Skip to main content

The involvement of human organic anion transporting polypeptides (OATPs) in drug-herb/food interactions

Abstract

Organic anion transporting polypeptides (OATPs) are important transporter proteins that are expressed at the plasma membrane of cells, where they mediate the influx of endogenous and exogenous substances including hormones, natural compounds and many clinically important drugs. OATP1A2, OATP2B1, OATP1B1 and OATP1B3 are the most important OATP isoforms and influence the pharmacokinetic performance of drugs. These OATPs are highly expressed in the kidney, intestine and liver, where they determine the distribution of drugs to these tissues. Herbal medicines are increasingly popular for their potential health benefits. Humans are also exposed to many natural compounds in fruits, vegetables and other food sources. In consequence, the consumption of herbal medicines or food sources together with a range of important drugs can result in drug-herb/food interactions via competing specific OATPs. Such interactions may lead to adverse clinical outcomes and unexpected toxicities of drug therapies. This review summarises the drug-herb/food interactions of drugs and chemicals that are present in herbal medicines and/or food in relation to human OATPs. This information can contribute to improving clinical outcomes and avoiding unexpected toxicities of drug therapies in patients.

Background

Solute carrier transporters (SLCs) are transmembrane proteins that mediate the cellular uptake of endogenous and exogenous substances. The SLC superfamily consists of over 300 isoforms, many of which are localised at the plasma membrane of cells [1]. Organic anion transporter polypeptides (OATPs), encoded by the SLCO genes, are one of the most important SLC subfamilies that contribute to drug and endobiotic uptake into cells [2]. There are 11 known OATP isoforms, which are widely distributed across tissues such as the kidney, liver, intestine and brain [3].

Most OATP substrates are large hydrophobic anions with molecular weight > 350 Da. Some endogenous chemicals of physiological importance are substrates including steroids, bile acids, prostaglandins and thyroid hormones. OATPs also have important roles in the absorption, distribution and elimination of drugs such as anti-cancer agents, antibiotics, antivirals and statins [4]. Therefore, OATP function strongly influences the pharmacokinetics and pharmacodynamics of drugs and, in turn, their therapeutic outcomes. The dysfunction of OATPs may lead to suboptimal drug treatment and/or unexpected toxicities.

The transport mechanism of OATPs remains largely unclear. However, studies to date have suggested that substrate transport by OATPs is ATP and sodium independent. OATPs appear to act as electroneutral exchangers that couple substrate uptake to the cellular efflux of a counterion, such as bicarbonate, glutathione, conjugated glutathione or glutamate [4, 5].

General information about the major OATP isoforms

OATP1A2, OATP1B1, OATP1B3 and OATP2B1 have been well studied for their important roles in the kidney and liver. These isoforms have been extensively investigated in relation to a wide range of drugs.

OATP1A2 is the first identified and best characterized human OATP isoform. This transporter is located at the apical membrane of the distal nephron, where it mediates urinary drug secretion and reabsorption. It is also expressed in hepatic cholangiocytes, where it modulates the secretion of molecules into the bile duct. It has also been discovered in the brain endothelial cells that comprise the blood–brain barrier, at the apical membrane of intestinal enterocytes and at the apical membrane of retinal epithelial cells [6,7,8]. OATP1A2 is known to transport a broad range of clinically important drugs like methotrexate, imatinib and fexofenadine. Overall, it plays an important role in the absorption, distribution and elimination of various drugs and substances [3, 5].

OATP1B1 is selectively expressed at the basolateral membrane of hepatocytes [5]. It can transport many widely used drugs such as statins and anti-viral agents, as well as endogenous substances like estrone-3-sulfate (ES) and bilirubin.

Like OATP1B1, OATP1B3 is also specifically located at the basolateral membrane of hepatocytes, but it is primarily expressed around the central vein [3]. The substrate specificities of OATP1B1 and OATP1B3 somewhat overlap [4]. However, OATP1B3 has relatively less impact on drug pharmacokinetics than OATP1B1 considering there are relatively less reports on the influence of OATP1B3 on drug performance [3].

OATP2B1 is highly expressed at the basolateral membrane of hepatocytes where it modulates drug influx into the liver. It is also located at the apical membrane of the enterocytes, where it influences drug absorption. In the kidney, it mediates the secretion of molecules into and/or reabsorption from urine. It has also been identified in the placental syncytiotrophoblasts, skeletal muscle and endothelial cells of the blood–brain barrier [6, 9,10,11,12]. Many clinically important drugs such as atorvastatin, pravastatin and glibenclamide, as well as the endogenous compounds like ES, bile acids, pregnenolone sulphate and prostaglandins, have been found to be OATP2B1 substrates [5, 13].

OATP regulations in health and disease

OATP transporters can be regulated at the transcriptional and post- translational levels by a number of signalling pathways and kinases [1, 2]. For instance, kinases like protein kinase C (PKC) modulate the trafficking of OATP1A2, OATP1B1, OATP1B3 and OATP2B1 between the plasma membrane and intracellular compartments [13,14,15]. Similarly, casein kinase-2 also influences the expression and function of OATP1A2 by altering its subcellular trafficking to the plasma membrane [1].

Literature reports have shown that the expression and/or function of OATPs is altered in cancer and other disease states. OATPs may be novel tumour biomarkers and may influence the progression of hormone-dependent cancers [16]. For example, OATP1A2 protein expression is almost tenfold higher in breast cancer patients than that in normal subjects [17, 18]. This could be clinically significant, because OATP1A2 increases the uptake of ES, the major precursor of biologically active estrogen, which may enhance tumour cell proliferation and survival. OATP1B1 is normally liver specific but is over-expressed in tumours of the colon, lung, breast, prostate, ovary and pancreas [19,20,21,22]. Similarly, OATP1B3 is found to be overexpressed in breast, prostate and colorectal cancers [21, 23]. An important OATP1B3 substrate is testosterone, which has been shown to decrease the survival of patients with androgen dependent prostate cancer [24]. OATP2B1 expression is increased in breast, thyroid, glioma, prostate and testicular cancers [6, 24, 25]. This may enhance the uptake of ES into estrogen receptor (ER)-positive breast tumour cells and that of dehydroepiandrosterone into prostate cancer cells [23]. Again, OATP2B1 could promote tumorigenesis by enhancing cancer cell proliferation due to these effects [26]. Additionally, the expression of the orphan OATP transporter OATP5A1 is shown to be elevated in metastatic cancers [16]. Therefore, the altered expression of OATPs in the cancers mentioned above is closely related to cancer progression.

Abnormal OATP expression and function has also been reported in inflammatory conditions such as fibrosis, inflammatory bowel disease, cholestasis and advanced liver diseases, greatly contributing to disease progression [27]. For instance, OATP1A2 mRNA was increased in the placentas of pregnant patients with intrahepatic cholestasis [28]. OATP2B1 expression was upregulated in the patients with Crohn’s disease and ulcerative colitis [29]. Inflammatory bowel disease has been found to be associated with the increased level of OATP2B1 and OATP4A1 in the ileum and colon [29]; OATP4A1 is also reported to be upregulated in polycystic ovarian syndrome [30]. In contrast, the expression of OATP2B1 was decreased in the placentas of patients with bacterial chorioamnionitis to ~ 50% of the age-matched controls [31]. And OATP1B1 was also shown to be decreased in the livers of patients with primary sclerosing cholangitis [32] or severe viral hepatitis [27, 33]. Previous findings also suggested that cytokines such as TNF-α, IL-6, IL-1β, IFN-γ and oncostatin M, may regulate the expression of OATPs in cells [34,35,36].

Post-translational glycosylation of OATP1B1, 1B3 and 2B1 was decreased in the livers of patients with severe non-alcoholic fatty liver disease [37]. This might impair the subcellular trafficking of OATPs. Proteomics analysis showed that OATP2B1 was found to be increased in hepatic cirrhosis induced by hepatitis C infection, although the expression of OATP1B1 and OATP1B3 was unchanged [38]. Cholestasis also decreased the hepatic expression of OATP1A2, OATP1B1 and OATP1B3 at mRNA level [39,40,41]. Together, these findings provide insights into how severe liver diseases complicate the effectiveness of drug therapies by modulating OATP function and expression and thus, cellular influx of their drug substrates.

The influence of OATP genetic polymorphisms on drug performance

Studies have shown that several OATP genetic variants are possibly associated with the pathogenesis of human diseases. For example, mutations in the SLCO1B1 and SLCO1B3 genes have been implicated in Rotor syndrome [42]. When both transporters are defective, bilirubin is taken up by the liver inefficiently leading to serum accumulation and jaundice [43]. There have also been reports that patients carrying specific OATP1B1 polymorphisms are at an increased risk of severe hyperbilirubinaemia [44,45,46]. And OATP1A2 genotypes were also found to be involved in the neurodegenerative disease progressive supranuclear palsy; while OATP2A1 polymorphisms are indicated to be associated with primary hypertrophic osteoarthropathy [4, 47].

Genetic polymorphisms of OATPs may have great impact on pharmacokinetic performance of drugs. It has been shown that the single nucleotide polymorphisms of OATP1B1 result in the altered disposition of statins [48, 49]. Systemic exposure to the anti-diabetic drug nateglinide [46] and to the HIV protease inhibitor lopinavir is also increased in these individuals [50]. OATP1A2 polymorphisms have also been found to be associated with impaired imatinib clearance [51] and OATP2B1 polymorphisms could impact on fexofenadine pharmacokinetics [52].

Sebastian et al. found that OATP5A1 is atypical as it does not transport classic OATP substrates [53]. Instead, it appears to be a determinant of cell shape, differentiation and motility. Microarray studies have shown that OATP5A1 is widely expressed in the fetal brain, prostate, skeletal muscle and thymus. Like other OATPs, defective OATP5A1 has been associated with human pathologies. Its genetic deletion appears to be related to Mesomelia synostoses syndrome, which is a congenital disease characterised by short limb and malformation [54].

The interactions of herb/diet-derived chemicals with OATPs

Herbal and dietary supplements are currently popular in treating a wide range of health conditions. For example, silymarin, an active ingredient of milk thistle, can be used to treat intoxication caused by ingestion of death cap mushrooms [55]. St.John’s Wort is a herbal supplement that has been applied to treat depression [56]. One of the most common dietary products—green tea—is commonly adopted in weight loss programs; and despite definitive supporting evidence of efficacy, to prevent cancer and cardiovascular diseases [57]. Although complementary therapies and food are relatively safe, evidence to the contrary is increasing. In particular, some herbal preparations or food consumption may elicit clinically significant adverse reactions when co-administered with conventional medicines [58]. This problem is exacerbated by the ready availability of herbal and dietary supplements as well as the fact that people may self-medicate.

OATP1A2, 1B1, 1B3 and 2B1 contribute extensively to the disposition of drugs in humans [4]. The co-administration of drugs or other molecules that compete with specific OATP isoforms has the potential to elicit pharmacokinetic interactions [59]. During the pre-clinical phase of drug development, regulatory authorities require the evaluation of potential interactions involving major OATPs. However, herbal medicines derived from plants, fruits and vegetables widely used in many countries, are not subject to regulatory approval. Precautions are essential when herbal medicines and conventional drugs are co-administered and elicit interactions. Drug-drug interactions involving OATPs have been widely described in literature [60,61,62]; however, drug-herb/food interactions associated with these transporters have not been extensively reviewed so far.

As mentioned, a wide range of herbal and dietary compounds are substrates and inhibitors of OATPs [27, 63,64,65,66,67]. Unexpected adverse effects may occur if a conventional drug with a narrow therapeutic index competes for specific OATPs with herb/food chemicals [27]. Accordingly, a greater appreciation of potential drug-herb/food interactions involving OATPs could contribute to improved efficacy and safety of drug therapies when co-administered with herbal or food supplements. Table 1 summarises the documented interactions of commonly used herbal medicines or food-derived chemicals with human OATP transporters.

Table 1 Potential drug-herb/food interactions involving OATPs

Food such as pomegranate and olives contain antioxidants and other naturally occurring chemicals [64]. Studies have implicated food-derived chemicals interacting with OATP influx transporters [63, 68]. For instance, consumption of fruit juices such as grapefruit, apple and orange juice, can significantly reduce the oral bioavailability of drugs due to the inhibition of intestinal OATPs [69].

In most cases, the interactions of herb/food-derived chemicals with OATPs involve competitive inhibition of substrate transport, or allosteric inhibition due to altered transporter conformation. Other potential mechanisms include altered transporter expression, subcellular localisation or stability, but definitive evidence for such mechanisms is sparse. In general terms, changes in protein expression and/or rates of protein degradation occur over a relatively long timeframe, while functional modulations due to altered trafficking or interference with substrate binding are more rapid [70]. In this review, we discuss the relationship of the intake of both herb- and food-derived chemicals with the impairment of OATP function. The application of such information may prevent adverse effects of conventional drug therapies and improve treatment outcomes.

Drug-herb interactions involving OATPs

A wide range of herbal compounds have been found to modulate the substrate uptake mediated by OATPs (Table 1). For example, kampo products have been used in Japan as traditional herbal medicines in treating inflammatory bowel disease, nausea, diarrhea and gastrointestinal tract disorders for over 1500 years. Kampo products consist of crude extracts from more than 98 sources, such as Bofutsushosan, Rhei Rhizoma, Perillae herba, Scuterallia Radix, Glyryrrhizae Radix, Moutan cortex, Paeoniae Radix, Tribuli Fructus, Saussurea Radix and Curcumae Rhizoma. Japanese physicians have also recommended these products to cancer patients as supplements to ongoing chemotherapy and radiotherapy [80].

Kampo extracts contain a range of bioactive chemicals that have been shown to modulate OATP transport function in cells that over-express these transporters [67, 71, 72, 81]. Thus, complex polyhydroxylated multi-ring systems like the catechins Epigallocatechin gallate (EGCG) and epicatechin gallate (ECG), flavonoids such as scutellarin and baicalin, as well as saponins like glycyrrhizic acid, have been reported to inhibit the transport function of OATP1B1 and 2B1 (Table 1). In contrast, the glycosylated flavonoid rutin increased the substrate transport via OATP1B1 and OATP2B1 in over-expressing HEK 293 and HeLa cells [71, 74]. Because these transporters are widely implicated in the cellular influx of drugs like fexofenadine, glibenclamide, statins and β-adrenergic antagonists, there is a considerable potential for pharmacokinetic interactions with Kampo-derived chemicals.

Baicalin and its aglycone baicalein are flavonoids that are active constituents of Chinese skullcap. The flavonoid scutellarin is enriched in breviscapine. Extracts from these natural sources have been found to impair the transport function of OATPs in the over-expressing cells [67, 72, 76]. Horny Goat Weed is a widely used Chinese medicine in Asian countries. It contains the prenylated flavanol glycoside icariin, which is used to treat osteoporosis and male sexual dysfunction. However, icariin has been found to modulate the transport function of OATP2B1, OATP1B1 and OATP1B3 [27, 63]. Milk thistle contains the flavonolignan silymarin—a polyhydroxylated ring-containing molecule, which potently inhibits estradiol-17β-glucuronide uptake via OATP1B1 and 1B3, as well as ES uptake via OATP2B1 [75]. There are other herbal extracts containing chemicals that can modulate the substrate uptake mediated via OATPs. The triterpenoid saponins astragaloside and ginsenoside Rc found in Radix astragali and Panax ginseng extracts, the naphthodianthrone 2-O-galloyl-hyperin enriched in Pyrola incarnate Fisch as well as the glycosylated flavonoid epimedin C present in Herba epimedii, have all been shown to inhibitors of OATP1B1 with moderate potency [72].

The capacity of molecules to modulate the rate of transporter-mediated substrate influx into cells has been reported but is under-explored. Again, the underpinning molecular mechanisms could be either substrate-dependent or substrate-independent. Chemicals that enhance the translocation of OATPs to the plasma membrane could activate the influx of chemicals in a substrate-independent fashion. Amiodarone (a potent natural compound used in the treatment of cardiac arrhythmias) and rutin have been shown to increase OAT2B1-mediated substrate uptake via increased translocation to the plasma membrane [82, 83].

Molecules that bind directly to specific domains or regions within OATPs could elicit rapid changes in substrate uptake [83]. For example, the steroid progesterone has been shown to activate OATP2B1 function by promoting a conformational change in the substrate-binding site [70]. Similar allosteric interactions have also been indicated to activate CYP monooxygenases and increase the rate of substrate oxidation [84].

Drug-food interactions involving OATPs

Therapeutic complications due to drug-food interactions involving OATPs have also been demonstrated (Table 1). Grapefruit and apple juices contain a range of flavonoids that have been found to impair the transport function of OATP1A2, 1B1, 1B3 and 2B1, some with IC50 and Ki values in a low micromolar range [67, 72, 77, 78]. In vivo studies in patients have suggested that co-administration of grapefruit, orange or apple juices decreased the systemic availability of fexofenadine by 65–75% and celiprolol by more than 80% [85, 86]. Such inhibitory effects of fruit juices are reversible as the removal of fruit juice restores OATP function [69]. Literature also reported that ingestion of fruit juices that contain high concentrations of naringin directly inhibits enteric OATP1A2 and decreases the oral bioavailability of fexofenadine [77, 85]. Accordingly, it has been suggested that the consumption of fruit juices should be avoided within 4 h of drug administration to minimise adverse effects.

Green and black teas are popular beverages that are widely consumed with potential health benefits. Catechins and polyhydroxylated flavonoids such as theaflavin, that are present in teas can impact on OATPs and reduce the systemic exposure to OATP drug substrates like rosuvastatin [87].

Other classes of natural compounds that are potent in modulating OATP activities, are present in vegetables and fruits widely used in health treatments. For instance, licorice contains the saponin glycyrrhizic acid, which has been in managing chronic hepatitis and gastric ulcers. Glycyrrhizic acid was reported to inhibit the cellular uptake of OATP substrates atorvastatin, fluvastatin and rosuvastatin [76]. Literature also indicated that chemically similar molecules like the triterpenoid saponins ursolic acid, oleanolic acid and betulinic acid, are potent OATP inhibitors [64, 68]. Ursolic acid is present in pomegranate with anti-mutagenic and anti-viral properties. It can inhibit the OATP1B1- and OATP1B3-mediated uptake of fluorescein-conjugated methotrexate analogues [88]. Similarly, biochanin A in peanuts and mulberrin in mulberries are flavonoids that have been found to inhibit OATP1B1 and OATP2B1, respectively, in a non-competitive manner [74, 76, 89]. Thus, food-derived chemicals, if ingested in enough amounts, have the potential to influence the safety and efficacy of co-administered drugs.

Clinical significance of drug-herb/food interactions mediated through OATPs

There is limited clinical evidence available regarding drug-food/herb interactions via OATPs [70]. In healthy Chinese volunteers, the flavonoid quercetin present in apples and other fruits, significantly decreased the bioavailability of pravastatin (an OATP1B1 drug substrate), most likely by decreasing intestinal absorption [72]. Another clinical study examined the influence of fruit juices on the disposition of the antihistamine fexofenadine. The oral pharmacokinetics of fexofenadine was assessed in ten healthy individuals received grapefruit, orange or apple juice (1.2 L over 3 h) in a randomized 5-way crossover study. It was found that grapefruit, orange, and apple juices decreased the fexofenadine area under the plasma concentration–time curve (AUC), the maximal plasma drug concentration (Cmax), and the urinary excretion values to only 30–40% of control [85]. Another study found that grapefruit juice decreased the Cmax of acebutolol (an OATP1A2 substrate) by 19%, although the AUC was essentially unchanged from control [81]. Similarly, the administration of orange juice was found to decrease the Cmax and AUC of the β-adrenoceptor antagonist atenolol (a common substrate of OATP1A2 and 2B1) by 49% and 40%, respectively [73]. There are also pharmacokinetic studies reporting that green tea decreased the absorption of rosuvastatin in healthy volunteers due to the inhibition of intestinal OATP1A2- and OATP2B1-mediated drug uptake [87, 90]. Additionally, green tea has been shown to interfere with pharmacokinetics of nadolol in Japanese volunteers, possibly due to a decrease in absorption via influencing OATP1A2 and OATP2B1 [90, 91].

Conclusions

OATP transporters are widely distributed in human tissues and mediate the influx of many drugs and endogenous substances. Food and herbs are important sources of nutrients with potential health benefits. Accordingly, the potential of drug-herb/food interactions due to concurrent use of food and herbal agents alongside conventional drugs is high, which may greatly impact on therapeutic outcomes and toxicities. Increased awareness of these interactions could inform precautions to avoid co-administration of herbal medicines or food with OATP drug substrates to improve clinical outcomes.

Availability of data and materials

All the data used to support the findings of this study are available from the corresponding author upon reasonable request.

Abbreviations

AUC:

Area under the plasma concentration–time curve

Cmax :

Maximal plasma drug concentration

EC:

Epicatechin

ECG:

Epicatechin gallate

EGC:

Epicatechin-3-gallate

EGCG:

Epigallocatechin gallate

ES:

Estrone-3-sulfate

OAT:

Organic anion transporter

OATP:

Organic anion transporting polypeptide

β-PGG:

1,2,3,4,6-penta-O-galloyl-β-d-glucose

SLCs:

Solute Carrier transporters

References

  1. Chan T, Cheung FS, Zheng J, Lu X, Zhu L, Grewal T, Murray M, Zhou F. Casein Kinase 2 Is a Novel Regulator of the Human Organic Anion Transporting Polypeptide 1A2 (OATP1A2) Trafficking. Mol Pharm. 2016;13(1):144–54.

    CAS  PubMed  Google Scholar 

  2. Murray M, Zhou F. Trafficking and other regulatory mechanisms for organic anion transporting polypeptides and organic anion transporters that modulate cellular drug and xenobiotic influx and that are dysregulated in disease. Br J Pharmacol. 2017;174(13):1908–24.

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Zhou F, Zhu L, Wang K, Murray M. Recent advance in the pharmacogenomics of human Solute Carrier Transporters (SLCs) in drug disposition. Adv Drug Deliv Rev. 2017;116:21–36.

    CAS  PubMed  Google Scholar 

  4. Kovacsics D, Patik I, Ozvegy-Laczka C. The role of organic anion transporting polypeptides in drug absorption, distribution, excretion and drug-drug interactions. Expert Opin Drug Metab Toxicol. 2017;13(4):409–24.

    CAS  PubMed  Google Scholar 

  5. Roth M, Obaidat A, Hagenbuch B. OATPs, OATs and OCTs: the organic anion and cation transporters of the SLCO and SLC22A gene superfamilies. Br J Pharmacol. 2012;165(5):1260–87.

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Bronger H, Konig J, Kopplow K, Steiner HH, Ahmadi R, Herold-Mende C, Keppler D, Nies AT. ABCC drug efflux pumps and organic anion uptake transporters in human gliomas and the blood-tumor barrier. Cancer Res. 2005;65(24):11419–28.

    CAS  PubMed  Google Scholar 

  7. Chan T, Zheng J, Zhu L, Grewal T, Murray M, Zhou F. Putative transmembrane domain 6 of the human organic anion transporting polypeptide 1A2 (OATP1A2) influences transporter substrate binding, protein trafficking, and quality control. Mol Pharm. 2015;12(1):111–9.

    CAS  PubMed  Google Scholar 

  8. Glaeser H, Bailey DG, Dresser GK, Gregor JC, Schwarz UI, McGrath JS, Jolicoeur E, Lee W, Leake BF, Tirona RG, et al. Intestinal drug transporter expression and the impact of grapefruit juice in humans. Clin Pharmacol Ther. 2007;81(3):362–70.

    CAS  PubMed  Google Scholar 

  9. Knauer MJ, Urquhart BL. Meyer zu Schwabedissen HE, Schwarz UI, Lemke CJ, Leake BF, Kim RB, Tirona RG: Human skeletal muscle drug transporters determine local exposure and toxicity of statins. Circ Res. 2010;106(2):297–306.

    CAS  PubMed  Google Scholar 

  10. Kobayashi D, Nozawa T, Imai K, Nezu J, Tsuji A, Tamai I. Involvement of human organic anion transporting polypeptide OATP-B (SLC21A9) in pH-dependent transport across intestinal apical membrane. J Pharmacol Exp Ther. 2003;306(2):703–8.

    CAS  PubMed  Google Scholar 

  11. Kullak-Ublick GA, Ismair MG, Stieger B, Landmann L, Huber R, Pizzagalli F, Fattinger K, Meier PJ, Hagenbuch B. Organic anion-transporting polypeptide B (OATP-B) and its functional comparison with three other OATPs of human liver. Gastroenterology. 2001;120(2):525–33.

    CAS  PubMed  Google Scholar 

  12. St-Pierre MV, Hagenbuch B, Ugele B, Meier PJ, Stallmach T. Characterization of an organic anion-transporting polypeptide (OATP-B) in human placenta. J Clin Endocrinol Metab. 2002;87(4):1856–63.

    CAS  PubMed  Google Scholar 

  13. Kock K, Koenen A, Giese B, Fraunholz M, May K, Siegmund W, Hammer E, Volker U, Jedlitschky G, Kroemer HK, et al. Rapid modulation of the organic anion transporting polypeptide 2B1 (OATP2B1, SLCO2B1) function by protein kinase C-mediated internalization. J Biol Chem. 2010;285(15):11336–47.

    PubMed  PubMed Central  Google Scholar 

  14. Hong M, Hong W, Ni C, Huang J, Zhou C. Protein kinase C affects the internalization and recycling of organic anion transporting polypeptide 1B1. Biochim Biophys Acta. 2015;1848(10 Pt A):2022–30.

    CAS  PubMed  Google Scholar 

  15. Powell J, Farasyn T, Kock K, Meng X, Pahwa S, Brouwer KL, Yue W. Novel mechanism of impaired function of organic anion-transporting polypeptide 1B3 in human hepatocytes: post-translational regulation of OATP1B3 by protein kinase C activation. Drug Metab Dispos. 2014;42(11):1964–70.

    PubMed  PubMed Central  Google Scholar 

  16. Brenner S, Klameth L, Riha J, Scholm M, Hamilton G, Bajna E, Ausch C, Reiner A, Jager W, Thalhammer T, et al. Specific expression of OATPs in primary small cell lung cancer (SCLC) cells as novel biomarkers for diagnosis and therapy. Cancer Lett. 2015;356(2 Pt B):517–24.

    CAS  PubMed  Google Scholar 

  17. Schwabedissen HEMZ, Tirona RG, Yip CS, Ho RH, Kim RB. Interplay between the Nuclear Receptor Pregnane X Receptor and the Uptake Transporter Organic Anion Transporter Polypeptide 1A2 Selectively Enhances Estrogen Effects in Breast Cancer. Cancer Res. 2008;68(22):9338–47.

    Google Scholar 

  18. Svoboda M, Mungenast F, Gleiss A, Vergote I, Vanderstichele A, Sehouli J, Braicu E, Mahner S, Jager W, Mechtcheriakova D, et al. Clinical significance of organic anion transporting polypeptide gene expression in high-grade serous ovarian cancer. Front Pharmacol. 2018;9:842.

    PubMed  PubMed Central  Google Scholar 

  19. Cui Y, Konig J, Nies AT, Pfannschmidt M, Hergt M, Franke WW, Alt W, Moll R, Keppler D. Detection of the human organic anion transporters SLC21A6 (OATP2) and SLC21A8 (OATP8) in liver and hepatocellular carcinoma. Lab Invest. 2003;83(4):527–38.

    CAS  PubMed  Google Scholar 

  20. Hamada A, Sissung T, Price DK, Danesi R, Chau CH, Sharifi N, Venzon D, Maeda K, Nagao K, Sparreboom A, et al. Effect of SLCO1B3 haplotype on testosterone transport and clinical outcome in caucasian patients with androgen-independent prostatic cancer. Clin Cancer Res. 2008;14(11):3312–8.

    CAS  PubMed  PubMed Central  Google Scholar 

  21. Lee W, Belkhiri A, Lockhart AC, Merchant N, Glaeser H, Harris EI, Washington MK, Brunt EM, Zaika A, Kim RB, et al. Overexpression of OATP1B3 confers apoptotic resistance in colon cancer. Cancer Res. 2008;68(24):10315–23.

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Muto M, Onogawa T, Suzuki T, Ishida T, Rikiyama T, Katayose Y, Ohuchi N, Sasano H, Abe T, Unno M. Human liver-specific organic anion transporter-2 is a potent prognostic factor for human breast carcinoma. Cancer Sci. 2007;98(10):1570–6.

    CAS  PubMed  Google Scholar 

  23. Liu T, Li Q. Organic anion-transporting polypeptides: a novel approach for cancer therapy. J Drug Target. 2014;22(1):14–22.

    PubMed  Google Scholar 

  24. Pressler H, Sissung TM, Venzon D, Price DK, Figg WD. Expression of OATP family members in hormone-related cancers: potential markers of progression. PLoS ONE. 2011;6(5):e20372.

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Al Sarakbi W, Mokbel R, Salhab M, Jiang WG, Reed MJ, Mokbel K. The role of STS and OATP-B mRNA expression in predicting the clinical outcome in human breast cancer. Anticancer Res. 2006;26(6C):4985–90.

    CAS  PubMed  Google Scholar 

  26. Matsumoto J, Ariyoshi N, Sakakibara M, Nakanishi T, Okubo Y, Shiina N, Fujisaki K, Nagashima T, Nakatani Y, Tamai I, et al. Organic anion transporting polypeptide 2B1 expression correlates with uptake of estrone-3-sulfate and cell proliferation in estrogen receptor-positive breast cancer cells. Drug Metab Pharmacokinet. 2015;30(2):133–41.

    CAS  PubMed  Google Scholar 

  27. Lu X, Xu C, Ng WV, Zhu L, Zhou F. The interaction of herbal compounds with human Organic anion/cation transporters. Journal of Pharmacogenomics pharmacoproteomics. 2014;5(5):142.

    Google Scholar 

  28. Cui T, Liu Y, Men X, Xu Z, Wu L, Liu S, Xing A. Bile acid transport correlative protein mRNA expression profile in human placenta with intrahepatic cholestasis of pregnancy. Saudi Med J. 2009;30(11):1406–10.

    PubMed  Google Scholar 

  29. Wojtal KA, Eloranta JJ, Hruz P, Gutmann H, Drewe J, Staumann A, Beglinger C, Fried M, Kullak-Ublick GA, Vavricka SR. Changes in mRNA expression levels of solute carrier transporters in inflammatory bowel disease patients. Drug Metab Dispos. 2009;37(9):1871–7.

    CAS  PubMed  Google Scholar 

  30. Plaza F, Gabler F, Romero C, Vantman D, Valladares L, Vega M. The conversion of dehydroepiandrosterone into androst-5-ene-3 beta,17 beta-diol (androstenediol) is increased in endometria from untreated women with polycystic ovarian syndrome. Steroids. 2010;75(12):810–7.

    CAS  PubMed  Google Scholar 

  31. Petrovic V, Kojovic D, Cressman A, Piquette-Miller M. Maternal bacterial infections impact expression of drug transporters in human placenta. Int Immunopharmacol. 2015;26(2):349–56.

    CAS  PubMed  Google Scholar 

  32. Oswald M, Kullak-Ublick GA, Paumgartner G, Beuers U. Expression of hepatic transporters OATP-C and MRP2 in primary sclerosing cholangitis. Liver. 2001;21(4):247–53.

    CAS  PubMed  Google Scholar 

  33. Stieger B, Geier A. Genetic variations of bile salt transporters as predisposing factors for drug-induced cholestasis, intrahepatic cholestasis of pregnancy and therapeutic response of viral hepatitis. Expert Opinion on Drug Metabolism & Toxicology. 2011;7(4):411–25.

    CAS  Google Scholar 

  34. Le Vee M, Gripon P, Stieger B, Fardel O. Down-regulation of organic anion transporter expression in human hepatocytes exposed to the proinflammatory cytokine interleukin 1beta. Drug Metab Dispos. 2008;36(2):217–22.

    PubMed  Google Scholar 

  35. Le Vee M, Jouan E, Moreau A, Fardel O. Regulation of drug transporter mRNA expression by interferon-gamma in primary human hepatocytes. Fundam Clin Pharmacol. 2011;25(1):99–103.

    PubMed  Google Scholar 

  36. Le Vee M, Jouan E, Stieger B, Lecureur V, Fardel O. Regulation of drug transporter expression by oncostatin M in human hepatocytes. Biochem Pharmacol. 2011;82(3):304–11.

    PubMed  Google Scholar 

  37. Clarke JD, Novak P, Lake AD, Hardwick RN, Cherrington NJ. Impaired N-linked glycosylation of uptake and efflux transporters in human non-alcoholic fatty liver disease. Liver Int. 2017;37(7):1074–81.

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Wang L, Collins C, Kelly EJ, Chu X, Ray AS, Salphati L, Xiao G, Lee C, Lai Y, Liao M, et al. Transporter Expression in Liver Tissue from Subjects with Alcoholic or Hepatitis C Cirrhosis Quantified by Targeted Quantitative Proteomics. Drug Metab Dispos. 2016;44(11):1752–8.

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Chen HL, Liu YJ, Chen HL, Wu SH, Ni YH, Ho MC, Lai HS, Hsu WM, Hsu HY, Tseng HC, et al. Expression of hepatocyte transporters and nuclear receptors in children with early and late-stage biliary atresia. Pediatr Res. 2008;63(6):667–73.

    CAS  PubMed  Google Scholar 

  40. Congiu M, Mashford ML, Slavin JL, Desmond PV. Coordinate regulation of metabolic enzymes and transporters by nuclear transcription factors in human liver disease. J Gastroenterol Hepatol. 2009;24(6):1038–44.

    CAS  PubMed  Google Scholar 

  41. Keitel V, Burdelski M, Warskulat U, Kuhlkamp T, Keppler D, Haussinger D, Kubitz R. Expression and localization of hepatobiliary transport proteins in progressive familial intrahepatic cholestasis. Hepatology. 2005;41(5):1160–72.

    CAS  PubMed  Google Scholar 

  42. van de Steeg E, Stranecky V, Hartmannova H, Noskova L, Hrebicek M, Wagenaar E, van Esch A, de Waart DR, Oude Elferink RP, Kenworthy KE, et al. Complete OATP1B1 and OATP1B3 deficiency causes human Rotor syndrome by interrupting conjugated bilirubin reuptake into the liver. J Clin Invest. 2012;122(2):519–28.

    PubMed  PubMed Central  Google Scholar 

  43. Keppler D. The roles of MRP2, MRP3, OATP1B1, and OATP1B3 in conjugated hyperbilirubinemia. Drug Metab Dispos. 2014;42(4):561–5.

    PubMed  Google Scholar 

  44. Buyukkale G, Turker G, Kasap M, Akpinar G, Arisoy E, Gunlemez A, Gokalp A. Neonatal hyperbilirubinemia and organic anion transporting polypeptide-2 gene mutations. Am J Perinatol. 2011;28(8):619–26.

    PubMed  Google Scholar 

  45. Ieiri I, Suzuki H, Kimura M, Takane H, Nishizato Y, Irie S, Urae A, Kawabata K, Higuchi S, Otsubo K, et al. Influence of common variants in the pharmacokinetic genes (OATP-C, UGT1A1, and MRP2) on serum bilirubin levels in healthy subjects. Hepatol Res. 2004;30(2):91–5.

    CAS  PubMed  Google Scholar 

  46. Zhang W, He YJ, Gan Z, Fan L, Li Q, Wang A, Liu ZQ, Deng S, Huang YF, Xu LY, et al. OATP1B1 polymorphism is a major determinant of serum bilirubin level but not associated with rifampicin-mediated bilirubin elevation. Clin Exp Pharmacol Physiol. 2007;34(12):1240–4.

    CAS  PubMed  Google Scholar 

  47. Zhang Z, Xia W, He J, Zhang Z, Ke Y, Yue H, Wang C, Zhang H, Gu J, Hu W, et al. Exome sequencing identifies SLCO2A1 mutations as a cause of primary hypertrophic osteoarthropathy. Am J Hum Genet. 2012;90(1):125–32.

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Niemi M, Neuvonen PJ, Hofmann U, Backman JT, Schwab M, Lutjohann D, von Bergmann K, Eichelbaum M, Kivisto KT. Acute effects of pravastatin on cholesterol synthesis are associated with SLCO1B1 (encoding OATP1B1) haplotype *17. Pharmacogenet Genomics. 2005;15(5):303–9.

    CAS  PubMed  Google Scholar 

  49. Zhang A, Wang C, Liu Q, Meng Q, Peng J, Sun H, Ma X, Huo X, Liu K. Involvement of organic anion-transporting polypeptides in the hepatic uptake of dioscin in rats and humans. Drug Metab Dispos. 2013;41(5):994–1003.

    CAS  PubMed  Google Scholar 

  50. Hartkoorn RC, Kwan WS, Shallcross V, Chaikan A, Liptrott N, Egan D, Sora ES, James CE, Gibbons S, Bray PG, et al. HIV protease inhibitors are substrates for OATP1A2, OATP1B1 and OATP1B3 and lopinavir plasma concentrations are influenced by SLCO1B1 polymorphisms. Pharmacogenet Genom. 2010;20(2):112–20.

    CAS  Google Scholar 

  51. Yamakawa Y, Hamada A, Shuto T, Yuki M, Uchida T, Kai H, Kawaguchi T, Saito H. Pharmacokinetic impact of SLCO1A2 polymorphisms on imatinib disposition in patients with chronic myeloid leukemia. Clin Pharmacol Ther. 2011;90(1):157–63.

    CAS  PubMed  Google Scholar 

  52. Akamine Y, Miura M, Sunagawa S, Kagaya H, Yasui-Furukori N, Uno T. Influence of drug-transporter polymorphisms on the pharmacokinetics of fexofenadine enantiomers. Xenobiotica. 2010;40(11):782–9.

    CAS  PubMed  Google Scholar 

  53. Sebastian K, Detro-Dassen S, Rinis N, Fahrenkamp D, Muller-Newen G, Merk HF, Schmalzing G, Zwadlo-Klarwasser G, Baron JM. Characterization of SLCO5A1/OATP5A1, a solute carrier transport protein with non-classical function. PLoS ONE. 2013;8(12):e83257.

    PubMed  PubMed Central  Google Scholar 

  54. Isidor B, Pichon O, Redon R, Day-Salvatore D, Hamel A, Siwicka KA, Bitner-Glindzicz M, Heymann D, Kjellen L, Kraus C, et al. Mesomelia-Synostoses Syndrome Results from Deletion of SULF1 and SLCO5A1 Genes at 8q13. Am J Hum Genet. 2010;87(1):95–100.

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Vo KT, Montgomery ME, Mitchell ST, Scheerlinck PH, Colby DK, Meier KH, Kim-Katz S, Anderson IB, Offerman SR, Olson KR, et al. Amanita phalloides Mushroom Poisonings - Northern California, December 2016. MMWR Morb Mortal Wkly Rep. 2017;66(21):549–53.

    PubMed  PubMed Central  Google Scholar 

  56. Nahrstedt A, Butterweck V. Lessons learned from herbal medicinal products: the example of St. John’s Wort (perpendicular). J Nat Prod. 2010;73(5):1015–21.

    CAS  PubMed  Google Scholar 

  57. Kuriyama S, Shimazu T, Ohmori K, Kikuchi N, Nakaya N, Nishino Y, Tsubono Y, Tsuji I. Green tea consumption and mortality due to cardiovascular disease, cancer, and all causes in Japan: the Ohsaki study. JAMA. 2006;296(10):1255–65.

    CAS  PubMed  Google Scholar 

  58. Stieger B, Mahdi ZM, Jager W. Intestinal and hepatocellular transporters: therapeutic effects and drug interactions of herbal supplements. Annu Rev Pharmacol Toxicol. 2017;57:399–416.

    CAS  PubMed  Google Scholar 

  59. Lu X, Chan T, Zhu L, Bao X, Velkov T, Zhou QT, Li J, Chan HK, Zhou F. The inhibitory effects of eighteen front-line antibiotics on the substrate uptake mediated by human Organic anion/cation transporters, Organic anion transporting polypeptides and Oligopeptide transporters in in vitro models. Eur J Pharm Sci. 2018;115:132–43.

    CAS  PubMed  Google Scholar 

  60. Ivanyuk A, Livio F, Biollaz J, Buclin T. Renal drug transporters and drug interactions. Clin Pharmacokinet. 2017;56(8):825–92.

    CAS  PubMed  Google Scholar 

  61. Konig J, Muller F, Fromm MF. Transporters and drug-drug interactions: important determinants of drug disposition and effects. Pharmacol Rev. 2013;65(3):944–66.

    PubMed  Google Scholar 

  62. Tamai I, Nakanishi T. OATP transporter-mediated drug absorption and interaction. Curr Opin Pharmacol. 2013;13(6):859–63.

    CAS  PubMed  Google Scholar 

  63. Li Z, Cheung FS, Zheng J, Chan T, Zhu L, Zhou F. Interaction of the bioactive flavonol, icariin, with the essential human solute carrier transporters. J Biochem Mol Toxicol. 2014;28(2):91–7.

    PubMed  Google Scholar 

  64. Li Z, Wang K, Zheng J, Cheung FS, Chan T, Zhu L, Zhou F. Interactions of the active components of Punica granatum (pomegranate) with the essential renal and hepatic human Solute Carrier transporters. Pharm Biol. 2014;52(12):1510–7.

    CAS  PubMed  Google Scholar 

  65. Roth M, Timmermann BN, Hagenbuch B. Interactions of green tea catechins with organic anion-transporting polypeptides. Drug Metab Dispos. 2011;39(5):920–6.

    PubMed  PubMed Central  Google Scholar 

  66. Shams T, Lu X, Zhu L, Zhou F. The inhibitory effects of five alkaloids on the substrate transport mediated through human organic anion and cation transporters. Xenobiotica. 2018;48(2):197–205.

    CAS  PubMed  Google Scholar 

  67. Xu F, Li Z, Zheng J, Gee Cheung FS, Chan T, Zhu L, Zhuge H, Zhou F. The inhibitory effects of the bioactive components isolated from Scutellaria baicalensis on the cellular uptake mediated by the essential solute carrier transporters. J Pharm Sci. 2013;102(11):4205–11.

    CAS  PubMed  Google Scholar 

  68. Oh Y, Jeong YS, Kim MS, Min JS, Ryoo G, Park JE, Jun Y, Song YK, Chun SE, Han S, et al. Inhibition of organic anion transporting polypeptide 1B1 and 1B3 by betulinic acid: effects of preincubation and albumin in the media. J Pharm Sci. 2018;107(6):1713–23.

    CAS  PubMed  Google Scholar 

  69. Dolton MJ, Roufogalis BD, McLachlan AJ. Fruit juices as perpetrators of drug interactions: the role of organic anion-transporting polypeptides. Clin Pharmacol Ther. 2012;92(5):622–30.

    CAS  PubMed  Google Scholar 

  70. Jiro Ogura HY. Nariyasu Mano: stimulatory effect on the transport mediated by organic anion transporting polypeptide 2B1. Asian J Pharm Sci. 2019;15(2):181–91.

    PubMed  PubMed Central  Google Scholar 

  71. Iijima R, Watanabe T, Ishiuchi K, Matsumoto T, Watanabe J, Makino T. Interactions between crude drug extracts used in Japanese traditional Kampo medicines and organic anion-transporting polypeptide 2B1. J Ethnopharmacol. 2018;214:153–9.

    PubMed  Google Scholar 

  72. Wu LX, Guo CX, Qu Q, Yu J, Chen WQ, Wang G, Fan L, Li Q, Zhang W, Zhou HH. Effects of natural products on the function of human organic anion transporting polypeptide 1B1. Xenobiotica. 2012;42(4):339–48.

    CAS  PubMed  Google Scholar 

  73. Lilja JJ, Raaska K, Neuvonen PJ. Effects of orange juice on the pharmacokinetics of atenolol. Eur J Clin Pharmacol. 2005;61(5–6):337–40.

    CAS  PubMed  Google Scholar 

  74. Wang XD, Wolkoff AW, Morris ME. Flavonoids as a novel class of human organic anion-transporting polypeptide OATP1B1 (OATP-C) modulators. Drug Metab Dispos. 2005;33(11):1666–72.

    CAS  PubMed  Google Scholar 

  75. Kock K, Xie Y, Hawke RL, Oberlies NH, Brouwer KL. Interaction of silymarin flavonolignans with organic anion-transporting polypeptides. Drug Metab Dispos. 2013;41(5):958–65.

    CAS  PubMed  PubMed Central  Google Scholar 

  76. Wen FJ, Shi MZ, Bian JL, Zhang HJ, Gui CS. Identification of natural products as modulators of OATP2B1 using LC-MS/MS to quantify OATP-mediated uptake. Pharmaceutical Biology. 2016;54(2):293–302.

    CAS  PubMed  Google Scholar 

  77. Bailey DG, Dresser GK, Leake BF, Kim RB. Naringin is a major and selective clinical inhibitor of organic anion-transporting polypeptide 1A2 (OATP1A2) in grapefruit juice. Clin Pharmacol Ther. 2007;81(4):495–502.

    CAS  PubMed  Google Scholar 

  78. Mandery K, Bujok K, Schmidt I, Keiser M, Siegmund W, Balk B, Konig J, Fromm MF, Glaeser H. Influence of the flavonoids apigenin, kaempferol, and quercetin on the function of organic anion transporting polypeptides 1A2 and 2B1. Biochem Pharmacol. 2010;80(11):1746–53.

    CAS  PubMed  Google Scholar 

  79. Kondo A, Narumi K, Ogura J, Sasaki A, Yabe K, Kobayashi T, Furugen A, Kobayashi M, Iseki K. Organic anion-transporting polypeptide (OATP) 2B1 contributes to the cellular uptake of theaflavin. Drug Metab Pharmacokinet. 2017;32(2):145–50.

    CAS  PubMed  Google Scholar 

  80. Iwase S, Yamaguchi T, Miyaji T, Terawaki K, Inui A, Uezono Y. The clinical use of Kampo medicines (traditional Japanese herbal treatments) for controlling cancer patients’ symptoms in Japan: a national cross-sectional survey. Bmc Complementary and Alternative Medicine. 2012;12:222.

    PubMed  PubMed Central  Google Scholar 

  81. Lilja JJ, Raaska K, Neuvonen PJ. Effects of grapefruit juice on the pharmacokinetics of acebutolol. Br J Clin Pharmacol. 2005;60(6):659–63.

    CAS  PubMed  PubMed Central  Google Scholar 

  82. Ogura J, Koizumi T, Segawa M, Yabe K, Kuwayama K, Sasaki S, Kaneko C, Tsujimoto T, Kobayashi M, Yamaguchi H, et al. Quercetin-3-rhamnoglucoside (rutin) stimulates transport of organic anion compounds mediated by organic anion transporting polypeptide 2B1. Biopharm Drug Dispos. 2014;35(3):173–82.

    CAS  PubMed  Google Scholar 

  83. Segawa M, Ogura J, Seki S, Itagaki S, Takahashi N, Kobayashi M, Hirano T, Yamaguchi H, Iseki K. Rapid stimulating effect of the antiarrhythmic agent amiodarone on absorption of organic anion compounds. Drug Metab Pharmacokinet. 2013;28(3):178–86.

    CAS  PubMed  Google Scholar 

  84. Atkins WM. Implications of the allosteric kinetics of cytochrome P450s. Drug Discov Today. 2004;9(11):478–84.

    CAS  PubMed  Google Scholar 

  85. Dresser GK, Bailey DG, Leake BF, Schwarz UI, Dawson PA, Freeman DJ, Kim RB. Fruit juices inhibit organic anion transporting polypeptide-mediated drug uptake to decrease the oral availability of fexofenadine. Clin Pharmacol Ther. 2002;71(1):11–20.

    CAS  PubMed  Google Scholar 

  86. Lilja JJ, Backman JT, Laitila J, Luurila H, Neuvonen PJ. Itraconazole increases but grapefruit juice greatly decreases plasma concentrations of celiprolol. Clin Pharmacol Ther. 2003;73(3):192–8.

    CAS  PubMed  Google Scholar 

  87. Kim TE, Ha N, Kim Y, Kim H, Lee JW, Jeon JY, Kim MG. Effect of epigallocatechin-3-gallate, major ingredient of green tea, on the pharmacokinetics of rosuvastatin in healthy volunteers. Drug Des Devel Ther. 2017;11:1409–16.

    CAS  PubMed  PubMed Central  Google Scholar 

  88. Gui C, Obaidat A, Chaguturu R, Hagenbuch B. Development of a cell-based high-throughput assay to screen for inhibitors of organic anion transporting polypeptides 1B1 and 1B3. Curr Chem Genomics. 2010;4:1–8.

    CAS  PubMed  PubMed Central  Google Scholar 

  89. Ismair MG, Stanca C, Ha HR, Renner EL, Meier PJ, Kullak-Ublick GA. Interactions of glycyrrhizin with organic anion transporting polypeptides of rat and human liver. Hepatol Res. 2003;26(4):343–7.

    CAS  PubMed  Google Scholar 

  90. Werba JP, Misaka S, Giroli MG, Shimomura K, Amato M, Simonelli N, Vigo L, Tremoli E. Update of green tea interactions with cardiovascular drugs and putative mechanisms. J Food Drug Anal. 2018;26(2S):S72–7.

    CAS  PubMed  Google Scholar 

  91. Misaka S, Yatabe J, Muller F, Takano K, Kawabe K, Glaeser H, Yatabe MS, Onoue S, Werba JP, Watanabe H, et al. Green tea ingestion greatly reduces plasma concentrations of nadolol in healthy subjects. Clin Pharmacol Ther. 2014;95(4):432–8.

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thanks to Wanbangde Pharmaceutical Pty Ltd for the support of USYD-WEPON post-graduate scholarship.

Funding

We appreciate the financial support received from the Equity fellowship of the University of Sydney (Grant No. 2019 Equity Fellowship), the Young Talent’s Subsidy Project in Science and Education of the Department of Public Health of Jiangsu Province (No. QNRC2016627), Six talent peaks project in Jiangsu Province (No. WSW-047), Six-one Scientific Research Project (No. LGY2019087).

Author information

Authors and Affiliations

Authors

Contributions

YA, TS, ZC, YL and WS did the literature search and drafted the manuscript. KW, XB, LZ, MM and FZ critically reviewed the literature and revised the manuscript. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Fanfan Zhou.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that 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/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ali, Y., Shams, T., Wang, K. et al. The involvement of human organic anion transporting polypeptides (OATPs) in drug-herb/food interactions. Chin Med 15, 71 (2020). https://doi.org/10.1186/s13020-020-00351-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13020-020-00351-9

Keywords