Iowa State University Capstones, Theses and Graduate Theses and Dissertations Dissertations

2011 Bioavailability and metabolism of botanical constituents and enhancement of intestinal barrier function by caffeic ca id derivatives in Caco-2 cells Zhiyi Qiang Iowa State University

Follow this and additional works at: https://lib.dr.iastate.edu/etd Part of the Nutrition Commons

Recommended Citation Qiang, Zhiyi, "Bioavailability and metabolism of botanical constituents and enhancement of intestinal barrier function by caffeic ca id derivatives in Caco-2 cells" (2011). Graduate Theses and Dissertations. 10126. https://lib.dr.iastate.edu/etd/10126

This Dissertation is brought to you for free and open access by the Iowa State University Capstones, Theses and Dissertations at Iowa State University Digital Repository. It has been accepted for inclusion in Graduate Theses and Dissertations by an authorized administrator of Iowa State University Digital Repository. For more information, please contact [email protected].

Bioavailability and metabolism of botanical constituents and enhancement of

intestinal barrier function by derivatives in Caco-2 cells

by

Zhiyi Qiang

A dissertation submitted to the graduate faculty in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY

Major: Toxicology

Program of Study Committee: Suzanne Hendrich, Major Professor Patricia Murphy Diane Birt Manju Reddy Michael Wannemuehler

Iowa State University Ames, Iowa 2011 Copyright © Zhiyi Qiang, 2011, All rights reserved. ii

All rights reserved

INFORMATION TO ALL USERS The quality of this reproduction is dependent upon the quality of the copy submitted.

In the unlikely event that the author did not send a complete manuscript and there are missing pages, these will be noted. Also, if material had to be removed, a note will indicate the deletion.

iii

TABLE OF CONTENTS

LIST OF FIGURES ⅵ

LIST OF TABLES ⅷ

ABSTRACT ⅸ

CHAPTER 1. GENERAL INTRODUCTION 1

1.1 Introduction 1

1.2 Dissertation Organization 4

CHAPTER 2. LITERATURE REVIEW 6

1. GENERAL INTRODUCTION OF BOTANICAL MEDICINE 6

2. OVERALL INTRODUCTION OF ECHINACEA SANGUINEA AND PALLIDA 6

2.1 Phytochemistry of E. sanguinea and E. pallida 7

2.2 Health benefit of Echinacea sanguinea and pallida 8

2.2.1 Anti-inflammatory effect 9

2.2.2 Anti-oxidant effect 10

2.2.3 Anti-viral effect 10

2.3 Bioavailability of Echinacea sanguinea and pallid chemical constituents 12

2.4 Metabolism of Echinacea sanguinea and pallida 13

3. OVERALL INTRODUCTION OF 15

3.1 Phytochemistry of Prunella vulgaris 16

3.2 Health benefit of Prunella vulgaris 17

3.2.1 Anti-oxidant effect 17

3.2.2 Anti-inflammatory effect 17

3.2.3 Anti-mutagenic effect 18

3.2.4 Anti-viral effect 19 iv

3.3 Bioavailability of Prunella vulgaris chemical constituents 20

3.4 Metabolism of Prunella vulgaris 22

4. OVERALL INTRODUCTION OF CAFFEIC ACID DERIVATIVES 23

4.1 Source and intake of caffeic acid derivatives 23

4.2 Anti-inflammatory activities of caffeic acid derivatives 25

5. INTESTINAL ABSORPTION, METABOLISM, INFLAMMATORY

RESPONSES, AND RELEVANT METHODOLOGY 27

5.1 Intestinal function 27

5.1.1. Absorption mechanisms 27

5.1.2. Metabolism 29

5.2 Inflammatory responses of the intestine 33

5.3 Relevant methodology to study intestinal function 37

5.3.1 Caco-2 cells 38

5.3.2 Paracellular permeability test 40

5.3.3 Western blot, immunofluorescence and ELISA 41

6. REFERENCES 43

CHAPTER 3. GLUCURONIDATION OF ALKAMIDES AND KETONES IN

ECHINACEA SANGUINEA AND ECHINACEA PALLIDA EXTRACTS BY CACO-2

CELLS AND THEIR MODULATION OF P-GLYCOPROTEIN TRANSPORTER 61

Abstract 61 Introduction 63 Materials and Methods 66 Results 70 Discussions 73 Figures 82 Tables 90 v

References 92

CHAPTER 4. METABOLISM AND PERMEABILITY OF ROSMARINIC

ACID IN PRUNELLA VULGARIS AND URSOLIC ACID IN SALVIA

OFFICINALIS EXTRACTS ACROSS CACO-2 CELL MONOLAYERS 98

Abstract 98 Introduction 100 Materials and Methods 102 Results 106 Discussions 108 Figures 114 Tables 118 References 120 CHAPTER 5. CAFFEIC ACID DERIVATIVES ENHANCED EPITHELIAL

BARRIER INTEGRITY AND EXHIBITED ANTI-INFLAMMATORY

ACTIVITIES IN P-GLYCOPROTEIN TRANSPORTER KNOCKDOWN

CACO-2 CELLS 125

Abstract 125 Introduction 127 Materials and Methods 129 Results 133 Discussions 136 Figures 143 Tables 149 References 155 CHAPTER 6. GENERAL CONCLUSIONS 163 General Discussion 163

Recommendations for Future Research 167

ACKNOWLEDGEMENTS 168 vi

LIST OF FIGURES

Chapter 2. Figure 1. Structure of common alkamides, ketones, and caffeic acid derivatives found in Echinacea . 7

Chapter 2. Figure 2. Phenolic acid, flavonoids, and triterpenoids in

Prunella vulgaris. 17

Chapter 2. Figure 3.Transporters in intestinal epithelial cells. 32

Chapter 2. Figure 4. Intestinal epithelial intercellular junctions. 35

Chapter 3. Figure 1. The chemical structures of Bauer alkamide 8, 10, 11

and ketone 24. 82

Chapter 3. Figure 2. Transport of pure Bauer alkamide 8, 10 and 11

across Caco-2 cells. 84

Chapter 3. Figure 3. Glucuronidation of Bauer alkamide 8, 10, 11 and

ketone 24 for pure compounds or from extracts by Caco-2 cells. 86

Chapter 3. Figure 4. Chromatograms of Bauer ketone 24 across Caco-2 cell

monolayers as pure compound or from Echinacea species. 88

Chapter 3. Figure 5. The effects of Bauer alkamide 8, 10, 11, ketone 24,

E. sanguinea and E. pallida on P-gp expression. 89

Chapter 4. Figure 1. The chemical structures of rosmarinic and ursolic acids. 114

Chapter 4. Figure 2. The transport rate of ursolic acid for pure compound

and S. officinalis extract across Caco-2 cell monolayer as a function of

concentration at 5–20 μM after 4 h uptake study. 115

Chapter 4. Figure 3. Characteristics of transfer of rosmarinic (RA) and ursolic acids (UA) for pure compounds or plant extracts across Caco-2 cells

before and after deconjugation of post-experimental basolateral solutions. 117 vii

Chapter 5. Figure 1. The chemical structures of caffeic acid, , and m-hydroxyphenylpropionic acid (mHPP). 143

Chapter 5. Figure 2. P-glycoprotein transporter activity in Caco-2 cells and multidrug resistance (MDR) gene knockdown (KD) Caco-2 cells. 144

Chapter 5. Figure 3.TEER in Caco-2 cell monolayersincubated with or without caffeic acid derivatives (50 µM) and quercetin (100 µM). 146

Chapter 5. Figure 4. Lucifer yellow flux across Caco-2 cell monolayers incubated with or without caffeic acid derivatives (50 µM) and quercetin (100 µM) for 48 h incubation. 148 viii

LIST OF TABLES

Chapter 3. Table 1. The concentrations of alkamides and ketone in Echinacea

ethanol extracts determined by HPLC. 90

Chapter 3. Table 2. Apparent permeability coefficients (Papp) for pure alkamides

and extracts across the Caco-2 monolayer. 91

Chapter 4. Table 1. The contents of rosmarinic and ursolic acids in ethanol

extracts of P. vulgaris and S. officinalis determined by HPLC. 118

Chapter 4. Table 2. Apparent permeability coefficients (Papp) for lucifer yellow,

rosmarinic acid, ursolic acid and extracts across the Caco-2 monolayer. 119

Chapter 5. Table 1. Immunoblot analysis of tight junction proteins in the whole

extracts of Caco-2 cells incubated with caffeic acid derivatives (50 µM) and

quercetin (100 µM) for 48 hs, quantitated by densitometric analysis. 150

Chapter 5. Table 2. Immunoblot analysis of tight junction proteins in the whole

extracts of MDR KD Caco-2 cells incubated with caffeic acid derivatives (50 µM) and quercetin (100 µM) for 48 hs, quantitated by densitometric analysis. 151

Chapter 5. Table 3. Caffeic acid derivatives (50 µM, 48 h incubation)

augmented the inflammatory responses after treating Caco-2 cells with the stimuli. 152

Chapter 5. Table 4. Caffeic acid derivatives (50 µM, 48 h incubation) attenuated the

inflammatory responses in MDR KD Caco-2 cells. 153

Chapter 5. Table 5. Caffeic acid derivatives (50 µM, 48 h incubation) attenuated

the inflammatory responses in MDR KD Caco-2 cells treated with the stimuli. 154

Chapter 5. Table 6. Transfer of quercetin (100 µM) and caffeic acid derivatives

(50 µM) across Caco-2 and MDR KD Caco-2 cells. 155

ix

ABSTRACT

Public interest in botanical supplements has increased greatly in recent years as various plant

materials might be used for anti-inflammatory, immunostimulatory, antioxidant and cancer

preventive effects. Our long term goal is to improve our understanding of the

characteristics of phytochemicals that contribute to human health benefits on gut functions,

and thereby pave the way for optimizing herbal supplements for study in future clinical trials.

In this dissertation, the overarching hypotheses were that major components of the ethanolic

extracts of Echinacea, alkamides and ketones, and caffeic acid derivatives in the ethanolic

extract of Prunella vulgaris, will be transferred by Caco-2 cell monolayers and caffeic acid

derivatives will enhance the intestinal epithelial barrier function.

Echinacea has long been used as phytotherapy for wound healing, pain relief and treatment

of the common cold. In the first study, Bauer alkamides, the key components contained in

Echinacea sanguinea and Echinacea pallida, transferred across the Caco-2 cell monolayer

via passive diffusion, independent of other constituents in plant extract. The apparent

-6 permeability coefficients (Papp) were 2.8 ± 1.5- 43.8 ± 11.2 cm/s ×10 for tested three alkamides and the order of the transfer of them across Caco-2 cells was increased paralleled with compound hydrophicility. Tested alkamides were seemingly N-glucuronidated and both Echinacea extracts stimulated apparent glucuronidation and basolateral efflux of alkamide metabolites. Bauer ketone 24 was totally metabolized to more hydrophilic metabolites as a pure compound, but not found in either Echinacea species. The addition of

Bauer alkamides (175-230 µM) as well as the ethanolic extracts of E. sanguinea at 1 mg/mL

(containing 85 µM of alkamide 8, 2 µM of alkamide 10, and 0.7 µM of alkamide 11) and E. pallida at 5 mg/mL (containing 215 µM of alkamide 8, 25 µM of alkamide 10, and 45 µM of alkamide 11) reduced the efflux of the P-glycoprotein transporter (P-gp) probe calcein-AM x

from Caco-2 cells. These results suggest that other constituents in the plant extract had a facilitating effect on the metabolism and efflux of alkamides and ketones from Echinacea, which would improve the therapeutic benefits of these extracts, and that alkamides and

Echinacea extracts might be useful in potentiating some chemotherapeutics which are substrates for P-gp.

Prunella vulgaris is a perennial herb known as self-heal used to treat sore throat, fever, and wounds. Rosmarinic acid is a caffeic acid derivative found in various botanicals, especially in P. vulgaris. Ursolic acid, a pentacyclic triterpene acid, is also found in P. vulgaris but especially concentrated in Salvia officinalis (sage), which has been traditionally used to treat inflammation in the oral cavity, and may also be of interest in inhibiting gastrointestinal inflammation which is relevant to colitis and colon cancer. In the second study, Papp for rosmarinic acid and rosmarinic acid in P. vulgaris extracts was 0.2 ± 0.05 × 10-6 cm/s, significantly increased to 0.9 ± 0.2 × 10-6 cm/s after β-glucuronidase/sulfatase treatment.

-6 Papp for ursolic acid and ursolic acid in S. officinalis extract was 2.7 ± 0.3 × 10 cm/s and 2.3

± 0.5 × 10-6 cm/s before and after β-glucuronidase/sulfatase treatment, respectively.

Neither compound was affected in permeability by the herbal extract matrix. These results indicate that rosmarinic acid and ursolic acid in herbal extracts had similar uptake as that found using the pure compounds, which may simplify the prediction of compound efficacy, but the apparent lack of intestinal glucuronidation/sulfation of ursolic acid is likely to further enhance the bioavailability of that compound compared with rosmarinic acid.

In the third study, the effects of caffeic acid and related compounds on intestinal barrier function were investigated using Caco-2 cells as a model. Caffeic acid, rosmarinic acid, chlorogenic acid and m-hydroxyphenylpropionic acid (mHPP, a microbial metabolite of xi

caffeic acid and rosmarinic acid) up-regulated the expression of zonula occludens (ZO -1),

ZO-2, claudin-1 and occludin in Caco-2 cells. In addition, chlorogenic acid and mHPP were effective against adverse effects induced by inflammatory stimuli (LPS, interferon-γ,

IL-1β, and TNF-α) on tight junction proteins in Caco-2 cells. Caffeic acid derivatives up-regulated claudin-4 in P-glycroprotein transporter MDR1- knockdown (KD) Caco-2 cells and only mHPP was effective against the changes in tight junction protein expression induced by inflammatory stimuli in MDR KD Caco-2 cells. Caffeic acid derivatives augmented TNF-α and IL-6 levels in Caco-2 cells under the stimulated condition, but significantly reduced both cytokines in MDR KD caco-2 cells plus the stimuli. These results indicate that caffeic acid derivatives enhanced barrier function in human intestinal

Caco-2 cells and mHPP exhibited greater enhancement of intestinal barrier than the parent compounds. P-gp plays an essential role in the anti-inflammatory activities of caffeic acid derivatives.

In conclusion, these data confirmed the overarching hypotheses and suggest that the effect of plant matrix on bioavailability and metabolism of the constituent is compound specific, depending on the transfer mechanism; and caffeic acid derivatives could be gut health promoting as a dietary constituent, but these compounds might exacerbate damage under inflammatory stimuli. Moreover, intake of caffeic acid derivatives might speed up mucosal recovery or provide protection to the small-intestinal mucosa against the inflammatory mediators when P-gp inhibitors are co-administered, which may be attractive from a therapeutic point of view. 1

CHAPTER 1. GENERAL INTRODUCTION

Introduction

Echinacea supplements were the most commonly used nonvitamin, nonmineral, natural products both for adults and children according to 2007 NHIS survey (National Health

Interview Survey, 2007). Species of Echinacea, such as E. angustifolia, E. pallida, E.

sanguinea and E. purpurea, have immune modulatory (Borchers et al., 2000), antiviral

(Pleschka et al., 2009), and antibacterial activities (Sharma et al., 2010). Several

alkamides from the roots of E. angustifolia and E. pallida have suppressed the expression

of cyclooxygenase 2 (COX-2) and prostaglandin E 2 (PGE2) formation in H4 human

neuroglioma cells (Hinz et al., 2007) and lipopolysaccharide (LPS) stimulated RAW264.7

macrophages. Bauer ketones 23 and 24 at 2.6 and 5 µM, respectively, were identified as significant contributors to the PGE (2) inhibition by E. pallida root extract at 25 µg/mL

(LaLone et al., 2007). E. pallida root and aerial extracts at 2 mg/mL exhibited ~50% inhibition of 2, 2-diphenyl-2-picrylhydrazyl (DPPH) radical and the order of the antioxidant activities of the phenolics in E. pallida extract was (EC50 = 6.6

μM) > cichoric acid (8.6 μM) > cynarin (11.0 μM) > chlorogenic acid (18.9 μM) > caffeic

acid (19.1 μM) > (20.5 μM). The efficacy of Echinacea species may be

related to the bioavailability of its constituents. In a randomized crossover single-dose

study, 10 volunteers were orally administered standardized tincture of E. purpurea or

twelve E. purpurea tablets, and both doses contained 0.07 mg of the major alkamides.

Alkamides derived from the Echinacea tincture were detectable in serum within 30 min

after administration and their concentration peaked at 0.4 ng/ml; alkamides in tablets

reached a level of 0.12 ng/ml in serum 45 min after ingestion (Woelkart et al., 2006).

However, caffeic acid conjugates (caftaric acid, cichoric acid and echinacoside) were not 2

detected in any plasma sample at any time after ingestion of Echinacea tablets

manufactured from an ethanolic liquid extract in a human study (Matthias et al., 2005).

The Prunella (), a perennial herb that is widely distributed throughout

Asia and Europe, is used primarily as a traditional botanical medicine to reduce

osteoarthritis, rheumatoid arthritis, and gingivitis (Song et al., 2007; Adámková et al., 2004;

Jung et al., 2001). Rosmarinic acid (RA), one of the key constituent contained in Prunella,

was found to exhibit anti-inflammatory and antioxidative properties (Huang et al., 2009;

Psotova et al., 2003). Rosmarinic acid might be poorly absorbed in the small intestine

because when extract containing 200 mg of RA was orally administered to six men,

RA in both plasma and urine was present predominantly as glucuronide and/or sulfate conjugated forms, at 0.6 ± 0.2% and 1.5 ± 0.4% of the total intake, respectively, within 48 h after ingestion (Baba et al., 2005). Gut microbes may metabolize RA to give phenolics such as caffeic acid, o-coumaric acid and m-hydroxyphenylpropionic acid, which are then absorbed by monocarboxylic acid transporter (MCT)-mediated active processes (Konishi and Kobayashi, 2005).

Until now, studies have not been carried out to investigate the effects of plant extracts on the bioavailability and metabolism of alkamides, ketones and caffeic acid derivatives, nor have the mechanisms of the protection of caffeic acid derivatives on intestinal function been investigated. In this whole project, we focused our research interest in these less explored areas. The overarching hypotheses were that alkamides, ketones and caffeic acid derivatives will be transferred by Caco-2 cell monolayers and caffeic acid derivatives will enhance the intestinal barrier function. We conducted these studies to test three individual hypotheses within two main research fields: bioavailability and metabolism of botanical 3

constituents in vitro (Hypothesis and objective 1 and 2) and modulation of caffeic acid derivatives on intestinal function (Hypothesis and objective 3).

1. In the first study, we hypothesized that Bauer alkamides and ketones would be

absorbed similarly as pure compounds or as part of complex plant extracts (E.

sanguinea and E. pallida), and that these Echinacea extracts could inhibit

P-glycoprotein (P-gp) transporter in Caco-2 human intestinal epithelial cells. Caco-2

cells were incubated with pure alkamides and ketone as well as with the two extracts

that contain the specific alkamides to determine the permeability and metabolism of

those key constituents both in pure solutions and in herbal extracts. Efflux of the P-gp

probe calcein-AM across Caco-2 cells was assessed to study the inhibtion of P-gp

function by pure compounds and two extracts.

2. In the second study, our hypotheses were that the absorbability and metabolism of

rosmarinic acid and ursolic acid was independent of the plant extract matrix. This study

was also designed to investigate the permeability of rosmarinic acid and ursolic acid as

pure compounds in comparison to the permeability of these compounds when added as

a complex mixture in the form of P. vulgaris and S. officinalis ethanol extracts across

Caco-2 cell monolayers. It is anticipated that the results of these studies will provide a

foundation for future studies of to assess the efficacy of these herbs as therapeutic

strategies to treat gastrointestinal inflammation.

3. In the third study, our hypothesis was that caffeic acid , rosmarinic acid, chlorogenic

acid and m-hydroxyphenylpropionic acid (mHPP), a microbial metabolite of caffeic

acid and rosmarinic acid, will alter the intestinal barrier function in Caco-2 human

intestinal epithelial cells and P-gp transporter, encoded by multidrug resistance (MDR1)

gene, knockdown (KD) Caco-2 cells. Stimuli composed of interleukin (IL)-1β, tumor 4

necrosis factor (TNF)-α, interferon (IFN)-γ and lipopolysaccharides (LPS) were added

to the cell system to represent acute inflammatory reactions and mimic inflammatory

bowel disease. Tight junction protein expression and immune modulator excretion in

both Caco-2 cell model and MDR knockdown (MDR KD) Caco-2 cells were measured.

Dissertation Organization

The contents of this thesis includes an introduction, which is comprised of a general introduction and a literature review that delves into the phytochemical profile, health benefits, bioavailability and metabolism of the key constituents in Echinacea and Prunella vulgaris ethanolic extracts. Chapter 3 consists of a manuscript that will be submitted to the Journal of Planta Medica entitled “Glucuronidation of alkamides and ketones in

Echinacea sanguinea and Echinacea pallida extracts by Caco-2 cells and their modulation of P-glycoprotein transporter”. Co-authors on this manuscript include: Basil J. Nikolau, mentored Ludmila Rizshsky and provided the phytochemical profile in Echinacea ethanolic extracts using GC-MS; Patricia A. Murphy and Cathy Hauck worked on the extraction process for the Echinacea extracts; Mark P. Widrlechner and Joe-Ann McCoy provided all Echinacea plant material; Manju B. Reddy, mentored Zhiyi Qiang on Caco-2 cell culture; Suzanne Hendrich, mentored Zhiyi Qiang and provided guidance and edits to all research conducted and manuscript drafts. Chapter 4 consists of a manuscript submitted to the Journal of Ethnopharmacology entitled “Metabolism and permeability of rosmarinic acid in Prunella vulgaris and ursolic acid in Salvia officinalis extracts across

Caco-2 cell monolayers”. Co-authors on this manuscript include: Patricia A. Murphy and Cathy Hauck worked on the extraction process for the P. vulgaris extracts; Mark P.

Widrlechner and Joe-Ann McCoy provided all P. vulgaris plant material; Manju B. Reddy, 5

mentored Zhiyi Qiang and Zhong Ye on Caco-2 cell culture; Suzanne Hendrich, mentored

Zhiyi Qiang and Zhong Ye and provided guidance and edits to all research conducted and manuscript drafts. Chapter 5 consists of a manuscript that will be submitted to the Journal

of Nutrition entitled “Caffeic acid derivatives enhanced epithelial barrier integrity and exhibited anti-inflammatory activities in P-glycoprotein transporter knockdown Caco-2

cells”. Co-authors on this manuscript include: Diane Birt, mentored Nan Huang and

provided the guidance on Western Blot analysis of tight junction proteins and ELISA

analysis of cytokines in Caco-2 cells; Kazuya Maeda, provided P-glycoprotein knockdown

Caco-2 cells; Manju B. Reddy mentored Zhiyi Qiang on Caco-2 cell culture; Suzanne

Hendrich, mentored Zhiyi Qiang and Li Li and provided guidance and edits to all research

conducted and manuscript drafts. Finally, chapter 6 contains a general conclusion and

recommendations for future research. 6

CHAPTER 2. LITERATURE REVIEW

1. GENERAL INTRODUCTION OF BOTANICAL MEDICINE

Botanical medicine, or phytomedicine, has become a promising area in US, an alternative to conventional medicine. Human use of botanical medicine has a long history, whose record can date back to about 60,000 years ago (Solecki, 1975). It is estimated that as much as 80% of the world population still depends on herbal medicines to meet primary healthcare, mainly for particular diseases or symptoms based on the properties of a particular botanical (van Wijk, 2000). Indeed, medicinal herbs play an essential role in drug development as ~30% of all current medical drugs sold worldwide contain compounds derived from plant material (FAO, 2004). In the US, herbal medicine is classified as dietary supplements, which are consumed without prescription. Notable examples of the botanical dietary supplements used widely in the US today include

Echinacea, garlic, ginkgo biloba, ginseng, and St. John’s wort with the sales over 100 million annually.

2. OVERALL INTRODUCTION OF ECHINACEA SANGUINEA AND PALLIDA

The genus of Echinacea is a perennial herb that is native to North America and grouped within the Asteraceae family. Historically, Echinacea preparations have been used by

Native Americans to relieve pain and to treat various conditions such as snakebites and wounds. Early European settlers learned of its medicinal effects, and expanded its uses in treatment of infectious diseases (Borchers et al., 2000). Today, Echinacea products are mainly used for reducing the duration and/or the symptoms associated with the common cold and other upper respiratory infections, although clinical trials have not established efficacy (Barrett et al., 2010; Turner et al., 2005). It is estimated that Americans spend over $100 million per year on Echinacea products and their sales make up 10% of the total 7

US herbal market (Islam and Carter, 2005). Echinacea supplement was the most commonly used nonvitamin, nonmineral, natural products both for adults and children according to 2007 NHIS survey (National Health Interview Survey, 2007). The nine species of Echinacea that have been characterized are E. angustifolia, E. pallida, E. purpurea, E. sanguinea, E. simulata, E. tennesseensis, E. laevigata, E. atrorubens, and E. paradoxa, with each possessing its own unique phytochemical components (Miller and Yu,

2005). The diversity, or in some cases similarity, of the constituents present in these

Echinacea species contributes to the diverse bioactivities associated with the botanical

Echinacea.

2.1 Phytochemistry of E. sanguinea and E. pallida

A wide range of phytochemicals have been found in Echinacea, some of which have been reported to have bioactivity. These include alkamides and ketones (Bauer, 2002;

Woelkart et al., 2005), caffeic acid derivatives, such as caffeic acid, chlorogenic acid, caftaric acid, echinacoside, etc. (Cheminat et al., 1988, Figure 1), polysaccharides (Orinda et al., 1973), and glycoproteins (Egert and Beuscher, 1992).

Figure 1 Structure of common alkamides, ketones, and caffeic acid derivatives found in

Echinacea species 8

E. angustifolia, E. sanguinea, E. purpurea, and E. tennesseensis ethanolic extracts (15

μg/mL of each) were analyzed by HPLC showing greater quantities of Bauer alkamides than ketones or caffeic acid derivatives. The concentration range of Bauer alkamides 3,

8-14 and 17 were 0.1-2.8 μM. Ketones were present around 0.1 μM. Cichoric acid and

caftaric acid were present around 0.07-0.28 μM and 0.01-0.04 μM, respectively (LaLone et

al., 2007). Another study investigated the metabolite distribution in E. pallida methanolic extract (15 μg/mL) and found that alkamide 1/2, 3 and 4 were contained at 1.5, 0.7 and 0.4

μM, but ketones, including ketone 19-28, were present at 30.5 -52.4 μM (Hou et al., 2010),

10 fold greater than the quantity of alkamides, but caffeic acid derivatives were not determined in this study. Pellati et al. (2004) report the total phenolic content for E. angustifolia, E. pallida, and E. purpurea to be 10.49 mg/g, 17.83 mg/g, and 23.23 mg/g, respectively. E. pallida methanolic extract (10 μg/mL) contained 2.5 μM, 20.3 μM and

1.7 μM of caftaric acid, echinacoside, and cichoric acid, respectively (Figure 1).

Although many of the active compounds of Echinacea have been identified, relative

potency, mechanism of action, bioavailability and metabolism are less known.

2.2 Health benefits of E. sanguinea and E. pallida

E. pallida is one of the most medicinally used Echinacea species in the US and Europe and is being introduced into other regions due to the increasing popularity in alternative herbal

remedies. E. sanguinea has been neglected in the scientific literature, however, Birt et al.,

(2008) reported that E. sanguinea ethanolic extract had greater anti-inflammatory activity

in terms of the significant reduction of prostaglandin E 2 (PGE2) production in

lipopolysaccharide (LPS) stimulated RAW264.7 macrophages, than E. angustifolia, E.

pallida, and E. simulata. In addition, our preliminary experiments showed that alkamide 8,

10, 11 and ketone 24 were only present in E. pallida and E. sanguinea ethanol extracts, but 9

not in E. angustifolia and E.simulata ethanolic assessions. Therefore, E. pallida and E.

sanguinea were choosen in this study to investigate their absorption, metabolism and health

benefits.

2.2.1 Anti-inflammatory effects

E. angustifolia, E. pallida, and E. simulata have been described to possess an anti-inflammatory activity which was differed from species, extracts, and fractionation and contributed to main constituents (LaLone et al., 2007; Birt et al., 2008). Treatment with echinacoside at 0.5 µM, a caffeic acid derivative from root extract of E. pallida at 100 mg/mL, inhibited the signs of inflammation injected by 1 ml of a gel (1% ethylcellulose) on the dorsal area of the rats for 48 and 72 h (Speroni et al., 2002). Bauer ketones 23 and 24 at 2.6 and 5 µM, respectively, were identified as significant contributors to the PGE (2) inhibition by E. pallida root extract at 25 µg/mL. Three different alkamides (alkamide 5, alkamide 7 and alkamide 8) from the root of E. angustifolia have suppressed the expression of cyclooxygenase 2 (COX-2) and PGE2 formation at sites of inflammation in H4 human neuroglioma cells (Hinz et al., 2007), but E. pallida root extract did not show any suppressive effect on COX-2 promoter activity at 50 µg/mL at both transcriptional and translational levels in 12-O-tetradecanoylphorbol -13-acetate (TPA) stimulated RAW

264.7 cells, human breast adenocarcinoma MCF-7 cell as well as mouse models (Hou et al.,

2010). Combinations of different alkamides in Echinacea extracts might play a synergistic role in the inhibition of PGE2 production in LPS-stimulated RAW264.7 mouse macrophage cells. Active extracts containing < 2.8 μM of combined alkamides in 15

μg/mL of E. angustifolia, E. pallida, E. simulata, and E. sanguinea extracts significantly have inhibited PGE2 production (LaLone et al., 2007). In a summary, different species of 10

Echinacea had an effective anti-inflammatory action and the anti-inflammatory activity was mainly contributed to alkamides or caffeic acid derivatives.

2.2.2 Anti-oxidant effects

Assessing free radical scavenging was mainly used to evaluate the antioxidant capacity in

medicinal and other nutritional antioxidant supplements. Echinacea extracts have

been reported to be free radical scavengers and possess anti-oxidant properties. The roots

of E. purpurea, E. angustifolia and E. pallida exhibited significant free radical scavenging capabilities by the inhibition of Fe2+-induced lipid peroxidation (Sloley et al., 2001). E.

pallida root and aerial extracts at 2 mg/mL exhibited ~50% inhibition of 2,

2-diphenyl-2-picrylhydrazyl (DPPH) radical and displayed ~55% of Fe2+-chealating power

(Orhan et al., 2009). Capacities to quench DPPH radicals of phenolic compounds in

Echinacea species indicated that echinacoside (EC50 = 6.6 μM) > cichoric acid (8.6 μM) > cynarin (11.0 μM) > chlorogenic acid (18.9 μM) > caffeic acid (19.1 μM) > caftaric acid

(20.5 μM), which might be due to the number of hydroxyl groups on the aromatic ring as

the more hydroxyl groups, the greater the radical scavenging activity (Wang et al., 2003).

Alkamide 2 and 8 inhibited Cu (II)-catalyzed oxidation of human low-density lipoprotein in vitro and showed a synergistic effect with caffeic acid derivatives, and polysaccharides

(Dalby-Brown et al., 2005). In a summary, Echinacea species had an effective antioxidant

capacity across species and extracts; the antioxidant capacity was attributed to ability to

scavenge free radical.

2.2.3 Anti-viral effects

Echinacea extracts and its metabolic constituents possessed antiviral activities (Cheminat et al., 1988), but the results are very controversial. Hydroalcoholic root extract of E. pallida (100 µg/mL) exhibited high levels of antiviral activity against both types of herpes 11

simplex virus types 1 and 2 (HSV-1, HSV-2) analyzed by plaque reduction assays

(Schneider et al., 2009). Interleukin-2 is required for the clonal expansion and activation

of T cells during viral infection and has also been linked to decreased symptoms and reduction of virus revival in experimental rhinovirus infection in human Jurkat T cells

(Sasagawa et al., 2006). Aerial extract of E. purpurea at 50 µg/mL inhibited IL-2 production by ~65% and Bauer alkylamide 9 and Bauer alkylamide 11 at ~2 µM were both capable of inhibiting IL-2 production in Jurkat E6.1 T cells (Sasagawa et al., 2006).

In a human study, supercritical carbon dioxide, 60 % ethanol, or 20 % ethanol extracts of E.

angustifolia root did not show any statistically significant effects on rhinovirus infection

and severity of symptoms in 437 volunteers, measured by the volume of nasal secretions,

polymorphonuclear leukocyte or interleukin-8 concentrations in nasal-lavage specimens,

as well as quantitative-virus titer (Turner et al., 2005). The conflicting results might be

because the dose of Echinacea extract was not be adequate to show protection against

rhinovirus infection. However, aqueous-ethanolic E. purpurea extract induced a

statistically significant increase in the survival rate, prolonged the mean survival time and

reduced lung consolidation and virus titer after 6 days of intranasal infection with Influenza

A virus in mice (Bodinet et al., 2002). Fusco et al. (2010) also showed that E. purpurea

polysaccharide extract had lower systemic and pulmonary keratinocyte chemoattractant

(KC) and interleukin (IL)-10 levels and lower systemic interferon-γ (IFN-γ) levels

following influenza infection compared with controls in mice. Therefore, more and larger human studies are needed for the anti-viral effect of Echinacea extracts, and the dose,

species and type of Echinacea extract, time to initiation of treatment, virus type as well as

immune competence of volunteers should be considered. Meanwhile, more investigations 12

are critically required on the absorption and metabolism data regarding the various constituents of different Echinacea species.

2.3 Bioavailability of E. sanguinea and E. pallida chemical constituents

A major weakness with most in vitro studies dealing with botanicals and bioactivity is accounting for how much of the particular extract or compound actually enters the cells or

the body of the model being studied. Therefore, it is essential to investigate the

bioavailability of the diverse constituents contained in Echinacea species. Most of the

published papers were focusing on the bioavailability and pharmacokinetics of different alkamides and caffeic acid derivatives in human and cell culture models.

Alkamides 8 (Figure 1) from Echinacea species has been reported to transport through

human adenocarcinoma colonic cell line Caco-2 monolayer at 30 minutes after apical

loading of 25 μg/ml and about 15 % of these alkamides were detectable on the basolateral side, which was not affected by lipopolysaccharide or phorbol 12-myristate-13-acetate in order to mimic an inflammatory state (Jager et al., 2002). In another study, twelve alkylamides were shown to readily cross Caco-2 monolayers (apparent permeability >10-6 cm/s), and 2, 4-dienes were more readily transported than the equivalent 2-ene alkamides

(Matthias et al., 2004). Among the tested caffeic acid derivatives, , caftaric acid, cichoric acid, echinacoside, and exhibited poor uptake through the

Caco-2 monolayers (less than 5% transported). Chicca et al. (2008) also used the Caco-2 system to study the permeability of ketones from E. pallida extract. Bauer ketone 22 and

24 were readily permeable through the cell system, with ketone 22 showing the highest apparent permeability at 32 ± 3 × 10-6 cm/s.

13

In a randomized crossover study, six healthy adults of both sexes consumed 4 g of

Echinacea powder mixture which contained E. angustifolia, E. purpurea, and E. pallida.

The 24 h urinary recovery of cichoric acid of E. purpurea and E. pallida were 0.36 ± 0.25

and 0.34 ± 0.31 % of ingested dose (Lee et al., 2006). However, another human study did

not detect caffeic acid conjugates (caftaric acid, cichoric acid and echinacoside) in any

plasma sample at any time after ingestion of Echinacea tablets manufactured from an

ethanolic liquid extract, whereas alkamides were detected in plasma 20 minutes after tablet

ingestion (Matthias et al., 2005). In a randomized crossover single-dose study, 10

volunteers were orally administered either 4 ml of the standardized E. purpurea tincture or

twelve E. purpurea tablets, and both doses contained 0.07 mg of the major alkamides.

Alkamides in Echinacea tincture in serum were present at 0.4 ng/ml 30 min after admnistration, whereas alkamides in tablets were 0.12 ng/ml serum after 45 min (Woelkart et al., 2006). Similar bioavailability of alkylamides was found from the liquids (200 and

300 mg/ml) and tablets (600 and 675 mg/tablet) Echinacea formulations in a two-way crossover study in humans (Matthias et al., 2007). Taken together, these results indicate that alkamides, as well as certain ketones, are likely to cross the intestinal barrier and be bioavailable in humans, and therefore, should be able to carry out their bioactivities in vivo,

but caffeic acid derivatives are less likely to do so.

2.4 Metabolism of E. sanguinea and E. pallida

Due to potential interactions with other herbal supplements or pharmaceuticals, it is

necessary to understand how bioactive constituents from Echinacea are metabolized in the

liver or in other tissues. Cytochrome P450 (CYP) may affect the metabolism and further

bioavailability of Echinacea constituents. Matthias et al. (2005) studied the metabolism

of the alkamide constituents from Echinacea premium liquid, a mixture of 200 mg/mL of E. 14

angustifolia and 300 mg/mL of E. purpurea 60% ethanol/water root extracts, in human liver microsomes and provided evidence that NADPH dependent cytochrome P450 metabolized the 2-ene and 2, 4-dienes in the ethanolic extracts of Echinacea.

Monohydroxylated, monoepoxidized, N-dealkylated metabolites were determined after incubation of parent Bauer alkamides with both NADPH and human liver microsomes.

Further results from this study showed that 60% ethanol/water extract of E. purpurea containing 16 μM of Bauer alkylamides 8/9 significantly suppressed IL-2 secretion by 47% in the Jurkat T cells, but the suppression of IL-2 secretion was significantly lessened when

NADH was added with the same treatment, indicating the importance of understanding the

metabolism of constituents found in Echinacea and how it affects different bioactivities.

Echinacea preparations have been found to affect metabolizing enzyme activities and transporter functions in vitro, implying the importance regarding herb–drug interactions.

E. purpurea extract inhibited CYP3A4 activity, tested by the inhibition of the metabolism of 7-benzyloxy-trifluoromethylcoumarin, 7- benzyloxyquinoline and testosterone, three substrates of CYP3A4, using fluorometric assays (Hansen et al., 2008). Another study showed that E. purpurea inhibited CYP 3A4 mediated metabolism in C-DNA baculovirus which expressed human CYP 3A4 supersomes based on the decreased formation of

6-OH-testosterone (Hellum et al., 2008). Several Echinacea ethanolic extracts (11.2 -

2447 µg/mL) have been shown to inhibit baculovirus expressed CYP 2C19, 2D6, and 3A4

by 20-100 %, and alkamides contributed to the inhibitory effects seen with Echinacea

preparations (Modarai et al., 2007). The interaction between Echinacea and drugs has

been investigated. E. purpurea root extract (400 mg) significantly increased the systemic

clearance of intravenous midazolam by 34% but reduced clearance of caffeine and

tolbutamide in twelve healthy subjects, indicating the inhibition on CYP3A at hepatic and 15

intestinal sites as well as improving of CYP 1A2 and CYP 2C9 activity (Gorski et al., 2004).

Therefore, alkamides in Echinacea extracts not only can be enzymatically metabolized, but also may contribute to the inhibition of of these same enzymes as shown by extracts from different Echinacea species.

A previous study showed that n-hexane root extracts from E. pallida, E. angustifolia and E. purpurea at 30 µg/mL inhibited multidrug transporter P-glycoprotein (P-gp) activity in a human proximal tubular cell line (Romiti et al., 2008). P-gp transporter plays a key role in drug absorption and distribution because it limits the permeability across the GI tract by active efflux of potentially toxic substances back into the intestinal lumen. The transporter is also expressed in other tissues such as liver, kidney and blood–brain barrier where it also plays an important excretory role. On the other hand, P-gp confers resistance to anticancer chemotherapy because it is over-expressed in cancer cells (O'Connor 2007).

The use of botanicals which have anticancer properties as ABC transporter inhibitors may allow a new paradigm of clinically useful drug resistance circumvention.

In summary, it is crucial to investigate the uptake and metabolism of key components found in various Echinacea species as well as the influences of Echinacea plant matrix on the bioavailability of these key constituents. Because Echinacea supplements are usually ingested on a chronic basis, their long term effects on metabolic enzymes and efflux transporters needs to be investigated and may be of importance regarding herb–drug interactions.

16

3. OVERALL INTRODUCTION OF PRUNELLA VULGARIS

Prunella vulgaris, a perennial herb known as self-heal, is widely distributed throughout

Asia and Europe, is used primarily as a traditional botanical medicine to reduce

osteoarthritis, rheumatoid arthritis, and gingivitis (Song et al., 2007; Adámková et al., 2004;

Jung et al., 2001). In China, P. vulgaris is called ‘Xia Ku Cao’ and has a long history in

therapeutic use as an antipyretic, and more recently for anti-keratitis purposes (Zheng et al.,

1990). In Turkey, P. vulgaris is used primarily as an expectorant, tonic, antidiarrhoeic,

against dysentery, for common cold, rheumatoid arthritis and cardiac arrhythmia (Pinkas et

al., 1994; Markova et al., 1997). More recently, increasing evidence suggests that organic preparations of P. vulgaris exhibited anti-estrogenic, anti-inflammatory, and antioxidative

properties in vitro cell models and in rats (Psotova et al., 2003).

3.1 Phytochemistry of Prunella vulgaris

It has been found that aqueous extract of P. vulgaris contains on an average alkaloid (1.1%),

saponins (0.4%), phenolic acids (0.06%), tannins (0.06%), carbohydrates (0.4%), proteins

(0.4%) and lipids (2.4%, Rasool et al., 2010). The amount of caffeic acid, chlorogenic

acid and rosmarinic acid (Figure 2) in P. vulgaris was about 0.29, 0.32, and 9.46 mg/g,

respectively, in dried plant (Sahin et al., 2011). Cheung and Zhang (2008) reported that the concentration range of oleanolic acid, ursolic acid, and betulinic acid (Figure 2) was

0.3-0.9, 1.5-4.0, and 4.7-5.5 mg/g, respectively, in P. vulgaris samples collected from different areas of China. Rutin, quercetin, hyperoside, isoquercitrin, luteolin, cinaroside, kaempferol, and kaempferol-3-O- glucoside, etc (Figure 2) were identified from the methanol extract of P. vulgaris and the total amount of flavonoids was in the range of

2.16-10.29 % (Liao et al., 2008). 17

Figure 2. Phenolic acids, flavonoids, and triterpenoids in Prunella vulgaris

3.2 Health benefits of Prunella vulgaris

3.2.1 Anti-oxidant effects

P. vulgaris ethanol extract, containing 6.1% of major hydroxycinnamic acids, showed

significant anti-oxidative activities in terms of free radical scavenger effect on DPPH and

the anti-oxidative activity was partly in relation to the content of rosmarinic acid in this

plant (Lamaison et al., 1991). The aqueous extracts of P. vulgaris inhibited rat erythrocyte

hemolysis by 81 % and lipid peroxidation by 95% in rat kidney and brain homogenates at

200 µg/mL, as well as demonstrated strong superoxide- and hydroxyl radical-scavenging

activity by 61% at 100 µg/mL, but the content of the major constituents of P. vulgaris was

not determined in this study (Liu and NG, 2000).

3.2.2 Anti-inflammatory effects

Ethanol extracts of P. vulgaris at 30 µg/mL, rather than water extracts, significantly

inhibited production of LPS-induced inflammatory mediators PGE2 and NO by RAW

264.7 mouse macrophages (Huang et al., 2009). In this study, rosmarinic acid in P. vulgaris extract at 2.67 µM was found to independently inhibit inflammatory response, but it only partially explained the extract’s activity. LPS-induced cyclooxyginase-2 (COX-2) 18

and nitric oxide synthase (iNOS) protein expression were both attenuated by P. vulgaris

ethanol extracts, while rosmarinic acid only inhibited COX-2 expression. These authors

suggest that the inhibition exists at the transcriptional level and there could be compounds

other than rosmarinic acid in the extracts inhibiting the expression and/or the activity of the

two critical enzymes in inflammation.

The anti-allergic and anti-inflammatory constituents in P. vulgaris have been identified to

be triterpenes such as betulinic acid, ursolic acid, 2α, 3α-dihydroxy -ursoic acid, and

2α-hydroxyursolic acid. 2α, 3α-Dihydroxyursolic acid exhibited significant inhibition on

the release of β-hexosaminidase, a marker molecule for examining the degranulation

process in RBL-2H3 mast cells, in a dose-dependent manner and the IC50 value was 57 µM.

Ursolic acid and 2α-hydroxyursolic acid exhibited strong inhibition on the production of

NO from RAW 264.7 cells and the IC50 was 17 and 27 µM, respectively (Ryu et al., 2000).

Therefore, major constituents contained in P. vulgaris contribute to the anti-inflammatory

activities and NF-κB activation and COX-2 inhibtion might be the underlying mechanism.

3.2.3 Anti-mutagenic effect

Several triterpenoids including oleanic acid, 2α, 3α, 19α-trihydroxyursolic acid, and

maslinic acid inhibited A549 (non small cell lung adenocarcinoma), SK-OV-3 (ovarian

cancer cells), SK-MEL-2 (skin melanoma), and HCT15 (colon cancer cells), using the

sulforhodamin B bioassay to test the cytotoxicity, with the ED50 range between 3.7 -28 µM

(Lee et al., 2008). A water extracts of P. vulgaris at 50 µg/mL had a moderate

antimutagenic activity against picrolonic acid and benzo[a]pyrene induced mutation using

the Salmonella/microsomal system (Lee and Lin, 1988). Polysaccharide P32 (5 and 10

g/kg) isolated from P.vulgaris extract, consisted of rhamnose, arabinose, xylose, mannose, 19

glucose and galactose in a molar ratio of 3.46:49.32:58.91:0.43:2.64: 3.11, decreased the

number of tumors and weight by 64% and 55%, respectively, and significantly increased

the relative thymus weight to the body weight by 1.4 fold and the relative spleen weight to

the body weight by 1.3 fold in tumor-bearing mice (p < 0.05), indicating anti-tumor and

immunomodulation effects (Feng et al., 2010). In addition, the protective activity of

rosmarinic acid from Perilla frutescens on liver injury induced by LPS in

D-GalN-sensitized mice was examined (Osakabe et al., 2002). In this study, Perilla

extract (PE, 50-200 mg/kg in feed) and rosmarinic acid (135 mg/kg in feed) treatments

significantly reduced the elevation of plasma asparatate aminotransferase levels by 75%, but PE and rosmarinic acid treatments did not significantly diminish iNOS mRNA expression or plasma nitrate/nitrite levels as well as TNF-α mRNA expression both in liver

and plasma, indicating that liver protection of RA is due to the scavenging or reducing

activities-superoxide or peroxynitirite rather than to inhibition of TNF-α production.

Therefore, rosmarinic acid, triterpenoids and polysaccharides are responsible for the overall anti-tumor activites of P. vulgaris and the immunomodulatory and anti-oxidative effects may contribute to the anti-adenocarcinoma activity.

3.2.4 Anti-viral effects

The anti-HIV activity of P. vulgaris was reported by Tabba et al. (1989). The water

extract of P. vulgaris (50 µg/mL) significantly inhibited HIV-1 replication by blocking cell

to cell transmission of the virus, preventing syncytium formation, and interfering with the

ability of the virus to bind to CD4 receptor, suggesting that P. vulgaris extract antagonized

HIV-1 infection of susceptible cells by preventing viral attachment to the CD4 receptor

(Tabba et al., 1989). Moreover, anionic polysaccharides and triterpenoids from P.

vulgaris had anti-HSV activity in vitro. The antiviral activity (EC50) was about 66 µM, 20 20

µM and 10 µg/mL for betulinic acid, 2α, 3α-dihydroxyursolic acid and polysaccharides, respectively, measured by plaque reduction assay (Xu et al., 1999). In summary, studies have shown that P. vulgaris possesses a variety of pharmacological properties, and these novel properties may be utilized for future potential therapeutic applications.

3.3 Bioavailability of Prunella vulgaris chemical constituents

Most of the published papers were focusing on the bioavailability and pharmacokinetics of caffeic acid derivatives, triterpenoids, and flavonoids in P. vulgaris in human and cell culture models. Konishi and Kobayashi (2005) investigated transepithelial transport of rosmarinic acid, the major bioactive constituent in P. vulgaris, in intestinal Caco-2 cell monolayers and found that rosmarinic acid transferred across the intestinal barrier mostly by paracellular diffusion, like other esterified phenoclic acids, i.e. chlorogenic acid. The apparent permeability coefficient, Papp, for rosmarinic acid in Salvia miltiorrhiza ethanolic extract was 0.523 ± 0.301 × 10-6 cm/s (Lu et al., 2008). The effects of pH and the direction of permeation on the absorption of phenolic acids in Caco-2 cell system have been investigated, i.e. caffeic acid exhibited a pH-dependent directional transport from the apical to the basolateral side (Konishi et al., 2004). However, Lu et al. (2008) showed that there was no significant difference between Papp of apical to basolateral and basolateral to apical side for rosmarinic acid, a caffeic acid derivative. Meanwhile, the permeation of rosmarinic acid from the apical to the basolateral side in the presence of a proton gradient

(pH gradient 6.0/7.4) was nearly the same as that in the absence of a proton gradient (pH gradient 7.4/7.4), implying that proton coupled polarized transport was not involved for rosmarinic acid. In in vivo studies, Jirovsky et al. (2007) showed that after 91 days of feeding in pigs with methanol extract of P. vulgaris, rosmarinic acid and its metabolites were found in plasma at >0.08 μM after the hydrolysis by β-glucuronidase/arylsulfatase to 21

release the parent compound. Taken together, intact rosmarinic acid is poorly absorbed in the small intestine.

The uptake of triterpenoids such as oleanolic acid and betulinic acid has been studied both

-6 in vitro and in vivo. Papp value of oleanolic acid in Caco-2 cells was 1.1–1.3 × 10 cm/s.

The absolute oral bioavailability was 0.7% for oral doses of 25 and 50 mg/kg of oleanolic acid in rats, indicating that oleanolic acid is also poorly absorbed and might be extensively metabolized in vivo (Jeong et al., 2007). The uptake kinetics of 2,3-hydroxybetulinic acid in Caco-2 cells followed a one-phase exponential association process, reached near saturation after approximately 60 min and a maximal intracellular concentration of about

906.7 uM/mg protein, indicating the facilitated absorption of this particular compound

(Zheng et al., 2007).

The absorption of flavonoids has been studied. Walgren et al. (1998) found out that the

-6 reverse basolateral to apical flux of quercetin (Papp 11.1 ± 1.2 x 10 cm/s) was almost 2-fold

higher than the apical to basolateral flux (p < 0.001). Quercetin 4-glucoside demonstrated

-6 no absorption but did demonstrate basal to apical flux (Papp 1.6 ± 0.2 x 10 cm/s).

Quercetin-3, 4'-diglucoside showed a low apical to basolateral transport (Papp 0.09 ± 0.03 x

10-6 cm/s) and its reverse, basolateral to apical, transport was, however, 4-fold higher (p <

0.05). These results suggest that facilitated absorption of quercetin through the human

intestinal epithelium might be involved. Wang et al. (2005) also reported that quercetin,

kaempferol and isorhamnetin displayed polarized transport, with the Papp, basolateral-apical being higher than the Papp apical-basolateral in Caco-2 cells.

Meanwhile, quercetin and kaempferol (50 µM for each) inhibited the ATPase activity, and isorhamnetin (50 µM) stimulated the ATPase activity (p < 0.05) and the concentrations of 22

these flavonoids were significantly increased when incubated with the P-gp inhibitor verapamil (p < 0.05). The results indicated that quercetin, kaempferol and isorhamnetin were substrates of P-gp and the P-gp type efflux pump might limit the bioavailability of these flavonoids.

3.4 Metabolism of Prunella vulgaris

A major limiting step in the utilization of bioactive compounds from P. vulgaris is their intestinal absorption and metabolism. Phase I and II metabolites are often, but not always, less bioactive than parent compounds. The absorption or metabolism of rosmarinic acid has been examined in vivo to a limited extent (Baba et al., 2004; Konishi et al., 2005; Baba et al., 2005). When Perilla extract containing 200 mg of rosmarinic acid was orally administered to six men, rosmarinic acid in both plasma and urine was present predominantly as glucuronide and/or sulfate conjugated forms, at 0.6 ± 0.2% and 1.5 ± 0.4% of the total intake, respectively, within 48 h after ingestion (Baba et al., 2005). Rosmarinic acid (1.5%), methylated rosmarinic acid (2.9%), caffeic acid (1.0%), (1.0%) and a trace of m-coumaric acid (0.02%) were detected in the urine after 48 h intake of the plant extract (Baba et al., 2004). Therefore, ingested rosmarinic acid was absorbed and metabolized into conjugated and/or methylated forms and degraded to other caffeic acid derivatives.

Gut microbes may metabolize rosmarinic acid to give phenolics such as caffeic acid, o-coumaric acid, 3, 4-dihydroxyphenylpropionic acid (DHPP) and m-hydroxyphenylpropionic acid (mHPP), in which m-coumaric acid and mHPP were absorbed by monocarboxylic acid transporter (MCT)-mediated active processes, whereas 23

DHPP was mainly permeated across Caco-2 cells via the paracellular pathway (Konishi

and Kobayashi, 2005).

The metabolism of the flavonoids has been investigated both in vitro and in vivo.

Quercetin was metabolized to several substances in human leukaemia (HL-60) cells, such as quercetin glucuronides, quercetin-sulfate, myricetin by hydroxylation, isorhamnetin by methylation, and quinone form via oxidation (Ludwig-Müller et al., 2005). Another study investigated the effect of quercetin and its metabolites on paraoxonase 2 levels in

RAW264.7 macrophages (Boesch-Saadatmandi et al., 2009). Intracellular antioxidant enzyme paraoxonase 2 (PON2) may have a protective function in the prevention of atherogenesis. An enhancement of PON2 activity by dietary factors including flavonoids is therefore of interest. In this study, supplementation of RAW264.7 macrophages in culture with quercetin at 1, 10, and 20 µM resulted in a significant increase in PON2 mRNA and protein levels, as compared to untreated controls. Unlike quercetin, its glucuronidated metabolite quercetin-3-glucuronide did not affect PON2 gene expression in

cultured macrophages. However the methylated quercetin derivative isorhamnetin

enhanced PON2 gene expression in RAW264.7 cells to a similar extent as quercetin.

These results indicate that circulating flavonoid metabolites possess biological properties different from their nonconjugated parent compounds, further proving the importance of

understanding the metabolic fate of the bioactive constituents via Phase I and II biotransformation as well as gut microbial metabolism.

24

4. OVERALL INTRODUCTION OF CAFFEIC ACID DERIVATIVES

4.1 Source and intake of caffeic acid derivatives

Caffeic acid (3, 4-dihydroxycinnamic acid), one of the most common phenolic acids,

frequently occuring in fruits, vegetables, grains, dietary supplements and traditional herbs as simple esters with or saccharides (caffeic acid derivatives). Caffeic acid is the main phenolic acid aglycone in potatoes, with contents varying from 0.3 to 3.6 mg/100 g and 18.8 to 28 mg/100 g in tubers and peels, respectively (Radtke et al., 1998).

Chlorogenic acid is an ester of caffeic and quinic acids in coffee and many types of fruits, coffee, tea and traditional herbs (Clifford, 1999). Coffee consumers ingest 0.5-1 g chlorogenic acid/d (Clifford et al., 1999). In another study, caffeic acid intake was found to be 206 mg/d, and the principal sources were coffee (92% of caffeic acid) and fruit and

fruit juices combined (Radtke et al., 1998). Indeed, coffee consumption was shown to be

positively correlated to caffeic acid and chlorogenic acid intake, indicating that they can be

used as biomarkers for the estimation of dietary intake (Mennen et al., 2006).

Rosmarinic acid is a caffeic acid derivative found in many Lamiaceae herbs used

commonly as culinary herbs such as lemon balm, rosemary, oregano, sage, thyme and

peppermint (Lamaison et al., 1991). The first plant cell cultures found to be accumulating

rosmarinic acid were derived from Coleus blumei (Razzaque and Ellis, 1977). Suspension

cultures from this species were the first ones used for an attempted biotechnological

production of rosmarinic acid since they were found to accumulate up to 21% rosmarinic

acid in the dry weight. Basil, lemon thyme, mint, oregano, rosemary, sage, and thyme

were investigated to contain 0.22-0.97% of rosmarinic acid (Park, 2011).

25

Caffeic acid derivatives were metabolized by both human and animal gut microbes.

Rosmarinic acid (Konishi and Kobayashi, 2005; Yoshikawa et al., 2005), chlorogenic acid

(Konishi and Kobayashi, 2005), and cichoric acid (Peppercorn et al., 1971) were cleaved by microbial esterase, hydrolyzed to caffeic acid and then reduced to dihydrocaffeic acid (3,

4dihydroxyphenylpropionic acid); furthermore the later was dehydroxylated to m-hydroxyphenylpropionic acid (mHPP). After administration of chlorogenic acid at 250

μmol/day in rats for 8 days, total urinary excretion of caffeic, ferulic, and isoferulic acids was 28.1 % of intake (mol/mol); urinary mHPP was 4.0 %. Plasma metabolite concentrations in rats fed caffeic acid (250 μmol/day) for 8 d were caffeic acid (41.3 μM), ferulic acid (7.3 μM), isoferulic acid (4.5 μM), hippuric acid (54.2 μM), and mHPP (1.4 μM)

(Gonthier et al., 2003). When Perilla extract containing 200 mg of rosmarinic acid (RA) was orally administered to six men, RA in both plasma and urine was present predominantly as glucuronide and/or sulfate conjugated forms, at 0.6 ± 0.2% and 1.5 ± 0.4% of the total intake, respectively, within 48 h after ingestion (Baba et al., 2005). Gut microbes metabolized RA to give phenolics such as caffeic acid, o-coumaric acid and m-hydroxyphenylpropionic acid, which are then absorbed by monocarboxylic acid transporter (MCT)-mediated active processes (Konishi and Kobayashi, 2005). In summary, caffeic acid derivatives are ingested by humans and metabolized to various metabolites, which may contribute to the health benefits.

4.2 Anti-inflammatory activities of caffeic acid derivatives

Caffeic acid phenethyl ester, at 25μg/mL, has been shown to suppress NF-κB activation by suppressing the binding of the p50-p65 complex directly to DNA in human histiocytic

U937 cells (Natarajan et al., 1996). Nuclear factor kappa-B (NF-κB) was a heterodimeric transcription factor with a pivotal role in orchestrating immune and inflammatory 26

processes. In this study, the activation of NF-κB by TNF-α was completely blocked by caffeic acid phenethyl ester in human histiocytic cell line U937 cells, which are used to study the behaviour and differentiation of monocytes (Natarajan et al., 1996). In another study, caffeic acid and the predominant isomer of chlorogenic acid, 5-caffeoylquinic acid, significantly inhibited the H2O2 or TNF-α induced interleukin (IL-8) secretion and its mRNA expression in a dose-dependent manner in the range of 0.25-2.00 mM in Caco-2 cells (Zhao et al., 2008). IL-8 is a central pro-inflammatory chemokine involved in the pathogenesis of IBD.

Chlorogenic acid (5-20 μM) also showed suppression of IL-1β-induced activation of

NF-κB, degradation of inhibitor of κB (IκB)-α, as well as matrix metalloproteinases

(MMP-1, MMP-3, MMP-13) expressions in chondrocytes isolated from rabbits (Chen et al.,

2011). IL-1β is a pivotal catabolic factor in inflammatory cascades and induces inflammatory mediators and MMPs, leading to disruption of the balance between biosynthesis and degradation of the extracellular matrix (Chen et al., 2011). Chlorogenic acid (50 mg/kg, intraperitoneally) suppressed hepatic mRNA expression of toll-like receptor 4 (TLR4), TNF-α and phosphorylation of NF-κB p65 subunit in LPS-induced liver injury in mice (Xu et al., 2010). TLR4, one of the members of the TLR protein family, is essential for LPS-mediated signaling. After binding to TLR4 and CD14, LPS triggers several critical intracellular signaling pathways, including NF-κB signaling cascade, and leads to the production of inflammatory cytokines including TNF-α, INF-γ and nitric oxide

(Takeuchi et al., 1999).

Rosmarinic acid, at 1μM, reduced serum levels of TNF-a, IL-6, high-mobility group box 1 protein, while increased serum level of IL-10 in rats induced by cecal ligation and puncture. 27

Amelioration of hemodynamics and decrease in myeloperoxidase in lung, liver, and small intestine were also observed after rosmarinic acid injection (Jiang et al., 2009). In another study, rosmarinic acid in P. vulgaris (2.67 µM) was found to independently inhibit

LPS-induced inflammatory mediators PGE2 and NO by RAW 264.7 mouse macrophages, via COX-2 inhibition (Huang et al., 2009). These results suggest that caffeic acid derivatives, including caffeic acid, chlorogenic acid and rosmarinic acid, have the capacity to inactivate inflammatory response and the anti-inflammatory mechanism may inhibit activation of the NF-κB pathway by inhibiting IκB kinase activity.

5. INTESTINAL ABSORPTION, METABOLISM, INFLAMMATORY

RESPONSES, AND RELEVANT METHODOLOGY

5.1 Intestinal function

5.1.1 Absorption mechanisms

Dietary bioactive compounds pass across the intestinal epithelium via passive and active mechanisms. In passive transport, the cells do not use any energy to move the molecules and the molecules move through the concentration gradient (higher concentration to lower concentration). Passive diffusion can be divided into either transcellular diffusion

(through the cells) or paracellular diffusion (intercellular spaces). Transcellular passage includes the partition between the aqueous contents of intestinal lumen, the phospholipid bilayer of the apical cell membranes, the aqueous contents of the enterocytes, and again, the phospholipid bilayer of the basolateral cell membranes (Borchardt, 1998). The compound via paracellular pathway is usually polar or hydrophilic and small (Hidalgo, 1996) and the transcellular pathway is more common route, since the cellular absorbing surface is over

1000 times greater than the area of the paracellular spaces (Pappenheimer and Reiss, 1987).

The chemical and physical properties of the compound are critical for the permeability 28

across the epithelial cells. For example, chlorogenic acid and rosmarinic acid are

considered to pass across the epithelial barrier via paracellular pathway, while alkamides

transport via transcellular pathway (Konishi and Kobayashi, 2005; Matthias et al., 2005;

Konishi and Kobayashi, 2004). Moreover, caffeic acid conjugates (caftaric acid, cichoric acid and echinacoside) were not detected in any plasma sample at any time after ingestion

of Echinacea ethanolic extract, whereas alkamides were in the range of 0.1-120 ng/mL in

plasma 20 minutes after ingestion (Matthias et al., 2004), which might be because that

alkamides are more hydrophobic than caffeic acid derivatives, thus more bioavailable.

Membrane transporters located in the cell membranes are responsible for active transport

of nutrients. Active transport is the energy-demanding transfer of a nutrient across a cell

membrane against its concentration gradient (lower concentration to higher concentration,

Gropper et al., 2008). The energy for active transport comes from ATP generated by

respiration (in mitochondria). The active transport can be classified into facilitated

diffusion, primary active and secondary active transport, based on the energy requirement

(Amidon and Sadée, 1999). Facilitated diffusion does not require energy for its function

and specific molecules move down a concentration gradient, passing through the

membrane via a specific carrier protein. Thus, rather like enzymes, each carrier has its

own shape and only allows one molecule (or one group of closely related molecules) to

pass through. Primary active transport is an energy demanding process and the energy is

gained from hydrolysis of ATP to ADP from the high-energy phosphate bond. Secondary

active transport uses energy from ion gradients (mostly Ca2+, Na+, and H+ -gradients) across the cell membranes generated by primary active ion pumps (for example

Na+/K+-ATPase), and are also called symporters or antiporters (Amidon and Sadée, 1999).

Characteristics for active transport are the saturability of the carrier protein and 29

temperature dependency. Several absorptive transport systems with the primary function

of transporting nutrients (amino acids, oligopeptides, monocarboxylic acids, monosaccharides, organic anions and cations, bile acids and several water soluble vitamins)

are present in the small intestine (Steffansen et al. 2004). Caffeic acid and its colonic microbial metabolites, such as o-coumaric acid , ferulic acid and m-hydroxyphenylpropionic acid (mHPP) transport across the intestinal epithelial cells via

monocarboxylic acid transporter (MCT, Konishi and Kobayashi, 2004), indicating that a

monoanionic carboxyl group and a non-polar side chain or aromatic hydrophobic moiety are thought to be necessary components of a substrate for MCTs. Moreover, the major intestinal glucose transporters, such as Na+-dependent glucose transporter SGLT1

(SLC5A1), the facilitative transporter GLUT2 (SLC2A2), and GLUT4 (SLC2A4), have

been regarded as specific carriers for the transport of some flavonoids, such as quercetin,

quercetin glycosides, myricetin, catechin-gallate, and genistein (Chen et al., 2007; Strobel, et al., 2005). Therefore, the membrane transfer of phytochemicals is a fundamental part of their bioavailability, and the uptake mechanisms of phytochemicals could be compound-specific, which depends on the compound structure and conjugated moieties.

5.1.2 Metabolism

Although the human small intestine serves primarily as an absorptive organ, it also has the

ability to metabolize nutrients, drugs and other molecules. Many of the enzymes involved in Phase I and II reactions in the human liver have also been detected within intestinal epithelial cells. Cytochrome P450 (CYP) enzymes are located at the villous tip just beneath the brush border at the apex of the enterocyte, the main site of absorption for orally administered compounds and CYP content in human small intestine ranges from 20 to 210 pmol/mg, implying that significant variability in intestinal CYP expression: total CYP 30

content slightly increases from the duodenum to the jejunum and then decreases toward the

ileum (Thelen and Dressman, 2009). The goblet cells and the epithelial cells of the crypts

between the villi contain lower levels of drug metabolizing enzymes. As mentioned

before, alkamides can be metabolized by CYPs to form epoxides, and N-dealkylated and

hydroxylated products (Matthias et al., 2004). The most active hepatic CYP isoforms

involved in the production of epoxides were CYP1A1, CYP1A2 and CYP2A13, whereas

for the N-dealkylation only CYP1A1, CYP1A2 and CYP1B1 showed significant activity.

Of these forms CYP 1A1 was detected in human intestinal samples, although at least at low

levels (3.6–7.7 pmol/mg) while CYP1A2, is an almost exclusively hepatic CYP enzyme,

being virtually undetectable in human small intestine (Paine et al., 2006; Shimadat et al.,

1994). More interestingly, the 2, 4-diene alkylamides were more rapidly metabolized

when examined as the pure compound rather than in the Echinacea extract, suggesting that

some of the other consistuents present in the extract might have exerted an inhibitory effect

towards their metabolism by CYPs.

Among Phase II enzymes, UDP-glucuronosyltransferase (UGT) and sulfotransferase

(SULT), which catalyze sulfation and glucuronidation, may play important roles in the

conjugation and ultimately excretion and elimination of many phytochemicals containing

hydroxyl functional group either present in the parent structure and/or after biotransformation by the phase I enzyme such as the CYPs (Rushmore and Kong, 2002).

Some flavonoids, such as baicalein, quercetin, and methylated quercetin, are substrates of

UGT1A3 and UGT 1A9 and the preferred substrates for UGT1A3 and UGT1A9 contain the hydroxyl group at the C7-position (Xie et al., 2010). Moreover, luteolin in Artemisia afra aqueous extract, regardless of its form (i.e. whether aglycone and 7-0-glucoside), was taken up better and more efficiently metabolized to luteolin monoglucuronide, sulfate and 31

methylated luteolin sulfate conjugates than the aglycone and 7-0-glucoside forms

administered as pure solutions in Caco-2 cells (Mukinda et al., 2010), implying the

necessity to investigate the effect of the plant matrix on the uptake and metabolism of

phytochemicals.

Therefore, the small intestine plays a key role in first-pass metabolism of orally ingested

phytochemicals, which is supported by the position of the small intestine as the first site of exposure of phytochemicals to metabolic systems, by the large surface area available in the small intestine for absorption and subsequent metabolism, and by the transpoters located both at apical and basolateral sides of the enterocytes involved in the transfer or efflux of

the substrates.

Transporters, such as P-glycoprotein (P-gp), multidrug resistance-associated protein

(MRP), and organic anion transporting polypeptide (OATP) are expressed in intestine,

where they provide a formidable barrier against drug penetration, and play crucial roles in

drug absorption, distribution, and excretion (Xu et al., 2005, Figure 3). Because P-gp and

MRP utilize ATP to transport substrate the cell membrane, they are called ATP binding

cassette (ABC) transporters, which import or export a broad range of substrates, such as

amino acids, ions, sugars, lipids, etc. In human, 46 ABC transporters have been identified

(Chang, 2003). All ABC transporters are composed of two nucleotide binding domains

(NBDs) and two transmembrane domains (TMDs). The role of TMD is to recognize and

mediate the passage of substrates across the cell membrane (Chang, 2003). First-pass

metabolism can be affected by the interaction between CYP isoforms, specifically CYP

3A4, and P-gp in enterocytes (Patel and Mitra, 2001, Figure 3). Since CYP3A4 and P-gp are both localized to mature enterocytes on the villus tip, the extensive overlap among the 32

substrates for P-gp and CYP3A4, and pregnane X receptor (PXR) mediates co-induction of

CYP3A4 and P-glycoprotein, thus P-gp and CYP3A4 could be coordinately regulated, and substrate for both P-gp and CYP3A4 may be absorbed directly into the systemic circulation, metabolized by CYP3A4 in the enterocyte or secreted back into the intestinal lumen by

P-gp (Patel and Mitra, 2001). Therefore, P-gp-mediated efflux and CYP3A4 metabolism may be functionally interrelated and act accordingly to limit the passage of the molecules across the enterocyte.

However, some flavonoids, such as quercetin (Scambia et al., 1994), myricetin (Choi et al.,

2010) and genistein (Sergent et al., 2009) have been shown to inhibit P-gp-mediated efflux and CYP3A4 both in Caco-2 cells and animal models. These results underline the importance of intestinal studies to assess transporter activity linked to the ingestion of phytochemical- enriched supplements.

Figure 3. Transporters in intestinal epithelial cells 33

Note: BCRP: breast cancer resistance protein ; LRP: lung resistance-related protein; MRP:

multidrug resistance- associated protein; P-gp: P-glycoprotein; MCT: monocarboxylic acid

transporter; hPepT1: human peptide transporter 1.

5.2 Inflammatory responses of the intestine

Besides the absorptive and metabolic capabilities, the intestine mucosa membrane

exercises a truly antitoxic and protective function with regard to the microbial poisons.

Changes to this mucosal barrier can affect the absorption of both exogenous and

endogenous nutrients. Many diseases, such as inflammatory bowel disease (IBD),

ischemic disease, and graft-versus-host disease, are associated with the loss of intestinal

mucosal barrier function (Clayburgh et al., 2004; Chen et al., 2003; Hill et al., 2000).

Therefore, an intact intestinal mucosal barrier is of great importance for both ensuring

adequate provision of dietary nutrients to the whole body and preventing gut-related diseases.

An intact mucosal barrier depends on intercellular junctions, which help to seal the space between adjacent epithelial cells (the paracellular space), and tight junctions, which are the key elements of the seal (Turner, 2006). Tight junctions play a key role in the intestinal physical barrier to the diffusion of pathogens, toxins, and allergens from the lumen into the circulatory system. Tight junctions are multi-protein complexes composed of transmembrane proteins (i.e. occludin, claudins and junctional adhesion molecule (JAM)), peripheral membrane (scaffolding) proteins (i.e. zonula occludens (ZO)), and regulatory molecules that include kinases (Figure 4).

The most important of the transmembrane proteins are members of the claudin family, which not only define tight junction permeability and affect barrier function, but also 34

enhance cell proliferation and regeneration, as might be necessary to compensate for cell

loss in colitis. For example, increased claudin-1 and claudin-2 expression has been

reported in intestinal epithelial cells of patients with IBD (Heller et al., 2005). Meanwhile,

claudin-1 has also been shown to enhance neoplastic transformation, tumor growth and

metastasis in experimental models (Dhawan et al., 2005). Weber et al. (2008) reported

that claudin-1 and claudin-2 expression was elevated in active IBD, adenomas, and

IBD-associated dysplasia, but not acute, self-limited colitis (ASLC). In contrast,

claudin-4 expression was elevated in both active IBD and ASLC. Thus, increased claudin

expression may contribute to carcinogenesis in IBD.

At least two routes allow transport across the tight junction: leaky pathway, which allows

paracellular transport of large solutes, such as limited flux of proteins and bacterial LPS, and small pore pathway that is defined by tight junction-associated claudin proteins, which are primary determinants of charge selectivity and exclude molecules larger than 4 Å

(Turner et al., 2009). Thus, tight junctions show both size selectivity and charge selectivity, and these properties may be regulated individually or jointly by physiological or pathophysiological stimuli.

Peripheral membrane proteins, such as ZO-1 and ZO-2, are crucial to tight junction assembly and maintenance, because recent work suggests that ZO-1, which interacts directly with actin, occludin, claudins and other proteins, may be an essential effector of perijunctional actomyosin ring-mediated tight junction regulation (Van et al., 2009). In deed, Ca2+-calmodulin-dependent serine–threonine protein kinase myosin light chain

kinase (MLCK), which phosphorylates myosin II regulatory light chain (MLC) within the

perijunctional actomyosin ring to activate myosin ATPase activity, is essential for

Na+-nutrient co-transport-induced tight junction regulation (Scharl et al., 2009). Occludin, 35

a transmembrane tight junction protein also interacts directly with claudins and actin.

However, although numerous in vitro studies demonstrated a role for occludin in tight junction function, occludin is an integral part of tight junction strands (Fujimoto, 1995), while not essential for strand assembly because intestinal barrier function is intact in occludin deficient embryonic stem cells and occludin-deficient mice (Saitou et al., 2000;

Saitou et al., 1998).

Figure 4. Intestinal epithelial intercellular junctions

Note: ZO: zonula occludens

In IBD, the paracellular space has increased permeability, and the regulation of tight junctions is defective (Wang et al., 2006; Bruewer et al., 2003). The hallmark of active

IBD is a pronounced infiltration into the lamina propria of innate immune cells (neutrophils, macrophages, dendritic cells, and natural killer T cells) and adaptive immune cells (B cells and T cells). Increased numbers and activation of these cells in the intestinal mucosa elevate local levels of tumor necrosis factor α (TNF-α), interleukin-1β (IL-1β), interferon-γ

(IFN-γ), and cytokines of the interleukin-23–T helper 17 pathway (Abraham and Cho,

2009). For example, in Crohn’s disease, there is increased production of the Th17 36

cytokine interleukin-17 and the Th1 cytokines interferon-γ and TNF-α in the intestinal

mucosa. In ulcerative colitis, by contrast, there is usually an increase in interleukin-17 and

Th2 cytokines, such as IL-4, 5 and 13 (Abraham and Cho, 2009).

The ability of cytokines, such as TNF-α and IFN-γ, to regulate the function of the tight

junction barrier was first described 20 years ago (Madara et al., 1989). Myosin light chain

kinase (MLCK) has been shown to have a central role in TNF-α induced epithelial and

endothelial barrier dysregulation. Similar to Na+-nutrient co-transport, TNF-α-induced

MLCK activation seems to increase paracellular flux through the leak pathway.

IFN-γ-induced occludin internalization is mediated by myosin ATPase-dependent

macropinocytosis and occludin cleavage may even modify the barrier to enhance

transepithelial migration of inflammatory cells in the lung (Chun et al, 2009). Moreover,

IL-1β levels are markedly elevated in intestinal tissues in patients with IBD (Turner et al.,

2005). Recent studies have shown that IL-1β at physiologically relevant concentrations

(1-10 ng/ml) causes an increase in intestinal epithelial tight junction permeability, which is modulated in part by regulation of MLCK (Al-Sadi et al., 2007). The IL-1β induced increase in Caco-2 tight junction permeability was preceded by an increase in MLCK mRNA and protein levels; and targeted inhibition of IL-1β induced MLCK mRNA or

MLCK protein expression prevented the increase in tight junction permeability (Al-Sadi et

al., 2008). The IL-1β induced increase in retinal pigment epithelial tight junction

permeability was associated with a down-regulation of occludin protein and increase in

claudin-1 expression (Abe et al., 2003). Other investigators have also shown that IL-1β

causes an increase in claudin-1 and claudin-2 expressions and decrease in occludin expression in human astrocytes , Caco-2 cells and rat hepatocytes (Duffy et al., 2000;

Yamamoto et al., 2004; Al-Sadi et al., 2008). Therefore, in vitro research on intestinal 37

epithelial cells has mainly focused on the involvement of aforementioned cytokines- IL-1β,

TNF-α and IFN-γ, which activate intracellular cascades, increase transcriptional activity and the secretion of other cytokines and chemokines, as well as increase the paracellular permeability though defects in tight junctioning or assembly.

Therefore, it is critical to investigate the roles of the epithelial cell barrier in health and disease, with particular emphasis on the junctional complexes, which have a crucial role in barrier regulation. Previous studies showed that some polyphenols, such as epigallocatechin gallate, genistein, myricetin, kaempferol and quercetin, enhance intestinal tight junctions in human intestinal Caco-2 monolayers (Wells et al., 1999; Suzuki and Hara,

2010; Suzuki et al., 2011). Caffeic acid phenethyl ester inhibited the down-regulation of occludin, a transmembrane tight junction protein, induced by TNF-α in astrocytes, possibly via NF-κB pathway because NF-κB could act as a negative regulator of occludin expression in astrocytes (Wachtel et al., 2001). But very few studies have investigated the influences of phenolic acids on gut barrier functions, which might be the underlying mechanism of the protective effects of caffeic acid derivatives against colitis or colon cancer in animal models (Ye, et al., 2009; Tanaka, et al., 1990; Feng et al., 2010). Thus, our research will focus on the modulation of caffeic acid derivatives from traditional herbs on intestinal barrier function.

5.3 Relevant methodology to study intestinal function

It is important to determine the mechanism of the intestinal absorption of the ingested molecules and the interaction between the targeted compound and the intestinal epithelial cells, and hence it is vital to have good predictive models which offer information about the 38

fraction of compound absorbed across the intestinal wall, the amount of metabolism during

absorption as well as the influence of the compound on the intestinal function.

5.3.1 Caco-2 cells

Caco-2 cell line was developed from a moderately well differentiated colon adenocarcinoma obtained from a 72-year-old male (Fogh et al. 1977). Caco-2 cells

differentiate spontaneously and exhibit structural and functional differentiation patterns

characteristic of mature enterocytes (Pinto et al. 1983). Caco-2 cells reach confluency

within 4-6 days and reach fully differentiated status within 20 days (Pinto et al. 1983).

The differentiated cells exhibit structural and functional differentiation patterns similar to

mature human enterocytes, such as microvilli of the brush border membrane, tight

intercellular junctions, and brush border associated enzymes, Phase I and II metabolizing

enzymes as well as transporters. The structural and functional differentiation of the

microvilli is associated with the polarization of the monolayer after confluency.

Some parameters, such as transepithelial electrical resistance (TEER) or transport of the

hydrophilic molecule lucifer yellow or mannitol have been used as indicators of the

epithelial integrity. Caco-2 monolayers have a TEER between 250 and 800 Ω•cm2 which

is similar to the TEER in the colon rather than that in the small intestine which is lower

(Biganzoli et al., 1999). Indeed, fully differentiated Caco-2 cells form an epithelial

membrane with a barrier function similar to the human colon (Artursson et al., 1993) but

express carrier proteins similar to the small intestine (Baker and Baker 1992; Hidalgo et al.

1989). An increase in the transport of hydrophilic paracellular marker indicates increases

in the cell monolayer permeability (Hubatsch et al., 2007). These parameters can be used

to evaluate cytotoxicity of nutrients, drug and other compounds. 39

Bidirectional studies (apical to basolateral and basolateral to apical transport) in Caco-2

cells enable the investigation of the mechanisms of passive paracellular and passive

transcellular permeability, carrier mediated absorptive transport and efflux, as well as

active transport (Artursson et al., 1993). The apical compartment is used to mimic the

intestinal lumen and the basal compartment mimicking the blood stream.

Caco-2 cells express phase I (i.e. CYPs) and phase II detoxification enzymes (i.e. gluthathione S-transferase (GST), UGT and ST). Because Caco-2 cells contain many of the metabolic enzymes found in the intestinal epithelium, they have been used to evaluate the metabolism of compounds during the transport across the intestinal barrier (Meunier et al., 1995). Meanwhile, Caco-2 monolayers also express several absorptive and efflux transporters, such as P-gp, MRP, OATP, monocarboxylic transporter (MCT), breast cancer resistance protein transporter (BCRP), etc (Laitinen, 2006). Thus the model can be useful to screen a large number of compounds liable to be substrates for the transporter, i.e. P-gp, and to define the role of the transporter in limiting the absorption of the applied compounds.

Caco-2 cells also have limitations. First, the high TEER due to the “tighter” monolayer compared with in vivo situation leads to very low permeability for hydrophilic, mostly paracellularly permeating compounds. Caco-2 cells are better to study transcellularly passively permeating drugs (Laitinen, 2006). Second, Caco-2 cells are a cancer-derived line and therefore may have properties different from normal cells. Third, Caco-2 cells lack the crypt-villus axis, which is important for fluid and ion transport in vivo, and lack the mucus producing goblet cells, which leads to the lack of a prominent mucus layer 40

(Artursson 1996). More recently, mixed cultures of Caco-2 cells with mucin-producing

cells, i.e. HT-29, have been used in attempts to overcome this limitation of the method.

In conclusion, the major advantage of cell culture models is that we don’t need to consider

about the interspecies differences in the morphological and physiological characteristics, which are normal issues in animal models. So the cell culture model (such as Caco-2) has been adopted enthusiastically by the pharmaceutical industry, for the screening of compounds for intestinal absorption. It can provide comparative information, on the intestinal absorption, metabolism and the mucosal toxicity of applied molecules.

Moreover, in vivo absorption studies performed with laboratory animals and humans are expensive and time consuming and can pose ethical challenge, thus cell model is a useful

alternative to study uptake and transport of important compounds.

5.3.2 Paracellular permeability test

Paracellular permeability was measured by assessing transepithelial electrical resistance

(TEER) and using the transfer of a paracellular marker, lucifer yellow (LY), across Caco-2

monolayers (Hubatsch et al., 2007). After 21 d post seeding the cells on the transwell, the

transepithelial electrical resistance (TEER) was tested by Millicell ERS meter to reflect the

tightness of intercellular junctions and only cells with TEER ≥ 250 Ω·cm2 were used for

permeability study. As mentioned before, Caco-2 monolayers have a TEER between 250

and 800 Ω•cm2 which is similar to the TEER in the colon, and greater than that in the small intestine (Biganzoli et al., 1999). The transfer of a paracellular transferred compound is correlated with the TEER change: the greater the TEER, the tighter the cell monolayer, the less transfer of the paracellularly transferred compound, i.e. rosmarinic acid and chlorogenic acid (Konishi and Kobayashi, 2004). Further, the paracellular marker, LY 41

dissolved in Hank's Buffered Salt Solution was added to the apical chamber. For LY

quantification, the apical and basolateral solutions were collected after 1h, transferred to a

96 well plate and read spectrophotometrically at 450 nm. Previous studies used both

TEER and LY transfer to investigate the enhancement of intestinal tight junctions in

Caco-2 cells by some polyphenols, such as epigallocatechin gallate, genistein, myricetin,

kaempferol and quercetin (Suzuki and Hara, 2010; Suzuki et al., 2011). Therefore, TEER

and LY are often used together to reflect the tightness of the monolayer and the paracelluar

permeability.

5.3.3 Western blot, immunofluorescence and ELISA

Western blot is a widely used analytical technique for detecting specific proteins in the

given sample of tissue homogenate or extract. Gel electrophoresis was used first to

separate denatured proteins by the size of the polypeptide. The proteins were then

transferred to a membrane (typically nitrocellulose or polyvinylidene fluoride, PVDF),

where they were probed (detected) using antibodies specific to the target protein. Tight junction protein expression has been studied extensively using Western blotting (Al-Sadi et al., 2008; Liu et al., 2010; Mukinda, et al., 2010; Suzuki et al., 2011). Meanwhile, immunofluorescence uses the specificity of antibodies to their antigen to target fluorescent dyes to specific biomolecule targets within a cell, and therefore allows visualisation of the distribution of the target molecule through the sample. Suzuki et al. (2009) and Amasheh

et al. (2008) performed Western blot and immunofluorescence to measure the enhancement

of intestinal barrier function by quercetin through the assembly of zonula occluden (ZO)-2,

occludin, and claudin-1 and the expression of claudin-4 in Caco-2 cells. Further, cytokine

secretion, before and after treatment with stimuli to initiate the inflammatory response, was

evaluated using a sandwich ELISA method and quantified using the standard provided with 42

the kit (Van De Walle et al., 2010). Results were subsequently expressed in relative terms to the negative control (untreated cells) to facilitate comparison between groups.

In conclusion, the overarching hypotheses in my study were that alkamides, ketones and caffeic acid derivatives were transferred by Caco-2 cell monolayers independent of plant matrix; and caffeic acid derivatives enhanced the intestinal barrier function. We used

Caco-2 cell with HPLC detection to investigate the permeabilities and Phase II biotransformation of botanical constituents, i.e. alkamides, ketones, rosmarinic acid and ursolic acid, as pure compounds and in plant extracts. Western blot and immunofluorescence were used together to study the targeted tight junction protein expression and distribution, and ELISA was used to measure the cytokine excretion by

Caco-2 cells with and without the stimuli treatment.

43

6. REFERENCES

Abe, T., E. Sugano, Y. Saigo & M. Tamai (2003) Interleukin-1beta and barrier function of

retinal pigment epithelial cells (ARPE-19): aberrant expression of junctional

complex molecules. Invest Ophthalmol Vis Sci, 44, 4097-104.

Abraham, C. & J. Cho (2009) Interleukin-23/Th17 pathways and inflammatory bowel

disease. Inflamm Bowel Dis, 15, 1090-100.

Amasheh, M., S. Schlichter, S. Amasheh, J. Mankertz, M. Zeitz, M. Fromm & J. D.

Schulzke (2008) Quercetin enhances epithelial barrier function and increases

claudin-4 expression in Caco-2 cells. J Nutr, 138, 1067-73.

Al-Sadi, R., D. Ye, K. Dokladny & T. Y. Ma (2008) Mechanism of IL-1beta-induced

increase in intestinal epithelial tight junction permeability. J Immunol, 180,

5653-61.

Al-Sadi, R. M. & T. Y. Ma (2007) IL-1beta causes an increase in intestinal epithelial tight

junction permeability. J Immunol, 178, 4641-9.

Amidon, G.L, Sadée, W. (1999) Membrane transporters as drug targets. p 31.

Artursson, P., K. Palm & K. Luthman (2001) Caco-2 monolayers in experimental and

theoretical predictions of drug transport. Adv Drug Deliv Rev, 46, 27-43.

Artursson, P., A. L. Ungell & J. E. Löfroth (1993) Selective paracellular permeability in

two models of intestinal absorption: cultured monolayers of human intestinal

epithelial cells and rat intestinal segments. Pharm Res, 10, 1123-9.

Baba, S., N. Osakabe, M. Natsume & J. Terao (2004) Orally administered rosmarinic acid

is present as the conjugated and/or methylated forms in plasma, and is degraded and

metabolized to conjugated forms of caffeic acid, ferulic acid and m-coumaric acid.

Life Sci, 75, 165-78. 44

Baba, S., N. Osakabe, M. Natsume, A. Yasuda, Y. Muto, K. Hiyoshi, H. Takano, T.

Yoshikawa & J. Terao (2005) Absorption, metabolism, degradation and urinary

excretion of rosmarinic acid after intake of Perilla frutescens extract in humans.

Eur J Nutr, 44, 1-9.

Baker, S. S. & R. D. Baker (1992) Antioxidant enzymes in the differentiated Caco-2 cell

line. In Vitro Cell Dev Biol, 28A, 643-7.

Bauer, R. (2002) New knowledge regarding the effect and effectiveness of Echinacea

purpurea extracts. Wien Med Wochenschr, 152, 407-11.

Biganzoli, E., L. A. Cavenaghi, R. Rossi, M. C. Brunati & M. L. Nolli (1999) Use of a

Caco-2 cell culture model for the characterization of intestinal absorption of

antibiotics. Farmaco, 54, 594-9.

Birt, D. F., M. P. Widrlechner, C. A. Lalone, L. Wu, J. Bae, A. K. Solco, G. A. Kraus, P. A.

Murphy, E. S. Wurtele, Q. Leng, S. C. Hebert, W. J. Maury & J. P. Price (2008)

Echinacea in infection. Am J Clin Nutr, 87, 488S-92S.

Bodinet, C., R. Mentel, U. Wegner, U. Lindequist, E. Teuscher & J. Freudenstein (2002)

Effect of oral application of an immunomodulating plant extract on Influenza virus

type A infection in mice. Planta Med, 68, 896-900.

Boesch-Saadatmandi, C., S. Wolffram, A. M. Minihane & G. Rimbach (2009) Effect of

apoE genotype and dietary quercetin on blood lipids and TNF-alpha levels in apoE3

and apoE4 targeted gene replacement mice. Br J Nutr, 101, 1440-3.

Borchardt, R.T. (1998) Integration of pharmaceutical discovery and development: case

histories. p122.

Borchers, A. T., C. L. Keen, J. S. Stern & M. E. Gershwin (2000) Inflammation and Native

American medicine: the role of botanicals. Am J Clin Nutr, 72, 339-47. 45

Bruewer, M., A. Luegering, T. Kucharzik, C. A. Parkos, J. L. Madara, A. M. Hopkins & A.

Nusrat (2003) Proinflammatory cytokines disrupt epithelial barrier function by

apoptosis-independent mechanisms. J Immunol, 171, 6164-72.

Chang, G. (2003) Multidrug resistance ABC transporters. FEBS Lett, 555, 102-5.

Cheminat, A., Zawatzky, R., Becker, H., & Brouillard, R. (1988) Caffeoyl conjugates from

Echinacea species: Structures and biological activity. Phytochemistry, 27,

2787–2794.

Chen, C. H., H. J. Hsu, Y. J. Huang & C. J. Lin (2007) Interaction of flavonoids and

intestinal facilitated glucose transporters. Planta Med, 73, 348-54.

Chen, W. P., J. L. Tang, J. P. Bao, P. F. Hu, Z. L. Shi & L. D. Wu (2011) Anti-arthritic

effects of chlorogenic acid in interleukin-1β-induced rabbit chondrocytes and a

rabbit osteoarthritis model. Int Immunopharmacol, 11, 23-8.

Chen, L. W., Egan, L., Li, Z. W., Greten, F. R., Kagnoff, M. F, Karin, M. (2003) The two

faces of IKK and NF-kappaB inhibition: prevention of systemic inflammation but

increased local injury following intestinal ischemiareperfusion. Nat Med, 9,

575–81.

Cheung, H. Y. & Q. F. Zhang (2008) Enhanced analysis of triterpenes, flavonoids and

phenolic compounds in Prunella vulgaris L. by capillary zone electrophoresis with

the addition of running buffer modifiers. J Chromatogr A, 1213, 231-8.

Chicca, A., F. Pellati, B. Adinolfi, A. Matthias, I. Massarelli, S. Benvenuti, E. Martinotti, A.

M. Bianucci, K. Bone, R. Lehmann & P. Nieri (2008) Cytotoxic activity of

polyacetylenes and polyenes isolated from roots of Echinacea pallida. Br J

Pharmacol, 153, 879-85.

Choi, D. H., C. Li & J. S. Choi (2010) Effects of myricetin, an antioxidant, on the

pharmacokinetics of losartan and its active metabolite, EXP-3174, in rats: possible 46

role of cytochrome P450 3A4, cytochrome P450 2C9 and P-glycoprotein inhibition

by myricetin. J Pharm Pharmacol, 62, 908-14.

Chun, J. & A. Prince (2009) Ca2+ signaling in airway epithelial cells facilitates leukocyte

recruitment and transepithelial migration. J Leukoc Biol, 86, 1135-44.

Clayburgh, D. R., L. Shen & J. R. Turner (2004) A porous defense: the leaky epithelial

barrier in intestinal disease. Lab Invest, 84, 282-91.

Clifford, M. N. (1999) Chlorogenic acids and other cinnamates--nature, occurence and

dietary burden. J Sci Food Agric, 79, 362–72.

Dalby-Brown, L., H. Barsett, A. K. Landbo, A. S. Meyer & P. Mølgaard (2005) Synergistic

antioxidative effects of alkamides, caffeic acid derivatives, and polysaccharide

fractions from Echinacea purpurea on in vitro oxidation of human low-density

lipoproteins. J Agric Food Chem, 53, 9413-23.

Duffy, H. S., G. R. John, S. C. Lee, C. F. Brosnan & D. C. Spray (2000) Reciprocal

regulation of the junctional proteins claudin-1 and connexin43 by interleukin-1beta

in primary human fetal astrocytes. J Neurosci, 20, RC114.

Egert, D. & N. Beuscher (1992) Studies on antigen specifity of immunoreactive

arabinogalactan proteins extracted from Baptisia tinctoria and Echinacea purpurea.

Planta Med, 58, 163-5.

Feng, L., X. Jia, M. M. Zhu, Y. Chen & F. Shi (2010a) Antioxidant activities of total

phenols of Prunella vulgaris L. in vitro and in tumor-bearing mice. Molecules, 15,

9145-56.

Feng, L., X. B. Jia, F. Shi & Y. Chen (2010b) Identification of two polysaccharides from

Prunella vulgaris L. and evaluation on their anti-lung adenocarcinoma activity.

Molecules, 15, 5093-103. 47

Fogh, J., J. M. Fogh & T. Orfeo (1977) One hundred and twenty-seven cultured human

tumor cell lines producing tumors in nude mice. J Natl Cancer Inst, 59, 221-6.

Food and Agricultural Organization (FAO) (2004). The state of food security in the world:

monitoring progress towards the world food summit and millennium development

goals. FAO, 6-10.

Fujimoto, K. (1995) Freeze-fracture replica electron microscopy combined with SDS

digestion for cytochemical labeling of integral membrane proteins. Application to

the immunogold labeling of intercellular junctional complexes. J Cell Sci, 108,

3443-9.

Fusco, D., X. Liu, C. Savage, Y. Taur, W. Xiao, E. Kennelly, J. Yuan, B. Cassileth, M.

Salvatore & G. A. Papanicolaou (2010) Echinacea purpurea aerial extract alters

course of influenza infection in mice. Vaccine, 28, 3956-62.

Gonthier, M. P., M. A. Verny, C. Besson, C. Rémésy & A. Scalbert (2003) Chlorogenic

acid bioavailability largely depends on its metabolism by the gut microflora in rats.

J Nutr, 133, 1853-9.

Gorski, J. C., S. M. Huang, A. Pinto, M. A. Hamman, J. K. Hilligoss, N. A. Zaheer, M.

Desai, M. Miller & S. D. Hall (2004) The effect of Echinacea (Echinacea purpurea

root) on cytochrome P450 activity in vivo. Clin Pharmacol Ther, 75, 89-100.

Gropper, S.S., Smith, J. L., Groff, J. L. (2008). Advanced nutrition and human metabolism.

p4.

Han, E. H., J. H. Choi, Y. P. Hwang, H. J. Park, C. Y. Choi, Y. C. Chung, J. K. Seo & H. G.

Jeong (2009) Immunostimulatory activity of aqueous extract isolated from

Prunella vulgaris. Food Chem Toxicol, 47, 62-9. 48

Hansen, T. S. & O. G. Nilsen (2008) In vitro CYP3A4 metabolism: inhibition by

Echinacea purpurea and choice of substrate for the evaluation of herbal inhibition.

Basic Clin Pharmacol Toxicol, 103, 445-9.

Heller, F., P. Florian, C. Bojarski, J. Richter, M. Christ, B. Hillenbrand, J. Mankertz, A. H.

Gitter, N. Bürgel, M. Fromm, M. Zeitz, I. Fuss, W. Strober & J. D. Schulzke (2005)

Interleukin-13 is the key effector Th2 cytokine in ulcerative colitis that affects

epithelial tight junctions, apoptosis, and cell restitution. Gastroenterology, 129,

550-64.

Hellum, B. H. & O. G. Nilsen (2008) In vitro inhibition of CYP3A4 metabolism and

P-glycoprotein-mediated transport by trade herbal products. Basic Clin Pharmacol

Toxicol, 102, 466-75.

Hidalgo, I. J., T. J. Raub & R. T. Borchardt (1989) Characterization of the human colon

carcinoma cell line (Caco-2) as a model system for intestinal epithelial permeability.

Gastroenterology, 96, 736-49.

Hidalgo, M. A., K. A. Shah, B. J. Fuller & C. J. Green (1996) Cold ischemia-induced

damage to vascular endothelium results in permeability alterations in transplanted

lungs. J Thorac Cardiovasc Surg, 112, 1027-35.

Hill, G. R. & J. L. Ferrara (2000) The primacy of the gastrointestinal tract as a target organ

of acute graft-versus-host disease: rationale for the use of cytokine shields in

allogeneic bone marrow transplantation. Blood, 95, 2754-9.

Hinz, B., K. Woelkart & R. Bauer (2007) Alkamides from Echinacea inhibit

cyclooxygenase-2 activity in human neuroglioma cells. Biochem Biophys Res

Commun, 360, 441-6.

Hou, C. C., C. H. Chen, N. S. Yang, Y. P. Chen, C. P. Lo, S. Y. Wang, Y. J. Tien, P. W.

Tsai & L. F. Shyur (2010) Comparative metabolomics approach coupled with cell- 49

and gene-based assays for species classification and anti-inflammatory bioactivity

validation of Echinacea plants. J Nutr Biochem, 21, 1045-59.

Huang, N., C. Hauck, M. Y. Yum, L. Rizshsky, M. P. Widrlechner, J. A. McCoy, P. A.

Murphy, P. M. Dixon, B. J. Nikolau & D. F. Birt (2009) Rosmarinic acid in

Prunella vulgaris ethanol extract inhibits lipopolysaccharide-induced

prostaglandin E2 and nitric oxide in RAW 264.7 mouse macrophages. J Agric Food

Chem, 57, 10579-89.

Hubatsch, I., E. G. Ragnarsson & P. Artursson (2007) Determination of drug permeability

and prediction of drug absorption in Caco-2 monolayers. Nat Protoc, 2, 2111-9.

Islam, J. & R. Carter (2005) Use of Echinacea in upper respiratory tract infection. South

Med J, 98, 311-8.

Jager, H., L. Meinel, B. Dietz, C. Lapke, R. Bauer, H. P. Merkle & J. Heilmann (2002)

Transport of alkamides from Echinacea species through Caco-2 monolayers.

Planta Med, 68, 469-71.

Jeong, D. W., Y. H. Kim, H. H. Kim, H. Y. Ji, S. D. Yoo, W. R. Choi, S. M. Lee, C. K. Han

& H. S. Lee (2007) Dose-linear pharmacokinetics of oleanolic acid after

intravenous and oral administration in rats. Biopharm Drug Dispos, 28, 51-7.

Jiang, W. L., X. G. Chen, G. W. Qu, X. D. Yue, H. B. Zhu, J. W. Tian & F. H. Fu (2009)

Rosmarinic acid protects against experimental sepsis by inhibiting

proinflammatory factor release and ameliorating hemodynamics. Shock, 32,

608-13.

Jirovský, D., P. Kosina, M. Myslínová, J. Stýskala, J. Ulrichová & V. Simánek (2007)

HPLC analysis of rosmarinic acid in feed enriched with aerial parts of Prunella

vulgaris and its metabolites in pig plasma using dual-channel coulometric detection.

J Agric Food Chem, 55, 7631-7. 50

Jung, Y. B., K. J. Roh, J. A. Jung, K. Jung, H. Yoo, Y. B. Cho, W. J. Kwak, D. K. Kim, K.

H. Kim & C. K. Han (2001) Effect of SKI 306X, a new herbal anti-arthritic agent,

in patients with osteoarthritis of the knee: a double-blind placebo controlled study.

Am J Chin Med, 29, 485-91.

Konishi, Y. & S. Kobayashi (2004) Transepithelial transport of chlorogenic acid, caffeic

acid, and their colonic metabolites in intestinal caco-2 cell monolayers. J Agric

Food Chem, 52, 2518-26.

LaLone, C. A., K. D. Hammer, L. Wu, J. Bae, N. Leyva, Y. Liu, A. K. Solco, G. A. Kraus,

P. A. Murphy, E. S. Wurtele, O. K. Kim, K. I. Seo, M. P. Widrlechner & D. F. Birt

(2007) Echinacea species and alkamides inhibit prostaglandin E(2) production in

RAW264.7 mouse macrophage cells. J Agric Food Chem, 55, 7314-22.

Lamaison, J. L., C. Petitjean-Freytet & A. Carnat (1991) Medicinal Lamiaceae with

antioxidant properties, a potential source of rosmarinic acid. Pharm Acta Helv, 66,

185-8.

Lee, H. & J. Y. Lin (1988) Antimutagenic activity of extracts from anticancer drugs in

Chinese medicine. Mutat Res, 204, 229-34.

Lee, I. K., d. H. Kim, S. Y. Lee, K. R. Kim, S. U. Choi, J. K. Hong, J. H. Lee, Y. H. Park &

K. R. Lee (2008) Triterpenoic acids of Prunella vulgaris var. lilacina and their

cytotoxic activities in vitro. Arch Pharm Res, 31, 1578-83.

Lee, L. S., A. S. Andrade & C. Flexner (2006) Interactions between natural health products

and antiretroviral drugs: pharmacokinetic and pharmacodynamic effects. Clin

Infect Dis, 43, 1052-9.

Liao, L., Q. S. Guo, L. Liu & M. Liu (2008) Study of flavonoids distribution in various

populations of Prunella vulgaris. Zhongguo Zhong Yao Za Zhi, 33, 651-3. 51

Laitinen, L. (2006) Caco-2 cell cultures in the assessment of intestinal absorption: Effects

of some co-administered drugs and natural compounds in biological matrices.

Academic Dissertation.

Liu, F. & T. B. Ng (2000) Antioxidative and free radical scavenging activities of selected

medicinal herbs. Life Sci, 66, 725-35.

Lu, T., J. Yang, X. Gao, P. Chen, F. Du, Y. Sun, F. Wang, F. Xu, H. Shang, Y. Huang, Y.

Wang, R. Wan, C. Liu, B. Zhang & C. Li (2008) Plasma and urinary tanshinol from

Salvia miltiorrhiza (Danshen) can be used as pharmacokinetic markers for

cardiotonic pills, a cardiovascular herbal medicine. Drug Metab Dispos, 36,

1578-86.

Ludwig-Müller, J., S. V. Tokalov, A. Franz & H. O. Gutzeit (2005) Quercetin metabolism

in vital and apoptotic human leukaemia cells. Biol Chem, 386, 279-83.

Madara, J. L. & J. Stafford (1989) Interferon-gamma directly affects barrier function of

cultured intestinal epithelial monolayers. J Clin Invest, 83, 724-7.

Marková, H., J. Sousek & J. Ulrichová (1997) Prunella vulgaris L.--a rediscovered

medicinal plant. Ceska Slov Farm, 46, 58-63.

Matthias, A., R. S. Addison, L. L. Agnew, K. M. Bone, K. Watson & R. P. Lehmann (2007)

Comparison of Echinacea alkylamide pharmacokinetics between liquid and tablet

preparations. Phytomedicine, 14, 587-90.

Matthias, A., R. S. Addison, K. G. Penman, R. G. Dickinson, K. M. Bone & R. P. Lehmann

(2005) Echinacea alkamide disposition and pharmacokinetics in humans after

tablet ingestion. Life Sci, 77, 2018-29.

Matthias, A., J. T. Blanchfield, K. G. Penman, I. Toth, C. S. Lang, J. J. De Voss & R. P.

Lehmann (2004) Permeability studies of alkylamides and caffeic acid conjugates

from Echinacea using a Caco-2 cell monolayer model. J Clin Pharm Ther, 29, 7-13. 52

Matthias, A., E. M. Gillam, K. G. Penman, N. J. Matovic, K. M. Bone, J. J. De Voss & R. P.

Lehmann (2005b) Cytochrome P450 enzyme-mediated degradation of Echinacea

alkylamides in human liver microsomes. Chem Biol Interact, 155, 62-70.

Mennen, L. I., D. Sapinho, H. Ito, S. Bertrais, P. Galan, S. Hercberg & A. Scalbert (2006)

Urinary flavonoids and phenolic acids as biomarkers of intake for polyphenol-rich

foods. Br J Nutr, 96, 191-8.

Meunier, V., M. Bourrié, Y. Berger & G. Fabre (1995) The human intestinal epithelial cell

line Caco-2; pharmacological and pharmacokinetic applications. Cell Biol Toxicol,

11, 187-94.

Miller, S.C., & H. C. Yu (2004) Echinacea: the genus Echinacea. p 95.

Modarai, M., J. Gertsch, A. Suter, M. Heinrich & A. Kortenkamp (2007) Cytochrome P450

inhibitory action of Echinacea preparations differs widely and co-varies with

alkylamide content. J Pharm Pharmacol, 59, 567-73.

Mukinda, J. T., J. A. Syce, D. Fisher & M. Meyer (2010) Effect of the plant matrix on the

uptake of luteolin derivatives-containing Artemisia afra aqueous-extract in Caco-2

cells. J Ethnopharmacol, 130, 439-49.

Natarajan, K., S. Singh, T. R. Burke, D. Grunberger & B. B. Aggarwal (1996) Caffeic acid

phenethyl ester is a potent and specific inhibitor of activation of nuclear

transcription factor NF-kappa B. Proc Natl Acad Sci U S A, 93, 9090-5.

National Health Interview Survey (2007). Complementary and alternative medicine use

among adults and children: United States. Available at

http://nccam.nih.gov/news/2008/ nhsr12.pdf

O'Connor, R. (2007) The pharmacology of cancer resistance. Anticancer Res, 27, 1267-72.

O'Hara, M., D. Kiefer, K. Farrell & K. Kemper (1998) A review of 12 commonly used

medicinal herbs. Arch Fam Med, 7, 523-36. 53

Orhan, I., F. S. Senol, A. R. Gülpinar, M. Kartal, N. Sekeroglu, M. Deveci, Y. Kan & B.

Sener (2009) Acetylcholinesterase inhibitory and antioxidant properties of

Cyclotrichium niveum, Thymus praecox subsp. caucasicus var. caucasicus,

Echinacea purpurea and E. pallida. Food Chem Toxicol, 47, 1304-10.

Orinda, D., J. Diederich & A. Wacker (1973) Antiviral activity of components of

Echinacea purpurea. Arzneimittelforschung, 23, 1119-20.

Osakabe, N., A. Yasuda, M. Natsume, C. Sanbongi, Y. Kato, T. Osawa & T. Yoshikawa

(2002a) Rosmarinic acid, a major polyphenolic component of Perilla frutescens,

reduces lipopolysaccharide (LPS)-induced liver injury in D-galactosamine

(D-GalN)-sensitized mice. Free Radic Biol Med, 33, 798-806.

Paine, M. F., H. L. Hart, S. S. Ludington, R. L. Haining, A. E. Rettie & D. C. Zeldin (2006)

The human intestinal cytochrome P450 "pie". Drug Metab Dispos, 34, 880-6.

Pappenheimer, J. R. & K. Z. Reiss (1987) Contribution of solvent drag through

intercellular junctions to absorption of nutrients by the small intestine of the rat. J

Membr Biol, 100, 123-36.

Park, J. B. (2011) Identification and quantification of a major anti-oxidant and

anti-inflammatory phenolic compound found in basil, lemon thyme, mint, oregano,

rosemary, sage, and thyme. Int J Food Sci Nutr.

Patel, J. & A. K. Mitra (2001) Strategies to overcome simultaneous P-glycoprotein

mediated efflux and CYP3A4 mediated metabolism of drugs. Pharmacogenomics,

2, 401-15.

Pellati, F., S. Benvenuti, L. Magro, M. Melegari & F. Soragni (2004) Analysis of phenolic

compounds and radical scavenging activity of Echinacea spp. J Pharm Biomed

Anal, 35, 289-301. 54

Peppercorn, M. A. & P. Goldman (1971) Caffeic acid metabolism by bacteria of the human

gastrointestinal tract. J Bacteriol, 108, 996-1000.

Pinkas, H., R. P. Dolan & P. G. Brady (1994) Successful endoscopic transpapillary

drainage of an infected pancreatic pseudocyst. Gastrointest Endosc, 40, 97-9.

Pinto, M., Robine-Leon, S., Appay, M.-D., Kedinger, M., Triadou, N., Dussaulx, E.,

Lacroix, B., Simon-Assmann, P., Haffen, K., Fogh, J., Zweibaum, A. (1983)

Enterocyte-like differentiation and polarization of the human colon carcinoma

cell line Caco-2 in culture. Biol. Cell, 47, 323-330.

Psotová, J., M. Kolár, J. Sousek, Z. Svagera, J. Vicar & J. Ulrichová (2003) Biological

activities of Prunella vulgaris extract. Phytother Res, 17, 1082-7.

Radtke, J., J. Linseisen & G. Wolfram (1998) Phenolic acid intake of adults in a Bavarian

subgroup of the national food consumption survey. Z Ernahrungswiss, 37, 190-7.

Rasool, R., B. A. Ganai, S. Akbar, A. N. Kamili & A. Masood (2010) Phytochemical

screening of Prunella vulgaris L. - an important medicinal plant of Kashmir. Pak J

Pharm Sci, 23, 399-402.

Razzaque, A., & Ellis, B.E., (1997) Rosmarinic acid production in Coleus cell cultures.

Planta, 137, 287-291

Rimet, O., Mirrione, A., & Barra, Y (1999). Multidrug-resistant phenotype influences the

differentiation of a human colon carcinoma cell line. Biochem Biophys Res

Commun, 259, 43-49.

Romiti, N., F. Pellati, P. Nieri, S. Benvenuti, B. Adinolfi & E. Chieli (2008)

P-Glycoprotein inhibitory activity of lipophilic constituents of Echinacea pallida

roots in a human proximal tubular cell line. Planta Med, 74, 264-6. 55

Rushmore, T. H. & A. N. Kong (2002) Pharmacogenomics, regulation and signaling

pathways of phase I and II drug metabolizing enzymes. Curr Drug Metab, 3,

481-90.

Ryu, S. Y., M. H. Oak, S. K. Yoon, D. I. Cho, G. S. Yoo, T. S. Kim & K. M. Kim (2000)

Anti-allergic and anti-inflammatory triterpenes from the herb of Prunella vulgaris.

Planta Med, 66, 358-60.

Sahin, S., C. Demir & H. Malyer (2011) Determination of phenolic compounds in Prunella

L. by liquid chromatography-diode array detection. J Pharm Biomed Anal.

Saitou, M., K. Fujimoto, Y. Doi, M. Itoh, T. Fujimoto, M. Furuse, H. Takano, T. Noda & S.

Tsukita (1998) Occludin-deficient embryonic stem cells can differentiate into

polarized epithelial cells bearing tight junctions. J Cell Biol, 141, 397-408.

Saitou, M., M. Furuse, H. Sasaki, J. D. Schulzke, M. Fromm, H. Takano, T. Noda & S.

Tsukita (2000) Complex phenotype of mice lacking occludin, a component of tight

junction strands. Mol Biol Cell, 11, 4131-42.

Sasagawa, M., N. B. Cech, D. E. Gray, G. W. Elmer & C. A. Wenner (2006) Echinacea

alkylamides inhibit interleukin-2 production by Jurkat T cells. Int

Immunopharmacol, 6, 1214-21.

Scambia, G., F. O. Ranelletti, P. B. Panici, R. De Vincenzo, G. Bonanno, G. Ferrandina, M.

Piantelli, S. Bussa, C. Rumi & M. Cianfriglia (1994) Quercetin potentiates the

effect of adriamycin in a multidrug-resistant MCF-7 human breast-cancer cell line:

P-glycoprotein as a possible target. Cancer Chemother Pharmacol, 34, 459-64.

Scharl, M., Paul, G., Barrett, K. E. McCole, D. F. (2009) Adenosine monophosphate

activated protein kinase mediates the interferon-γ-induced decrease in intestinal

epithelial barrier function. J. Biol. Chem. 284, 27952–27563. 56

Schneider, S., J. Reichling, F. C. Stintzing, S. Messerschmidt, U. Meyer & P. Schnitzler

(2010) Anti-herpetic properties of hydroalcoholic extracts and pressed juice from

Echinacea pallida. Planta Med, 76, 265-72.

Sergent, T., I. Dupont, E. Van der Heiden, M. L. Scippo, L. Pussemier, Y. Larondelle & Y.

J. Schneider (2009) CYP1A1 and CYP3A4 modulation by dietary flavonoids in

human intestinal Caco-2 cells. Toxicol Lett, 191, 216-22.

Shimada, T., H. Yamazaki, M. Mimura, Y. Inui & F. P. Guengerich (1994) Interindividual

variations in human liver cytochrome P-450 enzymes involved in the oxidation of

drugs, carcinogens and toxic chemicals: studies with liver microsomes of 30

Japanese and 30 Caucasians. J Pharmacol Exp Ther, 270, 414-23.

Sloley, B. D., L. J. Urichuk, C. Tywin, R. T. Coutts, P. K. Pang & J. J. Shan (2001)

Comparison of chemical components and antioxidants capacity of different

Echinacea species. J Pharm Pharmacol, 53, 849-57.

Solecki, J. (1975) [Brufen treatment is disseminated sclerosis]. Wiad Lek, 28, 1197-8.

Song, Y. W., E. Y. Lee, E. M. Koh, H. S. Cha, B. Yoo, C. K. Lee, H. J. Baek, H. A. Kim, Y.

I. Suh, S. W. Kang, Y. J. Lee & H. G. Jung (2007) Assessment of comparative pain

relief and tolerability of SKI306X compared with celecoxib in patients with

rheumatoid arthritis: a 6-week, multicenter, randomized, double-blind,

double-dummy, phase III, noninferiority clinical trial. Clin Ther, 29, 862-73.

Speroni, E., P. Govoni, S. Guizzardi, C. Renzulli & M. C. Guerra (2002)

Anti-inflammatory and cicatrizing activity of Echinacea pallida Nutt. root extract.

J Ethnopharmacol, 79, 265-72.

Steffansen, B., C. U. Nielsen, B. Brodin, A. H. Eriksson, R. Andersen & S. Frokjaer (2004)

Intestinal solute carriers: an overview of trends and strategies for improving oral

drug absorption. Eur J Pharm Sci, 21, 3-16. 57

Strobel, P., C. Allard, T. Perez-Acle, R. Calderon, R. Aldunate & F. Leighton (2005)

Myricetin, quercetin and catechin-gallate inhibit glucose uptake in isolated rat

adipocytes. Biochem J, 386, 471-8.

Suzuki, T. & H. Hara (2009) Quercetin enhances intestinal barrier function through the

assembly of zonula [corrected] occludens-2, occludin, and claudin-1 and the

expression of claudin-4 in Caco-2 cells. J Nutr, 139, 965-74.

Suzuki, T. & H. Hara (2011) Role of flavonoids in intestinal tight junction regulation. J

Nutr Biochem, 22, 401-8.

Suzuki, T., S. Tanabe & H. Hara (2011) Kaempferol enhances intestinal barrier function

through the cytoskeletal association and expression of tight junction proteins in

Caco-2 cells. J Nutr, 141, 87-94.

Tabba, H. D., R. S. Chang & K. M. Smith (1989) Isolation, purification, and partial

characterization of prunellin, an anti-HIV component from aqueous extracts of

Prunella vulgaris. Antiviral Res, 11, 263-73.

Takeuchi, O., K. Hoshino, T. Kawai, H. Sanjo, H. Takada, T. Ogawa, K. Takeda & S.

Akira (1999) Differential roles of TLR2 and TLR4 in recognition of gram-negative

and gram-positive bacterial cell wall components. Immunity, 11, 443-51.

Tanaka, T., A. Nishikawa, H. Shima, S. Sugie, T. Shinoda, N. Yoshimi, H. Iwata & H.

Mori (1990) Inhibitory effects of chlorogenic acid, reserpine, polyprenoic acid

(E-5166), or coffee on hepatocarcinogenesis in rats and hamsters. Basic Life Sci, 52,

429-40.

Thelen, K. & J. B. Dressman (2009) Cytochrome P450-mediated metabolism in the human

gut wall. J Pharm Pharmacol, 61, 541-58.

Turner, J. R. (2006) Molecular basis of epithelial barrier regulation: from basic

mechanisms to clinical application. Am J Pathol, 169, 1901-9. 58

Turner, R. B., R. Bauer, K. Woelkart, T. C. Hulsey & J. D. Gangemi (2005) An evaluation

of Echinacea angustifolia in experimental rhinovirus infections. N Engl J Med, 353,

341-8.

Van Itallie, C. M., A. S. Fanning, A. Bridges & J. M. Anderson (2009) ZO-1 stabilizes the

tight junction solute barrier through coupling to the perijunctional cytoskeleton.

Mol Biol Cell, 20, 3930-40.

Van Wijk, J. (2000) Phytobusiness requires social chemistry. Phytochemistry, 55, 93-5.

Wachtel, M., M. F. Bolliger, H. Ishihara, K. Frei, H. Bluethmann & S. M. Gloor (2001)

Down-regulation of occludin expression in astrocytes by tumour necrosis factor

(TNF) is mediated via TNF type-1 receptor and nuclear factor-kappaB activation. J

Neurochem, 78, 155-62.

Walgren, R. A., U. K. Walle & T. Walle (1998) Transport of quercetin and its glucosides

across human intestinal epithelial Caco-2 cells. Biochem Pharmacol, 55, 1721-7.

Wang, F., B. T. Schwarz, W. V. Graham, Y. Wang, L. Su, D. R. Clayburgh, C. Abraham &

J. R. Turner (2006) IFN-gamma-induced TNFR2 expression is required for

TNF-dependent intestinal epithelial barrier dysfunction. Gastroenterology, 131,

1153-63.

Wang, M., J. E. Simon, I. F. Aviles, K. He, Q. Y. Zheng & Y. Tadmor (2003) Analysis of

antioxidative phenolic compounds in artichoke (Cynara scolymus L.). J Agric Food

Chem, 51, 601-8.

Wang, Y., J. Cao & S. Zeng (2005) Involvement of P-glycoprotein in regulating cellular

levels of Ginkgo flavonols: quercetin, kaempferol, and isorhamnetin. J Pharm

Pharmacol, 57, 751-8. 59

Weber, C. R., S. C. Nalle, M. Tretiakova, D. T. Rubin & J. R. Turner (2008) Claudin-1 and

claudin-2 expression is elevated in inflammatory bowel disease and may contribute

to early neoplastic transformation. Lab Invest, 88, 1110-20.

Wells, C. L., R. P. Jechorek, K. M. Kinneberg, S. M. Debol & S. L. Erlandsen (1999) The

isoflavone genistein inhibits internalization of enteric bacteria by cultured Caco-2

and HT-29 enterocytes. J Nutr, 129, 634-40.

Woelkart, K., E. Marth, A. Suter, R. Schoop, R. B. Raggam, C. Koidl, B. Kleinhappl & R.

Bauer (2006) Bioavailability and pharmacokinetics of Echinacea purpurea

preparations and their interaction with the immune system. Int J Clin Pharmacol

Ther, 44, 401-8.

Woelkart, K., W. Xu, Y. Pei, A. Makriyannis, R. P. Picone & R. Bauer (2005) The

endocannabinoid system as a target for alkamides from Echinacea angustifolia

roots. Planta Med, 71, 701-5.

Xie, S. & S. Zeng (2010) Stereoselective glucuronidation of propafenone and its analogues

by human recombinant UGT1A9. Chem Pharm Bull (Tokyo), 58, 879-83.

Xu, C., C. Y. Li & A. N. Kong (2005) Induction of phase I, II and III drug

metabolism/transport by xenobiotics. Arch Pharm Res, 28, 249-68.

Xu, H. X., S. H. Lee, S. F. Lee, R. L. White & J. Blay (1999) Isolation and characterization

of an anti-HSV polysaccharide from Prunella vulgaris. Antiviral Res, 44, 43-54.

Xu, Y., J. Chen, X. Yu, W. Tao, F. Jiang, Z. Yin & C. Liu (2010) Protective effects of

chlorogenic acid on acute hepatotoxicity induced by lipopolysaccharide in mice.

Inflamm Res, 59, 871-7.

Yamamoto, T., T. Kojima, M. Murata, K. Takano, M. Go, H. Chiba & N. Sawada (2004)

IL-1beta regulates expression of Cx32, occludin, and claudin-2 of rat hepatocytes

via distinct signal transduction pathways. Exp Cell Res, 299, 427-41. 60

Yoshikawa, T., & Terao, J. (2005) Absorption, metabolism, degradation and urinary

excretion of rosmarinic acid after intake of Perilla frutescens extract in humans.

Eur J Nutr, 44(1), 1-9.

Ye, Z., Z. Liu, A. Henderson, K. Lee, J. Hostetter, M. Wannemuehler & S. Hendrich (2009)

Increased CYP4B1 mRNA is associated with the inhibition of dextran sulfate

sodium-induced colitis by caffeic acid in mice. Exp Biol Med (Maywood), 234,

605-16.

Zhao, Z., H. S. Shin, H. Satsu, M. Totsuka & M. Shimizu (2008) 5-caffeoylquinic acid and

caffeic acid down-regulate the oxidative stress- and TNF-alpha-induced secretion

of interleukin-8 from Caco-2 cells. J Agric Food Chem, 56, 3863-8.

Zheng, M. (1990) Experimental study of 472 herbs with antiviral action against the herpes

simplex virus. Zhong Xi Yi Jie He Za Zhi, 10, 39-41, 6.

Zheng, Y., G. Wang, X. Wu, J. Sun, X. Li, Y. Gu, H. Lv, Y. Liang, B. Yan & W. Ye (2007)

Quantitative analysis of 23-hydroxybetulinic acid, a novel anticancer substance, in

the cellular environment of a human colon adenocarcinoma cell line for cellular

pharmacokinetic studies. Rapid Commun Mass Spectrom, 21, 779-86. 61

CHAPTER 3. GLUCURONIDATION OF ALKAMIDES AND KETONES IN ECHINACEA

SANGUINEA AND ECHINACEA PALLIDA EXTRACTS BY CACO-2 CELLS AND THEIR

MODULATION OF P-GLYCOPROTEIN TRANSPORTER

A paper to be submitted to the Journal of Planta Medica

Zhiyi Qiang 1, 2, Ludmila Rizshsky2, 3, Cathy Hauck 1, 2, Joe-Ann McCoy 2, 4, Mark P.

Widrlechner 2,5,6, Basil J. Nikolau2, 3, Manju B. Reddy1, Patricia A. Murphy 1, 2, Suzanne

Hendrich1, 2

1 Department of Food Science and Human Nutrition, Iowa State University, Ames, IA,

USA

2 The Center for Research on Botanical Dietary Supplements, Iowa State University, Ames,

IA, USA

3 Department of Biochemistry, Iowa State University, Ames, IA, USA

4 Bent Creek Institute, The North Carolina Arboretum, Asheville, NC, USA

5 US Department of Agriculture-Agricultural Research Service, North Central Regional

Plant Introduction Station, Ames, IA, USA

6 Departments of Horticulture and Agronomy, Iowa State University, Ames, IA, USA

Abstract

The use of Echinacea as a medicinal herb is prominent in the United States and many studies have been done over the years to assess the effectiveness of Echinacea as an immunomodulator. In this study, Bauer alkamides and ketones were hypothesized to be absorbed similarly as pure compounds or from Echinacea sanguinea and Echinacea pallid 62

ethanol extracts, as well as that these Echinacea extracts could inhibit P-glycoprotein transporter in Caco-2 human intestinal epithelial cells. Using HPLC analysis, the order of the permeation rate of pure alkamides across Caco-2 cells was increased paralleled with compound hydrophicility with passive diffusion, independent of the plant matrix. Tested alkamides were seemingly N-glucuronidated both as pure compounds and as found in two

Echinacea species. Both Echinacea ethanol extracts stimulated apparent glucuronidation and basolateral efflux of alkamide metabolites. E. pallida ethanol extract inhibited the efflux of glucuronides of Bauer alkamide 8 and 11 apically (lumen) while inducing the transfer of the conjugates basolaterally (systemic). Bauer ketone 24 was totally metabolized to more hydrophilic metabolites when administered as a pure compound, but not as found in either Echinacea extracts. Bauer alkamide 8, 10 and 11 (175-230 µM) and

E. sanguinea (1 mg/mL, containing 85 µM of alkamide 8, 2 µM of alkamide 10, and 0.7

µM of alkamide 11) and E. pallida (5 mg/mL, containing 215 µM of alkamide 8, 25 µM of alkamide 10, and 45 µM of alkamide 11) ethanol extracts reduced the efflux of the

P-glycoprotein transporter probe calcein-AM from Caco-2 cells. These results suggest that other constituents in the plant extract had a facilitating effect on the metabolism and efflux of alkamides and ketones from Echinacea, which would improve the therapeutic benefits of these extracts, and that alkamides and Echinacea extracts might be useful in potentiating some chemotherapeutics which are substrates for P-gp.

Key words: Echinacea; Alkamides; Ketones; Permeability; P-glycoprotein; Caco-2;

63

Introduction

Plant materials from the genus Echinacea have been widely used for centuries in North

America and later in Europe for the treatment and prevention of upper respiratory tract

infections, such as the common cold and influenza (Barnes et al., 2005; Kligler, 2003).

Echinacea supplements were the most commonly used nonvitamin, nonmineral, natural

products both for adults and children according to 2007 NHIS survey (National Health

Interview Survey, 2007). Species of Echinacea, such as E. angustifolia, E. pallida, E.

sanguinea and E. purpurea, have immune modulatory, antiviral, and antibacterial activities

(Borchers et al., 2000; Pleschka et al. 2009; Sharma et al., 2010).

E. pallida is one of the most medicinally used Echinacea species in the US and Europe and is being introduced into other regions due to the increasing popularity in alternative herbal

remedies. E. sanguinea has been neglected in the scientific literature, however, Birt et al.,

(2008) reported that E. sanguinea ethanolic extract had greater anti-inflammatory activity

in terms of the significant reduction of prostaglandin E 2 (PGE2) production in

lipopolysaccharide (LPS) stimulated RAW264.7 macrophages, than E. angustifolia, E.

pallida, and E. simulata. Studies with pure compounds and plant extract fractions indicate

that Echinacea species are rich in bioactive chemicals of which lipophilic alkamides and

ketones, intermediately hydrophilic phenolic compounds (mainly caffeic acid derivatives)

and polysaccharides are the most recognized for their immunomodulatory properties

(Barnes et al., 2005). Bauer alkamides 8

((2E,4E,8Z,10Z)-N-isobutyldodeca-2,4,8,10-tetraenamide), 10

((2E,4E,8Z)-N-isobutyldodeca-2,4,8-trienamide) and 11

((2E,4E)-N-isobutyldodeca-2,4-dienamide) at 50 µM and ketone 24 64

(pentadeca-8Z-ene-11,13-diyn-2-one) at 5 µM (Figure 1) possess anti-inflammatory

properties because they have been shown to significantly reduce nitric oxide and

prostaglandin E (2) production in lipopolysaccharide stimulated RAW264.7 macrophages

(LaLone et al., 2007). These alkamides also inhibit Cu (II) - catalyzed oxidation of human

low-density lipoprotein in vitro (Dalby-Brown et al., 2005).

Despite many in vitro studies ascribing biological activities to both the alkamides and

ketones, these activities in vivo are possible only if they are absorbed. Woelkart et al.

(2005) found about 5% of the ingested dose of Bauer alkamide 8 and 1% of Bauer alkamide

10 (Figure 1) in human blood 3 h after oral administration of a 60% ethanolic extract of E.

angustifolia containing 0.4-2 mg of Bauer alkamide 8 and 10 (Figure 1). Jager et al. (2005)

and Matthias et al (2005) showed the Papp (Apparent permeability coefficients), ranging from 3 × 10-6 to 3 × 10-4 cm/s, for various alkamides through Caco-2 monolayers, which

was correlated to structural variations. But very few studies have focused on the

metabolism of the bioavailable alkylamides (Toselli et al., 2010) and it is essential to

understand the fate of alkylamides after ingestion because the metabolites may have

lessened bioactivities compared with the parent compound (Cech et al., 2006). Caco-2

cells are immortalized human epithelial colorectal adenocarcinoma cells and offer a standard rapid, reliable, and low-cost model for in vitro prediction of intestinal drug

permeability and absorption (Hubatsch et al., 2007). Moreover, the plant material matrix

may alter absorption or metabolism and, consequently, the bioavailability of

phytochemicals (Manach et al., 2004). Therefore, it is crucial to investigate the effect that

the plant matrix may have on the uptake of key components found in Echinacea species.

Neither in vitro studies have been done in association with the uptake of Bauer alkamides

and ketones in different species of Echinacea, nor have the mechanism and metabolism 65

been investigated for their influence on pure compound uptake. In addition, our preliminary experiments showed that alkamide 8, 10, 11 and ketone 24 were only present in

E. pallida and E. sanguinea ethanol extracts, but not in E. angustifolia and E.simulata ethanolic extracts (data not published). Therefore, E. pallida and E. sanguinea were chosen in this study to investigate their absorption and metabolism.

Previous study showed that n-hexane root extracts from E. pallida, E. angustifolia and E. purpurea (30 µg/mL) inhibited multidrug transporter P-glycoprotein (P-gp) activity in a human proximal tubular cell line (Romiti et al., 2008). P-gp transporter plays a key role in drug absorption and distribution because it limits the permeability across the GI

(gastrointestinal) tract (Hunter and Hirst, 1997) by active efflux of potentially toxic substances back into the intestinal lumen. The transporter is also expressed in other tissues such as liver, kidney and blood–brain barrier where it also plays an important excretory role. On the other hand, P-gp confers resistance to anticancer chemotherapy because it is over-expressed in cancer cells (O'Connor 2007). The elucidation of the mechanisms by which botanicals mediate anticancer activities (i.e. inhibition of ABC transporters) may allow a new paradigm of clinically useful drug resistance circumvention.

Our hypotheses were that the absorbability of Bauer alkamides and ketones was independent of the plant extract matrix, alkamides and ketones were glucuronidated, and that ethanolic extracts of E. pallida and E. sanguinea extracts containing alkamides would inhibit P-gp activities in Caco-2 cells. This study was conducted to investigate not only the effect on epithelial barrier permeabilities and metabolism of Bauer alkamides and ketones as pure compounds and in E. pallida and E. sanguinea ethanol extracts across

Caco-2 cell monolayers, but also the inhibitory effects of pure compounds and plant 66

extracts on P-gp, to facilitate future studies of the efficacy of these herbs against

inflammation and herb-drug interactions.

2. Materials and methods

2.1. Plant extraction

Plant samples were provided by the North Central Regional Plant Introduction Station

(NCRPIS, Ames, IA) of the Agricultural Research Service of the U.S. Department of

Agriculture. Roots of E. pallida (PI 631293) and E. sanguinea (PI 633672) were obtained

from the NCRPIS, harvested, dried and ground (Voucher records: Inventory sample: PI

631293 SD 97ncao01; PI 633672 SD 97ncao01). Dried Echinacea root per population (6 g) were extracted with 500 mL of 95% ethanol by Soxhlet percolation for 6 h, filtered, dried by rotary evaporation and lyophilized. Then the extracts were redissolved in 0.5 mL of ethanol and stored at -20 °C under nitrogen. Information about the specific provenance of both accessions obtained from the NCRPIS is available on the Germplasm Resources

Information Network database at http://www.ars-grin.gov/npgs/acc/acc_queries.html.

2.2. Bauer alkamide and ketone synthesis

Chemical synthesis of Bauer alkamides and ketones (Bauer et al., 1988) were conducted according to the procedures described by Wu et al. (2004) and Bae (2006). Alkamide and ketone concentrations were calculated after correcting for percent purity, yielding concentrations equivalent to 100 % pure synthetic constituent. Calculated percent purity before correction for Bauer alkylamide 8 was 70%, Bauer alkylamide 10 was 82 %, Bauer alkylamide 11 was 92 %, and Bauer ketone 24 was 99%. All synthetic Bauer alkamides and ketone were dissolved in DMSO and stored at -80 °C under argon gas. 67

2.3. HPLC analysis

Methanol and acetonitrile (HPLC grade) were obtained from Fisher Scientific (Pittsburgh,

PA). HPLC analysis was performed on a Beckman Coulter 126 HPLC, equipped with

photodiode array detector model 168 and a model 508 autosampler (Beckman Coulter, Inc.,

Brea, CA). The mobile phase was CH3CN/H2O at a flow rate of 1.0 ml/min following a

linear gradient of 40–80% CH3CN in H2O over 45 min. A reverse phase analytical YMC

pack-ODS C18 column (250 mm× 4.6 mm × 5 μm, Waters Corp., Milford, MA ) was used at

room temperature. The wavelength was 260 nm and the retention times were 25.4 min for

Bauer alkamide 8, 29.1 min for Bauer alkamide 10, 35.6 min for Bauer alkamide 11, and

17.6 min for Bauer ketone 24. The limit of detection (LOD) and limit of quantitation

(LOQ), defined as a signal/noise ratio ≥ 3 and ≥ 6, was 0.04 μM and 0.08 μM for Bauer

alkamide 8, 10 and 11, 0.5 μM and 1.0 μM for Bauer ketone 24. The linear regression

equation for Bauer alkamide 8 was y = 27.234x - 0.3108 with R2 of 0.9996, y = 75.084x +

0.3075 with R2 of 0.9995 for Bauer alkamide 10, y = 36.758x - 0.2493 with R2 of 0.9999

for Bauer alkamide 11 and y = 73.532x + 0.0716 with R2 of 0.9994 for Bauer ketone 24, where y is the concentration of pure compound, and x is the ratio of peak area of the standard to peak area of the internal standard (2E,4E-hexadienoic acid isobutylamide,

ChromaDex, Irvine, CA). The intraday and interday CVs were 6.1 ± 2.1 % and 2.1 ± 0.4 % for Bauer alkamide 8, and 8.4 ± 2.6 % and 5.4 ± 3.0 % for Bauer alkamide 10 , 5.8 ± 1.4 % and 2.0 ± 0.4 % for Bauer alkamide 11, and 8.4 ± 2.5 % and 3.6 ± 0.4 % for Bauer ketone 24 across 0.08-100 μM.

2.4 Transepithelial transfer experiment 68

Caco-2 cells were obtained from American Type Culture Collection (Manassas, VA) at

passage 18 and all experiments were performed from passages 30–35. The cells were

cultured according to Hubatsch et al. (2007). Cytotoxicity of pure compounds and

extracts was measured according to Nasser et al. (2008). Pure Bauer alkamides and

ketone, at 10-1000 μM, and E. pallida and E. sanguinea ethanol extracts, at 0.1- 50 mg/mL, were tested for cytotoxicity. DMSO in DMEM (Dulbecco's modified Eagle's medium,

0.3%, Gibco Invitrogen, Carlsbad, CA) was used as control.

After the cells in the flask grew to 90–100% confluency, cells were trypsinized and seeded

on collagen-coated polytetrafluroethylene membrane inserts (0.45 μm) fitted in bicameral

chambers (Transwell-COL, 24 mm ID, Corning Inc., Corning, NY) at 1.2 × 105 cells/cm2.

The transepithelial electrical resistance (TEER) was tested by Millicell ERS meter (Fisher

Sci., Pittsburgh, PA) to reflect the tightness of intercellular junctions and only cells with

TEER ≥ 250 Ω·cm2 were used for permeability study (Hubatsch et al., 2007). At 21 d post-seeding (90–100% confluence) onto the transwell membrane, pure Bauer alkamides

and ketone and extracts at non-cytotoxic concentrations, dissolved in Hank's Buffered Salt

Solution (HBSS, pH 7.4, Gibco Invitrogen, Carlsbad, CA), were added to the donor side

with the receiving chamber containing 1.5 mL of HBSS. After 15, 30, 60, and 90 min,

solutions were collected from the receiver side. Samples were collected from both sides at

the end and membrane on the transwell insert was placed in 1.5 mL of ice-cold 0.5 mol

sodium hydroxide/L and sonicated with a probe-type sonic dismembrator (Biologics Inc.,

Manassas, VA); pH was adjusted to 7.0 and all samples were injected directly to HPLC for

analysis. The transport experiment was performed both at 37 °C and at 4 °C. Total

cellular protein was determined by Coomassie (Bradford) assay (Pierce Laboratories,

Rockford, IL). 69

2.5. Transepithelial transfer of pure compounds and extracts after treating with

β-glucuronidase/sulfatase

Pure Bauer alkamides and ketone and plant extracts at non-cytotoxic concentrations were

applied to Caco-2 cells. Twenty µL of β-glucuronidase (Type H-5 from Helix pomatia, 40

units/L, Sigma-Aldrich Co., St. Louis, MO) were added to the post-experimental apical and

basolateral solutions and to cell homogenates and incubated overnight at 37 °C to release

the parent compounds. These samples were then injected directly to HPLC.

Apparent permeability coefficients (Papp) were determined using the equation (Hubatsch et

al., 2007): Papp = (dQ / dt) (1 / (A × C0)); dQ/dt was the permeability rate constant (μmol/s);

2 A was the surface area of the membrane (cm ); and C0 was the initial concentration of the

compound (μM). Recovery relative to the initial amount (%) was calculated as the proportion of the original amount that permeated through the monolayer, which was calculated as the amount transported divided by the initial amount in the donor side.

Permeation rate (µmol/min/mg protein) was calculated as the amount transported divided by incubation time and protein content.

2.6. P-gp assay

P-gp activity was evaluated with fluorimetric measurement of the intracellular accumulation of calcein produced by ester hydrolysis of the P-gp substrate calcein-AM using VybrantTM Multidrug Resistance Assay Kit (Gibco Invitrogen, Carlsbad, CA).

After the cells in the flask grew to 90–100% confluency, cells were trypsinized and seeded

on 96 well plates at 1.2 × 105 cells/cm2. Cells were preincubated for 15 min with pure compounds (Bauer alkamide 8, 10, 11 and ketone 24) or extracts at non-cytotoxic 70

concentrations, thereafter calcein-AM was added and the fluorescence measured after 1

hour using microtiter plate reader Bio-Tek ELX 808 (Bio-Tek instruments. Inc., Winooski,

VT) at 490 nm. The control group received vehicle alone (DMSO in DMEM, 0.3% v/v).

The known P-gp inhibitor verapamil (10 µg/mL) was used as a control.

2.7. Statistical analysis

Data are given as means ± S. D. Differences in cytotoxicity, Papp, transport kinetics,

permeation rate, and absorbance representing calcein efflux by P-gp among treatments

were evaluated statistically using two-sample t-test, ANOVA and Tukey’s multiple

comparison tests by SAS 9.1 (SAS Institute Inc., Cary, NC). Differences were considered

significant at p < 0.05 and p < 0.01.

3. Results

3.1. Determination of alkamides and ketone concentrations in E. sanguinea and E. pallida

ethanol extracts

Bauer alkamide 8, 10 and 11 as well as ketone 24 were all present in Echinacea species

studied, but with different profiles (Table 1). E. sanguinea ethanol extract had ~three fold

greater Bauer alkamide 8 and ~six fold greater Bauer ketone 24 than E. pallida accession

on mole basis. The amount of Bauer alkamide 11 in E. pallida was about six fold more than

E. sanguinea and Bauer alkamide 10 content was similar in two species. Bauer alkamide

8 and ketone 24 were more concentrated in both Echineacea species compared with Bauer alkamide 10 and 11.

3.2 Cytotoxicity test 71

The concentrations of pure Bauer alkamide 8 greater than 350 μM, 950 μM for pure Bauer

alkamide 10, 460 μM for pure Bauer alkamide 11, and 990 μM for pure Bauer ketone 24

were significantly cytotoxic compared with the control (0.3% DMSO in DMEM, p < 0.05).

Ethanol extracts of E.sanguinea and E. pallida up to 5 mg/mL (containing 425 µM of alkamide 8, 10 µM of alkamide 10, and 3.5 µM of alkamide 11) and 20 mg/mL (containing

860 µM of alkamide 8,100 µM of alkamide 10, and 180 µM of alkamide 11) were significantly cytotoxic to the cells, respectively. Therefore, 10, 25, 50, 100 μM of pure

Bauer alkamides, ketone 24, and E. sanguinea and E. pallida ethanol extracts containing the same concentrations of each compound were used for permeability studies.

3.3. Transepithelial transfer of pure compounds and extracts before and after treating with

β-glucuronidase/sulfatase

Figure 2A shows the relationship between the initial permeation rate of Bauer alkamides and concentration. The rate of membrane permeation was calculated for Bauer alkamides as pure compounds. The uptake of Bauer alkamide 8, 10 and 11 increased linearly and exhibited non-saturable transport across the tested concentrations (10- 100 µM, Figure 2A).

Bidirectional permeation of three Bauer alkamides across Caco-2 cell monolayers was examined. Transport rate of Bauer alkamide 8, 10 and 11 in the basolateral to apical

(BL–AP) direction was nearly the same as that in the apical to basolateral (AP–BL) direction at each tested time point (Figure 2B, p > 0.05). The transepithelial transport in the two directions was not saturated within 90 min as shown in Figure 2B. The transport of Bauer alkamide 8, 10 and 11 across Caco-2 cell monolayers was investigated at both

37 °C and 4 °C to evaluate the effect of temperature on the transport of three alkamides

(AP-BL, Figure 2C). No significant differences were found in the permeation rates of three alkamides at 37 °C compared with that at 4 °C (p < 0.05). 72

Further, there are differences in permeability that correlate with variations in alkamide structure and the rank of the permeation of three alkamides was Bauer alkamide 8 > 10 > 11

(Table 2, Figure 2A, B and C). No Bauer alkamide 11 was found in the basolateral side when 10 µM of the pure compound was applied to the apical side, which was due to the lesser permeability of Bauer alkamide 11 compared with Bauer alkamide 8 and 10. No significant differences were found between alkamides as pure compounds compared to alkamides present in both E. sanguinea or E. pallida extracts (p > 0.05) in apparent permeability coefficients before deconjugation with β-glucuronidase (Papp, Table 2).

Formation of glucuronide conjugates during transcellular transport of the alkamides was investigated. After addition of the alkamides to the apical side of Caco-2 monolayers,

glucuronide metabolites of three alkamides in the E. sanguinea extract were found in all

compartments, including apical and basolateral sides as well as cell lysates, compared with

only in apical side for pure Bauer alkamide 8, in apical side and cells for pure Bauer

alkamide 10 as well as in basolateral side and cells for pure Bauer alkamide 11 (Figure 3A,

B and C). Interestingly, glucuronide conjugates were only detected basolaterally for

Bauer alkamide 8 and 10 from E. pallida extract (Figure 2 A and B), and only in cells for

Bauer alkamide 11 from E. pallida extract (p < 0.05, Figure 3C).

The Papp for Bauer ketone 24 in E. sanguinea and E. pallida ethanol extracts were 17.1 ±

7.8 and 12.2 ± 3.2 cm/s ×10-6 with no significant difference (p > 0.05, Table 2), while Bauer

ketone 24 was not found either apically or basolaterally nor in cell lysates at each time point

(15-90 min) across tested concentrations (10-100 µM). Two unknown peaks were

detected (retention time 13.3 and 15.4 min) in both apical and basolateral chambers and 73

cells after treating the Caco-2 cells with pure ketone 24, but not for the compound as found

in the two plant extracts (Figure 4B, C and D), which implies that pure Bauer ketone 24 was metabolized to more hydrophilic metabolites but this biotransformation was inhibited by other constituents in Echinacea extracts. After β-glucuronidase treatment, Bauer ketone 24 recovery relative to the applied amount was significantly increased both basolaterally and in cells for the compound contained in E. sanguinea and E. pallida extracts (Figure 3D). Papp was significantly increased for three alkamides and Bauer

ketone 24 in E. sanguinea extract and for Bauer alkamide 8, 10 and ketone 24 in E. pallida

extract compared with pure compounds after deconjugation with β-glucuronidase (p < 0.05,

Table 2).

3.4. Pgp inhibition by pure alkamides, E. sanguinea and E. pallida ethanol extracts

In the present study, the modulatory effects on P-gp activity of ethanol extracts of

E. sanguinea and E. pallida as well as pure alkamides and ketone were investigated. As

shown in Figure 5, After purity correction, pure Bauer alkamide 8 at 175 µM, Bauer

alkamide 8 at 205 µM, Bauer alkamide 11 at 230 µM significantly inhibited P-gp activity

(p < 0.05), but Bauer ketone 24 was not active across the tested concentrations (1-250 µM).

The ethanolic extracts of E. sanguinea and E. pallida significantly inhibited P-gp at 1

mg/mL (containing 85 µM of alkamide 8, 2 µM of alkamide 10, and 0.7 µM of alkamide 11)

and 5 mg/mL (containing 215 µM of alkamide 8, 25 µM of alkamide 10, and 45 µM of

alkamide 11), respectively (Figure 5).

74

4. Discussion

This study investigated the uptake and metabolism of alkamides and ketone, when introduced as pure compounds or as components of the complex ethanolic extracts of

Echinacea to cultures of the human intestinal epithelial cell line, Caco-2. The objective was to explore the effect that the compound structure and the plant matrix may have on the intestinal uptake and metabolism of alkamides and ketone from a traditional medicinal plant. We also used Caco-2 cells to study the interaction of alkamides and Echinacea ethanol extracts with P-gp, to identify compounds that could potentially inhibit P-gp.

The transport of pure alkamides and the accumulation from both the apical and basolateral sides of the cells were studied. Because Bauer alkamide 8 was considered to be absorbed via passive diffusion (Jager et al., 2002), it was expected that other alkamides would also be absorbed via similar pathway across Caco-2 cells. This is supported by the finding that the rate of membrane permeation of three alkamides increased linearly with concentration, was not saturable during the tested incubation period (15-90 min) and was not different between

AP-BL and BL-AP direction (Figure 2A and B). Further, the transport of three alkamides was temperature independent (Figure 2C), which indicates passive diffusion through the

Caco-2 cells. Because an active mechanism is involved, the transport of the molecule is almost completely abolished at 4°C compared with 37°C (Said et al., 2003).

The order of the uptake of three alkamides was Bauer alkamide 8 > 10 > 11, consistent with

Papp for Bauer alkamide 8, ~2.5 fold and ~15 fold greater than that of Bauer alkamide 10 and 11, respectively (Table 2). This trend for increases in hydrophilicity giving rise to 75

increases in the apparent permeability was also seen for the family of alkamides (Matthias et al., 2005).

Monohydroxylated, monoepoxidized, N-dealkylated metabolites were reported after incubation of parent Bauer alkamide 8 with both NADPH and human liver microsomes

(Matthias et al., 2005b). But no studies have investigated Phase II biotransformation of alkamides. Various amides, including aliphatic amides, cyclic amides and sulphonamides, have been reported to undergo N-glucuronidation as a major elimination pathway (Sanna,

2010). In our study, after treating the post-experimental apical and basolateral solutions and cell lysates with β-glucuronidase, the amount of the parent compound increased by

~1.5 to 4.2 fold for Bauer alkamide 8, 10 and 11 in apical side, basolateral side or cell lysates, indicating that alkamides were seemingly N-glucuronidated to some extent based on the amide structure (O-glucuronide might be hard to form because of the steric hinderance, Figure 1), revealing another important metabolic pathway for alkamides besides cytochromes P450. E. sanguinea constituents increased total glucuronide formation for all alkamides, suggesting allosteric activation by plant constituents other than alkamides, given the short time frame for the observed apparent stimulation of UDP glucuronosyltransferases (UGTs). Caffeic acid (0.6 mmol), quercetin, and apigenin (2-38 nmol), which were all found in Echinacea species, induced UGTs expression both in animal models and in cell lines (Svehlíková et al., 2004; Debersa et al., 2001; Galijatovic et al., 2000).

It is critical to investigate the effect of the plant matrix on the bioavailability of their bioactive compounds because the complex constituents may alter absorption or metabolism and, consequently, the bioavailability of phytochemicals by the influences on Phase I and II 76

enzymes and transporters. The plant extracts did not affect transport of pure alkamides before deconjugation with β-glucuronidase (Table 2), establishing that the uptake of alkamides was independent of the of the plant extract matrix for the two Echinacea species studied. This observation might be due to the transport mechanism of alkamides—passive diffusion across Caco-2 cell monolayers, without transporter mediated uptake involved for the parent compound. Our study is the first to report the effect of Echinacea species on the metabolism of alkamides and ketone. Interestingly, the recovery relative to the initial amount applied to the cells after β-glucuronidase hydrolysis was significantly increased in all compartments of the system, including apical and basolateral side as well as cell lysates, for all three alkamides studied in E. sanguinea extract. These results suggest that other constituents contained in Echinacea species might regulate the expression of MRP

(multidrug resistance-associated protein) or OATP (organic anion transporter protein) transporters which in turn affect the permeation of glucuronidated alkamides, because anionic conjugates (glutathione, glucuronide or sulfate) cannot exit cells unless an MRP or

OATP transporter is present (Peng et al. 1999). E. pallida ethanol extract inhibited the efflux of glucuronides of Bauer alkamide 8 and 10 apically (lumen) while enhancing the transfer of the conjugates basolaterally (systemic), which was not found for Bauer alkamide 11, indicating that alkamide conjugates may have different structural affinity to the efflux transporters (i.e. MRPs) and that greater amounts of Bauer alkamide 8 and 10 from plant sources may enter systemic circulation, compared with Bauer alkamide 11, after deconjugation by microbial β-glucuronidase leading to stronger immune modulatory effects.

The permeability and metabolism of Bauer ketone 24 were investigated as pure compound and the compound contained in Echinacea species. Pure Bauer ketone 24 might be 77

metabolized by CYP 450 based on the diene structure and ketone group. It has been

shown that there is a strong overlap between substrates for gut metabolism by CYPs and

gut active efflux by P-gp, suggesting that these two processes may work together to affect

the intestinal absorption of a variety of drugs (Wacher et al., 1995). Therefore, P-gp

transporter, located along the apical membrane of the enterocytes, might be involved in the

efflux of Bauer ketone 24 metabolites. The transport of Bauer ketone 24 metabolites to

the basolateral side might be due to passive diffusion along the concentration gradient.

More interestingly, about 5% of the Bauer ketone 24 was recovered in the basolateral side

for the compound contained in two Echinacea species, and the Papp for Bauer ketone 24 was

similar to that of Bauer alkamide 10 (Table 2). The reason for the difference in permeability of Bauer ketone 24 as pure compound or that in Echinacea species might be due to the effects of other constituents contained in the herb extract, because several

Echinacea species (11.2 – 2447 µg/mL) have been shown to inhibit CYP 2C19, 2D6, and

3A4 by 20-100 %, and alkamides contributed to the inhibitory effects seen with Echinacea preparations (Modarai et al., 2007). Pure Bauer alkamide 8, 10 and 11 were incubated with ketone 24 across 10-100 µM, but ketone 24 was still totally metabolized as pure compound (data not shown). This implies that the CYP isoforms inhibited by those alkamides were not involved in the biotransformation of ketone 24, although we can not rule out the possibility that the doses of those alkamides were too low to exhibit inhibitory effects. Bioactivity of these metabolites should be examined to determine the influence of intestinal metabolism on bioavailability of the alkamides and ketones. How the plant constituents alter glucuronidation also needs to be understood as this has implications for herb/drug interactions.

78

P-gp actively effluxes a wide range of structurally diverse anticancer agents and

P-gp-mediated multidrug resistance (MDR) has been associated with inhibition of

caspase-dependent tumor cell apoptosis (Smyth et al., 1998). One strategy for reversal of

MDR in cells expressing ABC transporters is combined use of anticancer drugs with

modulators. Inhibitors of ABC transporters, like P-gp and MRP1, can be used to enhance

effects of anticancer drugs in patients with MDR. In our study, E. sanguinea (1 mg/mL,

containing 85 µM of alkamide 8, 2 µM of alkamide 10, and 0.7 µM of alkamide 11) and E.

pallida (5 mg/mL, containing 215 µM of alkamide 8, 25 µM of alkamide 10, and 45 µM of

alkamide 11) significantly inhibited P-gp-mediated efflux of calcein AM, a substrate for

P-gp, indicating that alkamides were potentiated in the inhibition on P-gp by other plant

extract constituents. For example, quercetin at 100 µM inhibited 30 % of P-gp mediated

efflux of [3H]-taxol in Caco-2 cells (Hayeshi et al., 2006). There might be some

interactions between glucuronidation and P-gp modulation, because coregulation of UGTs

and ABC transporters by AhR (Ah-receptor) ligands has been studied in Caco-2 cells

(Walle et al., 1999; Bock et al., 2000; Bock-Hennig et al., 2002) showing that AhR, PXR

(pregnane X receptor), and CAR (constitutive androstane receptor) found in

promoter/enhancer region of the UGT gene might also be involved in the regulation of

ABC transporters (Bock and Bock-Hennig, 2010). Further, the concentrations of

Echinacea plant extracts that inhibited of P-gp in our study were greater than in a published

paper showing that extracts of roots of Echinacea species inhibited P-gp in human

proximal tubular kidney cells, and E. pallida extract was the most active at 3 μg/mL

compared with E. angustifolia and E. purpurea both at 30 μg/mL (Romiti et al. 2008).

This might be due to the difference in the cell lines used or the preparation of the extracts

(extraction ratio (raw material weight: final solution volume) was 12:1 in our study and

125:1 in Romiti’s study). The concentrations of Echinacea extracts showing inhibitory 79

activities of P-gp were greater than the concentrations of E. sanguinea and E. pallida

containing 10-100 µM of Bauer ketone 24 (0.08-0.8 mg/mL and 0.3-3.4 mg/mL, respectively), thus P-gp might not be inhibited and was able to efflux pure Bauer ketone 24

metabolites in our permeability study (Table 2 and Figure 4). Romiti et al. (2008) found that the lipophilic ketones isolated from E. pallida roots, namely

8-hydroxypentadeca-(9E)-ene- 11,13-diyn-2-one, 8-hydroxypentadeca- (9E,13Z)- dien-11-yn-2-one, pentadeca-(9E)-ene-11,13-diyne-2,8-dione, pentadeca-(9E,13Z)- dien-11-yne-2,8-dione and pentadeca-(8Z,13Z)-dien-11-yn-2-one, in the range of 200-800

µM, inhibited P-gp. Our study is the first report of pure Bauer alkamide 8, 10 and

11(greater than the amount contained in E. sanguinea at 1 mg/mL or E. pallida at 5 mg/mL) inhibiting P-gp mediated efflux in Caco-2 cells, which may have implications for anti-colitic or colon cancer drug interactions with P-gp. Because we have shown that alkamides are apparently glucuronidated, even though they are probably most absorbed in the small intestine, the glucuronides will be effluxed to the lumen to some extent, although not for the alkamides in E. pallida, and they would also be susceptible to biliary excretion into the intestine. Although N-glucuronides are more slowly hydrolyzed by

β-glucuronidase, produced by gut bacteria, than O-glucuronides or S-glucuronides

(Casarett and Doull, 2008), the parent alkamides would still be available to act in the large intestine. In the current study we analyze each chemically synthesized preparation prior to its use as an inhibitor of P-gp function, allowing for impurities in the alkamide preparations, and normalizing to a concentration at 100% purity; this normalization allowed for greater sensitivity in our screening of P-gp inhibition using the caco-2 intestinal cell line. In addition, alkamides were highly permeable, which permits their interaction with P-gp, since its binding sites are either inside the bilayer or at the inner leaflet of the cell membrane (c and Clarke, 2005). Although glucuronidation may prevent tissue uptake of 80

the alkamides, glucuronidase produced by liver and neutrophils (O'Leary et al., 2001;

Bartholomé et al., 2010) may convert the glucuronide conjugates to parent compounds and

alkamides may enter the hepatic portal vein at relative high concentrations, possibly

permitting alkamide inhibitory activity of P-gp in tissues throughout the body.

In conclusion, Bauer alkamide 8, 10 and 11 were transferred across Caco-2 cells,

independent of extract matrix, thus results on efficacy of the pure compounds may

reasonably be extrapolated to results using plant materials containing these compounds.

Alkamides were seemingly N-glucuronidated to some extent and Bauer ketone 24 was

transferred and metabolized to glucuronide conjugates as the compound from plant extracts

but not as pure compound, strongly suggesting that alkamides and Bauer ketone 24 in

Echinacea species had a more extensive metabolism in the Caco-2 cells than pure

compounds, and that plant matrix may have a facilitating effect on the metabolism of

alkamides and ketones from Echinacea. Because the Caco-2 cell model is considered to

be a good predictive model for human gastro-intestinal absorption, the results obtained

should also closely reflect the in vivo intestinal uptake and metabolism of alkamides and

ketones expected when Echinacea plant extracts and supplements are given to humans.

Bauer alkamide 8, 10 and 11 as well as E. sanguinea and E. pallida extracts inhibited P-gp

mediated efflux in Caco-2 cells. Considering the increasing knowledge about the role of

P-gp in cancer resistance and the worldwide increasing use of Echinacea preparations,

further studies in humans are required to specifically investigate the uptake and metabolism

of alkamides and ketones derived from natural plant products in humans. The reason for the enhanced metabolism by other accompanying constituents in plant extracts should also be established. Because Echinacea supplements are usually ingested on a chronic basis, 81

their long term effects on the metabolizing enzymes and expression of efflux transporters needs to be investigated and may be of importance regarding herb–drug interactions.

Acknowledgements

This publication was made possible by grant number P01 ES012020 from the National

Institute of Environmental Health Sciences (NIEHS) and the Office of Dietary

Supplements (ODS), National Institutes of Health (NIH) and grant number

9P50AT004155-06 from the National Center for Complementary and Alternative

Medicine (NCCAM) and ODS, NIH.

82

Figures

Figure 1. The chemical structures of Bauer alkamide 8, 10, 11 and ketone 24. 83

84

Figure 2. Transport of pure Bauer alkamide 8, 10 and 11 across Caco-2 cells. A.

Concentration dependency of the transport of three alkamides. B. Non- saturable transport of three alkamides at 25 µM during 90 min incubation period with no difference between apical to basolateral (AP-BL) and basolateral to apical (BL-AP) direction at each time point. C. The effect of temperature on the transport of the three alkamides (37°C vs

4°C). Means bearing different letters were significantly different among three alkamides by ANOVA and Tukey’s multiple comparison (p < 0.05). Data are the mean ± S. D.

(n=6).

85

86

Figure 3. Glucuronidation of Bauer alkamide 8, 10, 11 and ketone 24 for pure compounds

or from plant extracts by Caco-2 cells. Before, prior to deconjugation with

β-glucuronidase. After , subsequent to deconjugation with β-glucuronidase. The concentration of three Bauer alkamides and ketone 24 was 25 µM both for pure compounds or as found in the two Echinacea extracts. A. Bauer alkamide 8. B. Bauer alkamide 10.

C. Bauer alkamide 11. D. Bauer ketone 24. * Significantly different compared with the

recovery relative to the initial amount before the enzyme treatment by two-sample t-test (p

< 0.05). Data are the mean ± S. D (n=6). 87

88

Figure 4. HPLC chromatograms of Bauer ketone 24 across Caco-2 cell monolayers as pure compound or from Echinacea species. A. Bauer ketone 24 standard at 100 µM

(retention time 17.6 min). B. Bauer ketone 24 (100 µM) was not detected in the basolateral supernatant fluid, but two more hydrophilic metabolites were detected

(retention time 13.3 and 15.4 min). C. Bauer ketone 24 was shown in the basolateral side after apically applied E. sanguinea containing 100 µM of ketone 24 (retention time 17.6 min). D. Bauer ketone 24 was shown in the basolateral side after apically applied E. pallida containing 100 µM of ketone 24 (retention time 17.6 min).

89

Figure 5. The effects of Bauer alkamide 8, 10, 11, ketone 24 on P-glycoprotein transporter

activity. Control was 0.3% DMSO in PBS. Verapamil was used as positive control at

10 µg/mL. * Significantly different compared with control by two-sample t-test (p < 0.05).

** Significantly different compared with control by two-sample t-test (p < 0.01). Data are

the mean ± S. D (n=6). The ethanolic extract of E. sanguinea at 1 mg/mL contained 85

µM of alkamide 8, 2 µM of alkamide 10, and 0.7 µM of alkamide 11, and 5 mg/mL of E.

pallida extract containecd 215 µM of alkamide 8, 25 µM of alkamide 10, and 45 µM of alkamide 11.

90

Table 1. The concentrations of alkamides and ketone in Echinacea ethanol extracts

determined by HPLCa.

a E. sanguinea and E. pallida ethanolic extracts were diluted by methanol for HPLC

Bauer Bauer Bauer Bauer

alkamide 8 alkamide10 alkamide11 ketone 24

g/L(mM) g/L(mM) g/L(mM) g/L(mM)

E. sanguinea 3.2 ± 0.2 0.1 ± 0.01 0.03 ± 0.01 3.8 ± 0.5

(PI 633672, 154.3 g/Lb) (13.1 ± 0.7) (0.3 ± 0.02) (0.1 ± 0.04) (17.3 ± 2.5)

E. pallida 1.1 ± 0.1 0.1 ± 0.01 0.2 ± 0.03 0.6 ± 0.04

(PI 631293, 99.0 g/L ) (4.3 ± 0.3) (0.5 ± 0.02) (0.9 ± 0.1) (2.9 ± 0.2)

analysis (n=9). All values are means± SD. b PI, assession number. 91

Table 2. Apparent permeability coefficients (Papp) for pure alkamides and extracts across

the Caco-2 monolayera.

-6 Papp (cm/s ×10 )

Bauer Bauer Bauer Bauer

alkamide 8 alkamide 10 alkamide 11 ketone 24

Before d After Before After Before After Before After

Pure 43.8 ± 39.8 ± 17.7 ± 18.3 ± 2.8 ± 9.7 ± ND e ND compound b 11.2 a 13.4 b 8.8 a 5.9 b 1.5 a 2.3*a

c 54.6 ± 76.8 ± 12.9 ± 28.4 ± 3.7 ± 8.2 ± 12.2 ± 22.9 E. sanguinea 13.2 a 11.7 *a 5.6 a 4.1*a 1.9 a 2.4*a 3.2 a ±6.3*a

41.3 ± 65.2 ± 14.7 ± 22.3 ± 5.4 ± 4.3 ± 17.1 ± 28.4 ± E. pallida 7.4 a 6.8 *a 4.2 a 7.9*a 2.5 a 1.9 b 7.8 a 6.3*a a All values are means ± SD (n=9). TEER was 460-573 Ω·cm2. Means bearing different

letters were significantly different in each column (p < 0.05). b Bauer alkamide 8, 10, 11 and ketone 24 were 10-100 μM for pure compounds.

c Two herb extracts were diluted to contain same concentrations of Bauer alkamide 8, 10,

11 or ketone 24 as the pure compounds.

d Before = before deconjugation; After = after deconjugation.

e Bauer ketone 24 was not detected in the receiver side for pure compound.

* Significantly different compared with the Papp before deconjugation with β-glucuronidase

by ANOVA and Tukey’s multiple comparison. 92

References

Bae, J (2006) Synthesis of the natural compounds in Echinacea. Iowa State University

Thesis.

Barnes, J., L. A. Anderson, S. Gibbons & J. D. Phillipson (2005) Echinacea species

(Echinacea angustifolia (DC.) Hell., Echinacea pallida (Nutt.) Nutt., Echinacea

purpurea (L.) Moench): a review of their chemistry, pharmacology and clinical

properties. J Pharm Pharmacol, 57, 929-54.

Bartholomé, R., G. Haenen, C. H. Hollman, A. Bast, P. C. Dagnelie, D. Roos, J. Keijer, P.

A. Kroon, P. W. Needs & C. W. Arts (2010) Deconjugation kinetics of

glucuronidated phase II flavonoid metabolites by beta-glucuronidase from

neutrophils. Drug Metab Pharmacokinet, 25, 379-87.

Bauer, R., Jurcic. K, Puhlmann, J & Wagner, H (1988) Immunological in vivo and in vitro

examinations of Echinacea extracts. Arzneim. Forsch/Drug Res, 38, 276-281.

Binns, S. E., J. Hudson, S. Merali & J. T. Arnason (2002) Antiviral activity of characterized

extracts from echinacea spp. (Heliantheae: Asteraceae) against herpes simplex

virus (HSV-I). Planta Med, 68, 780-3.

Birt, D. F., M. P. Widrlechner, C. A. Lalone, L. Wu, J. Bae, A. K. Solco, G. A. Kraus, P. A.

Murphy, E. S. Wurtele, Q. Leng, S. C. Hebert, W. J. Maury & J. P. Price (2008)

Echinacea in infection. Am J Clin Nutr, 87, 488S-92S.

Bock, K. W. & B. S. Bock-Hennig (2010) UDP-glucuronosyltransferases (UGTs): from

purification of Ah-receptor-inducible UGT1A6 to coordinate regulation of subsets

of CYPs, UGTs, and ABC transporters by nuclear receptors. Drug Metab Rev, 42,

6-13. 93

Bock, K. W., T. Eckle., M. Ouzzine., S. Fournel-Gigleux (2000) Coordinate induction by

antioxidants of UDPglucuronosyltransferase UGT1A6 and the apical export pump,

MRP2 (multidrug resistance protein 2), in Caco-2 cells. Biochem Pharmacol, 59,

467–470.

Borchers, A. T., C. L. Keen, J. S. Stern & M. E. Gershwin (2000) Inflammation and Native

American medicine: the role of botanicals. Am J Clin Nutr, 72, 339-47.

Casarett, L. J., J. Doull, C. D. Klaassen. (2008) Casarett and Doull's toxicology: the basic

science of poisons. Seventh edition. p 257.

Cech, N. B., K. Tutor, B. A. Doty, K. Spelman, M. Sasagawa, G. M. Raner & C. A. Wenner

(2006a) Liver enzyme-mediated oxidation of Echinacea purpurea alkylamides:

production of novel metabolites and changes in immunomodulatory activity. Planta

Med, 72, 1372-7.

Dalby-Brown, L., H. Barsett, A. K. Landbo, A. S. Meyer & P. Mølgaard (2005) Synergistic

antioxidative effects of alkamides, caffeic acid derivatives, and polysaccharide

fractions from Echinacea purpurea on in vitro oxidation of human low-density

lipoproteins. J Agric Food Chem, 53, 9413-23.

Debersac, P., M. F. Vernevaut, M. J. Amiot, M. Suschetet & M. H. Siess (2001) Effects of

a water-soluble extract of rosemary and its purified component rosmarinic acid on

xenobiotic-metabolizing enzymes in rat liver. Food Chem Toxicol, 39, 109-17.

Dietz, B., J. Heilmann & R. Bauer (2001) Absorption of

dodeca-2E,4E,8Z,10E/Z-tetraenoic acid isobutylamides after oral application of

Echinacea purpurea tincture. Planta Med, 67, 863-4.

Galijatovic, A., U. K. Walle & T. Walle (2000a) Induction of

UDP-glucuronosyltransferase by the flavonoids and quercetin in Caco-2

cells. Pharm Res, 17, 21-6. 94

Guiotto, P., K. Woelkart, I. Grabnar, D. Voinovich, B. Perissutti, S. Invernizzi, M.

Granzotto & R. Bauer (2008) Pharmacokinetics and immunomodulatory effects of

phytotherapeutic lozenges (bonbons) with Echinacea purpurea extract.

Phytomedicine, 15, 547-54.

Hayeshi, R., C. Masimirembwa, S. Mukanganyama & A. L. Ungell (2006) The potential

inhibitory effect of antiparasitic drugs and natural products on P-glycoprotein

mediated efflux. Eur J Pharm Sci, 29, 70-81.

Hubatsch, I., E. G. Ragnarsson, P. Artursson (2007) Determination of drug permeability

and prediction of drug absorption in Caco-2 monolayers. Nat Protoc. 2, 2111-2119.

Hunter, J., & Hirst, B.H. (1997) Intestinal secretion of drugs. The role of P-glycoprotein

and related drug efflux systems in limiting oral drug absorption. Adv. Drug Deliv.

Rev. 225, 129–157.

Jager, H., L. Meinel, B. Dietz, C. Lapke, R. Bauer, H. P. Merkle & J. Heilmann (2002)

Transport of alkamides from Echinacea species through Caco-2 monolayers.

Planta Med, 68, 469-71.

Kligler, B (2003). Echinacea. Am Fam Physician, 67, 77-80.

LaLone, C. A., K. D. Hammer, L. Wu, J. Bae, N. Leyva, Y. Liu, A. K. Solco, G. A. Kraus,

P. A. Murphy, E. S. Wurtele, O. K. Kim, K. I. Seo, M. P. Widrlechner & D. F. Birt

(2007) Echinacea species and alkamides inhibit prostaglandin E(2) production in

RAW264.7 mouse macrophage cells. J Agric Food Chem, 55, 7314-22.

Loo, T. W. & D. M. Clarke (2005) Do drug substrates enter the common drug-binding

pocket of P-glycoprotein through "gates"? Biochem Biophys Res Commun, 329,

419-22.

Nasser, B., Moustaid, K., Moukha, S., Mobio, T.A., Essamadi, A., Creppy, E.E., 2008.

Evaluation of the cytotoxicity and genotoxicity of extracts of mussels originating from 95

Moroccan Atlantic coast, in human colonic epithelial cells Caco-2. Environmental

Toxicology, 23, 539-547.

Manach, C., J. L. Donovan. (2004) Pharmacokinetics and metabolism of dietary

flavonoids in humans. Free Radic Res, 38, 771-785.

Matthias, A., K. G. Penman, N. J. Matovic, K. M. Bone, J. J. De Voss & R. P. Lehmann

(2005) Bioavailability of Echinacea constituents: Caco-2 monolayers and

pharmacokinetics of the alkylamides and caffeic acid conjugates. Molecules, 10,

1242-51.

Modarai, M., J. Gertsch, A. Suter, M. Heinrich & A. Kortenkamp (2007) Cytochrome P450

inhibitory action of Echinacea preparations differs widely and co-varies with

alkylamide content. J Pharm Pharmacol, 59, 567-73.

O'Connor, R. (2007) The pharmacology of cancer resistance. Anticancer Res 27, 1267-72.

O'Leary, K. A., A. J. Day, P. W. Needs, W. S. Sly, N. M. O'Brien & G. Williamson (2001)

Flavonoid glucuronides are substrates for human liver beta-glucuronidase. FEBS

Lett, 503, 103-6.

Patricia, M., B. B. Barnes., L. N. Richard (2007) Complementary and alternative medicine

use among adults and children: United States, National Health Statistics Reports.

Peng, K. C., F. Cluzeaud, M. Bens, J. P. Duong Van Huyen, M. A. Wioland, R. Lacave & A.

Vandewalle (1999) Tissue and cell distribution of the multidrug

resistance-associated protein (MRP) in mouse intestine and kidney. J Histochem

Cytochem, 47, 757-68.

Pleschka, S., M. Stein, R. Schoop & J. B. Hudson (2009) Anti-viral properties and mode of

action of standardized Echinacea purpurea extract against highly pathogenic avian

influenza virus (H5N1, H7N7) and swine-origin H1N1 (S-OIV). Virol J, 6, 197. 96

Romiti, N., F. Pellati, P. Nieri, S. Benvenuti, B. Adinolfi & E. Chieli (2008)

P-Glycoprotein inhibitory activity of lipophilic constituents of Echinacea pallida

roots in a human proximal tubular cell line. Planta Med, 74, 264-6.

Said, H. M., A. Ortiz & T. Y. Ma (2003) A carrier-mediated mechanism for pyridoxine

uptake by human intestinal epithelial Caco-2 cells: regulation by a PKA-mediated

pathway. Am J Physiol Cell Physiol, 285, 1219-25.

Sanna, K. (2010) N-glucuronidation of drugs and other xenobiotics, Academic dissertation.

Division of Pharmaceutical Chemistry Faculty of Pharmacy University of Helsinki,

Finland.

Saracino, M. R. & J. W. Lampe (2007) Phytochemical regulation of UDP-

glucuronosyltrans- ferases: implications for cancer prevention. Nutr Cancer, 59,

121-41.

Senchina, D. S., D. A. McCann, J. M. Asp, J. A. Johnson, J. E. Cunnick, M. S. Kaiser & M.

L. Kohut (2005) Changes in immunomodulatory properties of Echinacea spp. root

infusions and tinctures stored at 4 degrees C for four days. Clin Chim Acta, 355,

67-82.

Sharma, S. M., M. Anderson, S. R. Schoop & J. B. Hudson (2010) Bactericidal and

anti-inflammatory properties of a standardized Echinacea extract (Echinaforce):

dual actions against respiratory bacteria. Phytomedicine, 17, 563-8.

Smyth, M. J., E. Krasovskis, V. R. Sutton & R. W. Johnstone (1998) The drug efflux

protein, P-glycoprotein, additionally protects drug-resistant tumor cells from

multiple forms of caspase-dependent apoptosis. Proc Natl Acad Sci U S A, 95,

7024-9. 97

Svehlíková, V., S. Wang, J. Jakubíková, G. Williamson, R. Mithen, Y. Bao (2004)

Interactions between sulforaphane and apigenin in the induction of UGT1A1 and

GSTA1 in CaCo-2 cells. Carcinogenesis, 25, 1629-37.

Toselli, F., A. Matthias, K. M. Bone, E. M. Gillam & R. P. Lehmann (2010) Metabolism of

the major Echinacea alkylamide N-isobutyldodeca-2E,4E,8Z,10Z-tetraenamide by

human recombinant cytochrome P450 enzymes and human liver microsomes.

Phytother Res, 24, 1195-201.

Wacher, V. J., C. Y. Wu & L. Z. Benet (1995) Overlapping substrate specificities and tissue

distribution of cytochrome P450 3A and P-glycoprotein: implications for drug

delivery and activity in cancer chemotherapy. Mol Carcinog, 13, 129-34.

Walle, U. K., A. Galijatovic & T. Walle (1999) Transport of the flavonoid chrysin and its

conjugated metabolites by the human intestinal cell line Caco-2. Biochem

Pharmacol, 58, 431-8.

Woelkart, K., C. Koidl, A. Grisold, J. D. Gangemi, R. B. Turner, E. Marth & R. Bauer

(2005) Bioavailability and pharmacokinetics of alkamides from the roots of

Echinacea angustifolia in humans. J Clin Pharmacol, 45, 683-9.

Wu, L., Bae, J. Kraus, G & Wurtele, E (2004) Diacetylenic isobutylamides of Echinacea:

Synthesis and natural distribution. Phytochemistry, 65, 2477-2484. 98

CHAPTER 4. METABOLISM AND PERMEABILITY OF ROSMARINIC ACID IN

PRUNELLA VULGARIS AND URSOLIC ACID IN SALVIA OFFICINALIS

EXTRACTS ACROSS CACO-2 CELL MONOLAYERS

A paper submitted to the Journal of Ethnopharmacology

Zhiyi Qiang a, Zhong Ye b, Cathy Hauck a, Patricia A. Murphy a, Joe-Ann McCoy c, Mark P.

Widrlechner d, e, Manju B. Reddy a, Suzanne Hendrich a

a Department of Food Science and Human Nutrition, Iowa State University, Ames, IA,

USA

b Department of Nutrition, University of California, Davis, CA, USA

c Bent Creek Institute, The North Carolina Arboretum, Asheville, NC, USA

d US Department of Agriculture-Agricultural Research Service, North Central Regional

Plant Introduction Station, Ames, IA, USA

e Departments of Horticulture and Agronomy, Iowa State University, Ames, IA, USA

Abstract

Ethnopharmacological relevance Rosmarinic acid (RA), a caffeic acid derivative found

in high concentrations in Prunella vulgaris (self-heal), and ursolic acid (UA), a pentacyclic

triterpene acid concentrated in Salvia officinalis (sage), have been traditionally used to treat

inflammation in the oral cavity, and may also be of benefit to gastrointestinal health in

general. 99

Aim of the study To investigate the permeability and metabolism of RA and UA as pure compounds and in P. vulgaris and S. officinalis ethanol extracts across human intestinal epithelial Caco-2 cell monolayers.

Materials and methods The permeability and Phase II biotransformation of RA and UA as pure compounds and in herbal extracts that contained same concentration as the pure compounds were compared using Caco-2 cells with HPLC detection.

Results The apparent permeability coefficient (Papp) for RA and RA in P. vulgaris extracts

was 0.2 ± 0.05 × 10-6 cm/s, significantly increased to 0.9 ± 0.2 × 10-6 cm/s after

β-glucuronidase/sulfatase treatment. Papp for UA and UA in S. officinalis extract was 2.7 ±

0.3 × 10-6 cm/s and 2.3 ± 0.5 × 10-6 cm/s before and after β-glucuronidase/sulfatase treatment, respectively. Neither compound was affected in permeability by the herbal extract matrix.

Conclusion RA and UA in herbal extracts had similar uptake as that found using the pure

compounds, which may simplify the prediction of compound efficacy. The apparent lack

of intestinal glucuronidation/sulfation of UA is likely to further enhance the bioavailability

of that compound compared with RA.

Key words: Prunella vulgaris; Salvia officinalis; Rosmarinic acid; Ursolic acid;

Permeability; Caco-2

100

Introduction

Rosmarinic acid (RA, Figure 1) is a caffeic acid (CA) derivative found in various

botanicals, especially in Prunella vulgaris, a perennial herb known as self-heal used to treat sore throat, fever, and wounds (Psotová et al., 2003). RA and P. vulgaris limit liver damage derived from a model of bacterial inflammation (Osakabe et al., 2002) and inhibit nervous system inflammation in another model (Swarup et al., 2007). Ursolic acid (UA,

Figure. 1), a pentacyclic triterpene acid, is also found in P. vulgaris but especially concentrated in sage leaves (Salvia officinalis), and inhibits inflammation-related changes in human gingival cells (Zdarilová et al., 2009) and in other models (Liu, 1995). This compound also has antimutagenic activities (Filipic et al., 1997). Both herbs have been used traditionally to treat inflammation in the mouth, and are of interest in inhibiting gastrointestinal inflammation which is relevant to colitis and colon cancer.

A major limiting step in the utilization of RA and UA is their intestinal absorption and metabolism. Both compounds contain hydroxyls and are likely to be glucuronidated or sulfated in intestinal cells, forms that are generally considered to be less bioactive than parent compounds. This metabolism has been demonstrated for RA but not for UA, and the plant matrix components might alter this metabolism but this has not been studied yet.

The absorption or metabolism of RA has been examined in vivo to a limited extent (Baba et al., 2004; Konishi et al., 2005; Baba et al., 2005). When Perilla extract containing 200 mg

of RA was orally administered to six men, RA in both plasma and urine was present predominantly as glucuronide and/or sulfate conjugated forms, at 0.6 ± 0.2% and 1.5 ± 0.4% of the total intake, respectively, within 48 h after ingestion (Baba et al., 2005). Gut microbes may metabolize RA to give phenolics such as caffeic acid, o-coumaric acid and 101

m-hydroxyphenylpropionic acid, which are then absorbed by monocarboxylic acid

transporter (MCT)-mediated active processes (Konishi and Kobayashi, 2005). After the

oral administration of Sambucus chinensis ethanol extract (40g/kg to rats, containing 80 mg

UA/kg) about 0.6% of ingested UA was recovered in plasma based on estimated blood

volume and plasma area under curve of this compound, suggesting poor absorption or

extensive metabolism and distribution to other body tissues (Liao et al., 2005). Both RA

and UA as constituents of P. vulgaris may contribute to its bioactivities, and the effects of

plant matrix on uptake and metabolism of its key bioactive compounds are of interest.

Caco-2 cells are immortalized human epithelial colorectal adenocarcinoma cells and offer a

standard rapid, reliable, and low-cost model for in vitro prediction of intestinal drug

permeability and absorption (Hubatsch et al., 2007). Caffeic acid derivatives, such as RA

and chlorogenic acid, have been studied using Caco-2 model and are proposed to transfer

across the intestinal barrier by paracellular diffusion (Konishi and Kobayashi, 2005;

Konishi and Kobayashi, 2004). The plant material matrix may alter absorption or

metabolism and, consequently, the bioavailability of phytochemicals (Manach et al., 2004).

P. vulgaris and S. officinalis contain sugars, steroids, alkaloids, essential oils, flavonoids,

polyphenols, triterpenoids, and saponins (Rasool et al., 2010; Cheung and Zhang, 2008; Lu

and Foo, 2000; Loizzo et al., 2008). Therefore, it is crucial to investigate the effect that the

plant matrix may have on the uptake of RA and UA found in P. vulgaris and S. officinalis.

In vitro studies have not been done in association with the permeation of RA in P. vulgaris extracts using Caco-2 cells nor have different sources of P. vulgaris plant material been compared for their influence on RA uptake. Uptake of UA in the Caco-2 cell model also has not been investigated, either as a pure compound or from plant extracts. Our hypotheses were that the absorbability and metabolism of RA and UA was independent of 102

the plant extract matrix and this study was conducted to investigate the permeabilities of

RA and UA as pure compounds and in P. vulgaris and S. officinalis ethanol extracts across

Caco-2 cell monolayers, to facilitate future studies of the efficacy of these herbs against gastrointestinal inflammation.

Materials and methods

Plant extraction

All P. vulgaris plant samples were provided by the U.S Department of Agriculture North

Central Regional Plant Introduction Station in Ames, IA. S. officinalis extract was

provided by Sabinsa Corporation (Payson, UT). from accessions P. vulgaris Ames

27664, 27665 (both originally collected in North Carolina), and 27748 (collected in

Missouri) were germinated in Petri plates at 25 °C (Voucher records: Herbarium specimen.

Taken by: McCoy, J., USDA, ARS. On: 08/24/2006. Located at: ISC. Inventory sample:

Ames 27664, 27665 and 27748. SD 04ncao01). The resulting seedlings, segregated by accession, were transferred to flats in a greenhouse (held at 20–25 °C). Upper flowering portions of 14-month-old plants were harvested at the time of peak flowering, dried, and ground. Four g aliquots of the ground samples were extracted with 500 mL of 95% ethanol by Soxhlet percolation for 6 h, filtered, dried by rotary evaporation and lyophilized.

Then the extracts were redissolved in 0.5 mL of ethanol and stored at -20 °C under nitrogen.

Information about the P. vulgaris accessions used for these experiments is available via the

Germplasm Resources Information Network database at http://www.ars-grin.gov/npgs/acc/acc_queries.html.

103

HPLC analysis

RA (90%) and UA (92%), both from Sigma-Aldrich Co. (St. Louis, MO) and 2, 4, 4'-

trihydroxybenzoin (THB, internal standard), synthesized in Dr. Hendrich’s laboratory

(Song et al., 1998) were dissolved in DMSO to use them as standards. Methanol,

acetonitrile (HPLC grade) and phosphoric acid (AR grade) were obtained from Fisher

Scientific (Pittsburgh, PA).

HPLC analysis was performed on a Beckman Coulter 126 HPLC, equipped with

photodiode array detector model 168 and a model 508 autosampler (Beckman Coulter, Inc.,

Brea, CA). The mobile phase was 1.25% phosphoric acid: acetonitrile (15% acetonitrile

at 0 min, then increased to 84% at 15 min and held for 40 min; decreased to 15% at 65 min)

at a flow rate of 0.5 ml/min. A reverse phase analytical YMC pack-ODS C18 column (250

mm× 4.6 mm × 5 μm, Waters Corp., Milford, MA ) was used at room temperature. The

wavelength was 210 nm and the retention times were 21.5 min for RA and 26.5 min for UA.

The limit of detection (LOD) and limit of quantitation (LOQ), defined as a signal/noise

ratio ≥ 3 and ≥ 6, was 0.05 μM and 0.1 μM both for RA and UA. The linear regression equation for RA was y = 95.256x - 0.0188 with R2 of 0.9995 and y = 236.07x + 0.0014 with

R2 of 0.9996 for UA, where y is the concentration of RA or UA, and x is the ratio of peak

area of the standard to peak area of the internal standard (THB). The intraday and interday

CVs were 7.3 ± 2.5 % and 3.4 ± 1.6 % for RA, and 9.1 ± 3.8 % and 5.7 ± 2.1 % for UA across 0.1-100 μM.

Transepithelial transfer experiment

Caco-2 cells were obtained from American Type Culture Collection (Manassas, VA) at passage 18 and all experiments were performed from passages 25–30. The cells were 104

cultured according to Hubatsch et al. (2007). Cytotoxicity of RA, UA, and extracts was

measured according to Nasser et al. (2008). RA, UA, P. vulgaris ethanol extracts

containing the same concentrations of RA, and S. officinalis extract containing the same

concentrations of UA, at 1, 10, 20, 50, and 100 μM in DMEM, were tested for cytotoxicity.

DMSO in DMEM (Dulbecco's modified Eagle's medium, 0.3%, Gibco Invitrogen,

Carlsbad, CA) was used as control.

After the cells in the flask grew to 90–100% confluency, cells were trypsinized and seeded

on collagen-coated polytetrafluroethylene membrane inserts (0.45 μm) fitted in bicameral

chambers (Transwell-COL, 24 mm ID, Corning Inc., Corning, NY) at 1.2 × 105 cells/cm2.

The transepithelial electrical resistance (TEER) was tested by Millicell ERS meter (Fisher

Sci., Pittsburgh, PA) to reflect the tightness of intercellular junctions and only cells with

TEER ≥ 250 Ω·cm2 were used for permeability study. At 20-21 d post seeding (90–100%

confluence) on the transwell, RA, UA and extracts at non-cytotoxic concentrations or 100

μM of the paracellular marker, lucifer yellow (LY, Sigma-Aldrich Co., St. Louis, MO),

dissolved in Hank's Buffered Salt Solution (HBSS, pH 7.4, Gibco Invitrogen, Carlsbad,

CA), were added to the apical chamber, then basolateral solutions were collected after 0.5,

1, 2, and 4 h. After 4 h, apical solutions were collected and membrane on the transwell

insert was placed in 1.5 mL of ice-cold 0.5 mol sodium hydroxide/L and sonicated with a

probe-type sonic dismembrator (Biologics Inc., Manassas, VA); pH was adjusted to 7.0

and all samples were injected directly to HPLC for analysis. For LY quantification, the

apical and basolateral solutions were transferred to a 96 well plate and read

spectrophotometrically at 450 nm. Total cellular protein was determined by Coomassie

(Bradford) assay (Pierce Laboratories, Rockford, IL).

105

Transepithelial transfer of single compounds and extracts after treating with

β-glucuronidase/sulfatase

Ten μM RA as a single compound and P. vulgaris ethanol extracts diluted to contain 10 μM

RA or 20 μM UA and S. officinalis extract diluted to contain 20 μM UA were applied to

Caco-2 cells. Twenty µL of β-glucuronidase/sulfatase (Type H-2 from Helix pomatia, 85 units/L of glucuronidase and 7.5 units/L of sulfatase, Sigma-Aldrich Co., St. Louis, MO) were added to the post-experimental apical and basolateral solutions and to cell homogenates and incubated overnight at 37 °C to release the parent compounds. These samples were then injected directly to HPLC.

Apparent permeability coefficients (Papp) were determined using the equation (Hubatsch et

al., 2007): Papp = (dQ / dt) (1 / (A × C0)); dQ/dt was the permeability rate constant (μmol/s);

2 A was the surface area of the membrane (cm ); and C0 was the initial concentration of the

compound (μM). Basolateral recoveries (%) were calculated as the proportion of the original amount that permeated through the monolayer, which was calculated as the amount transported divided by the initial amount in the apical chamber. Transport rate

(µM/h/cm2) was calculated as the amount transported divided by incubation time and the area of the membrane.

Statistical analysis

Data are given as means ± S. D. Differences in cytotoxicity, Papp, transport kinetics,

basolateral recoveries and transport rate of the pure compounds and extracts were

evaluated statistically using ANOVA and Tukey’s multiple comparison tests by SAS 9.1

(SAS Institute Inc., Cary, NC). Differences were considered significant at p < 0.05.

106

Results

Determination of RA and UA concentrations in P. vulgaris and S. officinalis ethanol

extracts

The amounts of RA in ethanol extracts varied ~sixfold across the three P. vulgaris

accessions; UA content in P. vulgaris extract was ~four- to twenty-fold less on a molar basis than was UA in S. officinalis extract (Table 1). The ethanolic extract of P. vulgaris

27748 had threefold greater RA and tenfold more UA than the two other accessions studied.

No RA was found in the S. officinalis extract.

Cytotoxicity test

For RA as a pure compound, concentrations greater than 50 μM were significantly cytotoxic compared with the control (0.3% DMSO in DMEM, p < 0.05). Ethanol extracts of P. vulgaris accessions containing ≥ 20 μM RA showed significant cytotoxicity.

Concentrations ≥ 20 μM of UA as a pure compound and S. officinalis extract containing same amount of UA were toxic to the Caco-2 cells (data not shown). Therefore, 1, 2, 5, 10

μM of RA as a single compound or in P. vulgaris ethanol extracts, and 2, 5, 10, 20 μM of

UA and S. officinalis extract containing these amounts of UA were used for permeability

studies.

Transepithelial transfer of single compounds and extracts before and after treating with

β-glucuronidase/sulfatase

The basolateral recovery of paracellular marker, LY, was 0.5 ± 0.1%. In the basal

chamber after 4 h, 1.3 ± 0.3, 1.2 ± 0.2, 1.4 ± 0.6 and 0.9 ± 0.4 % of parent RA was

transferred for pure compound, or for RA in P. vulgaris 27664, 27665 and 27748 extracts, 107

respectively (no significant differences, p > 0.05). The apical recoveries of RA were 86.7

± 2.9, 85.8 ± 3.6, 81.2 ± 5.7 % and 79.1 ± 11.2 % for the pure compound, or for RA in P.

vulgaris 27664, 27665 and 27748 extracts, respectively, with no significant differences (p >

0.05, data not shown).

The permeation rate was calculated for UA as a pure compound and UA in S. officinalis extract at the same doses. The uptake of UA increased linearly and significantly from 0.03

± 0.01 to 0.2 ± 0.04 µM/h/cm2 (p < 0.01) and was not saturable across the tested

concentrations (5- 20 µM, Figure 2). It was also not saturable across tested time-points

(0.5-4 h) both for pure compounds and UA contained in S. officinalis extract (Figure 3A).

Due to LOD of UA and the lesser amount of it in P. vulgaris extracts compared with RA,

UA was not detected for apically-applied P. vulgaris extracts containing 1-10 μM RA.

Basolateral recoveries of RA as pure compound and RA in P. vulgaris 27664 increased

over time (RA in P. vulgaris 27665 and 27748 are not shown), although RA was not detected basolaterally at 0.5 and 1h before deconjugation (Figure 3A). The basolateral transfer of UA was significantly greater than RA at each time point both for pure compounds or compounds contained in plant extracts (p < 0.01, Figure 3A). No significant differences were found between RA as a pure compound and RA in P. vulgaris extracts or UA as a pure compound and UA in S. officinalis extract (p > 0.05) in apparent permeability coefficients (Papp, Table 2).

After deconjugation using β-glucuronidase/sulfatase, RA recoveries significantly increased from 1.3 ± 0.3% to 6.1 ± 1.4%, 1.2 ± 0.2% to 6.5 ± 2.7 %, 1.2 ± 0.03% to 4.9 ± 0.6 %, 1.1 ±

0.2% to 5.5 ± 1.2 % (p < 0.01) in basal chamber for the single compound, or for RA in P.

vulgaris 27664, 27665 and 27748 extracts, respectively (Figure 3C, RA in P. vulgaris 108

27665 and 27748 are not shown). The transfer of RA glucuronide/sulfate conjugates was

not saturable during 4 h and increased linearly with the incubation time (p < 0.01, Figure

3B). Both RA as a single compound and RA in P. vulgaris extracts showed the same

increase in Papp after deconjugation (Table 2). To check the mass balance, the total

recoveries (apical + basolateral) of RA increased from 80-88% before deconjugation to

84-93% after deconjugation for pure compound and RA in extracts, with no new peaks

detected. The basolateral recoveries of UA were 19.0 ± 4.2% or 15.9 ± 3.2% for pure

compound before or after incubation with β-glucuronidase/sulfatase, and 17.6 ± 2.5% or

14.5 ± 4.3% for S. officinalis extract, not statistically different (p > 0.05, Figure 3C). No

significant differences were found in the apical recoveries of RA or UA (p > 0.05) and

neither RA nor UA was found in cells before or after deconjugation. Papp for UA or UA in

S. officinalis extract did not change after deconjugation reaction (Table 2).

Discussion

Because in vivo absorption studies performed with laboratory animals and humans are expensive and time consuming and can pose ethical challenges, Caco-2 cells are a useful

alternative to study uptake and transport of important compounds from plant materials

(Walgren et al., 1998). Plant matrices may influence phytochemicals by altering transporters of some compounds or of compound metabolites (Mukinda et al., 2010).

Phytochemical biotransformation may also be altered by other components of the plant matrix. To establish the effect that the plant matrix had on the in vitro uptake of RA and

UA found in P. vulgaris and S. officinalis, two herbs that may be important for gut health, the uptake of pure solutions of the two compounds and extracts of two herbs that are major sources of these compounds were compared in Caco-2 cell system. 109

RA in P. vulgaris and UA in S. officinalis extracts had similar efficient absorption as that in

pure solutions of each compound in Caco-2 cells (Figure 3A and Table 2, p > 0.05); this was not influenced by varied contents of RA across three accessions collected from

-6 different sources. The Papp for RA in Salvia miltiorrhiza was 0.5 ± 0.3 × 10 cm/s in

Caco-2 cells (Lu et al., 2008), consistent with our results. The results are also consistent with in vivo studies showing poor absorption of intact RA (Baba et al., 2004), suggesting that the absorption of RA is independent of other constituents contained in P. vulgaris ethanolic extracts. The impact of plant matrices on the bioavailability of phytochemicals is highly dependent on the compound of interest. For example, the absorption of epicatechin was not influenced by the ingredient composition of beverage food matrices in vitro (Neilson et al., 2009), but Hypericum perforatum L. product matrices affected the transport of rutin, hyperoside and isoquercitrin across Caco-2 cells (Gao et al., 2010), probably due to differences in matrix phytochemical composition and transport characteristics, i.e. paracellular transfer, carrier mediated or active transport. In our study, both RA and UA were major components of their respective plant material (~3% or greater by weight, and ~ 5-50 mM of ethanolic plant extract). It could be useful to compare RA or

UA uptake from plant matrices in which these components were less concentrated.

Studies are also needed to investigate the influences of preservation or processing of plant extract on the bioavailability of RA and UA both in vitro and in vivo.

If the basolateral movement of the paracellular marker, LY, is less than 0.7% and TEER value is greater than 250 Ω·cm2, the established monolayer is considered to be tight enough

for permeability experiments (Ohashi et al., 2009). In our permeability studies, we tested

the basolateral recoveries at time points and interval times chosen for a hydrophilic 110

paracellularly transferred compound (RA) (Hubatsch et al., 2007). RA uptake was not

saturable during 4 h incubation period (Figure 3A), indicating that passive diffusion

occurred for this parent compound. We only measured the uptake of RA at 10 µM due to

its cytotoxicity; RA was not detected at 0.5 and 1 h because of its LOD and the low

concentration applied to the apical chamber. Unidirectional transport (apical to

basolateral transfer) was investigated for RA in our study and the pH of both apical and basolateral solutions was 7.4 (to simulate small intestine pH), because Konishi and

Kobayashi (2005) showed that permeation of RA from the apical to the basolateral side is

similar to that from basolateral to apical side; the uptake in the presence of a proton

gradient (pH gradient 6.0/7.4) was nearly the same as that in the absence of a proton

gradient (pH gradient 7.4/.4), implying that proton coupled polarized transport was not

involved for RA. The transepithelial flux of RA was inversely correlated with TEER in

our experiment (data not shown), consistent with restriction of intestinal absorption of RA when the epithelial tight junction is tight enough, a finding similar to that shown for another caffeic acid derivative, chlorogenic acid (Konishi and Kobayashi, 2004).

Our finding of increased basolateral RA recoveries after deconjugation (Figure 3C) is the first direct report of this type of intestinal metabolism of RA; glucuronide/sulfate conjugates were transferred to the basolateral side (toward the circulation) rather than

apical side (the intestinal lumen), but this finding is consistent with in vivo observations of

RA glucuronide/sulfate conjugates as major forms of this compound in rat and human

plasma (Konishi et al., 2005; Baba et al., 2005). These results suggest that RA is absorbed

via both paracellular and transcellular diffusion and MRP (multidrug resistance protein) or

OATP (organic anion transporter protein) transporters might be involved in the permeation

of glucuronidated or sulfated RA, because anionic conjugates (glutathione, glucuronide or 111

sulfate) cannot exit cells unless an MRP transporter is present (Peng et al. 1999).

Intestinal metabolism of RA seemingly differs from that of its possible metabolites,

glucuronide/sulfate conjugates of caffeic acid, ferulic acid and o-coumaric acid which were

found only apically (Kern et al., 2003). This might be due to the specificity of MRP

transporters to the glucuronide/sulfate conjugates of different caffeic acid derivatives and

the locations of various MRP in apical or basolateral sides of the enterocytes (Peng et al.,

1999).

The plant extracts did not affect transfer of either UA or RA, and likewise the pure

compounds and the compounds from the plant extracts behaved similarly in their apparent

biotransformation as well (Figure 3A, B and C and Table 2), establishing that the transfer and metabolism of RA and UA was independent of the of the plant extract matrix for the two herbs studied. The transfer of UA seemed simpler than that observed for RA. The rate of membrane permeation of UA increased linearly with concentration and was not saturable at the tested time-points (Figure 2, 3A and 3B), indicating uptake by passive diffusion. Because the basolateral transfer of UA was greater than RA at each time point

(Figure 3A), and the partition coefficient (Log P) of RA was 0.2, compared with Log P of

UA of 6.4 (Konishi et al. 2005; Bérangère et al., 2004), more hydrophobic UA was more absorbable than RA. This result is consistent with an interpretation of previous in vivo findings in rats that the low plasma concentration of UA was not due to poor absorption but to extensive uptake of UA by other tissues (Liao et al. 2005). UA was absorbed and transferred by Caco-2 cells apparently with little glucuronidation/sulfation (Figure 3C and

Table 2), which is likely to further enhance the bioavailability of UA compared with RA.

This lack of UA biotransformation might be due to the differences in glucuronidation and sulfation of aliphatic alcohol (UA) and phenol alcohol (RA) or because of the distinct 112

structures of the two compounds: triterpenoid (UA) vs. caffeic acid derivative (RA). But

two studies showed that there was wide variability in affinity for

UDP-glucuronosyltransferase (UGT) and sulfotransferase (ST) across both general classes

of substrates (Chen et al., 1996; Ebner and Burchell, 1993). One related triterpenoid,

glycyrrhetinic acid (50-300 µM), was found to inhibit UGT2B7 activity in human liver

microsomes (Nakagawa et al., 2009), which might also occur for UA. To our knowledge,

no other in vivo or in vitro study has been done on the metabolism of ursolic acid. It is

possible that UA may undergo additional glucuronide or sulfate conjugation in the liver.

Future studies are required to elucidate the structures of RA and UA metabolites, which affect compound transfer mechanism and efficacy.

In conclusion, RA was transferred across Caco-2 cells almost entirely in conjugated form, but UA was absorbed and transferred mostly intact, both independent of extract matrix.

Our results predict no effect of plant matrix on the efficacy of either compound. It will be important to measure the uptake of gut microbial metabolites of RA and effects of RA on expression of efflux transporters and tight junction proteins which would particularly affect a compound transferred paracellularly such as RA. Cellular models of oral as well as intestinal uptake, metabolism and anti-inflammatory activity of RA and P. vulgaris and UA and S. officinalis will be of interest as well, given traditional uses of these herbs.

Acknowledgements

This publication was made possible by grant number P01 ES012020 from the National

Institute of Environmental Health Sciences (NIEHS) and the Office of Dietary

Supplements (ODS), National Institutes of Health (NIH) and grant number 113

9P50AT004155-06 from the National Center for Complementary and Alternative

Medicine (NCCAM) and ODS, NIH.

114

Figures

Rosmarinic acid Ursolic acid

Figure 1. The chemical structures of rosmarinic and ursolic acids.

115

Figure 2. The transport rate of ursolic acid for pure compound and S. officinalis extract across Caco-2 cell monolayer as a function of concentration at 5–20 μM after 4 h uptake study. S. officinalis extract was diluted to apply the same concentrations to the cells as were used for the pure compounds. Data are the mean ± S. D (n=9). Means bearing different letters were significantly different by ANOVA and Tukey’s multiple comparison (p <

0.01).

116

117

Figure 3. Characteristics of transfer of rosmarinic (RA) and ursolic acids (UA) for pure compounds or plant extracts across Caco-2 cells before and after deconjugation of post-experimental basolateral solutions. A. Before deconjugation. No RA was found at

0.5 or 1 h. Means bearing different letters were significantly different by ANOVA and

Tukey’s multiple comparison (p < 0.01). B. After deconjugation with

β-glucuronidase/sulfatase; for RA and RA in P. vulgaris 27664. C. Total basolateral recovery of RA and UA after deconjugation over 4 h incubation. * The basolateral recoveries of RA after β-glucuronidase/sulfatase incubation were significantly different from that before the enzyme treatment by two-sample t-test (p < 0.01). Basolateral recoveries were calculated as the amount transported divided by the initial amount in the apical chamber during 4 h uptake study. Data are the mean ± S. D (n=9).

118

Table 1. The contents of rosmarinic and ursolic acids in ethanol extracts of P. vulgaris and

S. officinalis determined by HPLC.

RA UA

g/L (mM) g/L (mM)

P. vulgaris 27664 (174.4 g/L) 8.4 ± 0.6 (23.3 ± 1.7) 0.1 ± 0.0 (0.2 ± 0.0)

P. vulgaris 27665 (66.2 g/L) 3.2 ± 0.3 (9.0 ± 0.8) 0.1 ± 0.0 (0.1 ± 0.0)

P. vulgaris 27748 (117.0 g/L) 19.9 ± 1.1 (55.2 ± 3.1) 0.6 ± 0.1 (1.0 ± 0.2)

S. officinalis (10 g/L) ND* 2.6 ± 0.4 (5.8 ± 0.9)

* Not detected.

119

Table 2. Apparent permeability coefficients (Papp) for lucifer yellow, rosmarinic acid, ursolic acid and extracts across the Caco-2 monolayer relative to deglucuronidation.

Before deconjugation After deconjugation c

LY 0.3 ± 0.1 -- RAa 0.2 ± 0.0 0.9 ± 0.2†

RA in P. vulgaris 27664 0.2 ± 0.0 1.0± 0.4†

RA in P. vulgaris 27665 0.2 ± 0.1 0.7 ± 0.1† RA in P. vulgaris 27748 0.1± 0.1 0.9± 0.1† UAb 2.8 ± 0.1* 2.4 ± 0.4* UA in S. officinalis 2.5 ± 0.4* 2.2 ± 0.6*

a RA was 10 μM for pure compound and P. vulgaris extracts containing the same content of

b RA. Papp of UA as a pure compound and in S. officinalis extract was calculated using

non-toxic concentrations (5, 10 and 20 μM).

c The Papp of rosmarinic and ursolic acids after treating with β-glucuronidase/sulfatase.

* Significantly different compared with Papp of LY and RA by two-sample t-test (p < 0.01).

† Significantly different before and after deconjugation by two-sample t-test (p < 0.01).

--: not determined.

120

References

Baba, S., Osakabe, N., Natsume, M., Terao, J., 2004. Orally administered rosmarinic acid

is present as the conjugated and/or methylated forms in plasma, and is degraded and

metabolized to conjugated forms of caffeic acid, ferulic acid and m-coumaric acid.

Life Sciences 75, 165-178.

Baba, S., Osakabe, N., Natsume, M., Yasuda, A., Muto, Y., Hiyoshi, K., Takano, H.,

Yoshikawa, T., Terao, J., 2005. Absorption, metabolism, degradation and urinary

excretion of rosmarinic acid after intake of Perilla frutescens extract in humans.

European Journal of Nutrition 44, 1-9.

Bérangère, C., Caussarieu, N., Morin, P., Morin-Allory, L., Lafosse, M., 2004. Rapid

analysis of triterpenic acids by liquid chromatography using porous graphitic carbon

and evaporative light scattering detection. Journal of Separation Science 27, 964-970.

Chen, G., Banoglu, E., Duffel, M.W., 1996. Influence of substrate structure on the catalytic

efficiency of hydroxysteroid sulfotransferase STa in the sulfation of alcohols.

Chemical Research in Toxicology 9, 67-74.

Cheung, H.Y., Zhang, Q.F., 2008. Enhanced analysis of triterpenes, flavonoids and

phenolic compounds in Prunella vulgaris L. by capillary zone electrophoresis with the

addition of running buffer modifiers. Journal of Chromatography A 1213, 231-238.

Ebner, T., Burchell, B., 1993. Substrate specificities of two stably expressed human liver

UDP-glucuronosyltransferases of the UGT1 gene family. Drug Metabolism and

Disposition 21, 50-55.

Filipic, M., Baricevic, D., 1997. Antimutagenic activity of Salvia officinalis extracts

against UV-induced mutations in E. coli strains. Mutation Research 379, 182-186.

121

Gao, S., Jiang, W., Yin, T., Hu, M., 2010. Highly variable contents of phenolics in St.

John's Wort products affect their transport in the human intestinal Caco-2 cell model:

pharmaceutical and biopharmaceutical rationale for product standardization. Journal

of Agricultural and Food Chemistry 58, 6650-6659.

Hubatsch, I., Ragnarsson, E.G., Artursson, P., 2007. Determination of drug permeability

and prediction of drug absorption in Caco-2 monolayers. Nature Protocols 2,

2111-2119.

Kern, S.M., Bennett, R.N., Needs, P.W., Mellon, F.A., Kroon, P.A., Garcia-Conesa, M.T.,

2003. Characterization of metabolites of hydroxycinnamates in the in vitro model of

human small intestinal epithelium caco-2 cells. Journal of Agricultural and Food

Chemistry 51, 7884-7891.

Konishi, Y., Kobayashi, S., 2004. Transepithelial transport of chlorogenic acid, caffeic acid,

and their colonic metabolites in intestinal caco-2 cell monolayers. Journal of

Agricultural and Food Chemistry 52, 2518-2526.

Konishi, Y., Hitomi, Y., Yoshida, M., Yoshioka, E., 2005. Pharmacokinetic study of

caffeic and rosmarinic acids in rats after oral administration. Journal of Agricultural

and Food Chemistry 53, 4740-4746.

Konishi, Y., Kobayashi, S., 2005. Transepithelial transport of rosmarinic acid in intestinal

Caco-2 cell monolayers. Bioscience, Biotechnology, and Biochemistry 69, 583-591.

Liao, Q., Yang, W., Jia, Y., Chen, X., Gao, Q., Bi, K., 2005. LC-MS determination and

pharmacokinetic studies of ursolic acid in rat plasma after administration of the

traditional Chinese medicinal preparation Lu-Ying extract. Yakugaku Zasshi 125,

509-515.

Liu, J., 1995. Pharmacology of oleanolic acid and ursolic acid. Journal of

Ethnopharmacology 49, 57-68.

122

Loizzo, M.R., Saab, A.M., Tundis, R., Statti, G.A., Menichini, F., Lampronti, I., Gambari,

R., Cinatl, J., Doerr, H.W., 2008. Phytochemical analysis and in vitro antiviral

activities of the essential oils of seven Lebanon species. Chemistry and Biodiversity 5,

461-470.

Lu, Y., Foo, L.Y., 2000. Flavonoid and phenolic glycosides from Salvia officinalis.

Phytochemistry 55, 263-267.

Lu, T., Yang, J., Gao, X., Chen, P., Du, F., Sun, Y., Wang, F., Xu, F., Shang, H., Huang, Y.,

Wang, Y., Wan, R., Liu, C., Zhang, B., Li, C., 2008. Plasma and urinary tanshinol

from Salvia miltiorrhiza (Danshen) can be used as pharmacokinetic markers for

cardiotonic pills, a cardiovascular herbal medicine. Drug Metabolism and Disposition

36, 1578-1586.

Manach, C., Scalbert, A., Morand, C., Remesy, C., Jimenez, L., 2004. Polyphenols: food

sources and bioavailability. The American Journal of Clinical Nutrition 79, 727-747.

Mukinda, J.T., Syce, J.A., Fisher, D., Meyer, M., 2010. Effect of the plant matrix on the

uptake of luteolin derivatives-containing Artemisia afra aqueous-extract in Caco-2

cells. Journal of Ethnopharmacology 130, 439-449.

Nakagawa, N., Katoh, M., Yoshioka, Y., Nakajima, M., Yokoi, T., 2009. Inhibitory effects

of Kampo medicine on human UGT2B7 activity. Drug Metabolism and

Pharmacokinetics 24, 490-499.

Nasser, B., Moustaid, K., Moukha, S., Mobio, T.A., Essamadi, A., Creppy, E.E., 2008.

Evaluation of the cytotoxicity and genotoxicity of extracts of mussels originating from

Moroccan Atlantic coast, in human colonic epithelial cells Caco-2. Environmental

Toxicology 23, 539-547.

Neilson, A.P., George, J.C., Janle, E.M., Mattes, R.D., Rudolph, R., Matusheski, N.V.,

Ferruzzi, M.G., 2009. Influence of chocolate matrix composition on cocoa flavan-3-ol

123

bioaccessibility in vitro and bioavailability in humans. Journal of Agricultural and

Food Chemistry 57, 9418-9426.

Ohashi, A., Fukumuro, M., Sawabe, K., Mamada, K., Sugawara, Y., Matsuoka, H.,

Hasegawa, H., 2009. Transcellular relocation of tetrahydrobiopterin across Caco-2

cells: a model study of tetrahydrobiopterin absorption through epithelial cells of

intestinal mucosa. Journal of Inherited Metabolic Disease 32, 73-78.

Osakabe, N., Yasuda, A., Natsume, M., Sanbongi, C., Kato, Y., Osawa, T., Yoshikawa, T.,

2002. Rosmarinic acid, a major polyphenolic component of Perilla frutescens,

reduces lipopolysaccharide (LPS)-induced liver injury in D-galactosamine

(D-GalN)-sensitized mice. Free Radical Biology and Medicine 33, 798-806.

Peng, K.C., Cluzeaud, F., Bens, M., Duong Van Huyen, J.P., Wioland, M.A., Lacave, R.,

Vandewalle, A., 1999. Tissue and cell distribution of the multidrug

resistance-associated protein (MRP) in mouse intestine and kidney. Journal of

Histochemistry and Cytochemistry 47, 757-768.

Psotová, J., Kolár, M., Sousek, J., Svagera, Z., Vicar, J., Ulrichová, J., 2003. Biological

activities of Prunella vulgaris extract. Phytotherapy Research 17, 1082-1087.

Rasool, R., Ganai, B.A., Akbar, S., Kamili, A.N., Masood, A., 2010. Phytochemical

screening of Prunella vulgaris L.,--an important medicinal plant of Kashmir. Pakistan

Journal of Pharmaceutical Sciences 23, 399-402.

Swarup, V., Ghosh, J., Ghosh, S., Saxena, A., Basu, A., 2007. Antiviral and

anti-inflammatory effects of rosmarinic acid in an experimental murine model of

Japanese encephalitis. Antimicrobial Agents and Chemotherapy 51, 3367-3370.

Song, T., Barua, K., Buseman, G., Murphy, P.A., 1998. Soy isoflavone analysis: quality

control and new internal standard. The American Journal of Clinical Nutrition 68,

1474-1479.

124

Walgren, R.A., Walle, U.K., Walle, T., 1998. Transport of quercetin and its glucosides

across human intestinal epithelial Caco-2 cells. Biochemical Pharmacology 55,

1721-1727.

Zdarilová, A., Svobodová, A., Simánek, V., Ulrichová, J., 2009. Prunella vulgaris extract

and rosmarinic acid suppress lipopolysaccharide-induced alteration in human gingival

fibroblasts. Toxicology in Vitro 23, 386-392.

125

CHAPTER 5. CAFFEIC ACID DERIVATIVES ENHANCED EPITHELIAL

BARRIER INTEGRITY AND EXHIBITED ANTI-INFLAMMATORY

ACTIVITIES IN P-GLYCOPROTEIN TRANSPORTER KNOCKDOWN CACO-2

CELLS

Zhiyi Qiang1, Nan Huang1, Kazuya Maeda2, Li Li1, Manju B. Reddy1, Diane Birt1,

Patricia A. Murphy1, Suzanne Hendrich1

A paper to be submitted to the Journal of Nutrition

1 Department of Food Science and Human Nutrition, Iowa State University, Ames, IA,

USA

2 Department of Medical Pharmaceutics, University of Tokyo, Tokyo, Japan

Abstract

Our hypothesis was that caffeic acid derivatives, namely caffeic acid , rosmarinic acid, chlorogenic acid and m-hydroxyphenylpropionic acid (a microbial metabolite of caffeic acid and rosmarinic acid) altered the intestinal barrier and immune-modulators toward an anti-inflammatory profile in Caco-2 human intestinal epithelial cells and P-glycoprotein

(P-gp) transporter, encoded by multidrug resistance (MDR1) gene, knockdown(KD)

Caco-2 cells. Stimuli composed of interleukin (IL)-1β, tumor necrosis factor (TNF)-α, interferon (IFN)-γ and lipopolysaccharides (LPS) were added to the cell system to represent the acute inflammatory reaction and mimic the inflammatory bowel disease.

Transepithelial electrical resistance (TEER) across the monolayers increased rapidly during incubation with quercetin (positive control) and m-hydroxyphenylpropionic acid

126

(mHPP) after 12 h, and only chlorogenic acid and mHPP both at 50 µM showed protection

against the reduction of TEER induced by the stimuli in Caco-2 cells. In MDR KD

Caco-2 cells, all treatment groups at 50 µM significantly augmented the TEER at each

time-point compared with the control both with and without the stimuli, which was

correlated with Lucifer yellow flux, a paracellular marker, across the cell monolayer.

Immunoblot analysis of tight junction proteins revealed that caffeic acid derivatives at 50

µM up-regulated ZO-1, ZO-2, claudin-1 and occludin in Caco-2 cells, and chlorogenic acid and mHPP were effective against the inflammatory stimuli induced reductionof ZO-2 and occludin, as well as induction of claudin-1 on tight junction proteins in Caco-2 cells.

Caffeic acid derivatives up-regulated claudin-4 in MDR KD Caco-2 cells and only mHPP was effective against the inflammatory stimuli induced changes of tight junction protein expressions in MDR KD Caco-2 cells. Caffeic acid derivatives at 50 µM significantly decreased the concentration of IL-6 in Caco-2 cells, but augmented TNF-α and IL-6 levels under the stimulated condition. In MDR KD caco-2 cells plus the stimuli, all treatment groups significantly reduced both TNF-α and IL-6 concentrations. These results suggest that P-gp plays an essential role in the anti-inflammatory activities of caffeic acid derivatives and mHPP exhibited greater enhancement of intestinal barrier function than the parent compounds.

127

Introduction

Dietary polyphenols in fruit and vegetables are antioxidants and epidemiologic studies have shown a correlation between an increased consumption of phenolic antioxidants and a reduced risk of cardiovascular disease and certain types of cancer (Bravo, 1998). Caffeic acid, one of the most common phenolic acids, frequently found in fruits, vegetables, grains and dietary supplements. Chlorogenic acid is combined from caffeic and quinic acids in coffee and many types of fruits. Rosmarinic acid is also a caffeic acid derivative found in many Lamiaceae herbs used commonly as culinary herbs such as rosemary, oregano and thyme. Phenolic acids such as caffeic acid, chlorogenic acidand and rosmarinic acid

(Figure 1) exhibit antioxidant, anti-inflammatory, anti-mutagenic and anti-carcinogenic effects in vitro (Rice-Evans et al., 1996; Scalbert, et al., 2002; Parnham and Kesselring,

1985), therefore are of interest in inhibiting gastrointestinal inflammation which is relevant to inflammatory bowel disease (IBD) and colon cancer.

Gastrointestinal epithelial cells provide a physical barrier to the diffusion of pathogens, toxins, and allergens from the lumen into the circulatory system. In healthy individuals, the intestinal barrier is constituted of an intact layer of epithelial cells, which are tightly connected by a surrounding system of tight junction strands. In IBD patients, epithelial barrier function is impaired (Gitter et al., 2001) with reduction of tight junction strands in ulcerative colitis and tight junction strand breaks in Crohn’s disease (Marin et al., 1983).

Tight junction strands are formed by the assembly of a multiple-protein complex, which regulates the paracellular movement of ions, solutes, and water across the intestinal epithelium. Four integral transmembrane proteins, occludin, claudins, junctional adhesion molecule, and tricellulin, have been identified (Furuse et al., 1993; Furuse et al.,

128

1998; Ikenouchi, et al., 2005). The intracellular domains of these transmembrane proteins

interact with scaffold proteins, such as zonula occludens (ZO) proteins, which in turn anchor the transmembrane proteins to the perijunctional actomyosin ring. Previous studies showed that some polyphenols, epigallocatechin gallate; genistein; myricetin; kaempferol and quercetin, enhance intestinal tight junctions in human intestinal Caco-2 monolayers (Suzuki and Hara, 2010; Suzuki et al., 2011). No studies have investigated the influences of phenolic acids on gut barrier functions, which might be the underlying mechanism of the protective effects of caffeic acid derivatives against colitis or colon

cancer in animal models (Ye, et al., 2009; Tanaka, et al., 1990; Feng et al., 2010).

Several inflammatory mediators are thought to be implicated in the development of IBD.

The in vitro effects of interleukin (IL)-1β, tumor necrosis factor (TNF)- α, interferon (IFN)-

γ and lipopolysaccharides (LPS) on intestinal epithelial cells have been shown to activate the intracellular cascades and increase the secretion of IL-6 (Vitkus et al., 1998), IL-8

(Schuerer-Maly et al., 1994), prostaglandin E2 (PGE-2 , Wright et al., 2004) and nitric oxide (NO, Forsythe et al., 2002), as well as increase of the paracellular permeability through defects in tight junction assembly (Al-Sadi et al., 2007). IL-1β, TNF-α, and IFN-

γ have been found in increased concentrations in the intestinal tissue of IBD patients

(Ligumsky et al., 1990; Reimund et al., 1996) and LPS plays a key role in the initiation of

IBD because microflora may cause an initial inappropriate inflammatory stimulus if the

intestinal cells are dysregulated, leading to an exaggerated cytokine presence and IBD

development (Caradonna et al., 2000).

P-gp transporter, encoded by multidrug resistance (MDR1) gene, plays a key role in drug

absorption and distribution because it limits the permeability across the GI tract (Hunter

129

and Hirst, 1997) by active efflux of potentially toxic substances back into the intestinal lumen. A great insight into the physiological role of P-gp protein in the GI tract has derived from the phenotype of the MDR1 gene knockout mice. This model was prompted by the finding that the MDR1 gene is present in a region of the human genome (7q21.1) that may harbour a disease gene involved in susceptibility to IBD (Lathrop et al., 1996).

Traditional herbs, such as Prunella vulgaris and Hypericum gentinoides, containing caffeic acid derivatives, have been found to act against the spontaneous colitis in MDR1 knockout mice in terms of increasing the survival percentage, elevating the anti-inflammatory cytokine and chemokines in serum and reducing typhlocolitis score (data not published), but the mechanism is unknown.

This study was conducted to investigate the regulation of caffeic acid, rosmarinic acid, chlorogenic acid and m-hydroxyphenylpropionic acid (mHPP), a microbial metabolite of ingested caffeic acid or rosmarinic acid (Konishi and Kobayashi, 2004b; Konishi and

Kobayashi, 2005), on tight junction proteins and immune modulators in both Caco-2 cell model and MDR knockdown (MDR KD) Caco-2 cells. Stimuli composed of IL-1β,

TNF-α, IFN- γ and LPS were added to Caco-2 cell system to represent the acute inflammatory reaction as it might occur during the active phase of IBD. Our hypothesis was that caffeic acid derivatives altered the intestinal barrier and immune-modulators toward an anti-inflammatory profile and that the protection was stronger in MDR KD cells compared with Caco-2 cells.

130

Materials and methods

Chemicals. Caffeic acid (98%), rosmarinic acid (90%), chlorogenic acid (95%), and m-hydroxyphenylpropionic acid (mHPP, 92%), a metabolite of caffeic acid derivatives, were obtained from Sigma-Aldrich Co. (St. Louis, MO). Rabbit anti-claudin-1, claudin-4,

ZO-1, ZO-2, and occludin were purchased from Zymed Laboratories (Gibco Invitrogen,

Carlsbad, CA). Mouse anti-α-tubulin was purchased from Sigma (St. Louis, MO).

HRP-conjugated anti-mouse and -rabbit IgG were purchased from Zymed Laboratories.

IL-1β, TNF-α, and LPS were purchased from Sigma-Aldrich Co. (St. Louis, MO), and

IFN- γ was obtained from Invitrogen (Carlsbad, CA).

Cell culture. Caco-2 cells were obtained from American Type Culture Collection (ATCC,

Manassas, VA) at passage 18 and all experiments were performed from passages 35–45.

MDR knockdown (MDR KD) Caco-2 cells (a gift from Dr. Kazuya Maeda, the University of Tokyo, constructed from Caco-2 cells from ATCC using siRNA technique, Watanabe et al., 2005) were used between passages 21 and 25. The cells were cultured according to

Hubatsch et al. (2007). The cells were seeded on collagen-coated polytetrafluro- ethylene membrane inserts (0.45 μm) fitted in bicameral chambers (Transwell-COL, 24 mm ID,

Corning Inc., Corning, NY) at 1.2 × 105 cells/cm2. All experiments were conducted on d

20-21 postseeding. Stimuli composed of IL-1β (25 ng/mL), TNF-α (50 ng/mL) and IFN-γ

(50 ng/mL) were applied at the basolateral side of the cells, and LPS (1 µg/mL) was applied at both sides, representing on the one hand the common luminal microflora and, on the other hand, the increased presence of bacterial components in serum during IBD (Walle et al., 2010).

131

P-gp assay. P-gp activity was evaluated with fluorimetric measurement of the intracellular

accumulation of calcein produced by ester hydrolysis of the P-gp substrate calcein-AM

using VybrantTM Multidrug Resistance Assay Kit (Gibco Invitrogen, Carlsbad, CA) in

Caco-2 and MDR KD Caco-2 cells. After the cells in the flask grew to 90–100%

confluency, cells were trypsinized and seeded on 96 well plates at 1.2 × 105 cells/cm2.

Cells were preincubated for 15 min with or without the acknowledged P-gp modulator

verapamil (10 µg/mL), thereafter calcein-AM was added and the fluorescence measured

after 1 hour using microtiter plate reader Bio-Tek ELX 808 (Bio-Tek instruments. Inc.,

Winooski, VT) at 490 nm.

Measurement of intestinal barrier function. Intestinal barrier function was evaluated by measurement of transepithelial electrical resistance (TEER) and unidirectional flux of

lucifer yellow (LY) across Caco-2 cell monolayers in transwells. Caffeic acid, rosmarinic

acid, chlorogenic acid, mHPP (10, and 50 µM) and quercetin (100 µM) were added to the

apical wells and the cells were incubated for 48 h. TEER was measured before and at 6, 12,

24 and 48 h after administration of the caffeic acid derivatives by Millicell ERS meter

(Fisher Sci., Pittsburgh, PA). Lucifer yellow (LY, 100 µM), a paracellular marker, was

added apically and the flux into the basolateral chamber was assessed before and after the

experiment. For LY quantification, the apical and basolateral solutions were transferred to a 96 well plate and read spectrophotometrically at 450 nm. Whole cell extracts were prepared for immunoblot analysis of tight junction proteins after incubation for 48 h as described below. Cytotoxicity of caffeic acid, rosmarinic acid, chlorogenic acid and mHPP was measured according to Nasser et al. (2008).

132

Assessment of Protein Expression by Western Blot Analysis. At the end of the

experimental period, Caco-2 monolayers on the transwell were immediately rinsed with

ice-cold PBS, and cells were lysed with lysis buffer (50 mM Tris-HCl, pH 7.5, 150 mM

NaCl, 10 mM NaF, 5 mM EDTA, 2 mM sodium orthovanadate, 1 M phenylmethylsulfonyl

fluorid, 3 mg/L leupeptin, 5 mg/L aprotinin, and 1% Triton X-100) and scraped, and the

cell lysates were placed in microfuge tubes. Cell lysates were centrifuged to yield a clear

lysate. Supernatant was collected, and total cellular protein was determined by Coomassie

(Bradford) assay (Pierce Laboratories, Rockford, IL). Laemmli gel loading buffer (6%

SDS, 30% glycerol, 15% 2-b-mercaptoethanol, and 0.02% bromophenol blue in 188 mM

Tris, pH 6.8) was added to the lysate containing 20 µg of protein and boiled for 7 minutes, after which proteins were separated on SDS-PAGE gel. Proteins from the gel were transferred to polyvinylidene difluoride membranes. Membranes were blotted for ZO-1,

ZO-2, occludin, claudin-1, claudin-4 and α-tubulin using specific antibodies in combination with HRP-conjugated anti-mouse IgG or anti-rabbit IgG antibodies. The blots were developed using the ECL chemiluminescence method (GE Healthcare,

Piscataway, NJ). Quantification was performed by using densitometric analysis of specific bands on the immunoblots.

Determination of TNF-α and IL-6 secretion. At the end of the experimental period, the extracellular media in the transwell were collected and centrifuged at 10,000 Xg for 10 min.

TNF-α and IL-6 secretions were evaluated using a sandwich ELISA method (BD

Biosciences Pharmingen, San Diego, CA) and quantified in pg/ml using the standard

provided with the kit, with sensitivity limits being of 7.8 and 4.7 pg/ml, respectively for the

TNF-α and IL-6 assay.

133

Statistical analysis Data are given as means ± S. D. Differences in the percentage of the

TEER relative to the initial value, lucifer yellow flux, tight junction protein density, ,

cytokine concentrations and fluorimetric intracellular accumulation of calcein were

evaluated statistically using ANOVA and Tukey’s multiple comparison tests by SAS 9.1

(SAS Institute Inc., Cary, NC). Differences were considered significant at p < 0.05.

Results

P-gp activity in Caco-2 cells and MDR KD Caco-2 cells. As shown in Figure 2,

P-gp-mediated efflux of calcein-AM, a substrate for P-gp, was significantly less in MDR

KD Caco-2 cells with or without the treatment of P-gp inhibitor, verapamil, and in Caco-2

cells incubated with verapamil, compared with Caco-2 cells (p < 0.05). This indicates that

the P-gp function in the MDR KD Caco-2 cells was almost completely inhibited.

Caffeic acid derivatives enhance tight junction integrity in Caco-2 cells. In Caco-2

cells, 100 µM quercetin and 50 µM mHPP significantly increased TEER after 6 h and 12 h

incubation, respectively, compared with control and caffeic acid, rosmarinic acid and

chlorogenic acid (p < 0.05). The TEER in all treatment groups was higher than that of

control after 48 h incubation period (p < 0.05, Figure 3A). This was correlated with the

paracellular transport of LY, which was significantly reduced in all treatment groups

compared with the control after 48 h incubation (data not shown). After adding the

stimuli, the TEER diminished up to 35% (data not shown) in controls compared with that without adding the stimuli (273 ± 15 vs 420 ± 11 Ω·cm2), and only chlorogenic acid and mHPP both at 50 µM showed protection against the adverse effects of the stimuli on the

TEER (Figure 3B). Fifty µM of caffeic acid, rosmarinic acid or chlorogenic acid exacerbated the diminishment of the TEER induced by the stimuli (Figure 3B), which was

134

correlated with LY flux across the Caco-2 cells (Figure 4A), indicating that quercetin,

caffeic acid, and rosmarinic acid increased the permeability of the Caco-2 cells when the

stimuli were added to initiate the inflammatory responses.

In MDR KD Caco-2 cells, all treatment groups significantly augmented the TEER at each

time-point compared with the control (p < 0.05, Figure 3C). Interestingly, the elevation of the TEER by caffeic acid derivatives-incubated cells was greater than quercetin in MDR

KD cells compared with Caco-2 cells (Figure 3A and C), supported by the flux of LY in to the basolateral side after 48 h incubation (data not shown). After adding the stimuli in to

MDR KD Caco-2 cells, the TEER was significantly reduced up to 29% (321 ± 16 vs 453 ±

23 Ω·cm2) in controls compared with that without adding the stimuli (data not shown), and

all treatment groups significantly increased the TEER relative to the initial value compared

with controls (p < 0.05, Figure 3D). Among the treatment groups, mHPP showed greatest

protection against the stimuli in the elevation of the TEER (p < 0.05), correlated with the

flux of LY across the cell monolayer (Figure 3D and 4B).

Caffeic acid induces the up-regulation of tight junction proteins. After adding the

stimuli, the expression of ZO-1, ZO-2 and occludin was all significantly decreased, and

that of claudin-1 and claudin-4 increased in controls compared with without adding the

stimuli (p < 0.05, Table 1 and 2). Without the stimuli, the expression of ZO-1, ZO-2,

claudin-1 and occludin in Caco-2 cells was increased by 50 µM of caffeic acid derivatives

(p < 0.05, Table 1), but not at 10 µM (data not shown). After adding the stimuli,

chlorogenic acid and mHPP exhibited reversal of the response to inflammation:

up-regulated ZO-1 (not for mHPP), ZO-2 and occludin, as well as down-regulated

claudin-1 expressions compared with controls. But this protection was not found for

135

claudin-4 in caffeic acid derivative treated Caco-2 cells, while quercetin at 100 µM attenuated the effect of the stimuli on the expression of claudin-4.

In MDR KD Caco-2 cells after adding the stimuli, similar trend of the changes of tight junction proteins in controls was found in MDR KD Caco-2 cells, compared with Caco-2 cells. No treatment effect was found for ZO-1, ZO-2 and occludin (p > 0.05). Quercetin, chlorogenic acid and mHPP increased claudin-1 expression and all treatment groups increased claudin-4 expression. Caffeic acid derivatives significantly increased the expression of ZO-1 (not for chlorogenic acid) and occludin, and reduced the expression of claudin-1 and claudin-4 induced by the stimuli. Only mHPP was effective in the elevation of the expression of ZO-2 in MDR KD cells with stimuli (Table 2).

Caffeic acid derivatives reduced TNF-α and IL-6 in MDR KD Caco-2 cells. TNF-α and IL-6 were not detected in Caco-2 cells both with or without treatment with caffeic acid or its derivatives (data not shown). After adding the stimuli, TNF-α and IL-6 increased to

357.6 ± 14.1 pg/mL and 6.3 ± 1.8 pg/mL in controls in Caco-2 cells. Rosmarinic acid and mHPP increased TNF-α and IL-6 concentrations compared with other treatments (Table 3).

In MDR KD Caco-2 cells, the level of TNF-α and IL-6 in control was greater than that in the Caco-2 cells (Table 4). After adding the stimuli, the level of TNF-α and IL-6 in controls further increased (Table 5). All treatment significantly reduced the concentration of TNF-α compared with control, while no treatment effect was found for IL-6 in MDR KD

Caco-2 cells without stimuli (Table 4). After adding the stimuli, all treatments significantly decreased both cytokine concentrations (Table 5).

136

Transepithelial transfer of caffeic acid derivatives across Caco-2 and MDR KD

Caco-2 cells. Basolateral recoveries of caffeic acid and mHPP were 4 fold and 3.3 fold

greater in MDR KD Caco-2 cells than that in Caco-2 cells. The amount of caffeic acid and

mHPP in MDR KD Caco-2 cell lysates was ~3 fold higher than that in Caco-2 cells. But

no significant difference was found for the amount of quercetin, rosmarinic acid and

chlorogenic acid in both basolateral side and cell lysates between Caco-2 and MDR KD

Caco-2 cells (Table 6). These results suggest that caffeic acid and mHPP, but not

quercetin, rosmarinic acid and chlorogenic acid, might be substrates for P-gp transporter.

Discussion

This study demonstrated that caffeic acid derivatives enhanced tight junction integrity

through the expression of tight junction proteins in both Caco-2 cells, which are widely

used to study the intestinal epithelium, and in MDR KD Caco-2 cells. The increment of

TEER induced by caffeic acid derivatives was evidenced by the increase of the tight

junction protein expression. Although the precise mechanisms underlying caffeic acid

-mediated promotive effect on tight junction integrity remain to be clarified, P-gp seems important in the regulation of both tight junction proteins and immunomodulators. mHPP exhibited greater enhancement of intestinal barrier function than other caffeic acid derivatives, indicating the key role of this metabolite formed by microbes on the protection of gut function in vivo.

Tight junction proteins bound to the actin cytoskeleton have a pivotal role in tight junction integrity. Tight junction protein levels correlate with the permeability of tight junction in

the epithelium (Betanzos et al., 2003; Basuroy, et al., 2006). The caffeic acid

137

derivatives-mediated increase of TEER was related to the increased ZO-1, ZO-2, occludin,

and claudin-1, but not claudin-4 in Caco-2 cells. The flux of LY through paracellular

pathway was reduced as expected, in conjuction with increased occludin which seemingly

made the paracellular barrier tighter. These results indicate that the assemblies of these

four proteins at the tight junction are responsible for the caffeic acid derivatives-mediated

promotion of intestinal barrier function.

The intestinal inflammatory response requires the correct integration of the various signals

that are released or received by the different cell types present in the gut (Van de Walle et

al., 2010). The regulated expression of a specific array of proinflammatory cytokines in

intestinal epithelial cell lines suggests that these cells are programmed to provide chemotactic and activating signals to adjacent and underlying immune and inflammatory

cells in the initial period following microbial invasion of epithelial cells and further serve to

amplify the mucosal inflammatory response (Jung et al., 1995). In our study, a cocktail

composed of IL-1β, TNF-α, IFN- γ and LPS was utilized to represent the acute

inflammatory reaction as it might occur during the active phase of IBD. After adding the

cocktail, the expressions of claudin-1 and 4 were significantly increased, while those of

ZO-1, ZO-2 and occludin were significantly reduced compared with that without stimuli

treatment (Table 1). Previous studies in cultured monolayers and animal models have

demonstrated that TNF-α which is central to Crohn disease pathogenesis, causes tight

junction barrier dysfunction, such as induction of the expression of claudins via a process

that requires myosin light chain kinase (MLCK) activation (Watterson et al., 2005;

Zolotarevsky et al., 2002) and reduction of occludins by endocytosis of the specific protein

(Wang et al., 2005), which in turn lead to the development of transient leaks and increases of the number of small pores (Weber et al., 2010). MLCK activation is observed in

138

intestinal epithelia of patients with active IBD (Blair et al., 2006) and increased claudin-1 and claudin-4 expression has also been reported in intestinal epithelial cells of patients with

IBD (Heller et al., 2005). Therefore, the use of a cocktail of inflammatory stimuli at plausible concentrations on differentiated Caco-2 cells was aimed to better reflect the in vivo situation during ongoing IBD.

Among the tested caffeic acid derivatives, only chlorogenic acid and mHPP showed protection of the intestinal barrier function in terms of acting against the suppression of

ZO-1, ZO-2 and occludin, as well as the induction of claudin-1 induced by the inflammatory stimuli in Caco-2 cells (Table 1). Our finding of the influence of caffeic acid derivatives on tight junction proteins is novel in showing this role of phenolic acid in gut barrier protection. Previous studies showed that two flavonoids, quercetin and kaempferol, promoted the assembly of tight junction proteins, claudin-1, occludin, and

ZO-2, and the expression of claudin-4 through the inhibition of PKC. In our study, caffeic acid derivatives, belonging to phenolic acids, were also able to promote the gut barrier function both with and without inflammatory stimuli, and the mechanism might be similar to the flavonoids, because caffeic acid at 70 µM inhibited phosphorylase kinase (PhK), protein kinase A (PKA), and PKC by about 90, 59, 57% in vitro, similar to quercetin

(Nardini et al., 2000). However, we can not rule out that caffeic acid derivatives may act differently compared with flavonoids as they are ionic in a neutral media (DMEM at pH 7.4 in our study), which might explain the different response of claudin-4 to phenolic acids vs. flavonoids, because claudin-4 creates an “electrostatic selectivity filter” controlling overall resistance and charge selectivity of the small pores (Anderson et al., 2009). Over expression of claudin-4 in Madin-Darby Canine Kidney (MDCK) epithelial cell lines decreased transmonolayer conductance by decreasing paracellular Na+ permeability

139

without affecting permeability for Cl– or flux for a noncharged solute. Therefore, overexpression of a claudin can confer paracellular ionic discrimination (Van Itallie et al.,

2001). Further, in our study, over-expressing claudin-4 did not stimulate an overall increase in the levels of other tight-junction components in MDR KD Caco-2 cells (Table

2), consistent with previous studies showing that in epithelial cells claudin-1 and claudin-4 can assemble into tight junction strands without the participation of either ZO-1 or occludin in MDCK cells (McCarthy et al., 2000; Van Itallie et al., 2001). One possible role for the

ZO proteins is to scaffold the integral membrane proteins and link them to the actin cytoskeleton (17). That levels of occludin and ZO measured do not diminish in the face of claudin-4 over-expression suggests that claudin-4 does not act to replace these integral membrane tight junction proteins, but adds to them (Van Itallie et al., 2001). Occludin is an integral part of tight junction strands (Fujimoto, 1995), but not essential for strand assembly because occludin deficient embryonic stem cells formed tight junctions (Saitou et al., 1998). Therefore, future studies should focus on the molecular mechanism of the regulation of caffeic acid derivatives on claudins because the claudin family not only define tight junction permeability and affect barrier function, but also enhance cell proliferation and regeneration, neoplastic transformation, tumor growth and metastasis in experimental models (Dhawan et al., 2005).

Of the several transport pumps investigated, MDR1 (i.e., P-gp) has been studied the most extensively in the context of overcoming chemo-resistance. Interestingly, P-gp was also proposed to function as a ‘gatekeeper’ at different tissue barriers. In our study, MDR KD

Caco-2 cells responded differently compared with Caco-2 cells after exposure to caffeic acid derivatives in that only claudins were up-regulated but not for ZO and occludins. The interaction between P-gp transporter and tight junction proteins in intestinal cells is

140

unknown yet; but in blood testis barrier, P-gp co-localizes with several tight junction

proteins such as occludin, claudin-11, JAM, ZO-1, N-cadherin and β-catenin. Meanwhile,

P-gp co-localized and co-immunoprecipitated with actin, as well as with ezrin, radixin and

moesin (ERM), a small family of adaptor proteins involved in linking tight junction proteins to actin filaments (Luciani et al., 2002). In our study, when MDR gene was knocked down, the up-regulation of caffeic acid derivatives on ZO and occludin proteins disappeared, while enhancement of claudin-4 occurred compared with Caco-2 cells with high expression of P-gp transporter (Table 2). This indicates that P-gp–actin interactions might be lost after the knockdown of MDR gene and that actin and ERM binding might be critical for caffeic acid derivatives to exhibit their protection, as well as that regulation of claudins, which are essential for forming tight junction strands, might be different with occludin and ZO proteins, which were indicated to be regulated by prolonged actin–myosin contraction produced contractile forces (Yao and Tsirka, 2011).

Intestinal epithelial cells produce various inflammatory mediators and participate in the development of intestinal inflammation. One of our key findings is that 50 µM of caffeic acid derivatives enhanced IL-6 and rosmarinic acid and mHPP augmented TNF-α secretion

in intestinal epithelial cells after inflammatory stimulation (Table 1). Caffeic acid did not

have significant inhibitory effects on stimulated cytokines in the present study, although

caffeic acid phenethyl ester and methyl ester, in the range of 25-100 µM, inhibited the

translocation of the transcription factor nuclear factor kappa B (NF-κB) and subsequent production and release of TNF-α in LPS induced RAW264.7 cells (Shin et al., 2004). This

may be due to lower membrane permeability of caffeic acid compared with its more

lipophilic esters. Previous studies showed that 3 µM rosmarinic acid down-regulated the

levels of TNF-α, IL-6, and high-mobility group box 1 protein, inhibited the IκB kinase

141

pathway, and modulated NF-κB in LPS induced RAW264.7 cells or human gingival

fibroblasts (Jiang et al., 2009; Zdarilová et al., 2009). In our study, rosmarinic acid and

mHPP both at 50 µM exaggerated rather than attenuated the excretion of TNF-α and IL-6

under the stimulated condition (Table 3). This difference might be due to the tested cell lines or the stimuli applied to the model. We utilized the endogenous mediators IL-1β,

TNF-α, and IFN- γ whose mucosal levels are increased during IBD as well as exogenous

LPS (Fais et al., 1991; Reimund et al., 1996; Caradonna et al., 2000) concomitantly affect the inflammatory response in intestinal epithelial cells. The concentration of LPS used in

our study (1 µg/mL) was greater than Jiang’s study (0.1 µg/mL alone, 2009). Thus, caffeic acid derivatives may excerbate the inflammatory situation when stimuli were presented at the intestinal epithelial cells.

More interestingly, when P-gp transporter, encoded by MDR gene, was knocked down in

Caco-2 cells, all treatments attenuated the inflammatory responses initiated by the stimuli in terms of the reduction of TNF-α and IL-6 levels (Table 4). There might be two reasons for the stronger inhibition of caffeic acid derivatives on the inflammatory responses. First, tested compounds might more easily enter the cells to exhibit their anti-inflammatory activities, which was proved by the finding that cellular content of caffeic acid and mHPP in the basolateral side and cell lysates of MDR KD Caco-2 system was ~3 fold greater than that in Caco-2 cells both with or without the stimuli (Table 6), but not for rosmarinic acid

and chlorogenic acid, whose transport across the cell monolayer was mainly via

paracellular pathway (Konishi and Kobayashi, 2004; Konishi and Kobayashi, 2005).

Secondly, several transcription factors, such as pregnane X receptor, constitutive androstane receptor, NF-κB, activator protein 1, etc., function as regulators for most ABC

transporter proteins, i.e. P-gp, multidrug resistance-associated protein, organic anion

142

transporter, etc. So knockdown of the MDR gene might affect transcriptional cascades that in turn affect inflammatory mediators.

In conclusion, caffeic acid derivatives enhanced barrier function in human intestinal

Caco-2 cells and mHPP exhibited greater enhancement of intestinal barrier than the parent compounds. P-gp plays an essential role in the anti-inflammatory activities of caffeic acid derivatives. The present findings would suggest that the caffeic acid derivatives could be gut health promoting as a dietary constituent, but the activities under inflammatory situation might need further mechanistic or in vivo studies. Moreover, intake of caffeic acid derivatives might speed up mucosal recovery or provide protection to the small-intestinal mucosa against the inflammatory mediators when P-gp inhibitors are co-administered, which are attractive from a therapeutic point of view. In view of the abundance and possible medicinal properties of plant or food derived polyphenols, this field deserves continuing studies to ascertain any actual therapeutic benefits in humans.

Acknowledgements

This publication was made possible by grant number P01 ES012020 from the National

Institute of Environmental Health Sciences (NIEHS) and the Office of Dietary

Supplements (ODS), National Institutes of Health (NIH) and grant number

9P50AT004155-06 from the National Center for Complementary and Alternative

Medicine (NCCAM) and ODS, NIH.

143

Figures

Figure 1. The chemical structures of caffeic acid, rosmarinic acid, chlorogenic acid and m-hydroxyphenylpropionic acid (mHPP).

144

Figure 2. P-glycoprotein transporter activity in Caco-2 cells and multidrug resistance

(MDR) gene knockdown (KD) Caco-2 cells. Verapamil was used as positive control at

10 µg/mL. Data are the mean ± S. D (n=6). Means bearing different letters were significantly different by ANOVA and Tukey’s multiple comparison (p < 0.05).

145

146

Figure 3. TEER in Caco-2 cell monolayers incubated with or without caffeic acid

derivatives (50 µM) and quercetin (100 µM). (A) Caco-2 cells; (B) Caco-2 cells treated

with the stimuli composed of IL-1β (25 ng/mL), TNF-α (50 ng/mL), IFN-γ (50 ng/mL) and

LPS (1 µg/mL) for 24hs; (C) MDR KD Caco-2 cells; (D) MDR KD Caco-2 cells treated with the stimuli composed of IL-1β (25 ng/mL), TNF-α (50 ng/mL), IFN-γ (50 ng/mL) and

LPS (1 µg/mL) for 24hs. Control was 0.3 % DMSO in DMEM. Values are expressed as

147

means ± SD, n = 4. Means bearing different letters at a time were significantly different by ANOVA and Tukey’s multiple comparison (p < 0.05).

148

Figure 4. Lucifer yellow flux across cell monolayers incubated with or without caffeic

acid derivatives (50 µM) and quercetin (100 µM) for 48 h incubation. (A) Caco-2 cells

treated with the stimuli composed of IL-1β (25 ng/mL), TNF-α (50 ng/mL), IFN-γ (50

ng/mL) and LPS (1 µg/mL) for 24 hs; (C) MDR KD Caco-2 cells treated with the stimuli

composed of IL-1β (25 ng/mL), TNF-α (50 ng/mL), IFN-γ (50 ng/mL) and LPS (1 µg/mL) for 24 hs. Control was 0.3 % DMSO in DMEM. Values are expressed as means ± SD, n

149

= 4. Means bearing different letters were significantly different by ANOVA and Tukey’s multiple comparison (p < 0.05). The dash line indicates the baseline of lucifer yellow flux across cell monolayers in each cell type.

150

Table 1. Immunoblot analysis of tight junction proteins in the whole extracts of Caco-2

cells incubated with caffeic acid derivatives (50 µM) and quercetin (100 µM) for 48 hs,

quantitated by densitometric analysis a.

Control Quercetin Caffeic Rosmarinic Cholorogenic mHPP acid acid acid ZO-1 2.2 ± 0.1d 2.3 ± 0.2d 2.9 ± 0.1b 3.3 ± 0.2a 2.5 ± 0.2c 2.9 ± 0.1b 1.3 ± 0.1f 1.1 ± 0.1g 1.4 ± 0.2f 1.5 ± 0.2f 1.7 ± 0.3e 1.3 ± 0.1f

ZO-2 0.4 ± 0.01b 0.5 ± 0.01a 0.5 ± 0.01a 0.5 ± 0.02a 0.5 ± 0.02a 0.5 ± 0.02a 0.3 ± 0.01c 0.3 ± 0.03c 0.3 ± 0.02c 0.4±0.03b 0.4±0.02b 0.5±0.01a

Claudin -1 0.6 ± 0.01f 0.7 ± 0.01cde 0.7 ± 0.01de 0.7 ± 0.02e 0.7 ± 0.02de 0.7 ± 0.02de 0.7 ± 0.02bcd 0.7 ± 0.01bc 0.7 ± 0.01b 0.8 ± 0.04a 0.6 ± 0.03f 0.6 ± 0.03f

Claudin -4 0.3 ± 0.02g 0.4 ± 0.01f 0.3 ± 0.01g 0.3 ± 0.01g 0.3 ± 0.01g 0.3 ± 0.03g 0.6 ± 0.01d 0.4 ± 0.02e 0.6 ± 0.01d 0.7 ± 0.02b 0.7 ± 0.04a 0.6 ± 0.01c

Occludin 11.7 ± 0.02f 12.3 ± 0.02e 12.9 ± 0.1c 14.6 ± 0.2b 12.6 ± 0.3d 14.8 ± 0.08a 10.6 ± 0.1h 11.7 ± 0.1f 7.6 ± 0.1j 9.4 ± 0.1i 11.7 ± 0.2g 14.6 ± 0.1b

a BOLD denotes cells treated with inflammatory cocktail, which was composed of IL-1β

(25 ng/mL), TNF-α (50 ng/mL), IFN-γ (50 ng/mL) and LPS (1 µg/mL) and incubated for

24 hs. The density values were normalized by the internal standard, α-tubulin. Control

was 0.3 % DMSO in DMEM. Values are means ± SD, n = 4. Significant changes within a

protein are noted by letters compared with untreated cells by ANOVA and Tukey’s

multiple comparison (p < 0.05).

151

Table 2. Immunoblot analysis of tight junction proteins in the whole extracts of MDR KD

Caco-2 cells incubated with caffeic acid derivatives (50 µM) and quercetin (100 µM) for 48

hs, quantitated by densitometric analysis a.

Control Quercetin Caffeic acid Rosmarinic Cholorogenic mHPP acid acid ZO-1 1.8 ± 0.03a 1.7 ± 0.2a 1.9 ± 0.1a 1.7 ± 0.1a 1.7 ± 0.1a 1.7 ± 0.1a 1.4 ± 0.04b 1.7 ± 0.1a 1.7 ± 0.1a 1.7 ± 0.1a 1.5 ± 0.1b 1.7 ± 0.2a

ZO-2 0.6 ± 0.01a 0.6 ± 0.01a 0.6 ± 0.01a 0.6 ± 0.05a 0.6 ± 0.07a 0.6 ± 0.02a 0.4 ± 0.01b 0.4±0.03b 0.4±0.01b 0.4±0.03b 0.4±0.01b 0.5±0.03a

Claudin -1 0.7 ± 0.01c 0.9 ± 0.06b 0.6 ± 0.02c 0.7 ± 0.01c 0.8 ± 0.02b 0.7 ± 0.04b 1.0 ± 0.2a 0.8 ± 0.1b 0.7 ± 0.04c 0.8 ± 0.03bc 0.7 ± 0.03c 0.7 ± 0.01c

Claudin -4 0.6 ± 0.04g 0.7 ± 0.01c 0.6 ± 0.01d 0.7 ± 0.01b 0.7 ± 0.02d 0.7 ± 0.01bc 0.7 ± 0.01bc 0.8 ± 0.02a 0.6 ± 0.02de 0.6 ± 0.01fg 0.6 ± 0.01ef 0.6 ± 0.01ef

Occludin 11.2 ± 0.08a 11.2 ± 0.06a 11.1 ± 0.5a 11.2 ± 0.3a 11.1 ± 0.3a 11.0 ± 0.04a 4.8 ± 0.3e 4.7 ± 0.1e 7.3 ± 0.02b 5.7 ± 0.1d 6.3 ± 0.1c 7.7 ± 0.1b

a BOLD denotes cells treated with inflammatory cocktail, which was composed of IL-1β

(25 ng/mL), TNF-α (50 ng/mL), IFN-γ (50 ng/mL) and LPS (1 µg/mL) and incubated for

24 hs. The density values were normalized by the internal standard, α-tubulin. Control

was 0.3 % DMSO in DMEM. Values are means ± SD, n = 4. Significant changes within a

protein are noted by letters compared with untreated cells by ANOVA and Tukey’s

multiple comparison (p < 0.05).

152

Table 3. Caffeic acid derivatives (50 µM, 48 h incubation) augmented the inflammatory responses after treating Caco-2 cells with the stimuli a.

TNF-α (pg/mL) IL-6 (pg/mL) Control 357.6 ± 14.1b 6.3 ± 1.8c

Quercetin 486.2 ± 95.4ab 16.8 ± 2.1a

Caffeic acid 517.0 ± 193.5ab 9.8 ± 0.3b

Rosmarinic acid 665.1 ± 8.6a 21.2 ± 6.6a

Chlorogenic acid 338.6 ± 1.7b 9.1 ± 5.2a

mHPP 604.4 ± 27.3a 18.3 ± 1.0a a Values are means ± SD, n = 4. Control was 0.3 % DMSO in DMEM. Means bearing different letters were significantly different within each column by ANOVA and Tukey’s multiple comparison (p < 0.05). Stimuli were composed of IL-1β (25 ng/mL), TNF-α (50 ng/mL), IFN-γ (50 ng/mL) and LPS (1 µg/mL) and incubated for 24 hs.

153

Table 4. Caffeic acid derivatives (50 µM, 48 h incubation) attenuated the inflammatory responses in MDR KD Caco-2 cells a.

TNF-α (pg/mL) IL-6 (pg/mL)

Control 63.2 ± 28.3a 15.2 ± 0.7ab

Quercetin ND 16.4 ± 0.2ab

Caffeic acid ND 18.1 ± 0.4a

Rosmarinic acid ND 20.7 ± 2.7 a

Chlorogenic acid ND 11.4 ± 0.1b

mHPP ND 20.3 ±4.5a

a Values are means ± SD, n = 4. Control was 0.3 % DMSO in DMEM. Means bearing

different letters were significantly different within each column by ANOVA and Tukey’s

multiple comparison (p < 0.05). ND= Not detected. The limit of detection was 7.8 and

4.7 pg/ml for the TNF-α and IL-6 assay, respectively.

154

Table 5. Caffeic acid derivatives (50 µM, 48 h incubation) attenuated the inflammatory responses in MDR KD Caco-2 cells treated with the stimuli a.

TNF-α (pg/mL) IL-6 (pg/mL)

Control 122.2 ± 12.3a 27.9 ± 4.6a

Quercetin ND 11.3 ± 5.4b

Caffeic acid ND 8.4 ± 0.2b

Rosmarinic acid ND 11.4 ± 1.2b

Chlorogenic acid ND 8.3 ± 2.1b

mHPP ND 8.4 ± 1.0b

a Values are means ± SD, n = 4. Control was 0.3 % DMSO in DMEM. Means bearing

different letters were significantly different within each column by ANOVA and Tukey’s

multiple comparison (p < 0.05). ND= Not detected. Stimuli were composed of IL-1β (25

ng/mL), TNF-α (50 ng/mL), IFN-γ (50 ng/mL) and LPS (1 µg/mL) and incubated for 24 hs.

The limit of detection was 7.8 and 4.7 pg/ml for the TNF-α and IL-6 assay, respectively.

155

Table 6. Transfer of quercetin (100 µM) and caffeic acid derivatives (50 µM) across

Caco-2 and MDR KD Caco-2 cells a.

Basolateral side Cell lysate

Caco-2 (%) MDR KD (%) Caco-2 (%) MDR KD (%)

Quercetin 3.3 ± 0.7 4.2 ± 1.4 0.2 ± 0.1 0.3 ± 0.1

Caffeic acid 2.4 ± 0.2 9.6 ± 0.3* 0.5 ± 0.3 1.8 ± 0.4*

Chlorogenic acid 1.3 ± 0.5 2.3 ± 1.1 ND ND

Rosmarinic acid 0.9 ± 0.2 1.1 ± 0.4 ND ND

mHPP 2.5 ± 0.6 8.4 ± 1.4* 0.7 ± 0.3* 2.2 ± 0.6* a Values are the mean ± SD (n=4). Control was 0.3 % DMSO in DMEM. * indicates significant difference between Caco-2 and MDR KD Caco-2 cells by two-sample t-test (p <

0.05). ND= Not detected. % represents the recovery relative to the initial amount.

156

References

Al-Sadi, R. M. & T. Y. Ma (2007) IL-1beta causes an increase in intestinal epithelial tight

junction permeability. J Immunol, 178, 4641-9.

Amasheh, S., T. Schmidt, M. Mahn, P. Florian, J. Mankertz, S. Tavalali, A. H. Gitter, J. D.

Schulzke & M. Fromm (2005) Contribution of claudin-5 to barrier properties in

tight junctions of epithelial cells. Cell Tissue Res, 321, 89-96.

Amasheh, M., Schlichter, S. Amasheh, S. Mankertz, J. Zeitz, M. Fromm & M. Schulzke, J.

D (2008) Quercetin enhances epithelial barrier function and increases claudin-4

expression in Caco-2 cells. J Nutr, 138, 1067-73.

Anderson, J. M. & C. M. Van Itallie (2009) Physiology and function of the tight junction.

Cold Spring Harb Perspect Biol, 1, a002584.

Basuroy, S., A. Seth, B. Elias, A. P. Naren & R. Rao (2006) MAPK interacts with occludin

and mediates EGF-induced prevention of tight junction disruption by hydrogen

peroxide. Biochem J, 393, 69-77.

Blair, S. A., S. V. Kane, D. R. Clayburgh & J. R. Turner (2006) Epithelial myosin light

chain kinase expression and activity are upregulated in inflammatory bowel disease.

Lab Invest, 86, 191-201.

Caradonna, L., L. Amati, T. Magrone, N. M. Pellegrino, E. Jirillo & D. Caccavo (2000a)

Enteric bacteria, lipopolysaccharides and related cytokines in inflammatory bowel

disease: biological and clinical significance. J Endotoxin Res, 6, 205-14.

Claude, P. (1978) Morphological factors influencing transepithelial permeability: a model

for the resistance of the zonula occludens. J Membr Biol, 39, 219-32.

Clayburgh, D. R., T. A. Barrett, Y. Tang, J. B. Meddings, L. J. Van Eldik, D. M. Watterson,

L. L. Clarke, R. J. Mrsny & J. R. Turner (2005) Epithelial myosin light chain

157

kinase-dependent barrier dysfunction mediates T cell activation-induced diarrhea in

vivo. J Clin Invest, 115, 2702-15.

Dhawan, P., A. B. Singh, N. G. Deane, Y. No, S. R. Shiou, C. Schmidt, J. Neff, M. K.

Washington & R. D. Beauchamp (2005) Claudin-1 regulates cellular

transformation and metastatic behavior in colon cancer. J Clin Invest, 115, 1765-76.

Fais, S., M. R. Capobianchi, F. Pallone, P. Di Marco, M. Boirivant, F. Dianzani & A.

Torsoli (1991) Spontaneous release of interferon gamma by intestinal lamina

propria lymphocytes in Crohn's disease. Kinetics of in vitro response to interferon

gamma inducers. Gut, 32, 403-7.

Feng, L., X. Jia, M. M. Zhu, Y. Chen & F. Shi (2010) Antioxidant activities of total phenols

of Prunella vulgaris L. in vitro and in tumor-bearing mice. Molecules, 15, 9145-56.

Forsythe, R. M., D. Z. Xu, Q. Lu & E. A. Deitch (2002) Lipopolysaccharide-induced

enterocyte-derived nitric oxide induces intestinal monolayer permeability in an

autocrine fashion. Shock, 17, 180-4.

Fujimoto, K. (1995) Freeze-fracture replica electron microscopy combined with SDS

digestion for cytochemical labeling of integral membrane proteins. Application to

the immunogold labeling of intercellular junctional complexes. J Cell Sci, 108 ( Pt

11), 3443-9.

Furuse, M., K. Fujita, T. Hiiragi, K. Fujimoto & S. Tsukita (1998) Claudin-1 and -2: novel

integral membrane proteins localizing at tight junctions with no sequence similarity

to occludin. J Cell Biol, 141, 1539-50.

Furuse, M., T. Hirase, M. Itoh, A. Nagafuchi, S. Yonemura & S. Tsukita (1993) Occludin:

a novel integral membrane protein localizing at tight junctions. J Cell Biol, 123,

1777-88.

158

Gitter, A. H., F. Wullstein, M. Fromm & J. D. Schulzke (2001) Epithelial barrier defects in

ulcerative colitis: characterization and quantification by electrophysiological

imaging. Gastroenterology, 121, 1320-8.

González-Mariscal, L., A. Betanzos, P. Nava & B. E. Jaramillo (2003) Tight junction

proteins. Prog Biophys Mol Biol, 81, 1-44.

Heller, F., P. Florian, C. Bojarski, J. Richter, M. Christ, B. Hillenbrand, J. Mankertz, A. H.

Gitter, N. Bürgel, M. Fromm, M. Zeitz, I. Fuss, W. Strober & J. D. Schulzke (2005)

Interleukin-13 is the key effector Th2 cytokine in ulcerative colitis that affects

epithelial tight junctions, apoptosis, and cell restitution. Gastroenterology, 129,

550-64.

Hubatsch, I., E. G. Ragnarsson & P. Artursson (2007) Determination of drug permeability

and prediction of drug absorption in Caco-2 monolayers. Nat Protoc, 2, 2111-9.

Ikenouchi, J., M. Furuse, K. Furuse, H. Sasaki & S. Tsukita (2005a) Tricellulin constitutes

a novel barrier at tricellular contacts of epithelial cells. J Cell Biol, 171, 939-45.

Jiang, W. L., X. G. Chen, G. W. Qu, X. D. Yue, H. B. Zhu, J. W. Tian & F. H. Fu (2009)

Rosmarinic acid protects against experimental sepsis by inhibiting

proinflammatory factor release and ameliorating hemodynamics. Shock, 32,

608-13.

Jung, H. C., L. Eckmann, S. K. Yang, A. Panja, J. Fierer, E. Morzycka-Wroblewska & M. F.

Kagnoff (1995) A distinct array of proinflammatory cytokines is expressed in

human colon epithelial cells in response to bacterial invasion. J Clin Invest, 95,

55-65.

Kausalya, P. J., S. Amasheh, D. Günzel, H. Wurps, D. Müller, M. Fromm & W. Hunziker

(2006) Disease-associated mutations affect intracellular traffic and paracellular

Mg2+ transport function of Claudin-16. J Clin Invest, 116, 878-91.

159

Konishi, Y. & S. Kobayashi (2004a) Microbial metabolites of ingested caffeic acid are

absorbed by the monocarboxylic acid transporter (MCT) in intestinal Caco-2 cell

monolayers. J Agric Food Chem, 52, 6418-24.

Ligumsky, M., P. L. Simon, F. Karmeli & D. Rachmilewitz (1990) Role of interleukin 1 in

inflammatory bowel disease--enhanced production during active disease. Gut, 31,

686-9.

Luciani, F., A. Molinari, F. Lozupone, A. Calcabrini, L. Lugini, A. Stringaro, P. Puddu, G.

Arancia, M. Cianfriglia & S. Fais (2002) P-glycoprotein-actin association through

ERM family proteins: a role in P-glycoprotein function in human cells of lymphoid

origin. Blood, 99, 641-8.

Manzano, S. & G. Williamson (2010) Polyphenols and phenolic acids from strawberry and

apple decrease glucose uptake and transport by human intestinal Caco-2 cells. Mol

Nutr Food Res, 54, 1773-80.

Marin, M. L., A. J. Greenstein, S. A. Geller, R. E. Gordon & A. H. Aufses (1983) A freeze

fracture study of Crohn's disease of the terminal ileum: changes in epithelial tight

junction organization. Am J Gastroenterol, 78, 537-47.

McCarthy, K. M., S. A. Francis, J. M. McCormack, J. Lai, R. A. Rogers, I. B. Skare, R. D.

Lynch & E. E. Schneeberger (2000) Inducible expression of claudin-1-myc but not

occludin-VSV-G results in aberrant tight junction strand formation in MDCK cells.

J Cell Sci, 113 Pt 19, 3387-98.

Nardini, M., C. Scaccini, L. Packer & F. Virgili (2000) In vitro inhibition of the activity of

phosphorylase kinase, protein kinase C and protein kinase A by caffeic acid and a

procyanidin-rich pine bark (Pinus marittima) extract. Biochim Biophys Acta, 1474,

219-25.

160

Osakabe, N., A. Yasuda, M. Natsume, C. Sanbongi, Y. Kato, T. Osawa & T. Yoshikawa

(2002a) Rosmarinic acid, a major polyphenolic component of Perilla frutescens,

reduces lipopolysaccharide (LPS)-induced liver injury in D-galactosamine

(D-GalN)-sensitized mice. Free Radic Biol Med, 33, 798-806.

Reimund, J. M., C. Wittersheim, S. Dumont, C. D. Muller, J. S. Kenney, R. Baumann, P.

Poindron & B. Duclos (1996a) Increased production of tumour necrosis

factor-alpha interleukin-1 beta, and interleukin-6 by morphologically normal

intestinal biopsies from patients with Crohn's disease. Gut, 39, 684-9.

Rice-Evans, C. A., N. J. Miller & G. Paganga (1996) Structure-antioxidant activity

relationships of flavonoids and phenolic acids. Free Radic Biol Med, 20, 933-56.

Saitou, M., K. Fujimoto, Y. Doi, M. Itoh, T. Fujimoto, M. Furuse, H. Takano, T. Noda & S.

Tsukita (1998) Occludin-deficient embryonic stem cells can differentiate into

polarized epithelial cells bearing tight junctions. J Cell Biol, 141, 397-408.

Satsangi, J., M. Parkes, E. Louis, L. Hashimoto, N. Kato, K. Welsh, J. D. Terwilliger, G. M.

Lathrop, J. I. Bell & D. P. Jewell (1996) Two stage genome-wide search in

inflammatory bowel disease provides evidence for susceptibility loci on

chromosomes 3, 7 and 12. Nat Genet, 14, 199-202.

Scalbert, A., C. Morand, C. Manach & C. Rémésy (2002) Absorption and metabolism of

polyphenols in the gut and impact on health. Biomed Pharmacother, 56, 276-82.

Schuerer-Maly, C. C., L. Eckmann, M. F. Kagnoff, M. T. Falco & F. E. Maly (1994)

Colonic epithelial cell lines as a source of interleukin-8: stimulation by

inflammatory cytokines and bacterial lipopolysaccharide. Immunology, 81, 85-91.

Shin, K. M., I. T. Kim, Y. M. Park, J. Ha, J. W. Choi, H. J. Park, Y. S. Lee & K. T. Lee

(2004) Anti-inflammatory effect of caffeic acid methyl ester and its mode of action

161

through the inhibition of prostaglandin E2, nitric oxide and tumor necrosis

factor-alpha production. Biochem Pharmacol, 68, 2327-36.

Suzuki, T., B. C. Elias, A. Seth, L. Shen, J. R. Turner, F. Giorgianni, D. Desiderio, R.

Guntaka & R. Rao (2009) PKC eta regulates occludin phosphorylation and

epithelial tight junction integrity. Proc Natl Acad Sci U S A, 106, 61-6.

Suzuki, T. & H. Hara (2011) Role of flavonoids in intestinal tight junction regulation. J

Nutr Biochem, 22, 401-8.

Suzuki, T., S. Tanabe & H. Hara (2011) Kaempferol enhances intestinal barrier function

through the cytoskeletal association and expression of tight junction proteins in

Caco-2 cells. J Nutr, 141, 87-94.

Swarup, V., J. Ghosh, S. Ghosh, A. Saxena & A. Basu (2007) Antiviral and

anti-inflammatory effects of rosmarinic acid in an experimental murine model of

Japanese encephalitis. Antimicrob Agents Chemother, 51, 3367-70.

Tanaka, T., A. Nishikawa, H. Shima, S. Sugie, T. Shinoda, N. Yoshimi, H. Iwata & H.

Mori (1990) Inhibitory effects of chlorogenic acid, reserpine, polyprenoic acid

(E-5166), or coffee on hepatocarcinogenesis in rats and hamsters. Basic Life Sci, 52,

429-40.

Van De Walle, J., A. Hendrickx, B. Romier, Y. Larondelle & Y. J. Schneider (2010)

Inflammatory parameters in Caco-2 cells: effect of stimuli nature, concentration,

combination and cell differentiation. Toxicol In Vitro, 24, 1441-9.

Van Itallie, C., C. Rahner & J. M. Anderson (2001a) Regulated expression of claudin-4

decreases paracellular conductance through a selective decrease in sodium

permeability. J Clin Invest, 107, 1319-27.

Vitkus, S. J., S. A. Hanifin & D. W. McGee (1998) Factors affecting Caco-2 intestinal

epithelial cell interleukin-6 secretion. In Vitro Cell Dev Biol Anim, 34, 660-4.

162

Wang, F., W. V. Graham, Y. Wang, E. D. Witkowski, B. T. Schwarz & J. R. Turner (2005)

Interferon-gamma and tumor necrosis factor-alpha synergize to induce intestinal

epithelial barrier dysfunction by up-regulating myosin light chain kinase expression.

Am J Pathol, 166, 409-19.

Watanabe, T., Onuki, R., Yamashita, S., Taira, K., & Sugiyama, Y (2005) Construction of a

functional transporter analysis system using MDR1 knockdown Caco-2 cells.

Pharm Res, 22, 1287-93.

Weber, C. R., D. R. Raleigh, L. Su, L. Shen, E. A. Sullivan, Y. Wang & J. R. Turner (2010)

Epithelial myosin light chain kinase activation induces mucosal interleukin-13

expression to alter tight junction ion selectivity. J Biol Chem, 285, 12037-46.

Wright, K. L., S. A. Weaver, K. Patel, K. Coopman, M. Feeney, G. Kolios, D. A. Robertson

& S. G. Ward (2004) Differential regulation of prostaglandin E biosynthesis by

interferon-gamma in colonic epithelial cells. Br J Pharmacol, 141, 1091-7.

Yao, Y. & S. E. Tsirka (2011) Truncation of monocyte chemoattractant protein 1 by

plasmin promotes blood-brain barrier disruption. J Cell Sci, 124, 1486-95.

Ye, Z., Z. Liu, A. Henderson, K. Lee, J. Hostetter, M. Wannemuehler & S. Hendrich (2009)

Increased CYP4B1 mRNA is associated with the inhibition of dextran sulfate

sodium-induced colitis by caffeic acid in mice. Exp Biol Med (Maywood), 234,

605-16.

Zdarilová, A., A. Svobodová, V. Simánek & J. Ulrichová (2009) Prunella vulgaris extract

and rosmarinic acid suppress lipopolysaccharide-induced alteration in human

gingival fibroblasts. Toxicol In Vitro, 23, 386-92.

Zolotarevsky, Y., G. Hecht, A. Koutsouris, D. E. Gonzalez, C. Quan, J. Tom, R. J. Mrsny &

J. R. Turner (2002) A membrane-permeant peptide that inhibits MLC kinase

163

restores barrier function in in vitro models of intestinal disease. Gastroenterology,

123, 163-72.

164

CHAPTER 6. GENERAL CONCLUSIONS

General Discussion

A major limiting step in the utilization of herbal constituents is their intestinal absorption and metabolism. Compound passes across the intestinal epithelium via passive diffusion, both paracellularly and transcellularly, and by active transport. After entering the cells, herbal constituents may undergo Phase I and II biotransformation by metabolizing enzymes; the parent compounds and the metabolites are effluxed to the apical side (luminal) or transported to the basolateral side (systemic) by various transporters located at both sides of the intestinal epithelial cells. In addition, herbal constituents might also have influences on the metabolizing enzyme activities or the expression and function of transporters, which in turn affects the bioavailability of other herbal constituents or dietary nutrients, indicating possible herb-herb and herb-drug interactions. Furthermore, caffeic acid derivatives from diet or tranditional herbs have been reported to have anti-inflammatory properties both in vitro and in vivo. Chapter 3, 4 and 5 are original projects which were performed on permeability and metabolism of herbal constituents and modulation of caffeic acid derivatives on intestinal function in Caco-2 cell line.

In Chapter 3, we have conducted the study of transport mechanism and metabolism of alkamides and ketone in Echinacea species and their inhibitory activites on P-glycoprotein transporter (P-gp) function. Tested alkamides transport the Caco-2 cell monolayer via passive diffusion, independent of the plant matrix, and were seemingly N-glucuronidated both as pure compounds and as found in E. saguinea and E. pallida extracts. Both

Echinacea extracts stimulated apparent glucuronidation and basolateral efflux of alkamide metabolites. Interestingly, E. pallida ethanol extract inhibited the efflux of glucuronides of Bauer alkamides apically (lumin) while inducing the transfer of the conjugates

165

basolaterally (systemic), and the phytochemical profile might contribute to the difference

between two Echincea species. It was not expected that Bauer ketone 24 was totally

metabolized to more hydrophilic metabolites as a pure compound, but not as found in either

Echinacea species, indicating the influences of plant matrix on ketone metabolism in

Caco-2 cells. We found that Bauer alkamide 8, 10 and 11 at 175-230 µM, E. sanguinea at

1 mg/mL (containing 85 µM of alkamide 8, 2 µM of alkamide 10, and 0.7 µM of alkamide

11) and E. pallida at 5 mg/mL (containing 215 µM of alkamide 8, 25 µM of alkamide 10,

and 45 µM of alkamide 11) significantly inhibited P-gp function. Taken together, plant matrix had a facilitating effect on the metabolism and efflux of alkamides and ketones from

Echinacea, and that alkamides and Echinacea extracts might be useful in potentiating some

chemotherapeutics which are substrates for P-gp.

In Chapter 4, we have investigated the permeabilities and Phase II biotransformation of

rosmarinic acid and ursolic acid as pure compounds and in herbal extracts using Caco-2

cells with HPLC detection. We found increased basolateral rosmarinic acid recoveries

after deconjugation indicating that glucuronide/sulfate conjugates were excreted to the

basolateral side (toward the circulation) rather than apical side (the intestinal lumen). In

addition, the plant extracts, P. vulgaris and S. officinalis, did not affect transfer of either

ursolic acid or rosmarinic acid, and likewise the pure compounds and the compounds from

the plant extracts behaved similarly in their apparent biotransformation as well, which was

different compared with alkamide metabolism in pure solutions vs. Echinacea species.

This difference might be due to the transport mechanism (paracelluar diffusion for rosmarinic acid vs. transcellualr diffusion for alkamides) and phytochemical profile of P.

vulgaris and Echinacea species. Meanwhile, ursolic acid was absorbed and transferred by

166

Caco-2 cells apparently with little glucuronidation/sulfation, which is likely to further enhance the bioavailability of ursolic acid compared with rosmarinic acid.

In Chapter 5, we performed the regulation of caffeic acid, rosmarinic acid, chlorogenic acid and m-hydroxyphenylpropionic acid (mHPP), a microbial metabolite of ingested caffeic acid, chlorogenic acicd or rosmarinic acid, on tight junction proteins and immune modulators in both Caco-2 cell model and multidrug resistance (MDR) knockdown (MDR

KD) Caco-2 cells. Stimuli composed of IL-1β, TNF-α, IFN- γ and LPS were added to

Caco-2 cell system to represent the acute inflammatory reaction as it might occur during the active phase of IBD. We found that caffeic acid derivatives enhanced tight junction integrity through the expression of tight junction proteins in both Caco-2 cells and MDR

KD Caco-2 cells. Among the tested caffeic acid derivatives, we reported for the first time that mHPP exhibited greater enhancement of intestinal barrier than the parent compounds, indicating that mHPP deserves more studies. One of our key findings was that 50 µM of caffeic acid derivatives were found to be enhancers of IL-6 and rosmarinic acid and mHPP augmented TNF-α secretion in intestinal epithelial cells after stimulation. More interestingly, when P-gp transporter, encoded by MDR gene, was knocked down in Caco-2 cells, all cafffeic acid derivatives attenuated the inflammatory responses initiated by the stimuli in terms of the reduction of TNF-α and IL-6 levels, implying that knockdown of the

MDR gene might affect transcriptional cascades that in turn affect inflammatory mediators.

As a conclusion, alkamides, ketone, RA and UA in herbal extracts was transported across

Caco-2 cell monolayer with different permeability rates. The metabolism was a complex process, especially in different Echinacea species. Caffeic acid derivatives could be gut health promoting as dietary constituents, and mHPP was deserved the promising research

167

future in both anti-inflammatory and anti-oxidative areas, particularly in molecular

mechanism related to revealing anti-colitic pathway.

168

Recommendations for Future Research

In Caco-2 transport study, glucuronide conjugates of herbal alkamides, ketones, and caffeic

acid derivatives were found, and plant matrix had influences on the metabolism and efflux

of the metabolites of alkamides based on the current preliminary result. As a next step,

we need to further investigate the inducibility of glucuronidation/sulfation, by herbal

components as well as the expression of Phase I and II enzymes in Caco-2 cells after

treating different pure compounds or herbal extracts. The identification of the herbal

constituent or combination of various components that have major effects on metabolizing

enzymes and transporter functions are also required.

With respect to the enhancement of caffeic acid derivatives on intestinal epithelial function,

more mechanistic studies need to be carried out in the future. More specifically, myosin

light chain kinase (MLCK), protein kinase C (PKC), ezrin, radixin and moesin (ERM) and

NF-κB pathways might be targeted to reveal the underlying mechanism of the protection of caffeic acid derivatives on intestinal barrier function.

169

ACKNOWLEDGEMENTS

“Once a teacher, always a teacher”; I would like to express my sincere appreciation to my

major professor, Dr. Suzanne Hendrich, for her continuous guidance and encouragement,

tireless support and dedication, as well as unfailing friendship and patience throughout my

dissertation study. I would also like to thank the members of my program of study committee, Dr. Patricia Murphy, Dr. Diane Birt, Dr. Manju Reddy, and Dr. Michael

Wannemuehler for their professional opinions, great advices, and generous help through

my graduate studies.

I want to thank my colleagues from our lab: Sun-Ok Lee, Xianai Wu, Zhong Ye, Li Li, Kai

Ling Kong, Visalakshi Arumugam, Martin Mutambuka, Esther Haugabrooks and

Mi-Kyung Lee; from Dr. Murphy’s Lab: Cathy Hauck; from Dr. Reddy’s Lab: Yin Zhou;

from Dr. Birt's lab: Nan Huang and Xiaozhu Zhang; from Dr. Wannemuehler’s Lab:Kelley

M Haarberg; from Botainical Center at Iowa State University: Janice E Meyer; Ludmila

Rizshsky; from Department of Food Science and Human Nutrition: Jeanne M Stewart, Jean

G Tilley, Erica Burkheimer and Brenda J Emery; and from Interdepartmental Toxicology:

Linda M Wild. Their constant support, technical assistance, valuable suggestions and hand-on help were presious and unforgettable.

Last, but not least, I would like to thank my family. To my Mom and Dad, thank you for instilling in me through your example that thoughtfulness, hard work, and determination are the tools to make dreams come true. To my wife Bing, thank you for your love, understanding and patience, we share weal and woe. And to my son Shannor ZeYuan

Qiang, welcome to this world and enjoy your life!