chromatography–mass spectrometry with selected ion monitoring. Chemosphere. 2004;56:1113–9.

    Article  Google Scholar 

  139. Shehab ZN, Jamil NR, Aris AZ. Occurrence, environmental implications and risk assessment of Bisphenol A in association with colloidal particles in an urban tropical river in Malaysia. Sci Rep. 2020;10:20360.

    Article  Google Scholar 

  140. Ishihara K, Nakajima N. Improvement of marine environmental pollution using eco-system: decomposition and recovery of endocrine disrupting chemicals by marine phyto- and zooplanktons. J Mol Catal B-Enzym. 2003;23:419–24.

    Article  Google Scholar 

  141. Van Zijl MC, Aneck-Hahn NH, Swart P, Hayward S, Genthe B, De Jager C. Estrogenic activity, chemical levels and health risk assessment of municipal distribution point water from Pretoria and Cape Town, South Africa. Chemosphere. 2017;186:305–13.

    Google Scholar 

  142. Kamaludin R, Rasdi Z, Othman MHD, Abdul Kadir SHS, Mohd Nor NS, Khan J, et al. Visible-light active photocatalytic dual layer hollow fiber (DLHF) membrane and its potential in mitigating the detrimental effects of bisphenol A in water. Membranes. 2020;10:32.

    Article  Google Scholar 

  143. Mohan H, Lim JM, Lee SW, Cho M, Park YJ, Seralathan KK, et al. Enhanced removal of bisphenol A from contaminated soil by coupling Bacillus subtilis HV-3 with electrochemical system. Chemosphere. 2020;249:126083.

    Article  Google Scholar 

  144. Zhang W, Yin K, Chen L. Bacteria-mediated bisphenol A degradation. Appl Microbiol Biot. 2013;97:5681–9.

    Article  Google Scholar 

  145. Ohore OE, Zhang S. Endocrine disrupting effects of bisphenol A exposure and recent advances on its removal by water treatment systems. A review. Sci Afr. 2019;5:e00135.

    Google Scholar 

  146. Kristanti RA, Ningsih F, Yati I, Kasongo J, Mtui E, Rachana K. Role of fungi in biodegradation of bisphenol A: A review. Trop Aquat Soil Pollut. 2023;3:131–43.

    Article  Google Scholar 

  147. Das R, Yao P, Yin H, Liang Z, Li G, An T. BPA degradation using biogenic manganese oxides produced by an engineered Escherichia coli with a non-blue laccase from Bacillus sp. GZB. Chemosphere. 2023;326:138407.

    Article  Google Scholar 

  148. Xing H, Zhang J, Tan L, Li R, Chen J, Lan J, et al. Degradation of bisphenol A enhanced by dialysis membrane enclosed laccase catalysis. Environ Technol Inno. 2022;27:102435.

    Article  Google Scholar 

  149. Hirano T, Honda Y, Watanabe T, Kuwahara M. Degradation of bisphenol A by the lignin-degrading enzyme, manganese peroxidase, produced by the white-rot basidiomycete, Pleurotus ostreatus. Biosci Biotech Bioch. 2000;64:1958–62.

    Article  Google Scholar 

  150. Kimura M, Michizoe J, Oakazaki SY, Furusaki S, Goto M, Tanaka H, et al. Activation of lignin peroxidase in organic media by reversed micelles. Biotechnol Bioeng. 2004;88:495–501.

    Article  Google Scholar 

  151. Tsutsumi Y, Haneda T, Nishida T. Removal of estrogenic activities of bisphenol A and nonylphenol by oxidative enzymes from lignin-degrading basidiomycetes. Chemosphere. 2001;42:271–6.

    Article  Google Scholar 

  152. Sakuyama H, Endo Y, Fujimoto K, Hatana Y. Oxidative degradation of alkylphenols by horseradish peroxidase. J Biosci Bioeng. 2003;96:227–31.

    Article  Google Scholar 

  153. Sakurai A, Toyoda S, Sakakibara M. Removal of bisphenol A by polymerization and precipitation method using Coprinus cinereus peroxidase. Biotechnol Lett. 2001;23:995–8.

    Article  Google Scholar 

  154. Karim Z, Husain Q. Application of fly ash adsorbed peroxidase for the removal of bisphenol A in batch process and continuous reactor: Assessment of genotoxicity of its product. Food Chem Toxicol. 2010;48:3385–90.

    Article  Google Scholar 

  155. Karim Z, Husain Q. Guaiacol-mediated oxidative degradation and polymerization of bisphenol A catalyzed by bitter gourd (Momordica charantia) peroxidase. J Mol Catal B-Enzym. 2009;59:185–9.

    Article  Google Scholar 

  156. Li HM, Nicell JA. Biocatalytic oxidation of bisphenol A in a reverse micelle system using horseradish peroxidase. Bioresource Technol. 2008;99:4428–37.

    Article  Google Scholar 

  157. Xu J, Tang T, Zhang K, Ai S, Du H. Electroenzymatic catalyzed oxidation of bisphenol-A using HRP immobilized on magnetic silk fibroin nanoparticles. Process Biochem. 2011;46:1160–5.

    Article  Google Scholar 

  158. Watanabe C, Kashiwada A, Matsuda K, Yamada K. Soybean peroxidase-catalyzed treatment and removal of BPA and bisphenol derivatives from aqueous solutions. Environ Prog Sustain. 2011;30:81–91.

    Article  Google Scholar 

  159. Daassi D, Prieto A, Zouari-Mechichi H, Martinez MJ, Nasri M, Mechichi T. Degradation of bisphenol A by different fungal laccases and identification of its degradation products. Int Biodeter Biodegr. 2016;110:181–8.

    Article  Google Scholar 

  160. Zdarta J, Antecka K, Frankowski R, Zgoła-Grześkowiak A, Ehrlich H, Jesionowski T. The effect of operational parameters on the biodegradation of bisphenols by Trametes versicolor laccase immobilized on Hippospongia communis spongin scaffolds. Sci Total Environ. 2018;615:784–95.

    Article  Google Scholar 

  161. Rampinelli JR, Melo MP, Arbigaus A, Silveira M, Wagner TM, Gern RMM, et al. Production of Pleurotus sajor-caju crude enzyme broth and its applicability for the removal of bisphenol A. An Acad Bras Cienc. 2021;93:e20191153.

    Article  Google Scholar 

  162. Mahesh N, Shyamalagowri S, Nithya TG, Aravind J, Govarthanan M, Kamaraj M. Trends and thresholds on bacterial degradation of bisphenol-A endocrine disruptor — a concise review. Environ Monit Assess. 2022;194:886.

    Article  Google Scholar 

  163. Kang JH, Katayama Y, Kondo F. Biodegradation or metabolism of bisphenol A: From microorganisms to mammals. Toxicology. 2006;217:81–90.

    Article  Google Scholar 

  164. Melcer H, Klecka G. Treatment of wastewaters containing bisphenol A: state of the science review. Water environment research : a research publication of the Water Environment Federation. 2011;83:650–66.

    Article  Google Scholar 

  165. Lobos JH, Leib TK, Su TM. Biodegradation of bisphenol A and other bisphenols by a gram-negative aerobic bacterium. Appl Environ Microbiol. 1992;58:1823–31.

    Article  Google Scholar 

  166. Matsumura Y, Hosokawa C, Sasaki-Mori M, Akahira A, Fukunaga K, Ikeuchi T, et al. Isolation and characterization of novel bisphenol-A-degrading bacteria from soils. Biocontrol Sci. 2009;14:161–9.

    Article  Google Scholar 

  167. Sasaki M, Akahira A, Oshiman K, Tsuchido T, Matsumura Y. Purification of cytochrome P450 and ferredoxin, involved in bisphenol A degradation, from Sphingomonas sp. strain AO1. Appl Environ Microbiol. 2005;71:8024–30.

    Article  Google Scholar 

  168. Kang JH, Kondo F. Bisphenol A Degradation by Bacteria Isolated from River Water. Arch Environ Con Tox. 2002;43:0265–9.

    Article  Google Scholar 

  169. Masuda M, Yamasaki Y, Ueno S, Inoue A. Isolation of bisphenol A-tolerant/degrading Pseudomonas monteilii strain N-502. Extremophiles. 2007;11:355–62.

    Article  Google Scholar 

  170. Yamanaka H, Moriyoshi K, Ohmoto T, Ohe T, Sakai K. Degradation of bisphenol A by Bacillus pumilus isolated from kimchi, a traditionally fermented food. Appl Biochem Biotech. 2007;136:39–51.

    Article  Google Scholar 

  171. Suyamud B, Thiravetyan P, Gadd GM, Panyapinyopol B, Inthorn D. Bisphenol A removal from a plastic industry wastewater by Dracaena sanderiana endophytic bacteria and Bacillus cereus NI. Int J Phytoremediation. 2020;22:167–75.

    Article  Google Scholar 

  172. Eltoukhy A, Jia Y, Nahurira R, Abo-Kadoum MA, Khokhar I, Wang J, et al. Biodegradation of endocrine disruptor Bisphenol A by Pseudomonas putida strain YC-AE1 isolated from polluted soil, Guangdong, China. BMC Microbiol. 2020;20:11.

    Article  Google Scholar 

  173. Kang JH, Ri N, Kondo F. Streptomyces sp. strain isolated from river water has high bisphenol A degradability. Lett Appl Microbiol. 2004;39:178–80.

    Article  Google Scholar 

  174. Shin EH, Choi HT, Song HG. Biodegradation of endocrine-disrupting bisphenol A by white rot fungus Irpex lacteus. J Microbiol Biotechnol. 2007;17:1147–51.

    Google Scholar 

  175. Yamanaka H, Moriyoshi K, Ohmoto T, Ohe T, Sakai K. Efficient microbial degradation of bisphenol A in the presence of activated carbon. J Biosci Bioeng. 2008;105:157–60.

    Article  Google Scholar 

  176. Toyama T, Sato Y, Inoue D, Sei K, Chang YC, Kikuchi S, et al. Biodegradation of bisphenol A and bisphenol F in the rhizosphere sediment of Phragmites australis. J Biosci Bioeng. 2009;108:147–50.

    Article  Google Scholar 

  177. Vijayalakshmi V, Senthilkumar P, Mophin-Kani K, Sivamani S, Sivarajasekar N, Vasantharaj S. Bio-degradation of Bisphenol A by Pseudomonas aeruginosa PAb1 isolated from effluent of thermal paper industry: Kinetic modeling and process optimization. J Radiat Res Appl Sci. 2018;11:56–65.

    Google Scholar 

  178. Kyrila G, Katsoulas A, Schoretsaniti V, Rigopoulos A, Rizou E, Doulgeridou S, et al. Bisphenol A removal and degradation pathways in microorganisms with probiotic properties. J Hazard Mater. 2021;413:125363.

    Article  Google Scholar 

  179. Noszczynska M, Chodor M, Jalowiecki L, Piotrowska-Seget Z. A comprehensive study on bisphenol A degradation by newly isolated strains Acinetobacter sp K1MN and Pseudomonas sp BG12. Biodegradation. 2021;32:1–15.

    Article  Google Scholar 

  180. Endo Y, Kimura N, Ikeda I, Fujimoto K, Kimoto H. Adsorption of bisphenol A by lactic acid bacteria, Lactococcus, strains. Appl Microbiol Biot. 2007;74:202–7.

    Article  Google Scholar 

  181. Bala S, Garg D, Thirumalesh BV, Sharma M, Sridhar K, Inbaraj BS, et al. Recent strategies for bioremediation of emerging pollutants: A review for a green and sustainable environment. Toxics. 2022;10:484.

    Article  Google Scholar 

  182. Sharma I. Ch 12: Bioremediation techniques for polluted environment: concept, advantages, limitations, and prospects. In: Mario Alfonso MT, Hugo SN, Agnieszka S, editors. Trace Metals in the Environment - New Approaches and Recent Advances. Rijeka: IntechOpen; 2021.

    Google Scholar 

  183. Han Y, Dai H, Rong X, Jiang H, Xue Y. Research progress of methods for degradation of bisphenol A. Molecules. 2023;28:8028.

    Article  Google Scholar 

  184. Nath A, Biswas S, Pal A. A comprehensive review on BPA degradation by heterogeneous Fenton-like processes. Water Sci Technol. 2022;86:714–45.

    Article  Google Scholar 

  185. Bartolomeu M, Neves MGPMS, Faustino MAF, Almeida A. Wastewater chemical contaminants: remediation by advanced oxidation processes. Photoch Photobio Sci. 2018;17:1573–98.

    Article  Google Scholar 

  186. Nipane D, Thakare SR, Khati NT. Synthesis of novel ZnO having cauliflower morphology for photocatalytic degradation study. J Catal. 2013;2013:940345.

    Google Scholar 

  187. O’Shea KE, Dionysiou DD. Advanced oxidation processes for water treatment. J Phys Chem Lett. 2012;3:2112–3.

    Article  Google Scholar 

  188. Rajeshwar K. Solar energy conversion and environmental remediation using inorganic semiconductor-liquid interfaces: The road traveled and the way forward. J Phys Chem Lett. 2011;2:1301–9.

    Article  Google Scholar 

  189. Ohko Y, Ando I, Niwa C, Tatsuma T, Yamamura T, Nakashima T, et al. Degradation of bisphenol A in water by TiO2 photocatalyst. Environ Sci Technol. 2001;35:2365–8.

    Article  Google Scholar 

  190. Watanabe N, Horikoshi S, Kawabe H, Sugie Y, Zhao J, Hidaka H. Photodegradation mechanism for bisphenol A at the TiO2/H2O interfaces. Chemosphere. 2003;52:851–9.

    Article  Google Scholar 

  191. Bechambi O, Jlaiel L, Najjar W, Sayadi S. Photocatalytic degradation of bisphenol A in the presence of Ce–ZnO: Evolution of kinetics, toxicity and photodegradation mechanism. Mater Chem Phys. 2016;173:95–105.

    Article  Google Scholar 

  192. Hunge YM, Yadav AA, Khan S, Takagi K, Suzuki N, Teshima K, et al. Photocatalytic degradation of bisphenol A using titanium dioxide@nanodiamond composites under UV light illumination. J Colloid Interf Sci. 2021;582:1058–66.

    Article  Google Scholar 

  193. Ashiegbu DC, Moloto N, Potgieter H. Improved photocatalytic activity of ZnO-[10%]BiOI and ZnO-[10%]WO3 heterostructure in the destruction of 2-chlorobiphenyl. Environ Sci Adv. 2023;2:325–38.

    Article  Google Scholar 

  194. Xu L, Liu NP, An HL, Ju WT, Liu B, Wang XF, et al. Preparation of Ag3PO4/CoWO4 S-scheme heterojunction and study on sonocatalytic degradation of tetracycline. Ultrason Sonochem. 2022;89:106147.

    Article  Google Scholar 

  195. Lee J, Kim J, Choi W. TiO2 photocatalysis for the redox conversion of aquatic pollutants. In: Tratnyek PG, Grundl TJ, Haderlein SB editors. Aquatic Redox Chemistry. Washington, DC: ACS Publications; 2011, p. 199–222.

    Chapter  Google Scholar 

  196. Bhatkhande DS, Pangarkar VG, Beenackers AACM. Photocatalytic degradation for environmental applications – a review. J Chem Technol Biotechnol. 2002;77:102–16.

    Article  Google Scholar 

  197. Tsai WT, Lee MK, Su TY, Chang YM. Photodegradation of bisphenol-A in a batch TiO2 suspension reactor. J Hazard Mater. 2009;168:269–75.

    Article  Google Scholar 

  198. da Silva JC, Reis Teodoro JA, Afonso RJ, Aquino SF, Augusti R. Photodegradation of bisphenol A in aqueous medium: monitoring and identification of by-products by liquid chromatography coupled to high-resolution mass spectrometry. Rapid Commun Mass Spectrom. 2014;28:987–94.

    Article  Google Scholar 

  199. Abo R, Kummer NA, Merkel BJ. Optimized photodegradation of Bisphenol A in water using ZnO, TiO2 and SnO2 photocatalysts under UV radiation as a decontamination procedure. Drink Water Eng Sci. 2016;9:27–35.

    Article  Google Scholar 

  200. Katsumata H, Taniguchi M, Kaneco S, Suzuki T. Photocatalytic degradation of bisphenol A by Ag3PO4 under visible light. Catal Commun. 2013;34:30–4.

    Article  Google Scholar 

  201. Mu C, Zhang Y, Cui W, Liang Y, Zhu Y. Removal of bisphenol A over a separation free 3D Ag3PO4-graphene hydrogel via an adsorption-photocatalysis synergy. Applied Catalysis B: Environmental. 2017;212:41–9.

    Article  Google Scholar 

  202. Garg A, Singhania T, Singh A, Sharma S, Rani S, Neogy A, et al. Photocatalytic degradation of bisphenol-A using N, Co codoped TiO2 catalyst under solar light. Sci Rep. 2019;9:765.

    Article  Google Scholar 

  203. Yu F, Bai X, Yang C, Xu L, Ma J. Reduced graphene oxide–P25 nanocomposites as efficient photocatalysts for degradation of bisphenol A in water. Catalysts. 2019;9:607.

    Article  Google Scholar 

  204. Cordeiro DS, Cassio FL, Ciccotti L, Hewer TLR, Corio P, Freire RS. Photocatalytic activity of Pr-modified TiO2 for the degradation of bisphenol A. Sn Appl Sci. 2021;3:258.

    Article  Google Scholar 

  205. Reddy PVL, Kim KH, Kavitha B, Kumar V, Raza N, Kalagara S. Photocatalytic degradation of bisphenol A in aqueous media: A review. J Environ Manage. 2018;213:189–205.

    Article  Google Scholar 

  206. Zhang X, Wang DK, Lopez DRS, Diniz da Costa JC. Fabrication of nanostructured TiO2 hollow fiber photocatalytic membrane and application for wastewater treatment. Chem Eng J. 2014;236:314–22.

    Article  Google Scholar 

  207. Simone S, Galiano F, Faccini M, Boerrigter ME, Chaumette C, Drioli E, et al. Preparation and characterization of polymeric-hybrid PES/TiO2 hollow fiber membranes for potential applications in water treatment. Fibers. 2017;5:14.

    Article  Google Scholar 

  208. Hu W, Yin J, Deng B, Hu Z. Application of nano TiO2 modified hollow fiber membranes in algal membrane bioreactors for high-density algae cultivation and wastewater polishing. Bioresource Technol. 2015;193:135–41.

    Article  Google Scholar 

  209. Yeo MK, Kang M. Photodecomposition of bisphenol A on nanometer-sized TiO2 thin film and the associated biological toxicity to zebrafish (Danio rerio) during and after photocatalysis. Water Res. 2006;40:1906–14.

    Article  Google Scholar 

  210. Daskalaki VM, Fulgione I, Frontistis Z, Rizzo L, Mantzavinos D. Solar light-induced photoelectrocatalytic degradation of bisphenol-A on TiO2/ITO film anode and BDD cathode. Catal Today. 2013;209:74–8.

    Article  Google Scholar 

  211. Frontistis Z, Daskalaki VM, Katsaounis A, Poulios I, Mantzavinos D. Electrochemical enhancement of solar photocatalysis: Degradation of endocrine disruptor bisphenol-A on Ti/TiO2 films. Water Res. 2011;45:2996–3004.

    Article  Google Scholar 

  212. Rao KVS, Subrahmanyam M, Boule P. Immobilized TiO2 photocatalyst during long-term use: decrease of its activity. Appl Catal B-Environ. 2004;49:239–49.

    Article  Google Scholar 

  213. Dzinun H, Othman MHD, Ismail AF, Puteh MH, Rahman MA, Jaafar J. Photocatalytic degradation of nonylphenol by immobilized TiO2 in dual layer hollow fibre membranes. Chem Eng J. 2015;269:255–61.

    Article  Google Scholar 

  214. Katoch A, Kim JH, Kim SS. TiO2/ZnO inner/outer double-layer hollow fibers for improved detection of reducing gases. ACS Appl Mater Inter. 2014;6:21494–9.

    Article  Google Scholar 

  215. Fu X, Yang R, Zhou G, Chen X, Liu Y, Chi J, et al. New progress in photocatalytic degradation of bisphenol A as representative endocrine disrupting chemicals. Curr Opin Green Sust. 2022;35:100629.

    Article  Google Scholar 

  216. Mohamed Noor SH, Othman MHD, Khongnakorn W, Sinsamphanh O, Abdullah H, Puteh MH, et al. Bisphenol A removal using visible light driven Cu2O/PVDF photocatalytic dual layer hollow fiber membrane. Membranes. 2022;12:208.

    Article  Google Scholar 

  217. Sukaini K, Mohamed Noor SH, Mansur S, Jaffar FH, Kamaludin R, Othman MHD, et al. Photodegradation of bisphenol A in water via round-the-clock visible light driven dual layer hollow fiber membrane. Catalysts. 2023;13:816.

    Article  Google Scholar 

  218. Zakria HS, Othman MHD, Kamaludin R, Jilani A, Omar MF, Ayub M, et al. Removal of bisphenol A from synthetic and treated sewage wastewater using magnetron sputtered CuxO/PVDF thin film photocatalytic hollow fiber membrane. J Water Process Eng. 2023;51:103425.

    Article  Google Scholar 

  219. Husain Q, Qayyum S. Biological and enzymatic treatment of bisphenol A and other endocrine disrupting compounds: a review. Crit Rev Biotechnol. 2013;33:260–92.

    Article  Google Scholar 

  220. Zhang J, Wu H, Shi L, Wu Z, Zhang S, Wang S, et al. Photocatalysis coupling with membrane technology for sustainable and continuous purification of wastewater. Sep Purif Technol. 2024;329:125225.

    Article  Google Scholar 

Download references

Acknowledgements

The authors gratefully acknowledge the financial support granted by the Ministry of Education Malaysia under the Higher Institution Centre of Excellence Scheme (Project Number: R.J090301.7809.4J430). The authors also would like to acknowledge Universiti Teknologi Malaysia for the financial support under High Impact Research Grant (Q.J130000.2409.08G34) and UTM ICC Prototype Research Grant (R.J130000.7709.4J483). Special thanks are also given to Universiti Teknologi MARA for the research grant namely BESTARI grant scheme (Project Number: 600-IRMI/DANA 5/3/BESTARI (P) (005/2018)), and the LESTARI grant scheme (Project Number: 600-RMI/DANA 5/3/LESTARI (60/2015)). The authors also would like to express the appreciation to JICA Technical Cooperation Project for ASEAN University Network/Southeast Asia Engineering Education Development Network (JICA Project for AUN/SEED-Net) for the research funding via Alumni Support Program for Research (Project number: UTM ASP-R 2101/R.J130000.7309.4B651). The authors would also like to thank the Research Management Centre, Universiti Teknologi Malaysia for the technical support provided.

Author information

Authors and Affiliations

Authors

Contributions

RK and ZR wrote the main manuscript text. MHDO and SK are responsible for editing the article. All authors reviewed the manuscript.

Corresponding authors

Correspondence to Mohd Hafiz Dzarfan Othman or Siti Hamimah Sheikh Abdul Kadir.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

We confirm that the manuscript has been read and approved by all named authors. We further confirm that the order of authors listed in the manuscript has been approved by all of us. We understand that the Corresponding Author is the sole contact for the Editorial process. He is responsible for communicating with the other authors about progress, submissions of revisions and final approval of proofs.

Competing interests

The authors declare they have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kamaludin, R., Rasdi, Z., Othman, M.H.D. et al. Human metabolic effects of BPA and the application of a hybrid photocatalytic membrane for BPA contaminated water. Sustain Environ Res 34, 15 (2024). https://doi.org/10.1186/s42834-024-00219-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s42834-024-00219-9

Keywords