InteractionsGuide Index Page

 
Case Analysis Toolclose
Enter Each Substance:


Analysis Search Terms:

St. John's Wort

Botanical Name: Hypericum perforatum L.
Pharmacopoeial Name: Hyperici herba.
Common Names: St. John’s wort, Klamath weed (historic: Fuga daemonum, herba solus).

Summary Table
herb description

Family

Clusiaceae (Guttiferae, Hypericaceae).

Habitat and Cultivation

Perennial; native in Europe Asia and North Africa; naturalized in the United States and considered a noxious weed in many areas; widespread in temperate zones, favoring disturbed ground.

Parts Used

Flowering tops.

Common Forms

  • Dried Plant:   Flowering tops.

  • Tincture:   60% ethanol, 1:2 to 1:5 weight/volume.

  • Standardized Extract:   0.3% hypericin, 2.0% to 4.5% hyperforin.

  • Infused Oil:   Fresh flowers, for external use.

herb in clinical practice

Overview

A well-documented botanical medicine since Greco-Roman times, St. John's wort (SJW) has a long history of folk and traditional use as a vulnerary (“wound healer”) and for banishing mental afflictions, particularly melancholy. For example, Gerard 1 (1633) described its use as a balm for wounds, burns, ulcers, and bites as being without equal. The oil made from the macerated flowers was listed in the first Pharmacopoeia Londinensis (1618). Hypericum perforatum was proved and introduced into the homeopathic materia medica by Muller in the mid-1800s and has been included in the Homeopathic Pharmacopoeia of the United States since that era, with primary indications focusing on nerve pain and traumatic injuries (e.g., concussion, coccygeal impact, sequelae).

More recently, clinical trial evidence accumulated through the 1980s and 1990s established the efficacy and safety of standardized SJW extracts for treating mild to moderate depression, and the “natural antidepressant” label propelled the herb to second-best-selling supplement in the United States by the late 1990s. In 2000, reports of serious interactions with prescription drugs began to appear, and the resulting adverse publicity caused sales of the herb to fall significantly, although SJW remains one of the top-selling U.S. botanicals. It was approved by the German Commission E for “depressive moods” (internally) and “contused injuries” (externally) in 1984. 2 The pharmacology and clinical effects of the herb are currently the focus of considerable research interest and, because of rapid accumulation of data, relatively recent literature reviews (e.g., 1997 American Herbal Pharmacopoeia monograph) are in some respects dated. 3 More recent reviews of the extensive literature include the 2003 European Scientific Cooperative on Phytotherapy (ESCOP) monograph 4 and a comprehensive monograph by McKenna et al. 5

Historical/Ethnomedicine Precedent

Traditionally, SJW was used as a calming herb for symptoms of nervous tension, including anxiety and insomnia, as well as a restorative for melancholic conditions that might currently be diagnosed as depression. Folk use attributed the herb with properties of protection against enchantments, including demonic possession, and it was used for warding off evil spirits. Hypericum was characterized as “hot and dry” in the Galenic humoral system of medicine and has classically been associated with the liver and spleen, as well as the Sun. Historically considered a “woundwort,” SJW is still used both internally and externally for pain relief, particularly neuralgic pain, shingles, mild contusions, and burns to the skin. For external use, the fresh flowers, traditionally harvested on St. John's Day (immediately following Summer Solstice), are the basis of a macerated oil, which is usually red (by the dianthrone hypericin). This red color was considered an indication of its vulnerary nature (likened to blood) by the Doctrine of Signatures. Before the modern clinical trial–driven indications of the herb for “mild to moderate depression,” the nervous system indications were less clearly defined and included “psychovegetative” disorders, as well as such conditions as nocturnal eneuresis and night terrors. Its psychological effects were considered much less pronounced than those of prescription medications; Weiss 6 classified the herb as a “mild (i.e., gentle) psychotropic” agent.

Known or Potential Therapeutic Uses

Analgesic, antiviral, anti-inflammatory, anxiety, coccygeal impact, concussion depression (mild to moderate), hepatoprotection, herpes simplex infection (orofacial and genital), herpes zoster (shingles and postherpetic neuralgia), menopause-related psychological symptoms, psychosomatic and somatiform disorders (mild), nervousness, neuralgia, nocturnal eneuresis, photodynamic antitumor activity, premenstrual syndrome, restlessness, sacral irritation and spinal injuries, sciatica, seasonal affective disorder, tissue healing and wound repair.

Key Constituents

Characteristic napthodianthrones, including hypericin; phloroglucinols, including hyperforin and adhyperforin.

Flavonoids, including proanthocyanidin polymers of catechin and epicatechin; flavonols; phenylpropanoids; essential oil; amino acids; xanthones.

Therapeutic Dosing Range

  • Dried Plant:   2 to 5 g/day.

  • Tincture and Fluid Extract:   As 1:1 equivalents, 1 to 3 mL/day.

  • Standardized Extracts:   900 mg/day in divided doses.

  • Topical:   Oleum hyperici , oily macerate from fresh flowering tops (applied as needed).

Also used in ultradilute succussed preparations based on homeopathic indications.

interactions review

Strategic Considerations/Background

Although an old medicine, SJW has a pivotal place in the relatively recent field of herb-drug interactions. The publication of convincing reports of interactions between SJW and digoxin 7 in 1999 and cyclosporine 8 and indinavir 9 in 2000 was seminal, initiating a widespread reevaluation of the safety of this popular herb, previously considered to be benign, in the context of conventional medications. 10 It also propelled the issue of potential interactions between botanicals and pharmaceuticals into media prominence and research focus. The subsequent years have seen increased understanding of the pharmacology of SJW, and the herb is now known to be associated with a number of clinically significant pharmacokinetic interactions, as suggested by the original reports. These interactions are mediated by its effects on several key components of drug metabolism, including the cytochrome P450 (CYP450) mixed-oxidase system, various conjugases and transferases, as well as the transporter proteins that modulate drug efflux across intestinal, renal, and biliary epithelia. These systems compose what are now often referred to as phases (or stages) I, II, and III of drug metabolism/detoxification.

The initial reports of SJW interactions with narrow-therapeutic-range drugs prompted sweeping warnings in professional and consumer media about the dangers of SJW herb-drug interactions (and often of herb-drug interactions in general). At the time, however, the actual number of reports of documented SJW-related drug interactions was, and in fact remains, relatively small, with data of widely varying reliability. Surveying the available cases in 2001, Fugh-Berman and Ernst 11 found 54 published reports claiming SJW interactions. Of these, 29 were rejected as unclassifiable, and the remaining 25 were evaluated for reliability according to the authors’ “reliability rating score” system. Of these, 12 were classified as “unreliable,” 11 as “possible,” and only two as “likely.” More recently, Meyer et al. 12 analyzed six documented potential herb-drug interactions, including SJW-cyclosporine and SJW-digoxin, across a wide range of “tertiary sources” and found high variability in the reporting of the data, with only three sources even mentioning all six known interactions. Interestingly, as recently reviewed by Izzo, 13 clinical reports of SJW-drug interactions seem to be decreasing rather than increasing in frequency.

Mills et al. 14 recently conducted a systematic review of trials investigating SJW pharmacokinetic interactions with conventional drugs. The authors found the methodological quality of the studies was limited; in particular lacking accepted controls such as correct randomization, observance of established blinding procedures, and allowance for time-dependent effects. They also found that only 15 of the 22 available studies assayed the SJW content of the preparations used, and that varied dosing regimens and duration of exposure to the herb were common, without presenting a rationale for the tested dosing patterns. These limitations mean that most trials on SJW interactions do not appear to conform to the U.S. Food and Drug Administration (FDA)–recommended standards for safeguards against bias in pharmacokinetic trials. 15 This in turn results in questions about the interpretation and applicability of the available data that can only be resolved by more and better-designed studies, as well as consistent application of necessary standards in pharmacovigilance.

Official and regulatory reaction was also triggered by the initial SJW interaction reports. In 2000 the U.K. Committee on Safety of Medicines (CSM) 16 issued a general advisory letter on SJW interactions to all physicians and pharmacists. This included a fact sheet listing medications for which SJW might interact and advised patients to “stop taking St John's Wort,” while warning against immediate discontinuation in the event that drug levels might rise, causing serious adverse effects. Lists of drugs that might interact with SJW, causing “serious adverse interactions,” were provided, including selective serotonin reuptake inhibitors (SSRIs), anticonvulsants, and triptans. In 2001 the Irish Medical Board (IMB) 17 restricted SJW to physician prescription only, effectively removing the herb (along with ginkgo and several others) from general public access, citing the monoamine oxidase inhibitor (MAOI) activity of SJW as potentially interacting with tyramine foods and potentiating MAOI drugs, as well as claiming SJW caused phototoxicity and other (unspecified) adverse effects. The FDA issued an advisory to health care professionals warning about the SJW-indinavir interaction in 2000, also suggesting physicians alert patients about potential drug interactions involving “any drug metabolized via the cytochrome P450 pathway.” 18

Effects on Drug Metabolism and Bioavailability

Cytochrome P450

The complete spectrum of induction and inhibition effects of SJW on the CYP450 system in vivo in humans is not yet fully characterized. Possibly because of a number of differing investigative methodologies, as well as differences between the various types of extracts used, the available studies are inconclusive. In vitro evidence exists for inhibition effects by crude SJW extracts, its flavonoid components, and hypericin and hyperforin on CYP450 1A2, 2C9, 2C19, 2D6, and 3A4. 19,20In vivo studies using single probe drugs that are specific CYP substrates have found induction effects by SJW on 3A4, 21 and with multiprobe drug “cocktails,” for 3A4, 2E1, 1A2, and 2D6 22 and 2C19. 23 By contrast, no significant effects on 2D6 and 3A4 were found by two other groups, 24,25and a further probe cocktail study found no effect on 1A2, 2C9, or 2D6. 26 More recent studies have confirmed in vivo coordinate induction effects by SJW on hepatic and intestinal 3A4 and P-glycoprotein (P-gp). 27,28

Summarizing the data available at this time, SJW definitely induces human 3A4; probably induces 1A2, 2C19, and 2E1; and probably does not significantly affect 2C9 or 2D6. It also induces P-gp and possibly other, related transporters. There is a degree of tissue specificity, with induction of both hepatic and intestinal 3A4, as well as a possible biphasic effect, at least on 3A4 and P-gp, with short-term inhibition followed by an increasing induction of enzymes over 7 to 10 days. However, evidence from isolated constituent studies suggests that hyperforin plays the main role in induction activity. 23,29-34The initial inhibition may be caused by hypericin, but also by flavonoid constituents; a number of flavonoids are known to inhibit 3A4, with those from grapefruit and other citrus-derived flavonoids being the best-known examples. 35,36This “biphasic” effect of a short-term enzyme inhibition succeeded by longer-term induction has recently been demonstrated in a clinical study of voriconazole pharmacokinetics. This open-label study with 16 healthy male volunteers determined that that SJW coadministration with voriconazole (a substrate of CYP2C19) led to a short-term but clinically insignificant increase in the area under curve (AUC) of 22%, and after 15 days, AUC was reduced by 59% compared with controls. 37,38

Pregnane X Receptor

The recent finding that hyperforin, an active phloroglucinol constituent compound of SJW, acts as a high-affinity ligand for the orphan nuclear receptor pregnane X receptor (PXR) is highly significant. 39,40The PXR and related nuclear receptors, such as the constitutive androstane receptor (CAR) and the retinoid X receptor (RXR), have been described as “promiscuous” because of the unprecedented structural diversity of compounds that interact with their ligand-binding domain (LBD). 41-43Activation of the PXR leads to upregulation of genes controlling multiple aspects of xenobiotic metabolism, including phase I (CYP450 1A1, 1A2, 2B6, 2C9, and 3A4) mixed oxidases, phase II conjugases (uridine diphosphate [UDP] glucuronosyltransferases, glutathione- S-transferases, sulfonyltransferases), and phase III drug transporters (MDR1/P-gp, MDR2, organic anion-transporting polypeptides [OATPs]). 19 , 26 , 29 , 30 , 32 , 33 , 44-51

The implication is that the PXR and related nuclear receptors may effectively act to coordinate xenobiotic detoxification. 41-43,52-56The PXR itself is subject to a degree of genetically determined polymorphism, the importance of which remains to be clarified, but pronounced interspecies differences are known to exist in activator compounds, with marked differences among rodent, rabbit, and human ligands. 41,57,58Pascussi et al. 59 have aptly described expression of the genes controlling xenobiotic metabolism as a “tangle of networks of nuclear and steroid receptors, where receptors share partners, ligands, DNA response elements and target genes and where the different pathways exhibit cross-talk at several levels.”

A broader view of SJW emerges from these recent developments. The herb can be conceptualized as a master inducer of detoxification, or more accurately as a xenosensory activator, capable of triggering the complex adaptive system evolved to metabolically eliminate toxic compounds, both endogenous and xenobiotic. 60,61The downstream consequences of PXR activation on drug metabolism suggest that, to some extent, SJW interactions may be predicted (and thus managed) on the basis of whether a given coadministered drug is a substrate of the enzymes or transporters induced by PXR activation, particularly 3A4 and P-gp. 55,62

P-Glycoprotein

Induction of P-gp by SJW further complicates the picture and may confound attempts to predict interactions. P-glycoprotein is a membrane-associated, adenosine triphosphate (ATP)–dependent “pumping” protein that ejects foreign or toxic compounds from cells and mediates “multidrug resistance” when induced in cancer cells. Durr et al. 28 estimated the induction of intestinal P-gp by SJW at a 1.5-fold increase in healthy human volunteers. Ernst 63 noted drugs that are dual substrates of both P-gp and CYP3A4 likely present an increased risk of pharmacokinetic interaction as a result of co-induction by SJW. However, the relative contributions of P-gp and 3A4 to drug efflux appear to be complex and differ for different agents that are dual substrates. 64

The existence of several polymorphisms in P-gp phenotypes affects normal levels of expression of both hepatic and intestinal P-gp. These polymorphisms are known to exhibit variation with racial and gender characteristics. 65,66As with P450 enzymes, dietary food ingredients may also affect P-gp expression; known examples include piperine from black pepper and some citrus flavonoids. 67,68Alpha-tocopherol can also influence P-gp, probably through PXR activation. 52 Finally, the role of non–P-gp drug transporters, such as the OATP family, has recently emerged as another potential mechanism in controlling drug bioavailability, although modulating influences on OATP expression are not currently well characterized.

Overall, the interplay between CYP3A4 and P-gp (and other transporters) is not well understood, but this “drug-efflux metabolism alliance,” as aptly named by Benet and Cummins, 69 remains of a crucial research area for future elucidation of drug interactions. 70,71

Managing Pharmacokinetic Interactions

Numerous pharmaceuticals are metabolized by CYP3A4, which is a low-affinity, high-throughput P450 enzyme expressed primarily in the small intestinal mucosa and liver. This has led to suggestions that SJW may interact with more than 50% of all known drugs. Indeed, evidence is now rapidly accumulating from preclinical screening studies that confirms SJW induction effects on a range of drugs, particularly 3A4 substrates, often in the absence of any clinical interactions data. However, the magnitude of SJW induction effects is considerably less than that of other known PXR ligands, the best-known example being rifampin, a mainstay of conventional tuberculosis therapy. Rifampin is a coordinate inducer of P-gp and 3A4 with an induction effect on midazolam (a 3A4-specific substrate) that is 25 times that of SJW. 72 Red wine has similar order-of-magnitude effects as SJW on oral clearance of cyclosporine (a dual substrate). 73

Theoretical predictions should be confirmed by clinical data before an interaction can be assumed inevitable. For example, carbamazepine is a well-known substrate and inducer of 3A4. When SJW was given for 14 days to patients previously stabilized on carbamazepine, no effect of SJW on carbamazepine kinetics or drug levels was observed. 74 This suggests that close attention must be paid to the precise metabolic pathways involved for each specific drug and to the associated effects on induction or inhibition of P450, enzymes, transferases, and transporters. Unfortunately, older drugs were not always well characterized by their manufacturers in terms of their interaction with the P450 metabolizing enzymes, leading to obvious problems for prediction and management of metabolic interactions.

Proposed coadministration should also consider different temporal patterns of combining herb and pharmaceutical agents. Three alternative scenarios are possible. First, adding an inducer (SJW) to a substrate (drug) will induce a lowering of previously stable drug levels over 1 to 2 weeks through increased drug metabolism, risking consequent loss of therapeutic efficacy. Moreover, in the case of SJW, initial inhibition may complicate this pattern, creating an apparent biphasic effect. Second, if the substrate (drug) is added to inducer (SJW), standard drug-dosing levels may be inadequate and may result in failure of therapy. Notably, this would not apply to drugs whose level is established by monitoring and titration to a therapeutic endpoint (e.g., coumarin/INR value). Third, withdrawal of an inducer (SJW) from a regimen of previously stable coadministration with a substrate drug will reverse induction and possibly cause rebound toxicity from elevated drug levels. Theoretically, this series of patterns would be “reversed” if the drug concerned was a prodrug, depending on activation for the metabolic transformation by the CYP450 induced. Armstrong et al. 75 well describe this schema of possible pharmacokinetic interaction patterns among inducers, inhibitors, and substrates of CYP450 drug-metabolizing enzymes.

In summary, if appropriate data about metabolic pathways of a drug are available, the pharmacokinetics of any drug proposed for coadministration with SJW should be reviewed before prescription and, wherever possible, drugs metabolized by multiple routes selected. If this is not possible, and if compelling reasons exist for coadministration of the herb with the drug, precautionary measures should be adopted; this is mandatory for any drug with narrow therapeutic indices. Introduction or cessation of SJW should be ramped or tapered, respectively, and serum levels of the pharmaceutical need to be monitored to titrate drug levels and thus counter increased clearance rates. When factors such as financial cost or intermediate metabolite toxicity militate against compensatory increases in drug doses, avoidance of coadministration is the optimum management solution.

The literature on SJW interactions continues to expand, with persistent calls in secondary sources for large-scale in vitro screening of herbs to establish the “risk” of potential (pharmacokinetic) interactions with drugs. These calls ignore that drug disposition is unpredictably mediated by a wide variety of dietary 62 compounds, foods, herbs, beverages, and lifestyle products and also affected by a wide range of individual variables, from genomics through biological, lifestyle, and socioeconomic factors, all of which render meaningful screening virtually impossible.

One study analyzing responses of six different ethnic groups to SJW did not uncover significant differences in induction effects on CYP3A4 and P-gp. 76 However, Gurley et al. 77 examined CYP450 phenotypes in elderly versus younger subjects and found age-related differences in responsiveness to botanical agents regarding CYP3A4 induction, concluding that population vulnerabilities may exist in elders. The results of in vitro tests are often contradictory and may be at odds with clinical reality because of the inherent differences between experimental systems and the in vivo complexities of herbal administration; therefore these tests have limited predictive value. Butterweck et al. pointed out that logically, systematic screening for pharmacokinetic interactions should first be applied to narrow-therapeutic-index drugs. 78,79

Some argue that understanding and managing variability in drug responses would be better than scaremongering about overstated adverse effects of herbs. 80,81More recent mainstream papers suggest that the emphasis is beginning to shift in a more constructive direction. 62,82Equally, the development of “low-hyperforin” extracts of SJW may provide efficacy in antidepressant indications without invoking PXR-mediated downstream effects on drug disposition. 34,83However, hyperforin confers numerous other properties on SJW whole-plant extracts, including anti-inflammatory, antitumor, and antiangiogenic effects. 84

Pharmacodynamic Interactions

In addition to pharmacokinetic interactions, pharmacodynamic interactions based on the antidepressant activity of SJW have been widely suggested, principally when combined with the SSRI antidepressants. The evidence for pharmacodynamic interactions is more problematic than that supporting the metabolic interactions, partly related to the general unreliability of SJW case reports, as previously noted. 11,14Qualitative data sources such as postal surveys of psychiatrists have been used to suggest adverse reports and interactions that in effect are unassessable. 85 Safety and efficacy data from clinical trials of SJW suggest that adverse effects of the herb are an order of magnitude less (1%-3%) than those of pharmaceutical antidepressants. 86 Despite the known interactions issues, SJW remains a first-line treatment for mild to moderate depression in Europe. 87 Significantly, the adverse effect data from clinical trials of the herb suggest a completely different profile of adverse effects than with common antidepressant drugs. This correlates with current understanding of the underlying mechanisms of SJW's observed antidepressant effects. The herb is now believed to work through novel and apparently complex mechanisms, dissimilar to those of known pharmaceutical antidepressants.

Initial research presumed a typical druglike biogenic amine mechanism for SJW, but early in vitro data suggesting MAOI activity have not been substantiated by in vivo studies. Reports of hypertensive MAOI-SJW interactions are lacking, as are reliable reports of interactions between SJW and tyramine-containing food substances 88,89(see also Theoretical, Speculative, and Preliminary Interactions Research later). Extensive research in vitro and on animals has examined the effects of both full-spectrum SJW extracts and isolated constituents on neurotransmitter uptake for serotonin, dopamine, noradrenaline, gamma-aminobutyric acid (GABA), andL-glutamate. 90-100The emerging conclusion is that the phloroglucinol-derivative hyperforin acts as a synaptosomal uptake inhibitor for all five of these neurotransmitters. Müller 97 has described this effect as “broad-band” reuptake inhibition. The molecular mechanism of the “pseudo-nonselective reuptake” effect is thought to be related to activation by hyperforin of a sodium ion channel that causes an increase in intracellular sodium content, modifying the sodium gradient that is the common basis of all neuronal neurotransmitter transport proteins. 90,92,99,100

Although hyperforin appears to be unique in having an approximately equal inhibitory effect on all five neurotransmitters, its effects also are at least an order of magnitude less than that of pharmaceutical antidepressants when quantified in vitro. 97 The improbability of achieving in vivo concentrations of hyperforin from oral SJW consumption that could correspond to effects of synthetic neurotransmitter uptake inhibitors is rarely considered when suggestions of “serotonin syndrome” are made relating to SJW interactions. 89 Serotonin syndrome, first characterized by Sternbach 101 in 1991, was initially described as the result of the adverse interaction of SSRIs with MAOI drugs. The clinical concept of serotonin syndrome has been overused and frequently misapplied in the drug interactions literature. 102 The concept has been reviewed and revised by Radomski et al., 103 who found a high level of misdiagnosis and distinguished several subsets of the serotonin syndrome based on symptom severity, from transient mild symptoms to fatal toxic states. The latter must also be differentiated from neuroleptic malignant syndrome. 104

Although hyperforin appears to be the only constituent that can affect uptake of all five neurotransmitters, it cannot be considered responsible for all the observed antidepressant effects of SJW. In some animal behavioral models of depression (e.g., the Porsolt test), hyperforin-free extracts exhibited significant activity, suggesting that other constituents have an effect. Clinical trials with a low-hyperforin extract also demonstrated antidepressant activity against placebo, fluoxetine, and imipramine. 105,106Furthermore, methodological controversy continues to surround clinical trials comparing SJW with placebo and pharmaceutical antidepressants, particularly because of the well-documented, powerful placebo responses associated with these trials. 107-109

Despite the absence of definitive understanding of the mechanism of SJW antidepressant activity, caution regarding potential interactions with pharmaceutical antidepressants is more than warranted. Also, several classes of psychiatric drugs are substrates or inhibitors of CYP3A4 and P-gp, suggesting combined pharmacokinetic and pharmacodynamic interactions with SJW. Common agents likely to be encountered in general and psychiatric practice include triazolobenzodiazepines (alprazolam, estazolam, midazolam, triazolam), which are substrates of 3A4, as are the nonbenzodiazepine hypnotics zolpidem and zaleplon and the “atypical” anxiolytic buspirone.

herb-drug interactions
Alprazolam, Midazolam, and Related Triazolobenzodiazepines
Amitriptyline and Related Tertiary Tricyclic Antidepressants
Anesthesia, General
Antiretrovirals: Protease Inhibitors and Nonnucleoside Reverse-Transcriptase Inhibitors
Cyclosporine
Digoxin, Digitoxin, and Related Cardiac Glycosides
Etoposide and Related Topoisomerase II Inhibitors
Fexofenadine
Imatinib
Irinotecan
Omeprazole and Related Proton Pump Inhibitors
Oral Contraceptives and Related Estrogen-Containing and Synthetic Estrogen and Progesterone Analog Medications
Paclitaxel, Docetaxel: Taxane Microtubule-Stabilizing Agents
Paroxetine and Related Selective Serotonin Reuptake Inhibitor and Serotonin-Norepinephrine Reuptake Inhibitor (SSRI and SSRI/SNRI) Antidepressants and Nonselective Serotonin Reuptake Inhibitors (NSRIs)
Simvastatin and Related HMG-COA Reductase Inhibitors (Statins)
Tacrolimus
Verapamil and Related Calcium Channel Blockers
Voriconazole and Related Triazole Antifungal Agents
  • Evidence: Voriconazole (Vfend).
  • Extrapolated, based on similar properties: Fluconazole (Diflucan), itraconazole (Sporanox), posaconazole (Noxafil).
Potential or Theoretical Adverse Interaction of Uncertain Severity
Impaired Drug Absorption and Bioavailability, Precautions Appropriate

Probability: 2. Probable
Evidence Base: Preliminary

Effect and Mechanism of Action

A pharmacokinetic interaction between SJW and voriconazole, which are metabolized by CYP450 3A4, 2C9, and 2C19, resulting in a lowering of drug levels due to enzyme induction by SJW. The interaction is experimentally confirmed, but clinical reports are lacking to date.

Research

Rengelshausen et al. 37 examined the disposition of single doses of voriconazole in 16 healthy male volunteers stratified by CYP2C19 genotype on day 1 and day 15 of concomitant SJW administration (300 mg three times daily, extract LI160). They found an overall decrease in AUC at day 15 of up to 59%, broadly equivalent to the effect of SJW on tacrolimus and cyclosporine. In addition, they found that 2C19 wild type (extensive metabolizer) exhibited the lowest exposure to the antifungal drug. The clearance data revealed an initial but insignificant increase in plasma levels of the drug after onset of SJW administration. This is a predictable result of the “biphasic” effect of SJW flavonoids mechanistically inhibiting CYP450, followed by a more potent effect of induction (see Effects on Drug Metabolism and Bioavailability). 38

Clinical Implications and Adaptations

The novel antifungal voriconazole is used in patients with invasive aspergillosis and other serious fungal invasions. Therapeutic dosing is critical and, in SJW coadministration, should be avoided. The manufacturer's data on voriconazole do not recommend 2C19 phenotyping, although the wild-type polymorphism would appear to be at risk for lowered drug exposure.

Warfarin and Oral Vitamin K Antagonist Anticoagulants
theoretical, speculative, and preliminary interactions research, including overstated interactions claims
Buspirone
Chlorzoxazone
Loperamide
Monoamine Oxidase (MAO) Inhibitors
Photosensitizing Agents
Theophylline/Aminophylline
Citations
  • 1.Gerard J. The Herbal. Revised and Enlarged. In: Johnson T, ed. 1975 ed. New York: Dover Publications; 1633.
  • 2.St. John’s wort. In: Blumenthal M, Busse W, Goldberg A et al. The Complete German Commission E Monographs. Austin, Texas: American Botanical Council: Integrative Medicine Communications; 1998:214-215.
  • 3.Upton R. St. John’s wort. American Herbal Pharmacopoeia. Scotts Valley, Calif; 1997.View Abstract
  • 4.ESCOP. Hyperici herba. ESCOP Monographs: the Scientific Foundation for Herbal Medicinal Products. 2nd ed. Exeter, UK: European Scientific Cooperative on Phytotherapy and Thieme; 2003:257-281.
  • 5.McKenna D, Jones K, Hughes K, Humphrey S. St John’s wort. Botanical Medicines. 2nd ed. Binghamton, NY: Haworth Press; 2002:923-986.
  • 6.Weiss R. Herbal Medicine. Meuss A, Translator. 6th ed. Beaconsfield, UK: Beaconsfield Publishers Ltd; 1988.
  • 7.Johne A, Brockmoller J, Bauer S et al. Pharmacokinetic interaction of digoxin with an herbal extract from St John’s wort (Hypericum perforatum). Clin Pharmacol Ther 1999;66:338-345.
  • 8.Ruschitzka F, Meier PJ, Turina M et al. Acute heart transplant rejection due to Saint John’s wort [letter]. Lancet 2000;355:548-549.
  • 9.Piscitelli SC, Burstein AH, Chaitt D et al. Indinavir concentrations and St John’s wort [letter]. Lancet 2000;355:547-548.
  • 10.Ernst E. Second thoughts about safety of St John’s wort. Lancet 1999;354:2014-2016.
  • 11.Fugh-Berman A, Ernst E. Herb-drug interactions: review and assessment of report reliability. Br J Clin Pharmacol 2001;52:587-595.View Abstract
  • 12.Meyer JR, Generali JA, Karpinski JL. Evaluation of herbal-drug interaction data in tertiary resources. Hosp Pharm 2004;39:149-160.
  • 13.Izzo AA. Drug interactions with St. John’s wort (Hypericum perforatum): a review of the clinical evidence. Int J Clin Pharmacol Ther 2004;42:139-148.View Abstract
  • 14.Mills E, Montori VM, Wu P et al. Interaction of St John’s wort with conventional drugs: systematic review of clinical trials. BMJ 2004;329:27-30.
  • 15.US Food and Drug Administration (FDA). Guidelines for industry: in vivo drug metabolism/drug interaction studies—study design, data analysis and recommendations for dosing and labeling. 1999 vol: www.fda.gov/cber/gdlns/metabol.pdf; 7/4/04.
  • 16.Committee on Safety of Medicines (CSM). Message from Professor A Breckenridge, Chairman, CSM (UK); February 2000.
  • 17.Irish Medical Board (IMB). Herbal Medicines Project: final report; 2001. http://www.euroherb.com/.
  • 18.Lumpkin M, Alpert S. Risk of drug interactions with St. John’s Wort and indinavir and other drugs. FDA Public Health Advisory. CDER; 2000. http://www.fda.gov/cder/drug/advisory/stjwort.htm.
  • 19.Obach RS. Inhibition of human cytochrome P450 enzymes by constituents of St. John’s wort, an herbal preparation used in the treatment of depression. J Pharmacol Exp Ther 2000;294:88-95.View Abstract
  • 20.Budzinski JW, Foster BC, Vandenhoek S, Arnason JT. An in vitro evaluation of human cytochrome P450 3A4 inhibition by selected commercial herbal extracts and tinctures. Phytomedicine 2000;7:273-282.View Abstract
  • 21.Roby CA, Anderson GD, Kantor E et al. St John’s wort: effect on CYP3A4 activity. Clin Pharmacol Ther 2000;67:451-457.
  • 22.Gurley BJ, Gardner SF, Hubbard MA et al. Cytochrome P450 phenotypic ratios for predicting herb-drug interactions in humans. Clin Pharmacol Ther 2002;72:276-287.View Abstract
  • 23.Wang L-S, Zhou G, Zhu B et al. St John’s wort induces both cytochrome P450 3A4-catalyzed sulfoxidation and 2C19-dependent hydroxylation of omeprazole. Clin Pharmacol Ther 2004;75:191-197.
  • 24.Markowitz JS, DeVane CL, Boulton DW et al. Effect of St. John’s wort (Hypericum perforatum) on cytochrome P-450 2D6 and 3A4 activity in healthy volunteers. Life Sci 2000;66:L133-L139.View Abstract
  • 25.Gerwertz N, Ereshefsky B, Lam Y et al. Determination of SJW differential metabolism at CYP2D6 and CYP3A4 using dextromethorphan probe technology. Abstracts from the 39th Annual Meeting, New Clinical Drug Evaluation Unit. Boca Raton, Fla; 1999:130.
  • 26.Wang Z, Gorski JC, Hamman MA et al. The effects of St John’s wort (Hypericum perforatum) on human cytochrome P450 activity. Clin Pharmacol Ther 2001;70:317-326.
  • 27.Dresser GK, Schwarz UI, Wilkinson GR, Kim RB. Coordinate induction of both cytochrome P4503A and MDR1 by St John’s wort in healthy subjects. Clin Pharmacol Ther 2003;73:41-50.
  • 28.Durr D, Stieger B, Kullak-Ublick GA et al. St John’s wort induces intestinal P-glycoprotein/MDR1 and intestinal and hepatic CYP3A4. Clin Pharmacol Ther 2000;68:598-604.
  • 29.Wenk M, Todesco L, Krahenbuhl S. Effect of St John’s wort on the activities of CYP1A2, CYP3A4, CYP2D6, N-acetyltransferase 2, and xanthine oxidase in healthy males and females. Br J Clin Pharmacol 2004;57:495-499.
  • 30.Wang EJ, Barecki-Roach M, Johnson WW. Quantitative characterization of direct P-glycoprotein inhibition by St John’s wort constituents hypericin and hyperforin. J Pharm Pharmacol 2004;56:123-128.
  • 31.Markowitz JS, Donovan JL, DeVane CL et al. Effect of St. John’s wort on drug metabolism by induction of cytochrome p450 3A4 enzyme. Obstet Gynecol Surv 2004;59:358-359.View Abstract
  • 32.Komoroski BJ, Zhang S, Cai H et al. Induction and inhibition of cytochromes p450 by the St. John’s wort constituent hyperforin in human hepatocyte cultures. Drug Metab Dispos 2004;32:512-518.View Abstract
  • 33.Chen Y, Ferguson SS, Negishi M, Goldstein JA. Induction of human CYP2C9 by rifampicin, hyperforin, and phenobarbital is mediated by the pregnane X receptor. J Pharmacol Exp Ther 2004;308:495-501.
  • 34.Madabushi R, Frank B, Drewelow B et al. Hyperforin in St. John’s wort drug interactions. Eur J Clin Pharmacol 2006:1-9.View Abstract
  • 35.Bailey DG, Dresser GK, Bend JR. Bergamottin, lime juice, and red wine as inhibitors of cytochrome P450 3A4 activity: comparison with grapefruit juice. Clin Pharmacol Ther 2003;73:529-537.View Abstract
  • 36.Huang SM, Hall SD, Watkins P et al. Drug interactions with herbal products and grapefruit juice: a conference report. Clin Pharmacol Ther 2004;75:1-12.View Abstract
  • 37.Rengelshausen J, Banfield M, Riedel KD et al. Opposite effects of short-term and long-term St John’s wort intake on voriconazole pharmacokinetics. Clin Pharmacol Ther 2005;78:25-33.
  • 38.Xie HG, Kim RB. St John’s wort-associated drug interactions: short-term inhibition and long-term induction? Clin Pharmacol Ther 2005;78:19-24.
  • 39.Moore LB, Goodwin B, Jones SA et al. St. John’s wort induces hepatic drug metabolism through activation of the pregnane X receptor. Proc Natl Acad Sci U S A 2000;97:7500-7502.View Abstract
  • 40.Watkins RE, Maglich JM, Moore LB et al. 2.1 A crystal structure of human PXR in complex with the St. John’s wort compound hyperforin. Biochemistry 2003;42:1430-1438.View Abstract
  • 41.Jones SA, Moore LB, Shenk JL et al. The pregnane X receptor: a promiscuous xenobiotic receptor that has diverged during evolution. Mol Endocrinol 2000;14:27-39.
  • 42.Kliewer SA. The nuclear pregnane X receptor regulates xenobiotic detoxification. J Nutr 2003;133:2444S-2447S.View Abstract
  • 43.Waxman DJ. P450 gene induction by structurally diverse xenochemicals: central role of nuclear receptors CAR, PXR, and PPAR. Arch Biochem Biophys 1999;369:11-23.View Abstract
  • 44.Schwarz D, Kisselev P, Roots I. St. John’s wort extracts and some of their constituents potently inhibit ultimate carcinogen formation from benzo[a]pyrene-7,8-dihydrodiol by human CYP1A1. Cancer Res 2003;63:8062-8068.View Abstract
  • 45.Brockmoller J, Kirchheiner J, Meisel C, Roots I. Pharmacogenetic diagnostics of cytochrome P450 polymorphisms in clinical drug development and in drug treatment. Pharmacogenomics 2000;1:125-151.View Abstract
  • 46.Dorne JLCM, Walton K, Renwick AG. Human variability for metabolic pathways with limited data (CYP2A6, CYP2C9, CYP2E1, ADH, esterases, glycine and sulphate conjugation). Food Chem Toxicol 2004;42:397-421.
  • 47.Foster BC, Vandenhoek S, Hana J et al. In vitro inhibition of human cytochrome P450–mediated metabolism of marker substrates by natural products. Phytomedicine 2003;10:334-342.View Abstract
  • 48.Henderson L, Yue QY, Bergquist C et al. St John’s wort (Hypericum perforatum): drug interactions and clinical outcomes. Br J Clin Pharmacol 2002;54:349-356.
  • 49.Hennessy M, Kelleher D, Spiers JP et al. St John’s wort increases expression of P-glycoprotein: implications for drug interactions. Br J Clin Pharmacol 2002;53:75-82.
  • 50.Dresser GK, McDonald W, Kim RB, Bailey DG. Evaluation of herbal products as potential inhibitors of MDR1. Clin Pharmacol Ther 2004;75:P79.
  • 51.Zhou S, Lim LY, Chowbay B. Herbal modulation of P-glycoprotein. Drug Metab Rev 2004;36:57-104.View Abstract
  • 52.Traber MG. Vitamin E, nuclear receptors and xenobiotic metabolism. Arch Biochem Biophys 2004;423:6-11.View Abstract
  • 53.Maglich JM, Stoltz CM, Goodwin B et al. Nuclear pregnane X receptor and constitutive androstane receptor regulate overlapping but distinct sets of genes involved in xenobiotic detoxification. Mol Pharmacol 2002;62:638-646.View Abstract
  • 54.Moore LB, Maglich JM, McKee DD et al. Pregnane X receptor (PXR), constitutive androstane receptor (CAR), and benzoate X receptor (BXR) define three pharmacologically distinct classes of nuclear receptors. Mol Endocrinol 2002;16:977-986.View Abstract
  • 55.Moore JT, Kliewer SA. Use of the nuclear receptor PXR to predict drug interactions. Toxicology 2000;153:1-10.View Abstract
  • 56.Moore LB, Parks DJ, Jones SA et al. Orphan nuclear receptors constitutive androstane receptor and pregnane X receptor share xenobiotic and steroid ligands. J Biol Chem 2000;275:15122-15127.View Abstract
  • 57.Forman BM. Polymorphisms in promiscuous PXR: an explanation for interindividual differences in drug clearance? Pharmacogenetics 2001;11:551-552.
  • 58.Lamba J, Lamba V, Schuetz E. Genetic variants of PXR (NR1I2) and CAR (NR1I3) and their implications in drug metabolism and pharmacogenetics. Curr Drug Metab 2005;6:369-383.View Abstract
  • 59.Pascussi JM, Gerbal-Chaloin S, Drocourt L et al. The expression of CYP2B6, CYP2C9 and CYP3A4 genes: a tangle of networks of nuclear and steroid receptors. Biochim Biophys Acta 2003;1619:243-253.
  • 60.Treasure JE. Warding off evil in the 21st century: St. John’s wort as a xenosensory activator? J Am Herbalists Guild 2005;6:48-51.
  • 61.Janosek J, Hilscherova K, Blaha L, Holoubek I. Environmental xenobiotics and nuclear receptors: interactions, effects and in vitro assessment. Toxicol In Vitro 2006;20:18-37.View Abstract
  • 62.Pal D, Mitra AK. MDR- and CYP3A4-mediated drug-herbal interactions. Life Sci 2006;78:2131-2145.View Abstract
  • 63.Ernst E. St John’s wort supplements endanger the success of organ transplantation. Arch Surg 2002;137:316-319.
  • 64.Dresser GK, Schwarz UI, Wilkinson GR, Kim RB. Coordinate induction of both cytochrome P4503A and MDR1 by St John’s wort in healthy subjects. Clin Pharmacol Ther 2003;73:41-50.
  • 65.Lin JH. Drug-drug interaction mediated by inhibition and induction of P-glycoprotein. Adv Drug Deliv Rev 2003;55:53-81.View Abstract
  • 66.Lin JH, Yamazaki M. Role of P-glycoprotein in pharmacokinetics: clinical implications. Clin Pharmacokinet 2003;42:59-98.View Abstract
  • 67.Di Marco MP, Edwards DJ, Wainer IW, Ducharme MP. The effect of grapefruit juice and seville orange juice on the pharmacokinetics of dextromethorphan: the role of gut CYP3A and P-glycoprotein. Life Sci 2002;71:1149-1160.View Abstract
  • 68.Bhardwaj RK, Glaeser H, Becquemont L et al. Piperine, a major constituent of black pepper, inhibits human P-glycoprotein and CYP3A4. J Pharmacol Exp Ther 2002;302:645-650.View Abstract
  • 69.Benet LZ, Cummins CL. The drug efflux-metabolism alliance: biochemical aspects. Adv Drug Deliv Rev 2001;50 Suppl 1:S3-S11.View Abstract
  • 70.Cummins CL, Jacobsen W, Benet LZ. Unmasking the dynamic interplay between intestinal P-glycoprotein and CYP3A4. J Pharmacol Exp Ther 2002;300:1036-1045.
  • 71.Cummins CL, Wu CY, Benet LZ. Sex-related differences in the clearance of cytochrome P450 3A4 substrates may be caused by P-glycoprotein. Clin Pharmacol Ther 2002;72:474-489.View Abstract
  • 72.Backman JT, Olkkola KT, Neuvonen PJ. Rifampin drastically reduces plasma concentrations and effects of oral midazolam. Clin Pharmacol Ther 1996;59:7-13.View Abstract
  • 73.Tsunoda SM, Harris RZ, Christians U et al. Red wine decreases cyclosporine bioavailability. Clin Pharmacol Ther 2001;70:462-467.View Abstract
  • 74.Burstein AH, Horton RL, Dunn T et al. Lack of effect of St John’s wort on carbamazepine pharmacokinetics in healthy volunteers. Clin Pharmacol Ther 2000;68:605-612.
  • 75.Armstrong SC, Cozza KL, Sandson NB. Six patterns of drug-drug interactions. Psychosomatics 2003;44:255-258.View Abstract
  • 76.Xie R, Tan LH, Polasek EC et al. CYP3A and P-glycoprotein activity induction with St. John’s wort in healthy volunteers from 6 ethnic populations. J Clin Pharmacol 2005;45:352-356.View Abstract
  • 77.Gurley BJ, Gardner SF, Hubbard MA et al. Clinical assessment of effects of botanical supplementation on cytochrome P450 phenotypes in the elderly: St John’s wort, garlic oil, Panax ginseng and Ginkgo biloba. Drugs Aging 2005;22:525-539.
  • 78.Izzo AA, Ernst E. Interactions between herbal medicines and prescribed drugs: a systematic review. Drugs 2001;61:2163-2175.
  • 79.Butterweck V, Derendorf H, Gaus W et al. Pharmacokinetic herb-drug interactions: are preventive screenings necessary and appropriate? Planta Med 2004;70:784-791.
  • 80.Williamson EM. Interactions between herbal and conventional medicines. Expert Opin Drug Saf 2005;4:355-378.
  • 81.Treasure JE. MEDLINE and the mainstream manufacture of misinformation. J Am Herbalists Guild 2006;6:50-56.
  • 82.Singh YN. Potential for interaction of kava and St. John’s wort with drugs. J Ethnopharmacol 2005;100:108-113.View Abstract
  • 83.Arold G, Donath F, Maurer A et al. No relevant interaction with alprazolam, caffeine, tolbutamide, and digoxin by treatment with a low-hyperforin St John’s wort extract. Planta Med 2005;71:331-337.
  • 84.Medina MA, Martinez-Poveda B, Amores-Sanchez MI, Quesada AR. Hyperforin: more than an antidepressant bioactive compound? Life Sci 2006;79:105-111.
  • 85.Walter G, Rey JM, Harding A. Psychiatrists’ experience and views regarding St John’s wort and “alternative” treatments. Aust N Z J Psychiatry 2000;34:992-996.
  • 86.Schulz V. Incidence and clinical relevance of the interactions and side effects of Hypericum preparations. Phytomedicine 2001;8:152-160.View Abstract
  • 87.Di Carlo G, Borrelli F, Izzo AA, Ernst E. St John’s wort: Prozac from the plant kingdom. Trends Pharmacol Sci 2001;22:292-297.
  • 88.Bladt S, Wagner H. Inhibition of MAO by fractions and constituents of Hypericum extract. J Geriatr Psychiatry Neurol 1994;7 Suppl 1:S57-59.View Abstract
  • 89.Cott JM. In vitro receptor binding and enzyme inhibition by Hypericum perforatum extract. Pharmacopsychiatry 1997;30 Suppl 2:108-112.View Abstract
  • 90.Singer A, Wonnemann M, Muller WE. Hyperforin, a major antidepressant constituent of St. John’s wort, inhibits serotonin uptake by elevating free intracellular Na+1. J Pharmacol Exp Ther 1999;290:1363-1368.View Abstract
  • 91.Wonnemann M, Singer A, Siebert B, Muller WE. Evaluation of synaptosomal uptake inhibition of most relevant constituents of St. John’s wort. Pharmacopsychiatry 2001;34 Suppl 1:S148-S151.View Abstract
  • 92.Wonnemann M, Singer A, Muller WE. Inhibition of synaptosomal uptake of 3H-l-glutamate and 3H-GABA by hyperforin, a major constituent of St. John’s wort: the role of amiloride sensitive sodium conductive pathways. Neuropsychopharmacology 2000;23:188-197.View Abstract
  • 93.Simmen U, Burkard W, Berger K et al. Extracts and constituents of Hypericum perforatum inhibit the binding of various ligands to recombinant receptors expressed with the Semliki Forest virus system. J Recept Signal Transduct Res 1999;19:59-74.
  • 94.Muller WE, Singer A, Wonnemann M. Hyperforin: antidepressant activity by a novel mechanism of action. Pharmacopsychiatry 2001;34 Suppl 1:S98-S102.View Abstract
  • 95.Eckert GP, Muller WE. Effects of hyperforin on the fluidity of brain membranes. Pharmacopsychiatry 2001;34 Suppl 1:S22-S25.View Abstract
  • 96.Gobbi M, Moia M, Pirona L et al. In vitro binding studies with two Hypericum perforatum extracts—hyperforin, hypericin and biapigenin—on 5-HT6, 5-HT7, GABA(A)/benzodiazepine, sigma, NPY-Y1/Y2 receptors and dopamine transporters. Pharmacopsychiatry 2001;34 Suppl 1:S45-S48.View Abstract
  • 97.Muller WE. Current St John’s wort research from mode of action to clinical efficacy. Pharmacol Res 2003;47:101-109.
  • 98.Mennini T, Gobbi M. The antidepressant mechanism of Hypericum perforatum. Life Sci 2004;75:1021-1027.View Abstract
  • 99.Koch E, Chatterjee SS. Hyperforin stimulates intracellular calcium mobilisation and enhances extracellular acidification in DDT1-MF2 smooth muscle cells. Pharmacopsychiatry 2001;34 Suppl 1:S70-S73.
  • 100.Krishtal O, Lozovaya N, Fisunov A et al. Modulation of ion channels in rat neurons by the constituents of Hypericum perforatum. Pharmacopsychiatry 2001;34 Suppl 1:S74-S82.View Abstract
  • 101.Sternbach H. The serotonin syndrome. Am J Psychiatry 1991;148:705-713.View Abstract
  • 102.Hilton SE, Maradit H, Moller HJ. Serotonin syndrome and drug combinations: focus on MAOI and RIMA. Eur Arch Psychiatry Clin Neurosci 1997;247:113-119.View Abstract
  • 103.Radomski JW, Dursun SM, Reveley MA, Kutcher SP. An exploratory approach to the serotonin syndrome: an update of clinical phenomenology and revised diagnostic criteria. Med Hypotheses 2000;55:218-224.View Abstract
  • 104.Von Halling Laier MG, Gram LF. [Serotonin syndrome and malignant neuroleptic syndrome: a review based on the material from the National Board of Adverse Drug Reactions]. Ugeskr Laeger 1996;158:6933-6937.View Abstract
  • 105.Woelk H. Comparison of St John’s wort and imipramine for treating depression: randomised controlled trial. BMJ 2000;321:536-539.
  • 106.Schrader E. Equivalence of St John’s wort extract (Ze 117) and fluoxetine: a randomized, controlled study in mild-moderate depression. Int Clin Psychopharmacol 2000;15:61-68.
  • 107.Cott JM, Rosenthal N, Blumenthal M. St John’s wort and major depression. JAMA 2001;286:42; author reply 44-45.
  • 108.Walsh BT, Seidman SN, Sysko R, Gould M. Placebo response in studies of major depression: variable, substantial, and growing. JAMA 2002;287:1840-1847.View Abstract
  • 109.Barbui C, Cipriani A, Brambilla P, Hotopf M. “Wish bias” in antidepressant drug trials? J Clin Psychopharmacol 2004;24:126-130.
  • 110.Markowitz JS, Donovan JL, DeVane CL et al. Effect of St John’s wort on drug metabolism by induction of cytochrome P450 3A4 enzyme. JAMA 2003;290:1500-1504.
  • 111.Stockley I. Stockley’s Drug Interactions. 6th ed. London: Pharmaceutical Press; 2002.
  • 112.Johne A, Schmider J, Brockmoller J et al. Decreased plasma levels of amitriptyline and its metabolites on comedication with an extract from St. John’s wort (Hypericum perforatum). J Clin Psychopharmacol 2002;22:46-54.View Abstract
  • 113.Jakovljevic V, Popovic M, Mimica-Dukic N et al. Pharmacodynamic study of Hypericum perforatum L. Phytomedicine 2000;7:449-453.View Abstract
  • 114.Irefin S, Sprung J. A possible cause of cardiovascular collapse during anesthesia: long-term use of St. John’s wort. J Clin Anesth 2000;12:498-499.
  • 115.Choo EF, Leake B, Wandel C et al. Pharmacological inhibition of P-glycoprotein transport enhances the distribution of HIV-1 protease inhibitors into brain and testes. Drug Metab Dispos 2000;28:655-660.View Abstract
  • 116.Huang L, Wring SA, Woolley JL et al. Induction of P-glycoprotein and cytochrome P450 3A by HIV protease inhibitors. Drug Metab Dispos 2001;29:754-760.View Abstract
  • 117.De Maat MM, Hoetelmans RM, Math RA et al. Drug interaction between St John’s wort and nevirapine. AIDS 2001;15:420-421.
  • 118.Stebbing J, Bower M. Comparative pharmacogenomics of antiretroviral and cytotoxic treatments. Lancet Oncol 2006;7:61-68.View Abstract
  • 119.Hebert MF. Contributions of hepatic and intestinal metabolism and P-glycoprotein to cyclosporine and tacrolimus oral drug delivery. Adv Drug Deliv Rev 1997;27:201-214.View Abstract
  • 120.Kwei GY, Alvaro RF, Chen Q et al. Disposition of ivermectin and cyclosporin A in CF-1 mice deficient in MDR1a P-glycoprotein. Drug Metab Dispos 1999;27:581-587.
  • 121.Wu CY, Benet LZ, Hebert MF et al. Differentiation of absorption and first-pass gut and hepatic metabolism in humans: studies with cyclosporine. Clin Pharmacol Ther 1995;58:492-497.View Abstract
  • 122.Bauer S, Stormer E, Johne A et al. Alterations in cyclosporin A pharmacokinetics and metabolism during treatment with St John’s wort in renal transplant patients. Br J Clin Pharmacol 2003;55:203-211.
  • 123.Mandelbaum A, Pertzborn F, Martin-Facklam M, Wiesel M. Unexplained decrease of cyclosporin trough levels in a compliant renal transplant patient. Nephrol Dial Transplant 2000;15:1473-1474.View Abstract
  • 124.Mai I, Kruger H, Budde K et al. Hazardous pharmacokinetic interaction of Saint John’s wort (Hypericum perforatum) with the immunosuppressant cyclosporin. Int J Clin Pharmacol Ther 2000;38:500-502.
  • 125.Breidenbach T, Kliem V, Burg M et al. Profound drop of cyclosporin A whole blood trough levels caused by St. John’s wort (Hypericum perforatum). Transplantation 2000;69:2229-2230.View Abstract
  • 126.Barone GW, Gurley BJ, Ketel BL et al. Drug interaction between St. John’s wort and cyclosporine. Ann Pharmacother 2000;34:1013-1016.
  • 127.Karliova M, Treichel U, Malago M et al. Interaction of Hypericum perforatum (St. John’s wort) with cyclosporin A metabolism in a patient after liver transplantation. J Hepatol 2000;33:853-855.View Abstract
  • 128.Mueller SC, Uehleke B, Woehling H et al. Effect of St John’s wort dose and preparations on the pharmacokinetics of digoxin. Clin Pharmacol Ther 2004;75:546-557.
  • 129.Kullak-Ublick GA, Ismair MG, Stieger B et al. Organic anion-transporting polypeptide B (OATP-B) and its functional comparison with three other OATPs of human liver. Gastroenterology 2001;120:525-533.View Abstract
  • 130.Dresser GK, Bailey DG, Leake BF et al. Fruit juices inhibit organic anion transporting polypeptide–mediated drug uptake to decrease the oral availability of fexofenadine. Clin Pharmacol Ther 2002;71:11-20.View Abstract
  • 131.Mikkaichi T, Suzuki T, Onogawa T et al. Isolation and characterization of a digoxin transporter and its rat homologue expressed in the kidney. Proc Natl Acad Sci USA 2004;101:3569-3574.View Abstract
  • 132.Andelic S. [Bigeminy: the result of interaction between digoxin and St. John’s wort]. Vojnosanit Pregl 2003;60:361-364.
  • 133.Dresser GK, Bailey DG. The effects of fruit juices on drug disposition: a new model for drug interactions. Eur J Clin Invest 2003;33 Suppl 2:10-16.View Abstract
  • 134.Peebles KA, Baker RK, Kurz EU et al. Catalytic inhibition of human DNA topoisomerase IIα by hypericin, a naphthodianthrone from St. John’s wort (Hypericum perforatum). Biochem Pharmacol 2001;62:1059-1070.View Abstract
  • 135.Block KI, Gyllenhaal C. Clinical corner. Herb-drug interactions in cancer chemotherapy: theoretical concerns regarding drug metabolizing enzymes. Integr Cancer Ther 2002;1:83-89.View Abstract
  • 136.Wang Z, Hamman MA, Huang SM et al. Effect of St John’s wort on the pharmacokinetics of fexofenadine. Clin Pharmacol Ther 2002;71:414-420.
  • 137.Frye RF, Fitzgerald SM, Lagattuta TF, Egorin MJ. Effect of St. John’s wort on imatinib mesylate pharmacokinetics. Clin Pharmacol Ther 2004;75:P96.View Abstract
  • 138.Rochat B. Role of cytochrome P450 activity in the fate of anticancer agents and in drug resistance: focus on tamoxifen, paclitaxel and imatinib metabolism. Clin Pharmacokinet 2005;44:349-366.View Abstract
  • 139.Mathijssen RH, Loos WJ, Verweij J, Sparreboom A. Pharmacology of topoisomerase I inhibitors irinotecan (CPT-11) and topotecan. Curr Cancer Drug Targets 2002;2:103-123.View Abstract
  • 140.Mathijssen RH, van Alphen RJ, Verweij J et al. Clinical pharmacokinetics and metabolism of irinotecan (CPT-11). Clin Cancer Res 2001;7:2182-2194.View Abstract
  • 141.Rivory LP. Metabolism of CPT-11: impact on activity. Ann NY Acad Sci 2000;922:205-215.View Abstract
  • 142.Perry M. The Chemotherapy Source Book. 3rd ed. Philadelphia: Lippincott, Williams & Wilkins; 2001.
  • 143.Mansky PJ, Straus SE. St. John’s wort: more implications for cancer patients. J Natl Cancer Inst 2002;94:1187-1188.View Abstract
  • 144.Sparreboom A, Cox MC, Acharya MR, Figg WD. Herbal remedies in the United States: potential adverse interactions with anticancer agents. J Clin Oncol 2004;22:2489-2503.View Abstract
  • 145.Xie HG. Additional discussions regarding the altered metabolism and transport of omeprazole after long-term use of St John’s wort. Clin Pharmacol Ther 2005;78:440-441.
  • 146.Donovan JL, DeVane CL, Lewis JG et al. Effects of St John’s wort (Hypericum perforatum L.) extract on plasma androgen concentrations in healthy men and women: a pilot study. Phytother Res 2005;19:901-906.
  • 147.Hall SD, Wang Z, Huang SM et al. The interaction between St John’s wort and an oral contraceptive. Clin Pharmacol Ther 2003;74:525-535.
  • 148.Pfrunder A, Schiesser M, Gerber S et al. Interaction of St John’s wort with low-dose oral contraceptive therapy: a randomized controlled trial. Br J Clin Pharmacol 2003;56:683-690.
  • 149.Pregnancies prompt herb warning. BBC News; Feb 6, 2002.
  • 150.Yue QY, Bergquist C, Gerden B. Safety of St John’s wort (Hypericum perforatum). Lancet 2000;355:576-577.
  • 151.Barditch-Crovo P, Trapnell CB, Ette E et al. The effects of rifampin and rifabutin on the pharmacokinetics and pharmacodynamics of a combination oral contraceptive. Clin Pharmacol Ther 1999;65:428-438.View Abstract
  • 152.LeBel M, Masson E, Guilbert E et al. Effects of rifabutin and rifampicin on the pharmacokinetics of ethinylestradiol and norethindrone. J Clin Pharmacol 1998;38:1042-1050.View Abstract
  • 153.Komoroski BJ, Parise RA, Egorin MJ et al. Effect of the St. John’s wort constituent hyperforin on docetaxel metabolism by human hepatocyte cultures. Clin Cancer Res 2005;11:6972-6979.View Abstract
  • 154.Wada A, Sakaeda T, Takara K et al. Effects of St John’s wort and hypericin on cytotoxicity of anticancer drugs. Drug Metab Pharmacokinet 2002;17:467-474.
  • 155.Engels FK, Ten Tije AJ, Baker SD et al. Effect of cytochrome P450 3A4 inhibition on the pharmacokinetics of docetaxel. Clin Pharmacol Ther 2004;75:448-454.View Abstract
  • 156.Engels FK, Sparreboom A, Mathot RA, Verweij J. Potential for improvement of docetaxel-based chemotherapy: a pharmacological review. Br J Cancer 2005;93:173-177.View Abstract
  • 157.Cozza K, Armstrong S, Oesterheld J. Drug Interaction Principles for Medical Practice. 2nd ed. Washington, DC: American Psychiatric Publishing; 2003.
  • 158.Neuvonen PJ, Pohjola-Sintonen S, Tacke U, Vuori E. Five fatal cases of serotonin syndrome after moclobemide-citalopram or moclobemide-clomipramine overdoses. Lancet 1993;342:1419.View Abstract
  • 159.Gordon JB. SSRIs and St. John’s wort: possible toxicity? Am Fam Physician 1998;57:950, 953.View Abstract
  • 160.Nierenberg AA, Burt T, Matthews J, Weiss AP. Mania associated with St. John’s wort. Biol Psychiatry 1999;46:1707-1708.View Abstract
  • 161.Prost N, Tichadou L, Rodor F et al. [St. John’s wort–venlafaxine interaction]. Presse Med 2000;29:1285-1286.View Abstract
  • 162.Spinella M, Eaton LA. Hypomania induced by herbal and pharmaceutical psychotropic medicines following mild traumatic brain injury. Brain Inj 2002;16:359-367.View Abstract
  • 163.Sugimoto K, Ohmori M, Tsuruoka S et al. Different effects of St John’s wort on the pharmacokinetics of simvastatin and pravastatin. Clin Pharmacol Ther 2001;70:518-524.
  • 164.Bogman K, Peyer AK, Torok M et al. HMG-CoA reductase inhibitors and P-glycoprotein modulation. Br J Pharmacol 2001;132:1183-1192.View Abstract
  • 165.Bolley R, Zulke C, Kammerl M et al. Tacrolimus-induced nephrotoxicity unmasked by induction of the CYP3A4 system with St John’s wort. Transplantation 2002;73:1009.
  • 166.Hebert MF, Park JM, Chen Y-L et al. Effects of St. John’s wort (Hypericum perforatum) on tacrolimus pharmacokinetics in healthy volunteers. J Clin Pharmacol 2004;44:89-94.View Abstract
  • 167.Mai I, Stormer E, Bauer S et al. Impact of St John’s wort treatment on the pharmacokinetics of tacrolimus and mycophenolic acid in renal transplant patients. Nephrol Dial Transplant 2003;18:819-822.
  • 168.Tannergren C, Engman H, Knutson L et al. St John’s wort decreases the bioavailability of R- and S-verapamil through induction of the first-pass metabolism. Clin Pharmacol Ther 2004;75:298-309.
  • 169.Maurer A, Johne A, Bauer S et al. Interaction of St John’s Wort extract with phenprocoumon. European Journal Clinical Pharmacology 1999;55:(abstract 22).
  • 170.Keller C, Matzdorff AC, Kemkes-Matthes B. Pharmacology of warfarin and clinical implications. Semin Thromb Hemost 1999;25:13-16.View Abstract
  • 171.He M, Korzekwa KR, Jones JP et al. Structural forms of phenprocoumon and warfarin that are metabolized at the active site of CYP2C9. Arch Biochem Biophys 1999;372:16-28.
  • 172.Noldner M, Chatterjee S. Effects of two different extracts of St. John’s wort and some of their constituents on cytochrome P450 activities in rat liver microsomes. Pharmacopsychiatry 2001;34 Suppl 1:S108-S110.View Abstract
  • 173.Wadelius M, Sorlin K, Wallerman O et al. Warfarin sensitivity related to CYP2C9, CYP3A5, ABCB1 (MDR1) and other factors. Pharmacogenomics J 2004;4:40-48.
  • 174.Takahashi H, Echizen H. Pharmacogenetics of CYP2C9 and interindividual variability in anticoagulant response to warfarin. Pharmacogenomics J 2003;3:202-214.
  • 175.Scordo MG, Pengo V, Spina E et al. Influence of CYP2C9 and CYP2C19 genetic polymorphisms on warfarin maintenance dose and metabolic clearance. Clin Pharmacol Ther 2002;72:702-710.
  • 176.Kamali F, Khan TI, King BP et al. Contribution of age, body size, and CYP2C9 genotype to anticoagulant response to warfarin. Clin Pharmacol Ther 2004;75:204-212.
  • 177.Takahashi H, Wilkinson GR, Caraco Y et al. Population differences in S-warfarin metabolism between CYP2C9 genotype-matched Caucasian and Japanese patients. Clin Pharmacol Ther 2003;73:253-263.
  • 178.Jiang X, Williams KM, Liauw WS et al. Effect of St John’s wort and ginseng on the pharmacokinetics and pharmacodynamics of warfarin in healthy subjects. Br J Clin Pharmacol 2004;57:592-599.
  • 179.Wells PS, Holbrook AM, Crowther NR, Hirsh J. Interactions of warfarin with drugs and food. Ann Intern Med 1994;121:676-683.View Abstract
  • 180.Uehleke B. Hypericum interactions: an update. Sixth International ESCOP Scientific Symposium. Bonn Germany; 2001. http://www.phytotherapy.org/gphy/6IS-ESCOP.htm.
  • 181.Khawaja IS, Marotta RF, Lippmann S. Herbal medicines as a factor in delirium. Psychiatr Serv 1999;50:969-970.View Abstract
  • 182.Zullino D, Borgeat F. Hypertension induced by St. John’s wort: a case report. Pharmacopsychiatry 2003;36:32.View Abstract
  • 183.Patel S, Robinson R, Burk M. Hypertensive crisis associated with St. John’s wort. Am J Med 2002;112:507-508.View Abstract
  • 184.Hopfner M, Maaser K, Theiss A et al. Hypericin activated by an incoherent light source has photodynamic effects on esophageal cancer cells. Int J Colorectal Dis 2003;18:239-247.View Abstract
  • 185.Ladner DP, Klein SD, Steiner RA, Walt H. Synergistic toxicity of delta-aminolaevulinic acid–induced protoporphyrin IX used for photodiagnosis and Hypericum extract, a herbal antidepressant. Br J Dermatol 2001;144:916-918.View Abstract
  • 186.Kelty CJ, Brown NJ, Reed MW, Ackroyd R. The use of 5-aminolaevulinic acid as a photosensitiser in photodynamic therapy and photodiagnosis. Photochem Photobiol Sci 2002;1:158-168.View Abstract
  • 187.Nebel A, Schneider BJ, Baker RK, Kroll DJ. Potential metabolic interaction between St. John’s wort and theophylline [letter]. Ann Pharmacother 1999;33:502.
  • 188.Morimoto T, Kotegawa T, Tsutsumi K et al. Effect of St. John’s wort on the pharmacokinetics of theophylline in healthy volunteers. J Clin Pharmacol 2004;44:95-101.View Abstract