116
Nanoencapsulation of Tea Catechins in Casein Micelles: Effects on Processing and Biological Functionalities by Sanaz Haratifar A Thesis presented to The University of Guelph In partial fulfilment of requirements for the degree of Doctor of Philosophy in Food Science Guelph, Ontario, Canada © Sanaz Haratifar, November, 2012

Nanoencapsulation of Tea Catechins in Casein Micelles

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Nanoencapsulation of Tea Catechins in Casein Micelles

Nanoencapsulation of Tea Catechins in Casein Micelles: Effects on Processing

and Biological Functionalities

by

Sanaz Haratifar

A Thesis

presented to

The University of Guelph

In partial fulfilment of requirements

for the degree of

Doctor of Philosophy

in

Food Science

Guelph, Ontario, Canada

© Sanaz Haratifar, November, 2012

Page 2: Nanoencapsulation of Tea Catechins in Casein Micelles

ABSTRACT

Nanoencapsulation Of Tea Catechins In Casein Micelles: Effects On Processing and

Biological Functionalities

Sanaz Haratifar Advisor:

University of Guelph, 2012 Professor M. Corredig

This thesis focuses on the interactions between milk proteins (caseins) and tea catechins and the

consequences of the interactions on the renneting properties and digestion of casein micelles as

well as on the biological functionality of the mixture, as measured using intestinal cell models.

The binding of epigallocatechin-gallate (EGCG) to casein micelles was quantified using HPLC

and fluorescence quenching, and it was shown that a substantial amount of EGCG can be

incorporated in the casein micelles. At concentrations < 2.5 mg/ml milk, most of the EGCG

added to milk was associated with the caseins. The formation of EGCG-casein micelles

complexes not only delayed the gelling point of milk, but they also affected the structure

formation of the gels.

EGCG is known to have antiproliferative activity on colon cancer cells.

Although nanoencapsulation of EGCG in casein micelles may have some advantages, such as

protecting this bioactive compound from degradation, it may also affect the bioavailability of

EGCG. To test this hypothesis, the effect of nanoencapsulation of EGCG in casein micelles on

the biological functionality of EGCG was tested by evaluating the cytotoxity and proliferation

behaviour of HT-29 colon cancer cells. It was demonstrated that nanoencapsulation did not

affect the bioefficacy of EGCG.

Similar experiments were also carried out on rat colonic cells, a normal line and its cancerous

tranformed line. For this study, nanoencapulated EGCG was subjected to in vitro digestion

Page 3: Nanoencapsulation of Tea Catechins in Casein Micelles

before absorption. The results showed that EGCG-casein binding did not affect the digestion of

the milk proteins. In this case, the bioefficacy of EGCG was not diminished as well.

In addition, studies on normal cell lines demonstrated the specific effect of EGCG on cancer

cells, favoring normal cell survival whether EGCG was isolated or complexed with milk.

These experiments bring further evidence that milk can be employed as an appropriate platform

for the delivery of bioactive compounds.

Page 4: Nanoencapsulation of Tea Catechins in Casein Micelles

iv

Acknowledgement

I would like to thank my advisor, Dr. Milena Corredig, for the opportunity to work on this

project. I appreciate all of the advice, time, and effort that Milena provided during my journey as

a PhD student . Without her direction and assistance this thesis would not have been possible.

I am also grateful to my advisory commitee: Dr. Gopinadhan Paliyath, Dr. Kelly Meckling and

Dr. Lisa Duizer for their advice during my study. Their comments and guidance, were always

comprehensive and constructive.

I am very grateful to Edita Verespej for all her technical support and would like to also thank

all my friends and lab mates, who I have had the opportunity to work with in the dairy lab and at

Canadian Research Institute for Food Safety (CRIFS).

I would like to show my gratitude to the Ontario Dairy Council and the Natural Sciences and

Engineering Council of Canada, through the Industrial Research Chair program, and the Canada

Research Chair Program for financial support of my study.

To my beloved family, especially my father, thank you for your support and encouragements

during all the ups and downs in my life.

Finally, to my sister, Farnaz, I am indebted in every possible way for all the love and

confidence you have in me. I can not imagine my life without you.

This thesis is dedicated to my mother and grandmother, my two angels in heaven.

Page 5: Nanoencapsulation of Tea Catechins in Casein Micelles

v

Table of contents

CHAPTER 1 ............................................................................................................................................. 1

1.1. General Introduction ..........................................................................................................1

1.2. Research Hypotheses ..........................................................................................................4

CHAPTER 2 ............................................................................................................................................. 6

2.1. Polyphenols .........................................................................................................................6

2.1.1. Tea.................................................................................................................................8

2.1.2. Tea-bioavailability ....................................................................................................... 12

2.1.3. Nano encapsulation of Tea catechins ........................................................................... 15

2.2. Milk ................................................................................................................................... 17

2.2.1. Caseins ........................................................................................................................ 18

2.2.2. Whey Proteins.............................................................................................................. 21

2.2.3. Rennet induced gelation of milk ................................................................................... 21

2.3. Interactions between milk proteins and tea catechins ....................................................... 22

CHAPTER 3 ........................................................................................................................................... 28

INTERACTIONS BETWEEN TEA CATECHINS AND CASEIN MICELLES AND

THEIR IMPACT ON RENNETING FUNCTIONALITY .......................................................... 28

3.1 Abstract ............................................................................................................................. 28

3.2 Introduction ....................................................................................................................... 29

3.3 Materials and Method .................................................................................................. 31

3.3.1 Milk preparation and interaction studies ........................................................................ 31

3.3.2 Fluorescence spectroscopy ............................................................................................ 32

3.3.3 Quantification of polyphenol binding to casein micelles ................................................ 34

3.3.4 Rheological properties of rennet-induced gels ............................................................... 35

3.3.5 Release of casein macro- peptide (CMP) from the surface of the casein micelles .......... 35

3.3.6 Statistical analysis ......................................................................................................... 35

3.4 Results and Discussion ...................................................................................................... 36

3.4.1 Fluorescence spectroscopy studies ................................................................................ 36

3.4.2 Analysis of tea polyphenols by HPLC ........................................................................... 43

3.4.3 The effect of EGCG on rennet induced gelation of casein micelles ................................ 47

Page 6: Nanoencapsulation of Tea Catechins in Casein Micelles

vi

3.5 Conclusions ........................................................................................................................ 55

CHAPTER 4 ........................................................................................................................................... 56

Anti-proliferative activity of nanoencapsulated tea catechin in casein micelles, using HT29

colon cancer cells ..................................................................................................................... 56

4.1. Abstract ............................................................................................................................ 56

4.2. Introduction ...................................................................................................................... 57

4.3. Materials and Methods .................................................................................................... 60

4.3.1 Milk samples preparation .............................................................................................. 60

4.3.2. Cell viability assay ....................................................................................................... 61

4.3.3. Cell proliferation assay ................................................................................................ 61

4.3.4. Statistical analysis ........................................................................................................ 63

4.4. Results and Discussion ..................................................................................................... 63

4.5. Conclusions ....................................................................................................................... 71

CHAPTER 5 ........................................................................................................................................... 72

THE EFFECT OF NANOENCAPSULATION OF TEA CATECHINS IN CASEIN

MICELLES ON THEIR BIOACCESSIBILITY STUDIED USING NORMAL AND

CANCEROUS RAT COLONIC CELLS ........................................................................................ 72

5.1 Abstract ............................................................................................................................. 72

5.2 Introduction ....................................................................................................................... 73

5.3 Materials and Methods ..................................................................................................... 76

5.3.1 Sample preparation ....................................................................................................... 76

5.3.2 In vitro gastro-duodenal digestion ................................................................................ 78

5.3.3 SDS-PAGE analysis...................................................................................................... 79

5.3.4 Cell proliferation assays ................................................................................................ 80

5.3.5 Statistical analysis ......................................................................................................... 81

5.4 Results and Discussion ...................................................................................................... 82

5.4.1 SDS-PAGE of milk and tea/milk complexes during the gastric phase ............................ 82

5.4.2 Proliferation of normal (4D) and tumorous (v-src) rat colonic cells exposed to EGCG

and EGCG -milk fractions before and after invitro digestion .................................................. 85

Page 7: Nanoencapsulation of Tea Catechins in Casein Micelles

vii

5.5 Conclusion ......................................................................................................................... 92

CHAPTER 6 ........................................................................................................................................... 93

GENERAL CONCLUSIONS ............................................................................................................. 93

REFERENCES ...................................................................................................................................... 96

Page 8: Nanoencapsulation of Tea Catechins in Casein Micelles

viii

Table of Figures

Figure 2.1 Chemical structure of flavonoids. ...............................................................................7

Figure 2.2 Chemical structure of flavan-3-ols. .............................................................................7

Figure 2.3. Structures of four catechins in tea infusions. ..............................................................9

Figure 3.2 Stern-Volmer plot for skim milk ( ) and caseins (♦) re-diluted in permeate. ............ 41

Figure 3.3 Ratio of tea catechin recovered in the colloidal phase (casein micelles) of skim milk,

as a function of the initial concentration of EGCG present in milk. . ......................................... 44

Figure 3.4 Scatchard plot for EGCG-milk caseins .................................................................... 46

Figure 3.5 Rennet coagulation process monitored by rheometer of the four samples………………….48

Figure 3.6 % CMP-release versus time after addition of rennet. ................................................ 52

Figure 3.7 % CMP-release versus time after addition of rennet. ................................................. 54

Figure 4.2 Amount of EGCG associated with the casein micelles in skim milk as a function of

the initial concentration of EGCG in milk……………………………………………………….65

Figure 4.3 % Viable cells after 24h exposure to isolated EGCG (solid) and EGCG encapsulated

in casein micelles. ................................................................................................................... 67

Figure 4.4 Proliferation of HT-29 cells exposed to isolated EGCG or EGCG in milk serum

(permeate) (A); milk whey (B) or encapsulated with the casein micelles in skim milk (C) ....... 70

Figure 5.2 SDS electrophoresis of milk digests . ....................................................................... 83

Figure 5.3 % Change of Tumor cell proliferation with increasing levels of isolated EGCG

(solid bars) and complexed EGCG-milk fractions before in vitro digestion ............................... 87

Figure 5.4 % Change of Tumor cell proliferation with increasing levels of EGCG in digests of

isolated EGCG (solid bars) and complexed EGCG-milk fractions. ............................................ 91

Page 9: Nanoencapsulation of Tea Catechins in Casein Micelles

ix

LIST OF TABLES

Table 3.1. Gelling time and gel stiffness of different tea-milk samples.......................................50

Page 10: Nanoencapsulation of Tea Catechins in Casein Micelles

1

CHAPTER 1

1.1. General Introduction

Interest for health promoting, natural food compounds has expanded in recent years due to the

increase in consumer awareness and the promotion of healthy eating and lifestyle.

However, challenges remain to ensure that such compounds not only remain in their active form

during processing and storage, but that their bioavailability is also maximized (Seymour et al.,

2008). Indeed, to be able to deliver the nutritional functionality, food products need to have not

only a nutritionally significant dose of bioactive molecules, but these molecules must also

remain fully functional for as long as necessary (Garti, 2008). Polyphenols are amongst the most

bioactive plant metabolites used in food and nutraceutical applications. Tea catechins, the focus

of this thesis, have been shown in numerous studies to have important preventive and therapeutic

roles for the most common human diseases such as cancer, cardiovascular and neurodegenerative

diseases (Manach et al., 2004; Sharma et al., 2009; Yang, 2009; Yang, 2012; Yuan et al., 2011).

However, higher amounts must be consumed to benefit from their bioactivity, as tea catechins

have been reported to have relatively poor absorption efficiency in vivo (Stevenson et al., 2007;

Papadopoulou et al., 2004; Green et al., 2007; Ferruzzi et al., 2010). The relations between

bioefficacy, concentration in food products, bioavailability and cell absorption are not straight

forward, and are the most challenging aspect in the development of food products with health

enhancing properties.

Market trends indicate that milk is an ideal vehicle for the delivery of bioactive compounds

targeting lifestyle diseases, as milk is a natural, multi-component, and nutrient-rich beverage that

is widely consumed around the world (Sharma et al., 2009). Many countries often consume tea

combined with milk (traditional consumption of black tea) and the number of new tea based

Page 11: Nanoencapsulation of Tea Catechins in Casein Micelles

2

products containing dairy ingredients, such as ready-to-drink tea chai latte, has expanded in

recent years (Ferruzzi et al., 2005).

Dairy proteins are often used as ingredients in foods for their functionalities, for example,

gelling, foaming, emulsifying. However, in the context of delivery of bioactive compounds,

many structural and physicochemical properties of milk proteins point to their delivery functions

in nature. Milk caseins for example, have been designed by nature to deliver nutrients such as

calcium, phosphate and protein to the neonate (Livney, 2010). Studies have shown that casein

micelles bind to hydrophobic molecules such as curcumin and Vitamin D, which these

compounds penetrate into the core of the micelles. The encapsulation of Vitamin D provides

significant protection against UV light and in the case of curcumin the encapsulation does not

impair the cytotoxic effects of curcumin on HeLa cells compared to an equal dose of free

curcumin (Semo et al, 2007; Sahu et al., 2008; Livney et al., 2009).

Strong interactions between tea catechins and casein proteins have been reported in many studies

however, most studies have been so far conducted on isolated caseins and not with casein

micelles or skim milk (Arts et al., 2002; Jobstyl et al., 2004; Papadopoulou et al., 2005; Pascal et

al., 2007; Yan et al., 2009; Frazier et al., 2009; Yuksel et al., 2010). Interpretation of studies

using caseins in milk media has been difficult since, by the addition of milk, the product matrix

complexity increases, and obtaining accurate and reliable measurements for tea catechins

becomes more challenging. The effect of milk on the bioefficacy of polyphenols in tea is still a

matter of debate as some studies indicate that addition of milk has a negative impact on tea

catechins, for example, on the antioxidant properties, while some others indicate no such effect

(Langley-Evans, 2000; Arts et al., 2002; Kartosova et al., 2007; Sharma et al., 2009; Cilla et al.,

2009; Leenon et al., 2000; Keogh et al., 2007; Kyle, 2007; Green et al., 2007; Dubeau et al.,

Page 12: Nanoencapsulation of Tea Catechins in Casein Micelles

3

2011; van der Burg-Koorevaar et al., 2011). It is important to state that there are differences in

the study designs and methodologies of the above mentioned studies which may contribute to the

controversy of the subject. Most studies have overlooked the strong interactions that take place

between the proteins and polyphenols, considering effective (i.e., bioavailable) only the material

recovered in the soluble phase.

This thesis focuses on the interactions between milk proteins (especially caseins) and tea

catechins, and their effect on processing functionality and bioavailability, as measured using

intestinal cell models. Although few studies have been reported on the effect of casein-tea

catechin interactions on the bioavailability of the catechin, rather limited experimental data is

available on the functional properties of caseins. Such studies are also needed to elucidate the

effects of polyphenols on protein functional properties that may influence the fate of polyphenol

application in food processing. Indeed, it is important to determine whether the functional

benefits of both compounds (tea and milk proteins) may work synergistically in a mixture, or

change the effect of the functionality of the components taken in isolation. Our work aims to

explore some of the effects of the addition of polyphenols to milk proteins and their relevance to

the design of improved dairy products.

In the first result chapter the interactions between tea catechins and casein micelles were

quantified using HPLC as well as fluorescence spectroscopy. Most binding studies carried out

so far have used isolated proteins, and this work advances our knowledge on the interactions

between EGCG and casein micelles in their native form. The use of two separate methods will

allow for stronger data interpretation. In this chapter, effect of tea catechin-milk casein

interactions on the rennet induced gelling of milk was evaluated.

Page 13: Nanoencapsulation of Tea Catechins in Casein Micelles

4

The second part of the work (Chapters 3 and 4) focused on the effect of the milk casein and tea

catechin complexes on the biological functionality of tea polyphenols using cell culture models.

It was hypothesized that caseins, by binding to EGCG, will decrease the bioavailability of this

component. In Chapter 3 we report on the effect of the tea catechin-milk protein complexes on

the biological functionality of the tea catechin, by applying cell cytotoxity and proliferation

assays on HT-29 colon cancer cells. Different milk fractions were also used to better undersand

the role played by the various milk components in modulating bioactivivity. In Chapter 4, the

effect of the mixture of milk and EGCG on two parallel sets of cell lines was studied, to

determine the cytotoxicity and proliferation in normal versus cancer cell lines. Afterwards, an in

vitro digestion model mimicking the gastric stage and duodenal stage of human digestion was

employed to determine the significance of this association on the digestibility of the milk

proteins and on the bioaccessibility of the tea polyphenol epigalocatechin gallate (EGCG).

This thesis will be able to further our understanding of the interactions between EGCG and

casein micelles, and the impact of such interactions on the processing and biological

functionality of milk. Demonstration of the effective incorporation of bioactives into the milk

matrix is essential before using milk as a platform for delivery of polyphenols.

Hence, this research tested the following hypotheses:

1.2. Research Hypotheses

The main hypothesis of this study was that the interactions between EGCG and casein micelles

improve the bio-accessibility of EGCG without altering the functional properties of caseins as

well as their digestive breakdown.

Therefore, the main objectives studied in this research were:

Page 14: Nanoencapsulation of Tea Catechins in Casein Micelles

5

1. Evaluate the interactions between EGCG and casein micelles, in their native form: milk.

2. Determine the effect of the EGCG-casein complexes on the gelation properties of caseins.

3. Determine if the complexes formed affect the biological functionality of tea polyphenols using

cell culture models.

4. Determine if the interactions between EGCG and milk affect the digestion of milk caseins and

if the complexes formed modify the bioactivity of tea polyphenols once digested.

Page 15: Nanoencapsulation of Tea Catechins in Casein Micelles

6

CHAPTER 2

2.1. Polyphenols

Interest in polyphenols as health-promoting nutrients has expanded dramatically in recent years.

The main reasons for this interest are the recognition of their great importance in our diet, and

their likely role in the prevention of various diseases. Polyphenols are found in a variety of foods

such as: vegetables, fruits, wine, coffee and tea (Manach et al., 2004). Polyphenols are secondary

plants metabolites and are naturally involved in the defense mechanisms of the plant, for

example, against ultraviolet radiation or aggression by pathogens. Polyphenols may be classified

into different groups as a function of the number of phenol rings that they contain and of the

structural elements that bind these rings to one another. The main groups include phenolic acids,

flavonoids, stilbenes, and lignans. Polyphenols found in the daily diet mainly consist of

flavonoids, of which approximately 8000 molecular compounds have been reported. Flavonoids

share a common structure consisting of 2 aromatic rings that are bound together by 3 carbon

atoms that form an oxygenated heterocycle (Figure 2.1). They are usually divided into six main

subclasses; flavonols, flavones, isoflavones, flavanones, anthocyanidins, and flavan-3-ols

(catechins and proanthocyanidins) (Manach et al., 2004, Hakimmdun, 2006). The most common

group of flavonoids are flavan-3-ols (Figure 2.2) which in the human diet are considered

functional ingredients of beverages, fruits and vegetables, food grains, herbal remedies, dietary

supplements, and dairy products. Flavan-3-ols have been reported to exhibit several health

beneficial effects by acting as antioxidant, anticarcinogen, cardiopreventive, antimicrobial,

antiviral, and neuro-protective agents (Patel et al., 2005; Scalbert et al., 2005; Aron et al., 2008).

Page 16: Nanoencapsulation of Tea Catechins in Casein Micelles

7

Figure 2.1 Chemical structure of flavonoids (Manach et al., 2004).

Figure 2.2 Chemical structure of flavan-3-ols (Aron et al., 2008).

Page 17: Nanoencapsulation of Tea Catechins in Casein Micelles

8

2.1.1. Tea

One major dietary source of polyphenols is tea. Tea is one of the most concentrated sources of

polyphenols. These substances comprise up to 30% of its dry weight (Balentine et al., 1997;

Dubeua et al., 2010). Tea is one of the most widely popular beverages, consumed by over two-

thirds of the world's population due to its medicinal, refreshing and mild stimulant effects.

Drinking a cup of tea for pleasure or in times of stress is a part of daily life for millions of people

all over the world. The discovery and use of tea as a beverage has a long history. It may have

originated during the ―Shen-Nong‖ era of ancient China, around 5000 to 6000 years ago (Chen,

2002). Tea plant is a perennial evergreen plant with three varieties, Camellia sinensis, Camellia

assamica and Camellia cambodiensis (Karak et al., 2010; Voung et al., 2010).

There are many types of tea but three types of teas, green, black, and oolong are produced and

consumed abundantly in different regions of the world. Each of these types are obtained from the

same plant (C. sinensis) but are different because of their processing history and extent of

fermentation of the final product. Therefore, based on the tea manufacturing (fermentation)

process, teas from the genus Camellia can be divided into three categories: green tea

(unfermented), oolong tea (partially fermented), and black tea (fully fermented). Of the total

world tea production, 78% is black tea, 20% is green tea, and 2% is oolong tea (Sharma et al.,

2009; Wan et al., 2009).

Green tea has been regarded as a rich source of flavan 3-ols also known as catechins (Figure 2.3)

including (+)-catechin (C), (-)-epicatechin (EC), (-) -epigallocatechin (EGC), (-)-epicatechin

gallate (ECG), and (-)-epigallocatechin gallate (EGCG). The tea catechins present in Camellia

sinensis leaf are transformed to polymeric theaflavin and thearubigin by oxidation occurring

during fermentation of black tea.

Page 18: Nanoencapsulation of Tea Catechins in Casein Micelles

9

Figure 2.3 Structures of four catechins in tea infusions. Epicatechin: R1=H, R2=H;

epigallocatechin: R1=OH, R2=H; epicatechin gallate: R1=H, R2=gallate (3,4,5-

trihydroxybenzoyl); epigallocatechin gallate: R1=OH, R2 = gallate (3,4,5-trihydroxybenzoyl).

(Ronner et al., 1998).

Page 19: Nanoencapsulation of Tea Catechins in Casein Micelles

10

The distinctive reddish-black colour, reduced bitterness and astringency, and characteristic

flavour give black tea a marked distinction from green tea (Ferrara et al., 2001; Kim et al., 2011;

Sang et al., 2011). The catechins are colorless, astringent, water soluble compounds. They are

readily oxidizable, although their oxidation potentials vary. This property has been exploited

through their use as food antioxidants (Henning et al., 2003). More than 50% of the mass of the

catechins in green tea is composed of EGCG and it has been often suggested that EGCG is

responsible for the majority of the potential health benefits attributed to green tea consumption.

The potential health benefits attributed to green tea and EGCG include antioxidant effects,

cancer chemoprevention, improving cardiovascular health, enhancing weight loss, protecting the

skin from the damage caused by ionizing radiation, and others (Zaveri et al., 2006; Sharma et al.,

2009; Yang et al., 2009; Gonzalez et al., 2011; Citarasu et al., 2010).

Because of green tea’s alleged beneficial effects, green tea–based products have become the

fourth most commonly used dietary supplement in the United States. Exposure to green tea can

take place in several forms: consumed as dilute beverages or concentrated supplements or

applied topically (Chan et al., 2010). A recent search of the literature revealed more than 8000

citations that relate to the chemistry, bioactivity, production, and potential health benefits of

green tea (Nagle et al., 2006).

The possible preventive activities of green tea against cancer and cardiovascular diseases have

been studied extensively during the past three decades. Among the biological activities of tea, the

cancer-preventive activity has attracted the greatest attention. In particular, epigallocatechin-3-

gallate (EGCG), the main ingredient of green tea extract, has been shown to inhibit growth in a

number of tumor cell lines at different organ sites including the skin, oral cavity, esophagus,

stomach, intestine, lung, liver, pancreas, mammary gland, urinary bladder, and prostate.

Page 20: Nanoencapsulation of Tea Catechins in Casein Micelles

11

The cancer-preventive processes of EGCG may include decreasing cell proliferation, increasing

apoptosis and suppression of angiogenesis. A large number of epidemiological studies have

examined the association between tea consumption and risk of cancers at various organ sites.

Some human cancer prevention trials with green tea polyphenol preparations have shown

promising results, more specifically causing reduced risk of upper gastrointestinal tract

cancers with diets characterized by a high intake of green tea. These studies have led to the

conclusion that EGCG is the most active compound to impart the cancer preventive effect of

green tea (Li et al., 2001; Chow et al., 2005; Lambert et al., 2007a; Yang 2009; Shpigelman et

al., 2010;Yuan et al., 2010; Yang et al., 2012).

In recent years the possible cancer preventive activity of EGCG has been extensively studied

using animal models and different cell line cultures from various organs. The inhibitory activity

of EGCG on cancer cells has been demonstrated during both the initiation and progression stages

of carcinogenesis (Vaidyanathan et al., 2003; Chen et al.,2004; Chen et al., 2007; Yang, 2009).

In vitro cell culture models are valuable tools for screening of chemo preventive agents and

provide preliminary data for in vivo studies. However, the specific cancer prevention

mechanisms related to EGCG, and in general to the molecules present in green tea, are still not

known. Since broad cancer-preventive effects have been observed in a variety of cell lines and

animal models, it is assumed that multiple molecular mechanisms, rather than a single receptor

or molecular target, may be involved. It has been seen that some of cancer preventive effects

may be by direct involvement of EGCG, such as binding to molecular targets while other effects

may be indirect, such as the inhibition of enzyme activities in certain cell pathways (Yang et al.,

2006; Yang et al., 2009; Shpigelman et al., 2010; Visioli et al., 2011).

Page 21: Nanoencapsulation of Tea Catechins in Casein Micelles

12

2.1.2. Tea-bioavailability

To exert their biological properties, polyphenols must be bio-accessible and ultimately available

to some extent in the target tissue. Bioaccessibility is defined as the amount of ingested

compounds that are available for absorption in the gastrointestinal tract. Therefore, the

availability and luminal changes prior to absorption play a major role in ensuring bioavailability.

The amount of bioaccessible compounds can be equal to or less than the amount that is released

from the food matrix (Stahl et al., 2002; Aura, 2005; Saura-Calixto et al., 2007). Bioavailability

appears to differ greatly among the various polyphenols, and the most abundant ones in our diet

are not necessarily those that have the best bioavailability profile. Consequently, it is not only

important to know how much of a nutrient is present in a specific food or dietary supplement, but

also how much of it is bioavailable.

The bioavailability of dietary polyphenols is limited by a number of factors such as amount

consumed, physico-chemical properties of the molecule, enzyme and microbial-mediated

biotransformation and active reflux, instability under digestive conditions, intestinal absorption

and food matrix interactions (Stevenson et al., 2007; Papadopoulou et al., 2004; Green et al.,

2007; Ferruzzi et al., 2010).

Despite the epidemiological evidence that consumption of various teas is associated with positive

health benefits, the fate of tea catechins in the digestive tract is yet to be clearly identified. Some

studies have shown that, incubation of tea catechins at acid pH (as found in the stomach) had

little effect of the concentration of individual catechins. However, incubation at slightly alkaline

pH, similar to that found in the small intestine, resulted in a rapid decline in the concentrations of

EGC, EGCG and ECG but not as much to EC. However, the decline was not accompanied by a

Page 22: Nanoencapsulation of Tea Catechins in Casein Micelles

13

reduction of a similar magnitude in the catechins antioxidant activity which is probably due to

the formation of different compounds of catechins (Roura et al., 2001; Cilla et al., 2009).

Intestinal factors are one of the most important factors that affect the bioavailability of

polyphenols. Following the ingestion of polyphenols, the absorption of some but not all

components occurs in the small intestine. The fraction of polyphenols that is not absorbed in the

small intestine reaches the colon, where it undergoes substantial structural modifications by the

colonic microflora which degrades the polyphenol fractions to simple phenolic acids. This

activity is of great importance for the biological action of polyphenols because specific active

metabolites are produced in such a way. During the course of absorption and prior to passage

into the bloodstream, polyphenols undergo many structural modifications due to conjugation

processes, which mainly include methylation, sulfation, and glucuronidation in the small

intestine and later in the liver which typically, enhances their likelihood of excretion in urine or

faeces. These modifications could affect the bioavailability of polyphenols and, consequently,

their biological activity. Therefore, it is clear that, during the course of absorption, polyphenols

are extensively modified. The metabolites present in blood, resulting from digestive and hepatic

activity, usually differ from the native compounds. Consequently, the compounds that reach cells

and tissues are often chemically, biologically, and in many instances functionally different from

the original dietary form. Therefore, due to the extensive biotransformation of polyphenols

during digestion, it is not adequate to predict their health effects by considering only their native

structures and the bioefficacy of their digests must also be assayed and observed (Manach et al.,

2005; Roura et al., 2008; Visioli et al., 2011).

Tea catechins, have shown relatively poor absorption efficiency in vivo, which among the

catechins, absorption of non-gallated forms (EC and EGC) has been more efficient than the

Page 23: Nanoencapsulation of Tea Catechins in Casein Micelles

14

gallated ones (EGCG and ECG) from the same tea matrix (Aron et al., 2008; Henning et al.,

2008; Ferruzzi et al., 2010). Catechins, especially gallated ones, are expected to have the greatest

activity in the oral cavity and the colon, based on their low rates of absorption (Yang et al.,

2008). Catechin losses of approximately 80%, including almost total degradation of EGCG, have

been observed during simulated digestion of simple tea infusions (Green et al., 2007). As the

unabsorbed gallated EGCG goes through the colon, and even the absorbed EGCG is mostly

excreted into the intestine via the bile duct, intestinal environments are the optimal site for

studies observing cancer inhibition by tea polyphenols (Vaidyanathan et al., 2003; Yang et. al.,

2008; Visiolli et al., 2011). It may also be hypothesized that the intestine is one of the most

important sites of activity for polyphenols.

As bioavailability is an issue with polyphenolic compounds, the compounds or agents that are

not well absorbed systemically, will be in direct contact with the digestive tract and may be

important for the cancer-preventive activity at those sites. The intestine may actually be exposed

to high levels of polyphenols after ingestion of polyphenols. Therefore, in many studies

concerning the anti cancer effect of polyphenols, cell line cultures from the digestive tract such

as human colorectal adenocarcinoma cells HT-29 and CaCo-2 are used (Yang 1998; Hong et al.,

2002; Chen et al., 2003; Gonzalez et al., 2010). The proliferation inhibitory effects of tea

catechins extracts (mainly EGCG) against HT-29 or CaCo-2 growth have been noticeably strong

showing a potential source of unknown chemopreventive agents that could be incorporated into

future research on colon cancer (Gonzalez et al., 2010). The findings that tea catechins inhibit

growth in vitro may suggest a potential of this tea for the prevention of colorectal cancer in vivo.

Page 24: Nanoencapsulation of Tea Catechins in Casein Micelles

15

2.1.3. Nano encapsulation of Tea catechins

It is questionable whether individuals can consume a large enough quantity of tea to obtain the

levels of catechins needed for health benefits. Therefore, utilisation of tea extracts in foods has

been considered as an alternative way to provide the health benefits from tea catechins (Voung et

al., 2011). Manufacturers have attempted to produce and package ready-to-drink green tea

infusions, some enriched with (-)-epigallocatechin gallate (EGCG) and/or other polyphenols, in

order to respond to the increasing market for functional foods that have potential health benefits.

Green tea extracts are also used as ingredients in many dietary supplements, such as vitamins and

weight reduction pills. However, the production of green tea beverages has been seen to be

problematic as the stability of EGCG and more other green tea catechins in solutions and drinks

is always a challenge (Kim et al., 2007; Bazinet et al., 2010; Yang et al., 2012). Food product

developers are therefore left with the challenge to preserve the stability, bioactivity and

bioavailability of these bioactive molecules, not only during processing and storage of foods, but

also during digestion. Indeed, to be able to deliver the nutritional functionality, food products

need to have not only a nutritionally significant dose of bioactive molecules, but these molecules

must also remain fully functional for as long as necessary (Garti, 2008). Encapsulation of these

bioactives may provide a solution to these challenges. EGCG seems to be an important candidate

for nanoencapsulation and enrichment in the diet (Shpigelman et al., 2010).

In nanoencapsulation technology, bioactives and drugs are loaded within the core of submicron

capsules or particles, which protect them from degradation (Reis et al., 2006). Interest in

nanocarriers as potential drug carriers in cancer therapy has expanded in the recent years. Milk

proteins are known to be natural carriers, as their structure facilitates the delivery of essential

micronutrients (calcium and phosphate), building blocks (amino acids), and immune system

Page 25: Nanoencapsulation of Tea Catechins in Casein Micelles

16

components (immunoglobulins, and lactoferrin), from mother to the neonate. This unique

property of milk proteins has encouraged researchers to study possibilities of using these natural

nano carriers for the delivery of health-promoting and bioactive compounds such as polyphenols.

Milk proteins are natural, available, inexpensive raw materials with high nutritional value and

good sensory properties, and they have many structural properties and functionalities which

make them highly suitable as vehicles for delivering various bioactives (Livney 2010).

Researchers have suggested the potential of milk proteins to deliver nutraceuticals such as

vitamins D, curcumin, minerals (Fe2+

, Mg2+

), carotenoids (β–carotene), polyphenols (resveratrol,

caffeine), omega 3 oils and other health-promoting fatty acids, probiotics and other

microorganisms, cancer therapy drugs and even flavors or aromas. Studies have shown that both

whey proteins and caseins as well, usually in their isolated forms have been able to act as

delivery vehicles. For example whey proteins such as β-lactoglobulin are known to be carriers of

molecules such as vitamin D, omega-3 fatty acid DHA, or polyphenolic compounds like

resveratrol from grapes especially after β-lactoglobulin has been heated. α-lactalbumin can

apparently bind to retinol and palmitic acid, while bovine serum albumin is often proposed for

delivery tasks in the circulatory system for delivering chemotherapeutic drugs (5-fluorouracil).

Also, β-casein nanoparticles have been proposed to be used for the delivery of gastric cancer

treatment drugs (Livney, 2010; Rahimi Yazdi et al., 2012).

Recent developments have shown that curcumin can be delivered via caseins to tumour cells, and

that vitamin D can be protected from UV degradation by encapsulation in casein micelles,

showing the potential of these nanostructures in the delivery of neutraceuticals (Semu et al.,

2007; Sahu et al., 2008; Livney, 2010; Shpigelman et al., 2010).

Page 26: Nanoencapsulation of Tea Catechins in Casein Micelles

17

Traditionally, in many countries, tea is consumed with the addition of milk to improve the

sensory properties and reduce the astringenty sensation caused by polyphenols.

Also, development of new products with tea ingredients formulated on dairy bases such as ready-

to-drink (RTD) tea beverages, chai latte beverages and other products containing dairy products

have expanded over the years (Ferruzzi et al., 2006). Milk is a natural, multi-component,

nutrient-rich beverage and is generally recognized as a source of beneficial substances for

growth and health in children and adults (Kalkwarf et al., 2003; Valeille et al., 2006). Also, for

oral delivery applications, biocompatibility of milk proteins is usually very good. Therefore,

milk is an ideal platform for delivery of other bioactive compounds and their additional benefits

to human health. Functional dairy beverages can satisfy the growing market need for health and

taste, but significant knowledge of the bioactive components and their interaction with dairy

components is required (Sharma 2005).

2.2. Milk

Milk is a fluid secreted by all mammals to supply to the neonate the nutrients required for its

growth. Milk is a colloidal suspension consisting of milk fat globules, milk casein proteins and

an aqueous phase consisting whey proteins, lactose, organic and inorganic salts dissolved in

water (Fox et al., 2000). In this polydispersed system, fat globules can be easily separated by

gravity or centrifugation. The typical composition of bovine milk is approximately 87.7% water,

4.9% lactose (carbohydrate), 3.4% fat, 3.3% protein, and 0.7% minerals. Bovine milk has pH of

approximately 6.7 (Fox et al., 1998). The protein fraction of milk is composed of caseins (αs1-,

αs2-, β-, and κ-casein) and whey proteins (mostly β-lactoglobulin, α-lactalbumin and bovine

serum albumin), accounting for ~80% and ~20%, respectively (Fox et al., 1998). The milk

Page 27: Nanoencapsulation of Tea Catechins in Casein Micelles

18

serum, composed of lactose, small molecular weight components and salts, can be separated

from the proteins by filtering skim milk using an ultrafiltration membrane (Walstra et al., 2006).

2.2.1. Caseins

About 80% of the total proteins of bovine milk are caseins, defined as the proteins that

precipitate at pH 4.6 at 20 °C (Fox et al., 2000). There are four individual types of casein

molecules, αs1-, αs2-, β- and κ-casein, in approximate relative amounts of 4:1:3.5:1.5,

respectively. Caseins are small molecules about 20-25 kDa in molecular weight (Dalgleish,

1998). These protein assemblies are designed by nature to concentrate, stabilize and transport

essential nutrients, mainly calcium, phosphate and protein, for the neonate. These proteins are of

great importance in dairy technology, as their processing functionality (for example, their gel

forming abilities) are key to many dairy processes. The primary structure of caseins is

characterized by a high amount of proline, which impedes the formation of significant portions

of ordered structures. For this reason caseins are very flexible and are defined rheomorphic, as

they can adapt their conformation to changing environmental conditions (Holt et al., 1993).

There are four main caseins, αs1-, αs2-, β- and κ-casein, and they differ in some molecular aspects

such as, degree of phosphorylation, glycosylation of κ–casein, hydrophobicity and proline

content. All caseins are phosphorylated, most molecules of αs1-casein contain 8 or 9 PO4

residues, β-casein usually contains 5 PO4 residues, αs2-casein contains 10-13 PO4 residues and κ-

casein contains only 1 PO4 residue. The phosphate groups of the caseins are esterified as

monoesters of serine and most of the phosphoserine residues in the caseins occur in clusters. The

phosphate groups are very important from a nutritional view point but they also bind polyvalent

cations strongly. Therefore apart from κ-casein, all of the other caseins are calcium sensitive as

Page 28: Nanoencapsulation of Tea Catechins in Casein Micelles

19

they precipitate in the presence of calcium ions (Fox et al., 2003). κ-Casein is the only

glycosylated casein which contains galactose, galactosamine and N-acetylneuraminic (sialic)

acid, that occurs either as trisaccharides or tetrasaccharides attached to theronine residues in the

C-terminal region. Caseins, especially β -casein, contain a high level of proline (cyclic amino

acid); in β -caseins 35 of 209 amino acid residues are prolines that are uniformly distributed

throughout the protein molecule, κ –casein contains 20 proline residues out of 169 amino acids,

αs1 and αs2-caseins contain 17 and 10 proline residues from 199 and 207 amino acids,

respectively. The caseins are relatively hydrophobic and, in particular, hydrophobic, polar and

charged residues on the caseins are not uniformly distributed throughout the sequences but occur

as hydrophobic or hydrophilic patches, giving the caseins strongly amphipathic structures.

αs1-casein has two hydrophobic and one hydrophilic region, that includes the phosphoserine

cluster, αs2-caseins has two hydrophobic and two phosphoseryl clusters, β–caseins has only one

hydrophobic and one phosphoseryl group and κ-casein has a hydrophobic N-terminal region but

no phosphoseryl cluster (Fox et al 2003; Farrell et al., 2006; Horne 2006; Dalgleish, 2011).

Nearly all caseins in milk (95%) are organized into colloidal particles known as micelles.

In bovine milk, the casein micelles are spherical particles of diameters ranging between 80-400

nm, with an average of about 200 nm and containing between 3-4 g of water per g protein

(Dalgleish, 2011). They contain 94% protein and 6% low molecular weight species (mainly

calcium and phosphate but also magnesium and citrate), which are collectively known as

colloidal calcium phosphate.

Many models for the structure of the casein micelles have been proposed and yet in regards to

the interactions involved in their assembly and maintaining the integrity of these particles,

dispute amongst researchers remains. However, there is a general agreement that caseins are held

Page 29: Nanoencapsulation of Tea Catechins in Casein Micelles

20

together by an equilibrium between electrostatic and hydrophobic interactions (Horne, 2009;

Horne, 2006) as well as van der Waals forces (Dalgleish, 2011). The interior of the micelles is

mainly composed of αs1-, αs2-and β-caseins bound to calcium phosphate nanoclusters by their

phosphoserine domains. κ-casein exists mainly on the surface of the micelle with the hydrophilic

C-terminal part protruding into the solvent. Of particular relevance for milk coagulation, the

N-terminal of κ-casein (2/3 of the molecule, called para-κ-casein) is strongly hydrophobic,

whereas the C-terminus (called casein macropeptide, or CMP) is strongly hydrophilic and is

usually glycosilated.

κ-caseins are not evenly distributed on the surface, but are present in clusters that are easily

accessible for attack by chymosin (and related enzymes) (Fox et al., 2003; Holt et al.,1996; de

Kruif et al., 2003).

Milk proteins also contain tryptophan residues throughout their molecular structure where the

αs-caseins contain two tryptophan residues at positions 164 and 199 for αs1 and positions 109 and

193 for αs2, β- and κ-casein have just one tryptophan residue at position 143 and 76,

respectively. As described in the different models of casein micelles, it is widely accepted that

the αs- caseins and β-casein are mainly located in the interior hydrophobic core of the micelle,

whereas as κ- casein is mostly on the surface of the casein micelles (Dalgleish 2010). Therefore,

most of the tryptophan residues are located in the hydrophobic domains of the casein micelle.

Casein micelles are destabilized by the proteolytic activity of specific enzymes, by isoelectric

precipitation, by addition of an excess of calcium ions or ethanol which their destabilization is at

the basis of many dairy processes (de Kruif et al., 1996; de Kruif, 1999).

Page 30: Nanoencapsulation of Tea Catechins in Casein Micelles

21

2.2.2. Whey Proteins

The other 20% of the total proteins in milk are whey proteins (WP) which are globular proteins

with high level of structural organization, causing susceptibility to thermal denaturation. The

compact and stable structure is given by intramolecular disulfide bonds, formed between the high

amount of cysteine residues. Unlike the caseins, which predominantly exist as associative

colloidal particles, the whey proteins are soluble in the milk serum (Fox et al., 2000).

The whey proteins include 50% β-lactoglobulin (β-LG), 20% α-lactalbumin (α-LA), 10% (BSA)

and 10% immunoglobulins (Ig). However, the first two are present in the greatest proportion and,

therefore, are of technological significance. At milk pH, β-lactoglobulin is a dimer of molecular

weight of 36.4 kDa while α-lactalbumin is a monomer of 14 kDa. Whey proteins are non-

phosphorylated and insensitive to calcium ions, are soluble at pH 4.6, soluble after rennet

coagulation of caseins and non sedimentable by centrifugation (Fox et al., 2003).

Following heat denaturation, β-lactoglobulin exposes its cysteine residues and forms complexes

with other cysteine containing proteins, such as α-lactalbumin, bovine serum albumin or

κ-casein (Livney et al., 2003).

2.2.3. Rennet induced gelation of milk

Milk proteins (caseins, whey proteins) are used in food products not only for their nutritional

value but for their functional properties as well. Major functionalities of milk proteins include

gelation (acid or/and rennet induced) emulsification, and foaming properties.

Rennet induced gelation (enzymatic coagulation) is the first step of the cheese-making process

and particularly relevant in dairy processing. This specific functionality of casein micelles occurs

without changes in pH and is induced by the enzymatic modifications of the surface of the casein

Page 31: Nanoencapsulation of Tea Catechins in Casein Micelles

22

micelles. The process is composed of two stages, which partially overlap each other (Dalgleish

2010). During the first stage, the enzymatic action (by chymosin) destabilizes casein micelles,

while in the second stage the micelles aggregate, forming a coagulum.

During the first stage, the proteolytic enzyme chymosin cleaves κ-casein, which mostly resides

on the surface of the casein micelles. Chymosin is a proteolytic enzyme active on κ-casein,

specifically on the Phe105-Met106 peptide bond. While the hydrophobic para-κ-casein segment

(residues 1-105) remains associated with the micelle, the hydrophilic portion of κ-casein, the

casein macropeptide (CMP, residues 106-169) is released into solution (Dalgleish et al., 2012).

The polyelectrolyte layer of κ-casein stabilizes the micelles by generating steric repulsion, and

therefore, any process that removes the C-terminal region of κ-casein causes a decrease in the

colloidal stability of the micelles. When approximately more than 85% of the κ-casein is

hydrolyzed, the steric stabilization generated by the few remaining κ-casein hairs is insufficient

to keep the micelles apart, and they begin to aggregate and eventually gel. Therefore, gelling

occurs only after a nearly full release of CMP in solution. Once decreased steric repulsion allows

the micelles to approach one another closely, hydrophobic interactions cause bonding between

the particles, as does the amount of ionic calcium; increased calcium also allows aggregation at

progressively lower levels of CMP release (Horne, 1986; de Kruif, 1999; Horne 2006; Dalgleish

2012).

2.3. Interactions between milk proteins and tea catechins

The interaction between polyphenols and proteins is relevant to diverse fields of knowledge such

as medicine, toxicology, chemistry, food science, and agriculture. Hence, the reactions have been

Page 32: Nanoencapsulation of Tea Catechins in Casein Micelles

23

studied throughout the years by numerous technical approaches (Goncalves et al., 2011). The

interactions between polyphenols and proteins are driven by hydrogen bonding between phenolic

hydroxyl and peptide carbonyl, and hydrophobic ―stacking‖ interactions between hydrophobic

amino acid residues and the aromatic rings of the phenols (Hofmann et al., 2006; Pascal et al.,

2008; Yan et al., 2008). In addition, binding is clearly affected by protein characteristics

including protein structure and amino acid composition, with proline-rich proteins having a

particularly high relative affinity for polyphenols, as well as polyphenol structure (eg.,

glycosylated or not) and molecular size (Poncet-legrand et al., 2006; Richard et al., 2006; Soares

et al., 2007; Frazier et al., 2009; Dubeau, 2010) .

It has been widely demonstrated that the astringency sensation is a consequence of the

interactions between proline-rich saliva proteins and polyphenols (O’Connell 2001; Charlton

2002; Hofmann et al., 2006). In a recent study, it has been shown that saliva proteins and

polyphenols interacts strongly and the astringency sensation is due to the free polyphenols rather

than the bound ones (Schwarz et al., 2008). In general, it has been found that catechins readily

interact with proteins rich in proline, with an open and flexible structure (Benick et al., 2002;

Papadopoulou et al., 2004; Maiti et al., 2006). Caseins contain high numbers of proline

residues evenly distributed throughout their amino acid sequences and have relatively open

structural features (like the salivary proline rich proteins), and are avid binders of polyphenols.

For this reason, isolated caseins have often been used as model proteins in various polyphenol-

protein studies (Jobstyl et al., 2004; Pascal et al., 2007; Yan et al., 2009). Catechins bind

strongest to the caseins (β-casein> αs-casein> κ-casein), followed by the whey proteins α-

lactalbumin, β -lactoglobulin and bovine serum albumin (Luck et al., 1994; Kartosova et al.,

2007). β-casein contains 209 amino acid residues, of which 35 prolines evenly distributed

Page 33: Nanoencapsulation of Tea Catechins in Casein Micelles

24

through the sequence (Ribadeau et al., 1973). EGCG and β-casein interact by forming

hydrophobic interactions between the phenolic rings of the EGCG and prolines in β-casein,

wrapping the casein around the catechin (Jobstyl et al., 2006).

In agreement with recent studies on saliva proteins (Schwartz et al., 2008), sensory studies have

shown that caseins, especially β-caseins, due to their strong binding ability to tea catechins,

decrease the sensation of astringency caused by the tea polyphenols. It has also been suggested

that since binding reduces astringency, it may decrease the bioavailability of EGCG as well

(Jobstyl et al., 2006; Yan et al., 2009). Although, numerous studies have shown the strong

interactions of polyphenols and milk caseins in their isolated forms (most frequently β-casein),

few studies have been conducted to delineate the nature of association of polyphenols and casein

micelles in their natural matrix, milk.

Past studies present contradictory results regarding the effect of milk on the bio-availability of

tea catechins, whereas, the effect of the complexes formed between tea catechins and milk

proteins in terms of decreasing or increasing the bioefficacy of these bioactive molecules is still

controversial. Some studies have found a masking effect of the antioxidant capacity of catechins

in the presence of proteins, the extent of which depended on both polyphenol and protein types.

The tea polyphenol that showed the highest masking in combination with β-casein, was

epigallocatechin gallate (EGCG) (Arts et al., 2002; Kilmartin et al., 2003; Kartosova et al.,

2007). The masking effect was also seen in the blood among volunteers who drank tea with milk,

compared to those who drank tea without milk, which it has been suggested that the interactions

between milk proteins (especially caseins) and the tea catechins are responsible for the decline of

the antioxidant capacity of tea catechins (Langley-Evans, 2000). In contrast, other similar studies

showed no relevant effect of milk on the plasma antioxidant capacity of tea polyphenols (Leenon

Page 34: Nanoencapsulation of Tea Catechins in Casein Micelles

25

et al., 2000; Keogh et al., 2007; Kyle, 2007). In addition, it has been shown that different

methods to determine the effect of milk on the antioxidant capacity of milk may lead to different

conclusions (Dubeau et al., 2011). Also, it has been shown by many studies that tea catechins,

more specifically EGCG, may exert their biological health benefits by not only acting as an

antioxidant but as well as a pro-oxidant. Therefore, determining the antioxidant capacity of tea

polyphenols may not be the most suitable marker or indicator of their bioacessability or bio-

availability.

It is important to state that the above discrepancies concerning the effect of milk on tea might be

due to the different study designs, different methodologies, and types of tea used. Most studies

have overlooked the strong interactions that take place between the proteins and polyphenols,

considering effective (i.e., bioavailable) only the material recovered in the soluble phase.

For example, one study claims that the addition of milk to fruit juices significantly decreases

their total polyphenol content compared to non-fortified fruit beverages, with flavan-3-ols

(catechins) being the most affected compounds. The decrease was more significantly seen in the

polyphenol content of fruit juice–milk digests after in vitro digestion (Cilla et al., 2009).

Sharma et al., 2008 also showed a significant decrease in the free phenolic content of black tea

after the addition of milk. In contrast, more recent studies have shown that the stability of

catechins in green tea post digestion is very poor, however green tea samples formulated with

50% bovine, soy, and rice milk, significantly increased total catechin recovery (Green et. al.,

2007; Ferruzzi, 2010).

Another factor that must be taken account is that during the course of gastric and intestinal

digestion, polyphenols undergo many structural and biological modifications that usually differs

them from the original compounds indigested. These modifications result in a complex

Page 35: Nanoencapsulation of Tea Catechins in Casein Micelles

26

circulating profile of both free catechins, methylated, glucuronidated and sulfated catechin

metabolites which many of the metabolites are yet to be identified. Consequently, these

metabolites that reach cells and tissues, in many instances function differently from the original

dietary form. Therefore, due to the extensive biotransformation of polyphenols during digestion,

it is not adequate to predict their health effects by considering only their native structures and the

bioefficacy of their digests must also be assayed and observed. However, many studies test the

compound of little biological relevance when trying to identify the physiological mechanisms

involved in the health effects of polyphenols, rather than testing their active metabolites (Manach

et al., 2005; Roura et al., 2008; Visioli et al., 2011).

Considering the poor digestive stability of catechins from plain green tea infusions, the ability of

milk to potentially modulate the tea catechin profile must be further studied. The protein–

catechin interactions may provide a physical trapping of the reactive catechin species causing

more stability of the catechins which ultimately may impact the catechin bioavailability and

physiological profile in vivo.

Interpretation of such studies is difficult since by the addition of milk, the product matrix

complexity increases, and therefore accurate and reliable measurements of tea catechins becomes

more challenging (Ferruzzi et al., 2006).

To date most of the polyphenol-milk protein studies have been mainly focused on what impact

the complex may have on the bioavailibilty of polyphenols, whereas, the fate of milk proteins,

during digestion has not been observed. Therefore, it is of importance to understand the impact

of proteolytic activity on the polyphenol-milk complexes as several active peptides, which exert

various activities on the digestive, cardiovascular, immune and nervous systems, are produced

during the digestion of caseins (Korhonen et. al., 2006; Picariello et. al., 2010). Some studies

Page 36: Nanoencapsulation of Tea Catechins in Casein Micelles

27

have shown, polyphenols may have an inhibitory effect on certain digestive proteases

(Tagliazucchi et. al., 2005; McDougall et al., 2008) while some other studies have seen a

stimulating effect on the rate of protein digestion (Siewright et al., 2006; McDougall et al., 2005;

Goncalves et. al., 2011). Although inhibition of proteases by polyphenol components can occur

in vitro, there is still not enough evidence to show that inhibition of protease activity occurs in

the digestive system in vivo (McDougall et al., 2008). However, this may still be a crucial factor

in casein digestion that must be observed when tea and milk complexes are consumed.

Page 37: Nanoencapsulation of Tea Catechins in Casein Micelles

28

Chapter 3

Interactions between tea catechins and casein micelles and their impact on

renneting functionality

3.1 Abstract

Strong interactions have been reported between tea catechins and milk proteins in isolation, but

less is understood on the association of polyphenols with casein micelles, and more importantly,

the consequences of the interactions on processing and biological functionalities. It was

hypothesized that epigallocatechin-gallate (EGCG), the main catechin present in green tea forms

complexes with the casein micelles, and the association modifies the processing functionality of

casein micelles. In this study, the binding of EGCG to casein micelles was quantified using

HPLC and fluorescence quenching. It was demonstrated that the formation of catechin-casein

micelles complexes affected the rennet induced gelation of milk. Tea catechins affected both the

primary as well as the secondary stage of gelation. These experiments clearly identify the need

for a better understanding of the effect of tea polyphenols on the functionality of casein micelles,

before milk can be used as an appropriate platform for delivery of bioactive compounds.

Key words: Polyphenols, caseins, interactions, gelling, fluorescence

Page 38: Nanoencapsulation of Tea Catechins in Casein Micelles

29

3.2 Introduction

Tea (Camellia sinensis) polyphenols have been reported to exhibit several beneficial health

effects by acting as antioxidant, anticarcinogen and cardiopreventive agents (Ferruzzi et al.,

2006). This is of particular interest as tea is the second most consumed beverage after water.

The most abundant group of polyphenols in tea are catechins. These compounds are water-

soluble, colorless compounds which contribute to astringency and bitterness in tea.

Major tea catechins include epicatechin, epigallocatechin, epigallocatechin-gallate, and

epicatechin-gallate in which epigallocatechin-gallate (EGCG) is the major catechin component

and also the most biologically active (Vaidyanathan et al., 2003; Henning et al., 2003). These

natural chemicals, especially EGCG, exhibit excellent radical scavenging and hence

antioxidative properties, and have been widely employed as food antioxidants (Graham et al.,

1992; O’Connell et al., 2001).

Traditionally, in many countries, tea is consumed with the addition of milk to improve the

sensory properties, i.e., to reduce the astringency sensation caused by polyphenols (Bennick

2002). Also, development of new products with tea ingredients formulated on dairy bases such as

ready-to-drink teas, chai latte and other beverages containing dairy products have expanded over

the years (Ferruzzi et al., 2005). Milk is a natural, multi-component, nutrient-rich beverage and is

generally recognized as a source of beneficial substances for growth and health in children and

adults (Kalkwarf et al., 2003; Valeille et al., 2006). For this reason, milk is an ideal platform for

delivery of other bioactive compounds, to obtain a new generation of dairy products providing

additional benefits to human health. Functional dairy beverages can satisfy the growing market

need for health and taste, but significant knowledge of the bioactive components and their

interaction with dairy components is required (Sharma 2005). It is important to determine

Page 39: Nanoencapsulation of Tea Catechins in Casein Micelles

30

whether the addition of these compounds in milk may affect the processing functionality of the

casein micelles. Our work aims to explore some of the effects of the addition of polyphenols to

milk proteins and their relevance to the design of improved dairy products.

Strong interactions have been reported between tea catechins and purified milk proteins in model

solutions. Catechins bind strongest to the caseins, followed by whey proteins, namely α-

lactalbumin, β-lactoglobulin and bovine serum albumin (Luck et al., 1994; Kartosova et al.,

2007). Sensory studies have shown that caseins, especially β-caseins, due to their strong binding

ability to tea catechins, decrease the sensation of astringency caused by the tea polyphenols

(Lesschaeve et. al., 2005; Hofmann et. al., 2006; Soares et. al., 2007; Schwarz et. al., 2008).

It has also been suggested that since binding reduces the astringency, it may decrease the

bioavailability of EGCG (Jobstyl et al., 2006; Yan et al., 2009).

Catechins readily interact with proteins rich in proline, with an open and flexible structure

(Benick et al., 2002; Papadopoulou et al., 2004; Maiti et al., 2006). The astringency perceived

during in mouth processing of polyphenol rich foods is attributed to the free polyphenols, rather

than those bound to proline-rich saliva proteins (O’Connell, 2001, Charlton, 2002; Hofmann et

al., 2006; Schwarz et al., 2006). Caseins which constitute about 80% of the proteins of bovine

milk contain high numbers of proline residues evenly distributed throughout their amino

acid sequences and have relatively open structural features (like the salivary proline rich

proteins). For this reason, isolated caseins have often been used as model proteins in various

polyphenol-protein studies (Jobstyl et al., 2004; Pascal et al., 2007; Yan et al., 2009).

The effect of the interactions of polyphenols with casein micelles on the processing functionality

of milk proteins is not fully understood, apart from the fact that it has been recognized that the

Page 40: Nanoencapsulation of Tea Catechins in Casein Micelles

31

presence of polyphenols enhances the heat stability of milk (O’Connell et al., 1999; O’Connell et

al., 2001).

The main objective of the proposed research is to further our understanding of the impact of the

milk protein/polyphenol interactions on the functional properties of milk proteins.

Although many studies have acknowledged these interactions, few studies have focused on the

association of polyphenols with casein micelles in milk. Indeed, most studies have been limited

to the interaction of isolated caseins (most frequently β-casein) with a range of polyphenols, but

not the casein micelles. To date, the possible effects of the interactions on the functional

properties of milk proteins have been studied to a lesser extent than the effects on the

bioavailability of the polyphenols. The effects of the complexes formed with the casein micelles

on the functional properties of caseins, in particular their rennet-induced aggregation, have yet to

be reported.

3.3 Materials and Method

3.3.1 Milk preparation and interaction studies

Fresh milk was collected from the Ponsonby Research Station of the University of Guelph

(Guelph, Ontario, Canada) and sodium azide (0.02% w/v) was immediately added to prevent

bacterial growth. Milk was skimmed at 6000 g for 20 min, at 20°C, in a Beckman–Coulter

centrifuge (Model J2-21, Beckman Coulter, Mississauga, ON, Canada) and filtered three times

through Whatman glass fibre filters (Fisher Sci.). Milk permeate was prepared from skim milk,

using a laboratory scale ultrafiltration cartridge (Millipore CDUF001LG; Fisher Scientific), with

a nominal cut-off of 10,000 Da and a nominal area of 1 ft2 (0.095 m

2).

Page 41: Nanoencapsulation of Tea Catechins in Casein Micelles

32

A source of green tea polyphenol extract (containing minimum 90% EGCG) was chosen from

DSM Nutritional Products INC (Montreal, Canada). Different concentrations of EGCG was

obtained from a stock solution of EGCG in water (20 mg/ml water). When needed, dilutions

were carried out using permeate (milk serum), to reach high ratios of EGGC to caseins while

maintaining concentrations of EGCG below those critical for solubility.

Separation of the soluble proteins (whey) from the casein micelles was accomplished by

centrifugation (36,000 g), at 20°C, for 45 min (Beckman Coulter, Canada, Inc., Mississauga, On,

Canada), as previously described (Del Angel et al., 2006). Afterwards, the pellet containing the

micellar caseins was re-diluted in fresh ultrafiltration permeate (see above) for fluorescence

experiments.

3.3.2 Fluorescence spectroscopy

Fluorescence quenching has been widely studied as a source of information about biochemical

systems in which biochemical applications of quenching are due to the molecular interactions

that result in quenching (Rawel et al., 2006; Lakowicz, 2006). The binding of EGCG with milk

proteins was quantified by fluorescence spectrophotometry (Shimadzu RF-5301PC

spectrofluorometer, Shimadzu Corp., Tokyo, Japan). Milk was diluted 1:20 (v/v) in milk serum

(permeate) and the fluorescence spectra were collected at varying concentrations of EGCG. In

addition, similar experiments were also performed with casein micellar suspensions, prepared as

described above. The casein suspension was also diluted 1:20 (v/v) in milk permeate. Samples

were excited at wavelength of 280 nm and the emission spectra were recorded from 300 to 500

nm (with slit widths of 3 for both excitation and emission).

Page 42: Nanoencapsulation of Tea Catechins in Casein Micelles

33

The Stern-Volmer equation (equation 3.1), was used to plot F0/F for the skim milk and casein

samples as a function of EGCG concentration added. F0 is the fluorescence intensity of control

skim milk or control caseins without the addition of EGCG, F is the fluorescence intensity after

fluorescence quenching (after addition of EGCG), [Q] is the concentration of the quencher

(EGCG) and Ksv is the Stern-Volmer quenching constant (Lakowicz, 2006).

F0/F = 1+ Ksv [Q] Eq. 3.1

Also, using a modified form of the Stern-Volmer equation (equation 3.2), the ratio of F0/ΔF was

plotted as a function of the concentration of EGCG where, the fa, which is the fraction of the

fluorescence that is accessible to quencher, and the Ka, the Stern-Volmer quenching constant of

the accessible fraction were calculated (Lakowicz, 2006).

F0/ ΔF = 1/ faKa[Q] + 1/ fa Eq. 3.2

In these experiments, the concentration of quencher was normalized by the amount of protein

present in the sample, specifically, 32 mg/mL for skim milk and 26 mg/mL for the casein

micelles suspensions.

Page 43: Nanoencapsulation of Tea Catechins in Casein Micelles

34

3.3.3 Quantification of polyphenol binding to casein micelles

Different concentrations of EGCG were added to skim milk, and after approximately 20 min of

incubation at room temperature (22°C), the mixtures were centrifuged at 36000g for 45 min (see

above) to separate the colloidal phase (consisting of casein micelles) from the whey proteins and

soluble EGCG. After the separation, the amount of catechin present in the whey fraction (i.e.,

the centrifugal supernatant) was analyzed by reversed phase HPLC as previously published

(Ferruzzi et al., 2006). In brief, aliquots of whey were combined with 2% aqueous acetic acid

(1:1 v/v) and centrifuged at 14,000g for 5 min. Supernatants were collected and filtered through a

0.45 µm PTFE filter and immediately injected. A parallel set of tea catechin solutions containing

whey protein isolates (separated from skim milk as described above) were also prepared and

were used as standard curves to determine the effect of the presence of whey proteins in the

quantification of the EGCG. No differences were noted, when compared to samples containing

isolated EGCG fractions (without whey protein). Hence, the experiments were run using whey

protein fractions and EGCG fractions as standards.

To quantify the binding of EGCG in milk, Scatchard plot was used to determine the dissociation

constant (Kd) which is a useful measure to describe the strength of binding (or affinity) between

proteins and their ligands. Using the Scatchard equation (equation 3.3 ) the Kd between EGCG

and casein proteins was determined. Where , is the average number of ligand molecules bound

per protein molecule. Curved Scatchard plots can be decomposed into successive linear plots,

and the dissociation constant Kd can be derived from each of the slopes (Dickinson et. al., 2001).

/[L] = 1/Kd - /Kd Eq. 3.3

Page 44: Nanoencapsulation of Tea Catechins in Casein Micelles

35

3.3.4 Rheological properties of rennet-induced gels

A fresh rennet solution was prepared from double strength rennet (Chymostar, Danisco,

Cranberry, NJ, USA) and immediately added to milk and tea-milk samples yielding a final rennet

concentration of 3.14x10-4

IMCU/ml. Immediately after rennet addition at 30°C, twenty-

milliliter samples were placed in concentric cylinders in a controlled-stress rheometer AR1000

(TA Instruments, New Castle, DE). The development of the gel structure was followed using

0.01 constant strain, 1.0 Hz frequency and temperature of 30°C. The time of cross-over between

G’ (storage modulus) and G’’ (loss modulus) was considered the gelation point.

3.3.5 Release of casein macro- peptide (CMP) from the surface of the casein micelles

Renneted milk and tea-milk samples were pipetted to several test tubes in a waterbath maintained

at 30°C. The rennet reaction was stopped at specific time intervals by addition of 3% perchloric

acid with subsequent vortexing. After overnight refrigerated storage, the supernatant was

collected after centrifugation at 4500g for 15 minutes, filtered through 0.45 μm filter, and

analyzed by reversed phase HPLC as previously published (López-Fandiño et al, 1993).

Intregration of the peaks was done by using ChromQuest software v.4.1 (ThermoFinnigan,

Burlington, ON, Canada) and compared to the maximum peak area for skim milk without tea

polyphenol added (control milk).

3.3.6 Statistical analysis

The analyses were repeated at least three times and evaluated by their means and SD.

Page 45: Nanoencapsulation of Tea Catechins in Casein Micelles

36

3.4 Results and Discussion

3.4.1 Fluorescence spectroscopy studies

The fluorescence spectra and the values of maximum intensity as a function of EGCG

concentration are depicted in Figure 3.1 for samples of skim milk and samples of micellar

caseins. A control of EGCG solution measured under the same conditions yielded a very low

signal. In both skim milk and casein micelles suspensions (Figure 3.1 A and B, respectively),

with the increase of EGCG concentration, the maximum fluorescence intensity decreased until

reaching a plateau. This plateau value was statistically equivalent between skim milk and casein

micelles samples. As shown in the insets of Figure 3.1, the emission maximum for both skim

milk and casein micelle samples (approximately 335 nm) progressively shifted to longer

wavelengths with increasing EGCG concentrations, together with a decreasing in intensity.

Similar results were noted for micellar caseins as well as skim milk, suggesting that most of the

quenching effect of tryptophan can be attributed to the association of EGCG with the casein

proteins. Since the changes in emission maximum of tryptophan residues is closely related to the

polarity of the microenvironment around them, the observations suggest that a change in the

conformation of caseins took place as EGCG concentration was increased (Ghosh et al., 2008).

It is important to note that fluorescence measurements on purified caseins (α and β caseins),

showed a similar behavior with a shift to longer wavelengths but no peak broadening at the high

concentrations of polyphenols. Theses studies suggested that the upward shift of the protein

emission bands observed in the spectra of EGCG–casein is related to more exposure of

tryptophan residue and unfolding of protein structure, which was also shown in infrared results

that showed major secondary structural changes for caseins upon EGCG complexation (Hasni et

al., 2010). However, the binding of EGCG with whey protein isolates and β-lactoglobulin has

Page 46: Nanoencapsulation of Tea Catechins in Casein Micelles

37

been previously studied whereas, the addition of EGCG to whey proteins did not cause a shift in

the wavelength, therefore indicating that the shift seen in Figure 3.1 is not related to the spectrum

of EGCG, but it is specific to the EGCG interactions with caseins (Shpigelman et al., 2010).

Page 47: Nanoencapsulation of Tea Catechins in Casein Micelles

38

Figure 3.1 Fluorescence intensity and fluorescence spectra (inset) for (A) caseins and (B) skim

milk. Arrow indicates increasing concentrations of EGCG. Values are the average of three

independent experiments.

A

B

Page 48: Nanoencapsulation of Tea Catechins in Casein Micelles

39

In the case of skim milk, the intrinsic fluorescence intensity is usually attributed to the caseins in

milk (αs1-, αs2-, β-, and casein) although α-lactalbumin and β-lactoblobulin (whey proteins), have

2 and 4 tryptophan residues, respectively (Brew et al., 1967). While the αs- caseins contain two

tryptophan residues at positions (164 and 199 for αs1) and positions (109 and 193 for αs2), β- and

κ-casein have just one Trp residue at position 143 and 76, respectively. As described in the

different models of casein micelles, it is widely accepted that the αs- caseins and β-casein are

mainly located in the interior hydrophobic core of the micelle, whereas as κ-casein is mostly on

the surface of the casein micelles (Dalgleish 2010). Therefore, most of the tryptophan residues

are located in the hydrophobic domains of the casein micelle.

The interactions between polyphenols and proteins are driven by hydrogen binding between

phenolic hydroxyl and peptide carbonyl, and hydrophobic ―stacking‖ interactions between

hydrophobic amino acid residues and the aromatic rings of the phenols (Hofmann et al., 2006;

Pascal et al., 2008; Yan et al., 2008). Furthermore, a study using purified caseins and tea

catechins suggests that the binding of the polyphenols to caseins seems to be more hydrophobic

than hydrophilic (Hasni et al., 2010).

To quantify the extent of the interactions, the fluorescence spectra of skim milk proteins and the

isolated casein micelles with EGCG were analyzed as shown in Figure 3.2. As casein proteins

are in the form of micelles and most of the tryptophan residues are located in the hydrophobic

domains of the micelle, each residue can be differently accessible to the quencher. The deviation

of linearity in the Stern Volmer plots shown in Figure 3.2 suggests that EGCG may not be able

to access all the sites within the casein micelles, as full quenching is not obtained with increasing

concentrations of EGCG in milk and in casein micelles as well, as that a plateu of quenching was

reached at approximately 0.04 mg EGCG per mg milk proteins. Whereas, recent studies with

Page 49: Nanoencapsulation of Tea Catechins in Casein Micelles

40

curcumin have demonstrated that with this polyphenol molecule full quenching of the

tryptophans is achieved indicating the penetration of curcumin within the hydrophobic core of

the casein micelles (Rahimi Yazdi et al., 2012). It has been previously suggested that small

molecules have the ability to penetrate into the inner core of the casein micelles due to the

porosity of the micelles which contain water channels within their structure (Dalgleish, 2011).

Page 50: Nanoencapsulation of Tea Catechins in Casein Micelles

41

Figure 3.2 Stern-Volmer plot for skim milk ( ) and caseins (♦) re-diluted in permeate.

Values are the average of three independent experiments.

0.0

0.5

1.0

1.5

2.0

2.5

3.0

0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14

Fo

/F

EGCG concentration (mg/mg protein)

Milk

Caseins

Page 51: Nanoencapsulation of Tea Catechins in Casein Micelles

42

The values of the Stern-Volmer quenching constant of the accessible fraction (Ka) and the

accessible fraction to the quencher (fa) were calculated using a modified Stern Volmer equation

(see Eq. 3.2). For skim milk the calculated fa and Ka are 0.80± 0.04 and 1.8 (± 0. 1) × 104 M

-1

respectively and for casein samples fa and Ka are 0.77± 0.06 and 1.3 (± 0. 2) × 104 M

-1,

respectively. While K values obtained for pure fractions of α-casein and β-casein with EGCG

were found to be 7.4 (±0.4) × 104 M

-1 for α-casein and 1.59 (±0.2) × 10

4 M

1 for β-casein , with

fa of 1.5 for both purified caseins (Hasni et al., 2011). The lower values obtained in our results

for milk and casein micelles compared to the purified α-casein and β-casein, show that while in

the molecular caseins there is access of EGCG, in the case of casein micelles, there is not full

access of EGCG to all the binding sites available in the molecular state.

The results clearly demonstrated that there were no significant differences in the binding values

of skim milk and caseins (P> 0.05), suggesting that the main sites of tryptophan quenching are

within the casein micelles. The slightly higher value of Ka for skim milk compared to the casein

micelles is probably caused by the presence of whey proteins.

Page 52: Nanoencapsulation of Tea Catechins in Casein Micelles

43

3.4.2 Analysis of tea polyphenols by HPLC

To determine the binding of tea catechins to casein micelles in a wider range of concentration,

binding isotherms were obtained by analyzing the amount of EGCG recovered in the serum

phase of the mixtures after separation of the colloidal casein micellar phase by centrifugation.

Figure 3.3 illustrates the amount of catechins recovered in the colloidal phase as a function of

ligand per protein present in skim milk. It is important to note that whey proteins remained in

the serum phase, hence, any EGCG interacting with the whey was included in the fraction

recovered in the supernatant and subtracted from the total EGCG, allowing for an estimate of the

amount associated with the casein micelles. At low concentrations nearly all of the tea catechins

added to the milk (around 95±2%) were recovered with the casein micelles fraction (figure 3.3).

The high amounts of EGCG in the colloidal phase (pellet) of milk clearly indicated that

recoveries of the tea catechins in the serum phase of complex mixtures cannot be used as an

index of polyphenol bioavailability. As shown in Figure 3.3 the majority of the tea catechin was

recovered with the casein micelles up to a concentration of 0.06 mg/mg protein in milk, after

which the amount of unbound polyphenol started to increase. It is assumed that at this

concentration of EGCG per milk proteins, casein micelles are saturated with the tea catechin,

meaning there are no more binding sites on the caseins for interactions with EGCG. Therefore,

with increasing the concentration of EGCG, the recovery in the soluble phase (serum phase) was

also increased. Such binding isotherm for EGCG and casein micelles has never been reported

before.

Page 53: Nanoencapsulation of Tea Catechins in Casein Micelles

44

Figure 3.3 % Tea catechin recovered in the colloidal phase (casein micelles) of skim milk, as a

function of the initial concentration of EGCG present in milk. Values are the average of 3

independent experiments (i.e., 3 milk batches) and bars represent the standard deviation.

0

10

20

30

40

50

60

70

80

90

100

0.00 0.01 0.02 0.03 0.06 0.13 0.26 0.51

% T

ea c

atec

hin

Initial EGCG concentration (mg/ mg protein)

Page 54: Nanoencapsulation of Tea Catechins in Casein Micelles

45

Using the Scatchard equation (equation 3.3) the dissociation constant (Kd) between EGCG and

casein proteins was determined (figure 3.4). The ratio of bound/free EGCG continued to

decrease as a function of bound EGCG, and three dissociation constants were derived from the

slopes. The dissociation constants calculated were 2.7× 10-4

M , 6.6 ×10-4

M and 6.5×10

-3 M for

the first, second and third portion of the curve, respectively. These results indicated different

types of binding for EGCG with the casein micelles, decreasing in specificity and with a larger

value of the dissociation constant. Kd indicates the strength of binding between the ligand and

the protein in terms of how easy it is to separate the complex, therefore, the smallest Kd

calculated for the first type of binding (equivalent to low EGCG concentrations to EGCG

concentrations at the saturation point of the caseins, up to 0.06 mg EGCG per mg milk protein-

see figure 3.3), show the strongest binding compared to EGCG concentrations after saturation

point of the caseins. It is important to note that another study using small-angle X-ray scattering,

showed a similar concentration of EGCG for saturating casein micelles as indicated by our work

(Shukla et al., 2009).

The combination of the fluorescence and RP-HPLC techniques used in this study clearly

demonstrated that EGCG strongly interacts with the casein micelles, and a substantial amount of

EGCG is associated with the colloidal phase of the casein micelles. The RP-HPLC results

showed, that at approximately 0.06 mg EGCG per mg milk protein the casein micelles are

saturated however the fluorescence results indicated that at around 0.04 mg per mg milk proteins

the quenching reaches a plateau. These results indicate that up to approximately 0.04 mg per mg

milk proteins, EGCG is mainly penetrated within the hydrophobic core of the casein micelles

and thereafter, up to saturation of casein micelles (0.06 mg EGCG per mg milk protein ), is

mostly located at the surface of the casein micelles.

Page 55: Nanoencapsulation of Tea Catechins in Casein Micelles

46

Bound/Total protein

0 2 4 6 8

(Bo

un

d/T

ota

l p

rote

in)/

Fre

e (M

-1)

0

2000

4000

6000

8000

10000

12000

14000

Figure 3.4 Scatchard plot for EGCG-milk caseins (see eq 3.3). The bound was calculated as the

amount of EGCG (M) that was within the casein micelle and free EGCG was calculated as the

amount of EGCG (M) that was quantified in the soluble whey proteins by RP-HPLC (see figure

3.3).

Page 56: Nanoencapsulation of Tea Catechins in Casein Micelles

47

3.4.3 The effect of EGCG on rennet induced gelation of casein micelles

To determine if the presence of a complex with EGCG affects the processing functionality of the

casein micelles, their ability to form gels with rennet was tested. Rennet induced gelation was

selected as it is caused by a surface destabilization of the casein micelles via a specific hydrolysis

of the stabilizing layer of k-casein on the micelles. Rennet gelation is an enzymatic coagulation

particularly relevant in dairy processing, as it is the first step of the cheese-making process. It is

composed of two stages, which partially overlap each other. During the first stage, the enzymatic

action of the rennet (chymosin) destabilizes the casein micelles, while in the second stage the

k-casein depleted micelles aggregate, in the presence of calcium, and form a protein network

(Walstra, 1990).

Figure 3.5 shows the develelopment of the structure during rennet induced gelation for milk with

and without EGCG. In particular, the value of G’ (storage modulus), as a function of renneting

time for milk samples containing different amounts of tea catechins. Concentrations were

selected from the results of Figure 3.3, as low, medium and saturation concentrations. When the

tea catechin was added to skim milk at very low amounts (0.002 mg EGCG/mg protein, i.e.

much lower than the saturation of the casein micelles, see Figure 3.3) the gel time of the sample

slightly increased compared with the control skim milk sample (without the addition of tea

catechin). As the concentration of tea catechin in skim milk increased, there was a significant

increase in the gelling time. At concentrations higher than the saturation point, gelation was not

observed. The formation of catechin-casein micelles complexes not only delayed the gelling

point of milk, but they also affected the structure formation, as could be noted by the

development of the lower elastic modulus (G′) over time in the tea- milk mixtures compared to

Page 57: Nanoencapsulation of Tea Catechins in Casein Micelles

48

control milk. Table 3.1 summarizes the difference in the gelation parameters as a function of

differnet concentration of EGCG added.

Time (min)

0 30 60 90 120 150 180 210 240

G' (

Pa)

0

20

40

60

80

Figure 3.5 Rennet coagulation process monitored by rheometer of the four samples: control

skim milk (●), skim milk with 0.002 mg EGCG/ mg milk protein (▲) skim milk with 0.02 mg

EGCG/ mg milk protein (■), skim milk with 0.08 mg EGCG/ mg milk protein (saturation point

of caseins).

Page 58: Nanoencapsulation of Tea Catechins in Casein Micelles

49

Table 3.1. Gelling time and Gel stiffness of different tea milk samples

Gelling time

(min)

Gel stiffness (Pa)

After 80 min of

gelling

Milk (ctrl) 35 ± 5 65.9±7.2

0.002 mg EGCG/ mg milk

protein

41 ±2 36.4±2.6

0.02 mg EGCG/ mg milk protein 65±4 21.3±3

0.08 mg EGCG/ mg milk protein 170±7 0.17±0.04

Values are the means of three replicates.

The results shown in Table 3.1 clearly demonstrated that the presence of tea catechins strongly

affected the gelling behavior of the casein micelles. To further elucidate the origin of the impact

of these interactions, the release of casein macro peptide (CMP) from the surface of the casein

micelles was measured. This should allow determining if the actual enzymatic hydrolysis

occurred at the same rate in all the samples. Chymosin is a proteolytic enzyme active on

κ-casein, specifically on the Phe105-Met106 peptide bond. Following enzymatic cleavage,

para-κ-casein remains on the micelle and the soluble, hydrophilic moiety, casein macro peptide

(CMP) is released into the serum. Removal of CMP from κ-casein reduces steric and

electrostatic repulsion between micelles and increases the attractive forces between the micelles,

leading to aggregation (Anema et al., 1996). Once a sufficient amount (~ 85-90%) of CMP is

released, rapid aggregation of the micelles and a sharp increase in the values of the elastic

modulus are detected (Sandra et al., 2007; Gaygadzhiev et al., 2009).

Figure 3.6 depicts the effect of different concentrations of tea catechin (EGCG) on the release of

CMP from the surface of the casein micelles. The presence of EGCG led to a slower removal of

CMP as measured by the time taken to reach 90% release (as all samples contained the same

Page 59: Nanoencapsulation of Tea Catechins in Casein Micelles

50

concentration of casein micelles, the maximum CMP released from the control skim milk was

taken as 100%). The control sample reached 90% of CMP released in about 33 min, while skim

milk with an EGCG concentration lower than the saturation point of the caseins reached the

same point after 60 min. Whereas, the EGCG concentration at saturation point of the caseins did

not reach the 90% of CMP during the time frame of the work.

The fact that even small amounts of EGCG inhibit aggregation and that there is a concentration

dependence, may suggest that the tea polyphenols are affecting the structure of the micelle and

the accessibility of the enzyme to substrate, as full inhibition occurs at concentrations close to the

saturation point of the caseins. Rennet containing the enzyme chymosin, specifically cleaves the

Phe105-Met106 bond of κ-casein. However, it has been shown that other residues near the cleaved

bond are also involved in the hydrolytic reaction of rennet on κ-casein.The peptide

corresponding to the residues 98-111 of κ -casein (His-Pro-His-Pro-His-Leu-Ser-Phe-Met-Ala-

Ile-Pro-Pro-Lys) was found to provide complete requirement for hydrolysis. Also, as proline is a

cyclic amino acid, it causes bends in the polypeptide chain of κ -casein therefore, facilitating

accessibility of the substrate to the shallow cleft of the enzyme (Chitpinityol et al., 1998). It is

proposed by the results found in this study that due to the fact that catechins have a strong

attraction to proline residues, EGCG is probably binding with the four proline residues on each

side of the cleaved bond therefore reducing accessibility of the enzyme to the substrate.

However another possibility of the delay in reaching critical CMP (~90%) for the tea-milk

samples may be due to the fact that catechins are know to have metal chelating properties

(Kumamoto et al.,2001; Green et al., 2007). It has been shown that once κ-casein is cut off from

the surface and a decrease is caused in steric repulsion, the casein micelles approach one another

closely. At this point hydrophobic interactions that cause bonding between the particles play an

Page 60: Nanoencapsulation of Tea Catechins in Casein Micelles

51

important role, as does the amount of ionic calcium. Also, a partial loss of colloidal calcium may

reduce the charge interactions, causing the formation of a weaker, more flexible casein network

(Dalgleish, 2012).

Page 61: Nanoencapsulation of Tea Catechins in Casein Micelles

52

Figure 3.6 % CMP-release versus time after addition of rennet. Control skim milk (■), 0.02 mg

EGCG/ mg milk protein (●) 0.08 mg EGCG/ mg milk protein (▲).

0

20

40

60

80

100

0 20 40 60 80 100 120

% C

MP

Rel

ease

Time (min)

Page 62: Nanoencapsulation of Tea Catechins in Casein Micelles

53

To further test the mentioned hypotheses, another assembly of caseins was studied using sodium

caseinate. Sodium caseinate is made by the disruption of the casein micelle as the calcium

phosphate content is removed, therefore, caseins are no longer in the micellar form and now are

in monomeric forms. Sodium caseinate without the addition of EGCG was used as control as

well as another sample of sodium caseinate with the addition of EGCG at saturation point of the

casein micelles and the CMP release of these samples was analysed. As seen in Figure 3.7, the

results were similar as what was seen in the case of the CMP release of the skim milk and EGCG

samples (Figure 3.6), where the presence of EGCG led to a slower removal of CMP compared to

the control. These results not only confirmed that EGCG strongly interact with caseins but they

also inhibit the accessibility of the enzyme to the caseins, no matter what form the caseins may

be.

Page 63: Nanoencapsulation of Tea Catechins in Casein Micelles

54

Figure 3.7 % CMP-release versus time after addition of rennet. Control sodium caseinate (■),

and 0.08 mg EGCG/ mg sodium caseinate (▲).

0

10

20

30

40

50

60

70

80

90

100

110

0 20 40 60 80 100 120 140

% C

MP

rel

ease

Time (min)

Page 64: Nanoencapsulation of Tea Catechins in Casein Micelles

55

3.5 Conclusions

Numerous studies have shown that strong interactions take place between tea catechins and milk

proteins, which we also confirmed. However, most studies have been on isolated caseins and not

caseins in milk as well as the effect of polyphenols on functional properties of milk proteins is

not fully understood. In this study it was shown that the binding of EGCG to caseins in skim

milk reaches a saturation point where afterwards EGCG is then seen in whey proteins. As the

concentration of EGCG increases and saturation of caseins is reached, a more significant effect is

seen on the renneting properties of milk. It was shown with fluorescence quenching of

tryptophan that to a certain degree, EGCG is incorporated into the porous casein micelles and

thereafter may be on the surface of the micelles resulting in less accessibility of the enzyme to

the substrate (caseins) or may be due to chelating of calcium resulting in a weaker casein

network. This study clearly shows that demonstration of successful and effective incorporation

of bioactives into milk matrix is essential before using milk as a platform for delivery of

polyphenols.

Page 65: Nanoencapsulation of Tea Catechins in Casein Micelles

56

CHAPTER 4

Anti-proliferative activity of nanoencapsulated tea catechin in casein micelles,

using HT29 colon cancer cells

4.1. Abstract

The present study tested the ability of casein micelles to deliver biologically active

concentrations of polyphenols to colon cancer cells (HT-29). Epigallocatechin-gallate (EGCG),

the major catechin found in green tea, was employed as model molecule, as it has been shown to

have antiproliferative activity on colon cancer cells. Casein micelles in milk may be an ideal

platform for the delivery of bioactive molecules such as EGCG, as casein proteins have been

shown to have strong interactions with polyphenols. In the present work, it was hypothesized that

the strong binding of caseins with EGCG causes a change in the bioavailability of EGCG.

The cytotoxity and proliferation behaviour of HT-29 colon cancer cells when exposed to EGCG

was compared to that of EGCG nanoencapsulated in casein micelles, in their native form, in

skim milk. EGCG-casein micelle complexes decreased the proliferation of HT-29 cells,

demonstrating that bioavailability was not reduced by the nanoencapsulation. As casein micelles

may act as protective carriers for EGCG in foods, it was concluded that nanoencapsulation of tea

catechins in casein micelles may not diminish their anticancer activity compared to free tea

catechins.

Page 66: Nanoencapsulation of Tea Catechins in Casein Micelles

57

4.2. Introduction

Inrecent years, polyphenols have been recognized as important components in the diet, as their

consumption has increasingly been associated with the prevention of various chronic diseases

(Yang et al., 2009; Gonzalez et al., 2011; Yuan et al., 2011). The major sources of polyphenols

in the diet are vegetables, fruits, wine, coffee and tea (Manach et al., 2004). Tea (Camellia

sinensis), which is the second most consumed beverage in the world, is a rich source of

extractable polyphenols, more specifically of flavan-3-ols, commonly referred to as catechins.

The catechins in green tea represent up to 85% of the total polyphenols (Ferruzzi 2010). Recent

epidemiological and biological data have associated green tea consumption with a reduced risk

of cancer and diseases of the cardiovascular system. Major tea catechins include epicatechin,

epigallocatechin, epigallocatechin-gallate (EGCG), and epicatechin-gallate. EGCG is the major

catechin component in tea and it appears to be the most biologically active (Vaidyanathan et al.,

2003; Henning et al., 2004) (Figure 4.1). Human clinical intervention trials have demonstrated

positive links between green tea consumption and a reduced risk of cancer (Li et al., 2000;

Lambert et al., 2007a).

Scavenging of reactive oxygen and nitrogen species, chelating of redox active metals, inhibiting

cancer-related transcription factors, and inhibiting oxidative enzymes are amongst the reactions

involving catechins which may result in these beneficial health effects (Green et al., 2007).

Bioavailability is usually defined as the fraction of the bioactive constituents (or the metabolites)

that have been absorbed by the body and can exert their biological actions at specific target

tissue sites (Ferruzzi, 2010; Visioli et al., 2011). Bioavailability appears to differ greatly among

polyphenols, and for tea catechins it is generally considered to be poor (Ferruzzi, 2010).

Page 67: Nanoencapsulation of Tea Catechins in Casein Micelles

58

Figure 4.1 Structure of (-) epi-gallocatechin gallate

Page 68: Nanoencapsulation of Tea Catechins in Casein Micelles

59

Several factors may play a part in limiting the bioavailability of polyphenols, including

molecular structure, amount consumed, intestinal absorption and the degree of bioconversion in

the gut (Papadopoulou et al., 2004; Yang et al., 2008). In addition, the interactions between the

bioactive molecules and the food matrix are considered critical factors limiting or enhancing

bioavailability (Saura-Calixto et al., 2007).

Milk is known to be a multi-component and a nutritious source of compounds beneficial to

growth and health in children as well as adults. Studies have shown that both whey proteins and

caseins act as delivery vehicles for bioactive compunds (Livney 2010). Nanoencapsulation of

-lactoglobulin has been shown to protect EGCG against oxidative degradation (Shpigelman et

al., 2010). Bovine serum albumin has often been referred to as a carrier of chemotherapeutic

drugs in the circulatory system (Livney 2010). Recent studies have also demonstrated the

association of bioactive molecules such as resveratrol, vitamin D and curcumin (a polyphenol)

with casein micelles (Semu et al., 2007; Sahu et al., 2008; Rahimi Yazdi et al., 2012).

The encapsulation of vitamin D in casein micelles provided partial protection against UV-

induced degradation of vitamin D (Semu et al., 2007).

Polyphenols are known to have strong binding affinities with proteins rich in proline, and

proteins with an open and flexible structure such as caseins (Benick et al., 2002; Papadopoulou

et al., 2004). Studies have shown that the highest extent of binding of catechins with isolated

forms of caseins, was noted for β-casein, followed by αs-caseins and κ-casein (Luck et al., 1994;

Kartosova et al., 2007). In terms of bioaccessibility, researchers have shown that curcumin

encapsulated within casein micelles can be efficiently delivered to Hela tumour cell culture

(Sahu et al., 2008). However, the bioaccessibility and bioavailability of polyphenols in milk

mixtures is still a source of debate. Some studies have suggested that the presence of protein

Page 69: Nanoencapsulation of Tea Catechins in Casein Micelles

60

masks the antioxidant capacity of tea catechins (Arts et al., 2002; Kartosova et al., 2007).

This masking effect was also reported in human clinical trials comparing tea drank with or

without milk (Langley-Evans 2000). In contrast, other reports showed no significant effect of

milk on the plasma antioxidant capacity of tea polyphenols (Leenon et al., 2000 Keogh et al.,

2007; Kyle 2007).

The discrepancies may derive from the differences designs and methodologies of the studies. It is

also important to note that stability of catechins during storage and digestion may be poor, but

when present in complex matrices such as bovine milk, soymilk or rice milk the stability may

increase (Green et al., 2007; Ferruzzi 2010).

The objective of this work was to test the effect of nanoencapsulation of EGCG in casein

micelles on its anticancer activity. Casein micelles can load a substantial amount of catechins

within their structure, acting as carriers for the catechin and therefore may affect their

bioavailability. The antiproliferative activities of EGCG on HT 29 colon cancer cells was tested

with free EGCG and EGCG in various milk matrices, namely, milk serum, milk whey and when

encapsulated in casein micelles.

4.3. Materials and Methods

4.3.1 Milk samples preparation

Commercial skim milk was purchased from local super market (Parmalat Canada Inc., Toronto,

Canada).

Separation of the soluble proteins (whey) from the casein micelles was accomplished by

centrifugation (36,000 ×g) at 20°C for for 45 min (Beckman Coulter, Canada, Inc, with rotor

type 70.1 Ti, Mississauga, On, Canada) as previously described (del Angel et al., 2006).

Page 70: Nanoencapsulation of Tea Catechins in Casein Micelles

61

Milk permeate was prepared using a laboratory scale ultrafiltration cartridge (Millipore

CDUF001LG; Fisher Scientific), with a nominal cut-off of 10,000 Da and a nominal area of

1 ft2 (0.095 m

2).

A polyphenol extract from green tea (containing minimum 90% EGCG) was obtained from DSM

Nutritional Products Inc. (Montreal, Canada). Different concentrations of EGCG were obtained

from a stock solution of EGCG in water (20 mg/ml water). The amount of EGCG encapsulated

within the casein micelles was calculated using HPLC, as described in detail in chapter 3.

4.3.2. Cell viability assay

Human colorectal cancer cells (HT-29) were obtained from the American Type Culture

Collection (ATCC- HTB-38; Virginia, USA). HT-29 cells were maintained in Dulbecco's

Modified Eagle Medium (DMEM) supplemented with 10% fetal bovine serum, 2.2 g/l sodium

bicarbonate, 1% penicillin/ streptomycin and 1% L-glutamine (Sigma-Aldrich Canada Ltd.

Oakville, Ontario, Canada), and incubated at 37°C in 95% humidity and 5% CO2 is air (Forma

Series II Water-jacketed CO2 Incubator, model 3110 , Forma Scientific, California, USA).

Trypan blue dye exclusion is typically used to estimate cell viability. This large molecular

weight dye will be excluded unless the cell membrane has been damaged or its permeability

compromised in some way. Healthy cells exclude the dye while dead or dying cells take the dye

up. Therefore, this procedure is used to estimate the percentage of viable cells under specific cell

culture conditions (Ehrich et al., 2000).

HT-29 colon cancer cells were seeded and allowed to grow for 48 h, afterwhich different

concentrations of tea catechins, alone and in combination with milk, were added to the cells and

incubated for a further 24 h. Each sample was diluted in media at a ratio of 1:5.6 (sample:media)

Page 71: Nanoencapsulation of Tea Catechins in Casein Micelles

62

(v/v). Control samples were also tested, either with media only (for free EGCG) or with milk,

with no EGCG (for tea-milk samples). After a continuous incubation of 24 h, trypsinized cells

were suspended in culture medium, counted using a Neubauer haemocytometer and viability was

measured using Trypan blue dye exclusion (see Ehrich et al., 2000 for details).

Cell viability was expressed as the ratio between the viable cells for each treatment (i.e. EGCG-

milk; EGCG free) and the number of viable cells present in the corresponding control.

Cell viability tests were used in combination with the sulforhodamine B cytotoxicity assay (see

below), to better determine changes in cell proliferation.

4.3.3. Cell proliferation assay

The anti-proliferation effect of EGCG on HT-29 cancer cells was determined using the

Sulforhodamine B (SRB) dye binding assay (see Shekhan et al., 1990 for details). SRB is a

colorimetric assay for the quantification of the total protein synthesis rate of cells in response to

different concentrations of tea catechin. SRB is an anionic dye that binds to synthesized proteins.

The fixed dye, measured photometrically after solubilization, correlates with protein synthesis

and with cell proliferation (Vichai et al., 2006).

HT-29 colon cancer cells were seeded on 96 well plates and allowed to adhere for 24 h.

Afterwards, different concentrations of tea catechin (EGCG) alone and in combination with milk,

were added to the cells and incubated for for an additional 24 h. Each sample was diluted in

media at a ratio of 1:5.6 (sample:media) (v/v). Control samples were also tested, either with

media only (for free EGCG) or with milk, with no EGCG (for tea-milk samples). After a

continuous incubation of 24 h, TCA-fixed cells were stained with SRB dye. At the end of the

staining period, SRB was removed and cultures were quickly rinsed with 1% acetic acid to

Page 72: Nanoencapsulation of Tea Catechins in Casein Micelles

63

remove unbound dye. After drying, the bound dye was solubilized with 10 mM Tris base (pH

10.5).The adbsorbance of plates was measured using an automated 96-well plate reader (Multi

detetctor Microplate Reader, Biotek Synergy HT Model, Vermont , USA) at a wavelength of 570

nm. The results are reported as % proliferation, which is the OD of each fraction at 570 nm

divided by the OD of its control multiplied (at 570 nm)by 100 (Vichai et al., 2006).

4.3.4. Statistical analysis

The results presented are the average of at least three independent experiments. Each experiment

had at least 5 sub samples. The mean values, standard deviations, and statistical differences were

evaluated using analysis of variance (ANOVA). All statistics were performed by using R

software (R Project for Statistical Computing version 2.15.1).

4.4. Results and Discussion

Various studies have observed that tea polyphenols inhibit tumour formation and proliferation in

different cell lines, as well as in animal models (Yang et al., 2006; Yang et al.,2009).

The inhibitory activity can include decreasing cell proliferation, increasing apoptosis and/or

suppressing angiogenesis (Araújo et al., 2011). This study focused on the effect of the matrix on

the antiproliferative properties of EGCG on colon cancer cells. Figure 4.2 shows the EGCG

encapsulated in the native casein micelles as a function of the amount of EGCG added in skim

milk. At a concentration of 2.5 mg EGGC/ml of milk, most of the tea catechin is associated with

the colloidal phase of the micelles (see also Chapter 3). For the cell cytotoxicity and proliferation

Page 73: Nanoencapsulation of Tea Catechins in Casein Micelles

64

assays, EGCG concentrations below saturation (0.2 and 0.5 mg/ml), at saturation (2.5 mg/ml)

and above saturation (5 mg/ml) were used.

The study of the effectiveness of EGCG on the inhibition of colo-rectal carcinogenesis is of great

interest (Gonzales et al., 2010). The majority of the studies related to the anti-proliferative

properties of polyphenols on colon cancer cells has so far been based on the effect of individual

polyphenols or crude polyphenol extracts (Hong et al., 2002; Chen et al., 2003; Na et al., 2006;

Bazinet et. al., 2010), and much less has been reported on polyphenols bound to proteins such as

in milk. However, gut epithelial cells are more likely to be exposed to complex food matrices,

and this, as previously discussed, may affect the bioavailability of polyphenols in the gut (Saura-

Calixto et al., 2007).

Page 74: Nanoencapsulation of Tea Catechins in Casein Micelles

65

Figure 4.2 Amount of EGCG associated with the casein micelles in skim milk as a function of

the initial concentration of EGCG in milk. Results are the average of three independent

experiments and bars represent the standard deviation.

0

10

20

30

40

50

60

70

80

90

100

0.15 0.2 0.5 0.75 1 2 2.5 3.3 5

% E

nca

psu

late

d E

GC

G in

cas

ein

s

Initial concentration of EGCG in milk (mg/ml)

Page 75: Nanoencapsulation of Tea Catechins in Casein Micelles

66

Figure 4.3 illustrates the difference in the viability of HT-29 cancer cells after 24h of incubation

with different concentrations of EGCG, either dispersed in water or in nanoencapsulated in the

casein micelles in skim milk. The final concentrations of EGCG in the wells are indicated in the

figure, and were 0.03, 0.075 mg/ml 0.15 mg/ml, 0.4 mg/ml and 0.75 mg/ml, corresponding 0.2,

0.5, 1, 2.5 and 5 mg/ml in the original milk. The viability of the cells exposed to unencapsulated

EGCG after 24 h of incubation (solid bars, Figure 4.3) decreased as a function of the amount of

EGCG present. This result was in full agreement with other studies using similar concentrations

of EGCG (Salucci et al., 2002; Bazinet et al., 2010). When the cells were exposed to the mixture

of EGCG-milk samples (Figure 4.3, empty bars), the same trend was observed, a decrease in

viability with increasing EGCG.

While at concentrations, < 0.15 mg/ml there were no differences in viability between

unencapsulated and encapsulated treatments, at higher concentrations, above saturation, viability

was significantly higher than for samples with EGCG in water. The lower decrease in viability

in the presence of milk is most probably due to differences in the composition of the media,

containing milk components.

Page 76: Nanoencapsulation of Tea Catechins in Casein Micelles

67

EGCG concentration (mg/ml)

Ctrl 0.03 0.075 0.15 0.4 0.75

% V

iable

cel

ls

0

20

40

60

80

100

Figure 4.3 Viable cells after 24h exposure to isolated EGCG (solid) and EGCG encapsulated in

casein micelles. All samples were diluted in media 1:5.6 (sample: media). The control used for

isolated EGCG was media without EGCG and the control used for of EGCG-milk fractions was

milk without EGCG. Values are the average of at least 3 replicates, bars represent standard

deviations. Different letters within each concentration of isolated and encapsulated EGCG,

indicate statistical significant differences at p<0.05.

a a

a a

a

a

a

b

b

b

Page 77: Nanoencapsulation of Tea Catechins in Casein Micelles

68

To better identify the effect of the medium, the matrix and the concentrations dependence of

EGCG encapsulated in casein micelles, the anti-proliferative effect of EGCG on HT-29 cancer

cells was studied. The results are summarized in Figure 4.4. When EGCG was non encapsulated

(solid bars Figure 4.4), even at very low concentrations (i.e. ≤ 0.02 mg/ml media corresponding

≤ 0.1 mg/ml water) it was able to significantly decrease the proliferation of the cancer cells.

At higher concentrations (i.e. ≥ 0.03 mg/ml media corresponding ≥ 0.2 mg/ml milk), the

decrease in proliferation reached a plateau.

The same trend in the proliferation of the HT-29 colon cancer cells was seen when they were

exposed to different concentrations of the EGCG-milk samples. Figure 4.4 compares the

proliferation of the cells incubated with EGCG dispersed in permeate, milk whey and skim milk.

In all cases, as the EGCG concentration increased, there was a decrease in the cancer cell

proliferation, eventually plateauing at high concentrations. Although at higher concentrations of

EGCG (i.e. ≥ 0.15 mg/ml media corresponding ≥ 1 mg EGCG /ml milk) no significant

differences were noted in the anti-proliferative activity of EGCG (p> 0.05), whether isolated or

present in a milk matrix, the bioefficacy of isolated EGCG seemed to be affected in the presence

of milk mixtures, as at low concentrations (<0.015 mg/ml) the extent of proliferation differed.

However, this could not be attributed to the encapsulation of EGCG in casein micelles (Figure

4.3 C), as the same effect was noted also for EGCG in milk serum (Figure 4.3 A) and in milk

whey (Figure 4.3 B). At the lower concentrations, one hypothesis is that the difference in

proliferation are due to a difference in the media composition, as milk serum, milk whey and

skim milk may contain growth factors not present in the water used to resuspend EGCG, nor in

the media used for cells.

Page 78: Nanoencapsulation of Tea Catechins in Casein Micelles

69

At about 0.03 mg/ml (corresponding to 0.2 mg/ml milk), a significantly higher proliferation was

noted for the cells treated with EGCG in skim milk, compared to EGCG in whey, permeate or

control. Hence, the effect of the interactions between tea catechin and caseins may still be

apparent since the EGCG-milk fractions reach a plateau with some delay compared to the

EGCG-whey and EGCG- permeate fractions. When comparing the estimated IC50 values, 0.01

mg/ml, 0.016 mg/ml, 0.018 mg/ml and 0.02 mg/ml for isolated EGCG, EGCG-whey, EGCG-

permeate and EGCG-milk (caseins), respectively, it could be hypothesized that this decrease in

the apparent bioefficacy may be caused by the complexes formed between EGCG and casein

micelles. Therefore, the strong interactions between EGCG and milk caseins, may affect the

ability of EGCG to inhibit the proliferation of the cancer cells.

Page 79: Nanoencapsulation of Tea Catechins in Casein Micelles

70

% P

roli

fera

tio

n

0

20

40

60

80

100 EGCG

EGCG in milk serum

% P

roli

fera

tio

n

0

20

40

60

80

100EGCGEGCG in milk whey

EGCG concentration (mg/ml)

ctrl 8e-3 0.015 0.03 0.08 0.15 0.4 0.8

% P

roli

fera

tio

n

0

20

40

60

80

100EGCG

EGCG in skim milk

Figure 4.4 Proliferation of HT-29 cells exposed to isolated EGCG (solid bars) or EGCG in milk

serum (permeate) (A); milk whey (B) or encapsulated with the casein micelles in skim milk (C).

Values are the average of at least three independent experiments. Bars represent standard

deviation. Different letters within each concentration of EGCG, indicate statistical significant

differences at p<0.05.

a a a a a a a a a a

b

b

a

a

a a a a a a a a a a

a

a

b

b

a a a a a a a

a

a

b

b

b

A

B

C

a

a a

Page 80: Nanoencapsulation of Tea Catechins in Casein Micelles

71

4.5. Conclusions

EGCG is one of the most extensively investigated chemopreventive phytochemicals. By

modulate signal transduction pathways involved in cell proliferation, transformation,

inflammation, apoptosis, metastasis and invasion, EGCG could potentially block multiple stages

of carcinogenesis (Na et. al, 2006). However, in order to exert their biological health benefits in

vivo, polyphenols must be bioavailable. This study showed that in spite of the strong binding of

EGCG to the casein micelles, the bioefficacy of EGCG was not diminished, as cell proliferation

results showed that the EGCG-milk complexes were still able to significantly decrease the

proliferation of HT-29 cancer cells, similar to isolated EGCG. It is important to note though, that

the present studies were carried out on fresh EGCG. Mixed systems may ensure better stability

to tea catechins with storage and during digestion, ultimately improving their bioefficacy.

Page 81: Nanoencapsulation of Tea Catechins in Casein Micelles

72

Chapter 5

The effect of nanoencapsulation of tea catechins in casein micelles on their

bioaccessibility studied using normal and cancerous rat colonic cells

5.1 Abstract

Numerous studies have demonstrated that tea catechins form complexes with milk proteins,

especially caseins. Much less work has been conducted to understand the metabolic conversions

of tea/milk complexes during gastro-duodenal digestion. The objective of this study was to

determine the significance of this association on the digestibility of the milk proteins and on the

bioaccessibility of the tea polyphenol epigalocatechin galate (EGCG). An in vitro digestion

model mimicking the gastric and duodenal phases of the human gastrointestinal tract was

employed to follow the fate of the milk proteins during digestion, and determine the bioefficacy

of EGCG free or encapsulated with the caseins. The samples, before and after digestion were

tested using two parallel colonic epithelial cell lines, a normal line (4D/WT) and its cancerous

transformed counterpart (D/v-src). EGCG digests showed a specific effect on proliferation of

cancer cells. In normal cells, neither free or encapsulated EGCG affected cell proliferation, at

concentrations <0.1 mg/ml. At higher concentrations, both free and encapsulated produced

similar decreases in proliferation. On the other hand, the bioefficacy on the cancer cell line

showed some differences at lower concentrations. The results demonstrated that regardless of the

extent of digestion of the nanoencapsulated EGCG, the bioefficacy of EGCG was not

diminished, confirming that casein micelles are an appropriate delivery system for polyphenols.

Page 82: Nanoencapsulation of Tea Catechins in Casein Micelles

73

5.2 Introduction

Numerous studies have attributed many of the beneficial health effects associated with green tea,

to the catechins (Li et al., 2001; Lambert et al., 2007a, Yang 2008; Shpigelman et al., 2010).

In recent years the possible cancer preventive activity of epigallocatechin-3-gallate (EGCG) has

been extensively studied using animal models and different cell lines from various organs and

species. The inhibitory activity of EGCG has been demonstrated during both the initiation and

progression stages of carcinogenesis (Vaidyanathan et. al., 2003; Chen et. al., 2007; Yang 2009).

Since broad cancer-preventive effects in cell lines and animal models have been observed, it is

assumed that multiple molecular mechanisms, rather than a single receptor or molecular target,

may be involved (Jung et al., 2001; Hsu et al., 2003; Hwang et. al., 2007). For EGCG, binding to

molecular targets and inhibition of enzyme activities have been reported, as well as broader

spectrum antioxidant activities (Yang et. al., 2006; Yang et al., 2009; Li et al., 2010). Tea

catechins potently induce apoptotic cell death in tumor cells, and cancer cells are more

susceptible to the inhibitory and chemotherapeutic effects compared to normal and non-

transformed cells (Yamamoto et al., 2003; Friedman, et al., 2007; Yang 2009).

However, to exert their biological properties, polyphenols must be bio-accessible to the target

tissue. The bioavailability of dietary polyphenols is limited by a number of factors such as the

amount consumed, physiochemical properties of the molecule, enzyme and microbial-mediated

biotransformation and active reflux, instability under digestive conditions, intestinal absorption

and food matrix interactions (Papadopoulou et al., 2004; Green et al., 2007; Ferruzzi et.al.,

2010).

It has been shown that tea catechins are stable in the gastric lumen, under conditions of low pH

caused by the gastric juices of the stomach. However, there are many factors that influence the

Page 83: Nanoencapsulation of Tea Catechins in Casein Micelles

74

extent and rate of absorption of ingested compounds by the small intestine (Spencer et al., 2003;

Visioli et al., 2011). Upon transfer from the stomach to the small intestine the pH rises to values

around neutral. It is well known that polyphenolic compounds such as catechins, oxidize in

neutral and alkaline pH environments. EGCG has been observed to rapidly oxidize, however, the

oxidation of EGCG resulted in the formation of dimerized products, which were observed to

possess greater superoxide radical scavenging activity than EGCG itself and had powerful iron

chelating properties (Spencer et al., 2003). However, most studies on the fate of polyphenols in

the digestive tract have been limited to specific types of polyphenols or polyphenols in

combination but not on the fate of polyphenols when incorporated in food matrices (Saura-

Calixto et. al., 2007).

Milk proteins are known to be natural carriers, as their structure facilitates the delivery of

essential micronutrients (calcium and phosphate), from mother to the neonate. Researchers have

suggested the potential of milk proteins, mostly in their isolated forms, to deliver nutrients such

as vitamins D, curcumin, minerals (Fe2+, Mg2+), carotenoids (β–carotene), polyphenols

(resveratrol, caffeine), cancer therapy drugs and even flavors or aromas (Semu et al, 2007; Sahu

et al., 2008; Livney 2010). A recent study has shown that casein micelles can take up a

substantial amount of EGCG (see Chapter 3), and that nanoencapsulation of EGCG with milk

proteins provides considerable protection to EGCG against oxidative degradation compared to

unprotected (non-encapsulated) EGCG (Shpigelman et al., 2010).

It has been previously demonstrated that the interactions taking place between the tea catechin

and the milk caseins in native casein micelles does not affect the antiproliferative activity of

EGCG using HT-29 colon cancer cells. However, studies on catechin-milk protein interactions

with respect to catechin availability after simulated digestion have not been reported.

Page 84: Nanoencapsulation of Tea Catechins in Casein Micelles

75

Furthermore, the complexes formed between EGCG and caseins may affect the digestion of the

proteins.

The effects that the tea catechins may have on digestive enzymes, more specifically proteases,

must be considered as well. Recent studies have suggested an inhibitory effect of some

polyphenols on proteases (Siewright et al., 2006; McDougall et al., 2005; Goncalves et. al.,

2011). Although inhibition of proteases by polyphenol components can occur in vitro, there is

still not enough evidence to show that inhibition of protease activity occurs in the digestive

system in vivo (McDougall et al., 2008), and the results are still a source of debate (Tagliazucchi

et al., 2005; McDougall et al., 2008).

In this study, we aim to examine the effects of EGCG encapsulation in casein micelles on normal

and cancerous rat colonic cell lines, both before and after in vitro digestion.

Page 85: Nanoencapsulation of Tea Catechins in Casein Micelles

76

5.3 Materials and Methods

5.3.1 Sample preparation

Commercial skim milk was purchased from a local supermarket (Parmalat Canada Inc., Toronto,

Canada). Green tea polyphenol extract (containing a minimum of 90% EGCG) was purchased

from DSM Nutritional Products INC (Montreal, Canada).

Different concentrations of EGCG were generated from a stock solution of EGCG in water (20

mg/ml water). The amount of EGCG encapsulated within the casein micelles was calculated

using HPLC, as described in detail in chapter 3). Figure 5.1 shows the percentage of

encapsulated EGCG at different concentrations of EGCG in milk. The EGCG concentration is

shown as mg EGCG per ml milk. These results showed that the casein micelles reach a

saturation point at around 2.5 mg EGCG per ml milk. Therefore, for cell cytotoxicity and

proliferation assays, concentrations of EGCG in milk at concentrations below saturation of

casein micelles, at saturation and above saturation were used.

Page 86: Nanoencapsulation of Tea Catechins in Casein Micelles

77

Figure 5.1 Amount of EGCG associated with the casein micelles in skim milk as a function of

the initial concentration of EGCG in milk. Results are the average of three independent

experiments and bars represent the standard deviation.

0

10

20

30

40

50

60

70

80

90

100

0.15 0.2 0.5 0.75 1 2 2.5 3.3 5

% E

nca

psu

late

d E

GC

G i

n c

asei

ns

Initial concentration of EGCG in milk (mg/ml)

Page 87: Nanoencapsulation of Tea Catechins in Casein Micelles

78

5.3.2 In vitro gastro-duodenal digestion

An in vitro digestion model mimicking the gastric and duodenal stage of human digestion was

applied on isolated EGCG, skim milk and EGCG-milk mixture (Malaki Nik et al., 2010).

Briefly, simulated gastric fluid (SGF) was added to an amber jar containing pre-incubated

samples at 37 ºC for 15 min. SGF was prepared by dissolving the desired amount of pepsin in

salt solution as previously reported in detail (Malaki Nik et al., 2010). The final concentration of

3.2 mg of pepsin per mL of mixture (samples plus SGF) and final pH of 2 were employed. The

mixture was then incubated at 37 ºC in a shakingwater bath (New Brunswick Scientific Co., Inc.,

NJ) at 250 rpm for 60 minutes.

Samples from the gastric phase were collected, and the reaction was stopped by raising the pH of

mixture to 6.5 using 1 M NaHCO3.

Immediately after the gastric phase, the mixture was treated with simulated bile fluids and

simulated duodenal fluids at a final pH of 6.5. The mixture was incubated at 37 ºC and 250 rpm

for 1 h, the final mixture contained 8 mM bile salts and 5 mg mL−1

pancreatin as previously

reported in detail (Malaki Nik et al., 2010).

To deactivate the trypsin enzyme activity, all samples were immediately cooled on ice and kept

frozen at -20◦C until further use. For the samples destined to cell culture studies, the samples

were diluted and stored in media containing 10% Fetal Bovine Serum (FBS) (1:6 digest: media

v/v) (Sigma-Aldrich Canada Ltd. Oakville, Ontario, Canada) and frozen until use.

Page 88: Nanoencapsulation of Tea Catechins in Casein Micelles

79

5.3.3 SDS-PAGE analysis

To determine if the encapsulation affected the digestion behaviour of the casein micelles during

the gastric phase, Sodium dodecyl sulphate polyacrylamide gel electrophoresis (SDS-PAGE)

was conducted on milk controls as well as milk containing EGCG at a concentration above

saturation, 5 mg/ml (see Chapter 3). The samples were withdrawn from the digestion mixture at

5, 10, 20, and 30 min during in vitro simulated gastric digestion. SDS-PAGE under reducing

conditions was carried out using 15% resolving gels (Malaki Nik et al., 2011). Aliquots of

digested samples of milk and EGCG-milk were mixed with electrophoresis sample buffer,

containing 200 mM Tris–HCl, 2% SDS, 40% glycerol, 2% β-mercaptoethanol, and 1%

bromophenol blue. The solutions were then heated at 95 ◦C for 5min, centrifuged at 10,000 g for

5min using an Eppendorf centrifuge (Brinkmann Instruments, New York, USA). The resulting

subnatant was loaded (10µL) onto the gels and run at a constant voltage of 200V.After each run,

gels were immediately stained using Coomassie blue R-250 for 30min and destained with a

solution of acetic acid: water: methanol (1:4.5:4.5) for 2 h with two changes, and then destained

overnight with a destaining solution containing 22.5% methanol and 5% acetic acid. SHARP JX-

330 scanner (Amersham Biosciences, Quebec) was employed to scan the gels.

Page 89: Nanoencapsulation of Tea Catechins in Casein Micelles

80

5.3.4 Cell proliferation assays

The cell lines used in this research were the immortalized non-tumorigenic cell line referred to as

FRC TEX CL 4D (4D/WT) and the derived line transformed FRC TEX CL D/v-src, established

by transfection of a v-src sequence into the 4D/WT parental line (Cha et al., 2005). The cell

lines were a generous gift from Dr. Susan E. Pories, New England Deaconess Hospital and

Harvard Medical School.

4D/WT and D/v-src cells were cultured in Dulbecco's modified Eagle's medium with 10% fetal

bovine serum (FBS), hydrocortisone (0.02 g/ml), insulin (0.25 g/ml), transferrin (0.12 g/ml),

polybrene (1 g/ml) and penicillin-streptomycin (50 U/ml) (Sigma-Aldrich Canada Ltd.

Oakville, Ontario). in a humidified environment at 37°C in the presence of 10% CO2 in air.

Cell proliferation assays were conducted using the sulforhodamine B test (SRB). SRB is a

colorimetric assay for the quantification of the total protein synthesis rate of cells in response to

different concentrations of tea catechin. SRB is an anionic dye that binds to synthesized proteins.

The fixed dye, measured photometrically after solubilization, correlates with total protein

synthesis rate and therefore with cell proliferation (Vichai et al., 2006).

Cell proliferation assays were conducted using the sulforhodamine B dye-binding assay (SRB)

(Shekhan et al.,1990). The SRB assay was used to quantifythe total protein in cells, in response

to different concentrations of tea catechin. The fixed dye, measured photometrically after

solubilization, correlates with total protein and correlates well with cell proliferation (Vichai et

al., 2006).

4D/WT and D/v-src cells were seeded on 96 well plates and allowed to adhere for 24h,

afterwhich different concentrations of tea catechin alone and in combination with milk were

added to the cells and incubated for a further 24 h. Before digestion, all samples were diluted in

Page 90: Nanoencapsulation of Tea Catechins in Casein Micelles

81

media 1:5.6 (sample: media). The digests of isolated EGCG and EGCG-milk were further diluted

to a final of 1:16 (sample: media) and delivered the the cells in 96-well plates. The control used

for EGCG digests was media without EGCG and the control used for of EGCG-milk digests was

digested milk.

After a continuous incubation of 24 h, TCA-fixed cells were stained with SRB dye. At the end

of the staining period, SRB was removed and cultures were quickly rinsed with 1% acetic acid to

remove unbound dye. After drying, the bound dye was solubilized with 10 mM Tris base (pH

10.5). The absorbance of plates was measured using an automated 96-well plate reader (Multi

detetctor Microplate Reader, Biotek Synergy HT Model, Vermont , USA) at a wavelength of 570

nm. The results are presented as percent control proliferation, which is calculated by dividing the

absorbance of each fraction at 570 nm by the absorbance of its control and multiplied by 100

(Vichai et al., 2006).

5.3.5 Statistical analysis

The results presented are the average of at least three independent experiments. Each experiment

had at least 5 sub samples. The mean values, standard deviations, and statistical differences were

evaluated using analysis of variance (ANOVA). All statistics were performed by using R

software (R Project for Statistical Computing version 2.15.1).

Page 91: Nanoencapsulation of Tea Catechins in Casein Micelles

82

5.4 Results and Discussion

5.4.1 SDS-PAGE of milk and tea/milk complexes during the gastric phase

It is becoming clear that tea polyphenols, such as gallated catechins, have poor bioavailability

and/or stability in vivo (Ferruzzi 2010). Therefore, a large proportion of ingested polyphenols

from tea is not taken up into the circulation and passes through the gastrointestinal tract (GIT)

unabsorbed. Therefore the potential health benefits of catechins will be critically dependent on

their metabolic handling and delivery within the GIT.

The SDS-PAGE results of milk samples and milk-EGCG samples withdrawn during the gastric

phase are shown in Figure 5.1. The electrophoretic analysis of milk during simulated gastric

digestion demonstrated an extensive proteolysis of the caseins after 5 min of incubation. Only a

faint band was recovered for undigested β-lactoglobulin.

However, the objective of the study was to observe the effect of EGCG-casein complexes on

casein digestion which the SDS-PAGE results of milk proteins, whether complexed with EGCG

or alone (figure 5.2) show no differences in the digestion patterns of the milk proteins.

These results indicate that the complex formation between EGCG and milk caseins does not

make the caseins less susceptible to proteolysis. Although, it has been suggested by some studies

that polyphenols may inhibit protease activity of some gastric enzymes, such a situation was not

seen in our results (Siewright et al., 2006; McDougall et al., 2005; Goncalves et. al., 2011).

Page 92: Nanoencapsulation of Tea Catechins in Casein Micelles

83

Figure 5.2 SDS electrophoresis of milk digests at 5, 10, 20 and30 minutes of gastric phase (a,b,c

and d) and EGCG-milk digests at 5, 10, 20 and 30 minutes of gastric phase (e,f,g and h) .

Page 93: Nanoencapsulation of Tea Catechins in Casein Micelles

84

Our results are in accordance with the results obtained by van der Burg-Koorevaar et. al., 2011

when black tea –milk digestates were examined by SDS-PAGE.

Caseins, having a flexible structure, are sensitive to proteolysis; purified bovine caseins are

rapidly cleaved by digestive proteases during in vitro digestion in various models (Dupont et. al.,

2010). The extensive proteolysis of the caseins after 5 minutes in the skim milk samples may be

due to an excessive amount of enzyme, such as pepsin, used during the gastric process. It has

been shown that heat treatments of milk may cause protein denaturation which alters the

separation of proteins in SDS-PAGE therefore this may be the reason for the extensive

proteolysis of the caseins in our study, as the skim milk utilized was pasteurized (Dupont et. al.,

2010). Studies have noted that the rate of hydrolysis differs between caseins and whey proteins

in bovine milk, and that β-lactoglobulin is relatively highly resistant to digestion in bovine milk

samples (Inglingstad et. al., 2010 ; Dupont et. al., 2010 ; Picariello et. al., 2010).

Page 94: Nanoencapsulation of Tea Catechins in Casein Micelles

85

5.4.2 Proliferation of normal (4D) and tumorous (v-src) rat colonic cells exposed to EGCG

and EGCG -milk fractions before and after invitro digestion

As green tea catechins are very unstable in neutral and alkaline solutions and are known to have

poor absorption and therefore low bioavailability, using milk as a natural nano carrier may be

beneficial. However, the impact of milk formulation on the bioaccessibility and bioavailability of

EGCG must be considered. By studying the affect of EGCG, both in isolation and complexed

with milk, on cell proliferation, we can get a better understanding of the bioaccessability of

EGCG.

For this assay, the EGCG was either dissolved in water or in skim milk, hence associated with

the casein micelles (see Chapter 3). The concentrations used for EGCG whether in water

(isolated form) or in milk (encapsulated with caseins) were derived from figure 5.1; lower

concentrations of EGCG before saturation of casein micelles (i.e. 0.2 mg/ml, 0.5 mg/ml, 1

mg/ml), a concentration of EGCG at saturation of casein micelles (2.5 mg/ml) and a

concentration of EGCG above saturation (5 mg/ml).

However, in order to expose the cell line to the EGCG and EGCG-milk fractions, further dilution

with media took place on the samples before digestion. Each fraction was diluted in media with

a ratio of 1: 5.6 (fraction: media (v/v)). Therefore, the final concentrations of EGCG in media

that were delivereded to cells correspond to concentrations of EGCG before saturation of casein

micelles (i.e. ≤0.19 mg/ml media), EGCG at saturation (i.e. 0.38 mg/ml media) and after

saturation of casein micelles (0.75 mg/ ml media). Also, it is important to note that the control

for isolated EGCG fractions was media, however, for EGCG-milk fractions, the control was

milk.

Figure 5.3 A and B show the results for normal (4D) and tumorigenic (v-src) rat colonic cells

when exposed to increasing concentrations of isolated EGCG and encapsulated EGCG-milk

Page 95: Nanoencapsulation of Tea Catechins in Casein Micelles

86

before applying invitro digestion. The results showed in terms of cell proliferation of tumor cells,

a significant difference was seen at lower concentrations of EGCG (P < 0.05), as a lower effect

on inhibiting the proliferation of tumorigenic (v-src) cells was seen meaning that the bioefficacy

of EGCG at lower concentrations was affected by skim milk. However, at higher concentrations

of EGCG (≥ 0.09 mg/ml) similar effects were seen whether EGCG was isolated or bound to

milk. These results indicate that the strong binding of EGCG and milk casein not affect the bio

efficacy of the tea catechin, nor does it possibly affect the mechanism(s) responsible for the

inhibitory activity on tumor cells.

In the case of normal cells exposed to increasing concentrations of isolated EGCG and

encapsulated EGCG-milk before applying in vitro digestion (Figure 5.3 b), although a significant

difference was not seen between EGCG-milk fractions and fractions with isolated EGCG

(P> 0.05), the higher concentrations of EGCG were toxic for the normal cells (≥ 0.4 mg/ml).

However, the anti proliferative effect of EGCG on normal cells was much less than the effect on

tumor cells.The half maximal (50%) inhibitory concentration (IC50) of isolated EGCG on tumor

cells was 0.012 mg/ml and for EGCG-Mik was 0.02 mg/ml. For both isolated EGCG and

EGCG-milk fractions, the IC50 value was 0.19 mg/ml for normal rat colonic cells.

Page 96: Nanoencapsulation of Tea Catechins in Casein Micelles

87

EGCG concentration (mg/ml)

ctr 6e-3 0.01 0.02 0.05 0.09 0.2 0.4 0.8

% P

roli

fera

tion

0

20

40

60

80

100

120

140

Tumor-EGCG

Tumor-EGCG+Milk

EGCG concentration (mg/ml)

ctr 6e-3 0.01 0.02 0.05 0.09 0.2 0.4 0.8

% P

roli

fera

tio

n

0

50

100

150

200Normal-EGCG

Normal-EGCG+Milk

Figure 5.3 % Change of Tumor cell proliferation with increasing levels of isolated EGCG

(solid bars) and complexed EGCG-milk fractions before in vitro digestion (A). % Change of

Normal cell proliferation with increasing levels of isolated EGCG (solid bars) and complexed

EGCG-milk fractions before in vitro digestion. The control used for isolated EGCG was media

without EGCG and the control used for EGCG-milk fractions was milk without EGCG. Different

letters within each concentration of EGCG, indicate statistical significant differences at p<0.05.

a a a a

a a

a a

a

b b

a

a a a a

a a

a a a a

a a

a a a a

b

a

a

b

Page 97: Nanoencapsulation of Tea Catechins in Casein Micelles

88

However, these results indicate the bioaccessibility of EGCG when the caseins are still intact, as

they are fractions used before in vitro digestion. To more closely mimic the physiological case,

the digests of the fractions were also delivered to the normal and tumor cell lines. Many studies

have shown that polyphenols undergo biotransformation during their digestion and the products

may not necessarily have the same biological activities (Spencer et al., 2003; Auro 2005; Visioli

et al, 2011). Tea beverages are commonly consumed along with milk, so the effects of the tea-

milk complexes on the bioavailibilty of the digests should also be examined.

The concentrations used for EGCG digestion, whether in water (isolated form) or in milk

(encapsulated with caseins), were derived from figure 5.1 as explained above. However, in order

to expose the cell lines to the EGCG digests and EGCG-milk digests, further dilution with media

took place, with each digest being diluted in media with a ratio of 1: 16 (digest: media (v/v)).

Therefore, the final concentrations of EGCG in media that were delivered to cells corresponded

to concentrations of EGCG before saturation of casein micelles (i.e. ≤0.15 mg/ml media), EGCG

at saturation (i.e. 0.15 mg/ml media) and after saturation of casein micelles (0.3 mg/ ml media).

Figure 5.4 A and B shows the anti-proliferative activity of EGCG digests and EGCG-milk

digests on normal (4D) and tumorigenic (v-src) rat colonic cells. The inhibitory effect of EGCG

did not differ between cell lines (p>0.05), whether EGCG was bound to milk or in isolated form.

However, at higher concentrations of EGCG in normal cells (≥ 0.15 mg/ml), EGCG was seen to

be toxic to the normal (4D) cells as well.

In most studies, the effect of polyphenols on inhibiting the proliferation of cancer cells is

observed, whereas, fewer studies have studied the effects on normal cell lines. However, some

studies that have examined the cytotoxicity of polyphenols against both cancer and normal cell

Page 98: Nanoencapsulation of Tea Catechins in Casein Micelles

89

lines, have reported the selective cytotoxicity against the cancer cell line (Chen et. al., 2003;

Yamamoto et. al., 2003; Hakimuddin et al., 2006).

Although the main factors that inhibit tumor cell proliferation but do not affect normal cell

proliferation are not clearly clarified, some possible explanations for this phenomenon have been

proposed. One explanation may be that as cancer cells have a rapid growth rate, usually caused

by the over expression of a few oncogenes, and if EGCG blocks the activity of some of these

oncogenes, it will result in the inhibition of the cell division. Conversely, growth and survival of

normal cells depends on the expression of many pathways, some of which may be inhibited by

EGCG, but others may not be inhibited. Therefore, in this case cell survival would not be

significantly affected (Weinstein et. al., 2008). Another study has hypothesized that some cells

that are more commonly in contact with plant-derived polyphenols, such as cells found in the

epidermis, oral mucosa, and digestive tract, and have naturally developed mechanism to tolerate

higher concentrations of polyphenols. However, cells from other organs that do not posses the

same mechanism(s), or cancer cells that may have lost their protective mechanisms, show

sensitivity to high concentrations of polyphenols (Yamamoto et. al., 2003).

Other proposed cancer prevention mechanisms include the antioxidant and pro-oxidant actions of

green tea catechins such as EGCG. Green tea catechins have been shown to act as antioxidants

that scavenge free radicals to protect normal cells. However, it has been demonstrated that

EGCG, mainly at higher concentrations, can cause the generation of reactive oxygen species

(ROS) intermediates, inducing oxidant stress and activating stress signalling cascades resulting

in tumor cell apoptosis (Chen et. al., 2003; Ebling et al., 2005; Na et. al., 2006).

However, the pro-oxidant activity of green tea catechins at higher concentrations, which may

ultimately cause apoptosis, has been shown in both transformed and nontransformed human

Page 99: Nanoencapsulation of Tea Catechins in Casein Micelles

90

bronchial cells as well as some other normal cell lines (Hsu et al., 2002; Yamamoto et. al., 2003;

Yang et al., 2006). Some case reports have shown that when excessive amounts of tea extracts

were consumed, liver toxicity was reported and they concluded that it may be due to a pro-

oxidant mechanism induced by tea catechins (Bonkovsky et. al., 2006; Lambert et.al., 2007b).

This may be the reason toxicity at higher concentrations of EGCG was seen for the normal (4D)

rat colonial cells for both forms of EGCG; isolated and complexed with milk (Figure 5.4 B).

However, Chow et al. (2003) concluded that it is safe to take green tea polyphenol products in

amounts equivalent to the EGCG content of 8–16 cups of green tea, once a day. There results

showed that following high EGCG concentration consumption (800 mg EGCG once a day),

systemic availability of free EGCG after chronic green tea polyphenol administration was

increased by more than 60% .

Page 100: Nanoencapsulation of Tea Catechins in Casein Micelles

91

EGCG concentration (mg/ml)

ctrl 2e-3 5e-3 0.01 0.02 0.04 0.08 0.15 0.3

% P

roli

fera

tion

0

20

40

60

80

100

120

EGCG concentration (mg/ml)

ctrl 2e-3 5e-3 0.01 0.02 0.04 0.08 0.15 0.3

% P

roli

fera

tio

n

0

20

40

60

80

100

120

Figure 5.4 % Change of Tumor cell proliferation with increasing levels of EGCG in digests of

isolated EGCG (solid bars) and complexed EGCG-milk fractions (A). % Change of Normal cell

proliferation with increasing levels of EGCG in digests of isolated EGCG and complexed

EGCG-milk fractions (B).The control used for isolated EGCG was media without EGCG and the

control used for EGCG-milk fractions was digested milk without EGCG.Within each

concentration at each of the cell lines, no statistical significant differences was seen (p˃0.05).

A

B

Page 101: Nanoencapsulation of Tea Catechins in Casein Micelles

92

5.5 Conclusion

In spite of the complex formation of casein micelles and EGCG, the caseins are still susceptible

to proteolysis during in vitro gastric digestion. As casein micelles are recognized for their

valuable functionality in the delivery of bioactives, the cell proliferation results demonstrate that

EGCG-milk complexes before digestion have similar bioefficacy to isolated EGCG. However, in

tumor cells, at lower concentrations of EGCG, there was a lower extent of inhibition of cancer

cell proliferation by the complexes compared to the free EGCG (≤ 0.4 mg/ml). However, after

gastro-duodenal digestion, there were similarities between complexed and free EGCG. These

results indicate that the strong binding of EGCG and milk casein does not affect the bio efficacy

of the tea polyphenols. In addition there were clear differences in the bioefficacy of EGCG for

normal or cancer cells, showing that the mechanism(s) applied by EGCG for the selectivity of

inhibiting the proliferation of tumor cells rather than normal cells was still maintained even when

complexed with milk.

Overall, these findings support previous data, that casein micelles can be employed as an

appropriate platform for delivery of bioactive compounds. In addition to establishing the

inhibitory effects of EGCG-milk complexes on rat cancer cells, we have shown that they are

protective towards normal cells. Such components may be useful in adjuvant cancer therapeutics.

Developing effective prevention/treatment strategies using natural food sources may help solve

some of the other problems encountered with conventional chemotherapeutic treatments

(Hakimuddin et al. 2006).

Page 102: Nanoencapsulation of Tea Catechins in Casein Micelles

93

Chapter 6

General Conclusions

Numerous reports are available on the affinity of tea catechins for caseins; however, in most

studies isolated caseins were employed and not casein micelles (Arts et al., 2002; Jobstyl et al.,

2004; Yan et al., 2009; Frazier et al., 2009; Yuksel et al., 2010). Former studies have suggested

the potential use of casein micelles as nanocarriers of hydrophobic compounds such as vitamin

D2 and curcumin (Livney et al., 2009; Sahu et al., 2008). The incorporation of the bioactive

molecules in casein micelles may be beneficial, not only because a higher concentration of these

molecules could be carried in the food product, but also because they may be protected against

adverse chemical reactions and/or interactions with other components in the food. However,

possible changes in the bioefficacy of bioactives when encapsulated in casein micelles needs to

be studied.

The results from this study clearly demonstrated that tea catechins, and specifically EGCG was

incorporated in casein micelles, and when <2.5 mg/ml were added to milk, most of the EGCG

interacted with the casein proteins, as it was recovered in the colloidal phase of milk, after

centrifugation. Furthermore, it was demonstrated that the tea catechin-caseins interactions

affected the processing behaviour of the casein micelles. Rennet gelation of casein micelles was

selected as it is caused by a surface destabilization of the casein micelles via an enzyme, and no

changes in environmental conditions occur during gelation, making data interpretation less

challenging. The results showed the association of the EGCG with the casein proteins in the

micelles hindered to some extent the renneting reaction, and more importantly, it affected the

Page 103: Nanoencapsulation of Tea Catechins in Casein Micelles

94

formation of a gel. Indeed, at high enough concentrations (above saturation point) gelation was

hindered.

To evaluate differences in the bioaccessibility of EGCG, the effect of the tea catechin-milk

protein complexes on cell cytotoxity and proliferation of HT-29 colon cancer cells were tested.

The results demonstrated that the incorporation of EGCG in casein micelles did not hinder the

wanted decrease of proliferation of HT-29 cells. The differences in the extent of the decrease

between free EGCG and encapsulated were attributed to the difference in the matrix, as EGCG

suspended in milk serum and milk whey showed similar behaviors compared to EGCG

incorporated with the micelles.

However, to be able to indicate a similar bioefficacy for nanoencapsulated EGCG, a more

relevant testing model should be used, where the samples are tested not only before digestion,

but also after in vitro digestion. Although the relevance of the in vitro digestion to in vivo is still

under debate, due to the heterogeneities occurring during in vivo digestion, it is possible to

assume that the gastrointestinal cells will be exposed to a wide population of nanoencapsulated

EGCG, ranging from non digested to fully digested. Hence, the present study also observed the

effect of EGCG-casein micelles complexes before and after in vitro digestion. In the study, both

normal and cancerous cells were used.

It was demonstrated that the interactions between EGCG and milk caseins did not cause the

proteins to be less susceptible to gastric proteolysis. The disappearance of casein micelles during

pepsin digestion was similar with or without EGCG. The results for both normal (4D) and tumor

(v-src) rat colonic cells showed that the EGCG-milk digests, when compared to isolated EGCG

digests, did not show a significant difference in cell proliferation (P> 0.05). These results

indicated that the strong binding of EGCG and milk casein does not affect the bioefficacy of the

Page 104: Nanoencapsulation of Tea Catechins in Casein Micelles

95

tea catechin, nor does it possibly affect the mechanism (s) applied to inhibit tumor cell growth. In

addition there were clear differences in the bioefficacy of EGCG for normal or cancer cells,

showing a selectivity of EGCG in inhibiting the proliferation of tumor cells rather than normal

cells whether in isolated form, or as complexed with milk.

The present research project contributed to better understanding of two aspects of the protein-

polyphenol interactions; 1. The impact of EGCG-casein complexes on the caseins by studying

protein functionality (rennet gelation) and protein digestion (in vitro). 2. The impact of EGCG-

casein complexes on the bioefficacy of EGCG by applying assays on normal and tumor cell lines

before and after in vitro digestion.

Further research is needed to explain the practical aspect of application of casein micelles for

delivery of bioactive compounds.

Page 105: Nanoencapsulation of Tea Catechins in Casein Micelles

96

References

Anema S.G., Klostermeyer H., (1996). Zeta-Potentials of casein micelles from reconstituted

skim milk heated at 120 degrees C. International Dairy Journal 6, 673-687.

Araújo J.R., Gonçalves P.,Martel F.,(Chemopreventive effect of dietary polyphenols in

colorectal cancer cell lines. (2011). Nutrition Research, 31 (2) , 77-87.

Aron P.M., Kennedy J. A. (2008). Review: Flavan-3-ols: Nature, occurrence and biological

activity. Mol. Nutr. Food Res., 52, 79 – 104.

Arts M. J. T., Haenen R. M. M., Wilms L., Sasja A. J., Heijnen C. G., et al., (2002). Interactions

between Flavonoids and Proteins: Effect on the Total Antioxidant Capacity. J. Agric. Food

Chem., 50 (5), 1184-1187.

Auro A.M., (2005). In vitro digestion models for dietary phenolic compounds. Espoo. VTT

Publications 575, 107, 57.

Balentine D. A., Wiseman S.A., Bouwens L.C., (1997). The Chemistry of Tea Flavonoids.

Critical Reviews in Food Science and Nutrition, 37 (8), 693-704.

Bazinet L., Araya-Farias M., Doyen A., Trudel D., Têtu B., (2010). Effect of process unit

operations and long-term storage on catechin contents in EGCG-enriched tea drink. Food

Research International 43, 1692–1701.

Bennick.A., (2002). Interaction of plant polyphenols with salivary proteins. Crit. Rev. Oral Biol.

Med. 13(2), 184-196.

Bonkovsky H. L., (2006) Hepatotoxicity associated with supplements containing Chinese green

tea (Camellia sinensis). Ann. Intern. Med. 144, 68–71.

Brew K., Vanaman T.C. , Hill R.L. (1967). Comparison of amino acid sequence of bovine α-

lactalbumin and hen’s egg white lysozyme. J. Biol. Chem. 242, 3747-3749.

Cha M. C., Lin A. and Meckling K. A., (2005). Low dose docosahexaenoic acid protects normal

colonic epithelial cells from araC toxicity. BMC Pharmacology, 5(7).

Chan P. C., Ramot Y., Malarkey D.E., Blackshear P. et., al., (2010). Fourteen-Week Toxicity

Study of Green Tea Extract in Rats and Mice. Toxicologic Pathology, 38, 1070-1084.

Charlton A.J., Baxter N. J., Lokman Khan M., Moir A. J. G., Haslam E., Davies A. P., and

Williamson M. P., (2002). Polyphenol/Peptide Binding and Precipitation. J. Agric. Food Chem.,

50 (6), 1593-1601.

Chen M.I., (2002). Tea and Health—An Overview. In Tea: Bioactivity and Therapeutic

Potential. edited by Yong-Su Zhen

Page 106: Nanoencapsulation of Tea Catechins in Casein Micelles

97

Chen C., G. Shen, Hebbar V., Hu R., Owuor E. D. and Kong A.N., (2003). Epigallocatechin-3-

gallate-induced stress signals in HT-29 human colon adenocarcinoma cells. Carcinogenesis 24

(8), 1369-1378.

Chen D., Daniel K.G.,. Kuhn D.J., Kazi A., Bhuiyan M., Li L., Wang Z., et al., (2004). Green tea

and tea polyphenols in cancer prevention.Frontiers in Bioscience 9, 2618-2631.

Chen L., Zhang H.Y., (2007). Review: Cancer Preventive Mechanisms of the Green Tea

Polyphenol Epigallocatechin-3-gallate. Molecules 12, 946-957.

Chow H. H., Cai Y., Hakim I. A., Crowell J. A., Shahi F., Brooks C. A., et al. (2003).

Pharmacokinetics and safety of green tea polyphenols after multiple-dose administration of

epigallocatechin gallate and polyphenon E in healthy individuals. Clinical Cancer Research 9,

3312–3319.

Chow H-H. S., Hakim I.A, Vining D.R., Crowell J.A., Ranger-Moore J., et. al., (2005). Effects

of Dosing Condition on the Oral Bioavailability of Green TeaCatechins after Single-Dose

Administration of Polyphenon E in Healthy Individuals. Clin Cancer Res.,11(12).

Chitpinityol S., Crabbe M.J.,(1998). Review: Chymosin and aspartic proteinases. Food

Chemistry, 61 (4), 395-418.

Cilla A., González-Sarrías A., Tomás-Barberán F. A., Espín J.C, Barberá R., (2009). J. Agric.

Food Chem.114, 813–820.

Citarasu, T. (2010). Herbal biomedicines: a new opportunity for aquaculture industry. Aquacult.

Int. 18, 403-414.

Dalgleish, D.G. (1998). Casein micelles as colloids: Surface structures and stabilities.J.Dairy

Sci. 81, 3013-3018.

Dalgleish D.G. (2011). On the structural models of bovine casein micelles - review and possible

improvements. Soft Matter.7:7 (6), 2265-2272.

Dalgleish D.G., Corredig M., (2012). The Structure of the Casein Micelle of Milk and Its

Changes During Processing. Annu. Rev. Food Sci. Technol. 3, 449–67.

De Kruif C. G., Zhulina E. B., (1996). κ-casein as a polyelectrolyte brush on the surface of

casein micelles. Colloids and Surfaces A. 117, 151–159.

De Kruif C.G. (1998). Supra-aggregates of casein micelles as a prelude to coagulation. J.Dairy

Sci. 81:3019-3028. De Kruif, C. G., Holt C. (2003). Casein micelle structure, functions and

interactions. In PF Fox and PLH McSweeney (eds). Advnced Dairy Chemistry. London : Kluwer

Academic/Plenum. 233-276.

Page 107: Nanoencapsulation of Tea Catechins in Casein Micelles

98

del Angel C.R., Dalgleish D.G., (2006). Structures and some properties of soluble protein

complexes formed by the heating of reconstituted skim milk powder. Food Res. Int. 39, 472-479.

Dickinson E., Miller R., Relkin P.,Vermersh J., (2001). Binding properties of vanillin to whey

proteins: Effect on protein conformational stability and foaming properties. Food Colloids

Fundamentals of Formulation. Royal Society Chemistry. 282-292

Dubeau S., Samson G., Tajmir-Riahi H.A., (2011). Dual effect of milk on the antioxidant

capacity of green, Darjeeling, and English breakfast teas. Food Chemistry 122, 539–545.

Dupont D., Mandalari G., Molle D., Jardin J., Rolet-Repecaud O., Duboz G., Leonil J., Mills

C.N., and Mackie A.R., (2010). Food processing increases casein resistance to simulated infant

digestion. Mol. Nutr. Food Res.,54, 1677–1689.

Ehrich M., Sharova L., (2000). In Vitro Methods for Detecting Cytotoxicity. Current Protocols in

Toxicology. 3:2.6.1–2.6.27.

Elbling L., Weiss R., Teufelhofer O., Uhl M., Knasmueller S., Schulte-Hermann R., Berger R.,

Micksc M., (2005). Green tea extract and (–)-epigallocatechin-3-gallate, the major tea catechin,

exert oxidant but lack antioxidant activities. The FASEB Journal express article 10.1096/fj.04-

2915fje.

Farrell H.M Jr, Malin E.L., Brown E.M., Qi P.X., (2006) Casein micelle structure: What can be

learned from milk synthesis and structural biology? Current Opinion in Colloid & Interface

Science 11, 135–147.

Ferrara L., Montesano D., Senatore A.,(2001).The distribution of minerals and flavonoids in the

tea plant (Camellia sinensis) . Il Farmaco 56, Issues 5-7, 397-401.

Ferruzzi M. G., Green R. J., (2006). Analysis of catechins from milk–tea beverages by enzyme

assisted extraction followed by high performance liquid chromatography. J. Agric. Food Chem.

99, 484–491.

Ferruzzi M. G., (2010). The influence of beverage composition on delivery of phenolic

compounds from coffee and tea. Physiology & Behavior 100, 33–41.

Fox P.F., Mc Sweeny P.L.H., (1998). Dairy chemistry and biochemistry. 1st. Ed, Springer, New

York.

Fox P.F., Guinee T.P., Cogan T.M., McSweeney P.L.H., (2000). Fundamentals of cheese

science. Gaithersburg, MD: Aspen Pub.

Fox P.F., McSweeney P.L.H., (2003). Milk proteins: general and historical aspects. In Advanced

dairy chemistry. Vol. 1: Proteins. 3rd ed. New York, USA: Kluwer Academic/Plenum Publisher.

Page 108: Nanoencapsulation of Tea Catechins in Casein Micelles

99

Frazier R. A., Deaville E.R., Green R.J., Stringano E., Willoughby I.,Plant J., Mueller-Harvey I.,

(2009). Interactions of tea tannins and condensed tannins with proteins. J. Pharm. Biomed. Anal.,

doi:10.1016/j.jpba.2009.05.035

Friedman M., Mackey B., Kim H.J., Lee I.S., Lee K.R., Lee S.U., Kozukue E., KozukuE N.,

(2007). Structure−Activity Relationships of Tea Compounds against Cancer Cells. J. Agric. Food

Chem.,55,243−253.

Garti N.( 2008). Delivery and controlled release of bioactives in foods and nutraceuticals. CRC

Press, Woodhead: Boca Raton, Fla.: Cambridge. 1-7.

Gaygadzhiev Z., Alexander M., Corredig M., (2009). Sodium caseinate-stabilized fat globules

inhibition of the rennet-induced gelation of casein micelles studied by Diffusing Wave

Spectroscopy. Food Hydrocolloids 23 (4), 1134-1138.

Ghosh K.S, Sahoo B.K, Dasgupta S., (2008). Spectrophotometric studies on the interaction

between epigallocatechin gallate and lysozyme. Chemical Physics Letters 452, 193–197.

Goncalves R, Mateus N., Pianet I., Laguerre M., de Freitas V., (2011). Mechanisms of Tannin-

Induced Trypsin Inhibition: A Molecular Approach. Langmuir, 27, 13122–13129.

Gonzalez de Mejia E., Song S.Y., Heck C.A., Ramirez-Mares M.V., (2010).Yerba mate tea (Ilex

paraguariensis): Phenolics, antioxidant capacity and in vitro inhibition of colon cancer cell

proliferation. J. Functl. Foods, 2, 23 – 34.

Graham H. N., (1992). Green Tea Composition, Consumption, and Polyphenol Chemistry.

Preventive Medicine 21, 334-350.

Green R. J., Murphy A.S., Schulz B., Watkins B.A., Ferruzzi M. G., (2007). Common tea

formulations modulate in vitro digestive recovery of green tea catechins. Mol. Nutr. Food Res.,

51, 1152 – 1162.

Hakimuddin F., Paliyath G., Meckling K., (2006). Treatment of Mcf-7 Breast Cancer Cells with

a Red Grape Wine Polyphenol Fraction Results in Disruption of Calcium Homeostasis and Cell

Cycle Arrest Causing Selective Cytotoxicity. J. Agric. Food Chem., 54 (20), 7912–7923.

Hasni I., Bourassa P., Hamdani S., Samson G., Carpentier R., Tajmir-Riah H.A., (2011).

Interaction of milk α- and β -caseins with tea polyphenols. Food Chemistry. 126 (2), 630–639

Henning S. M., Fajardo-Lira C., Lee H. W., Youssefian A., Go V. L. W., and Heber D., (2003).

Catechin Content of 18 Teas and a Green Tea Extract Supplement Correlates With the

Antioxidant Capacity. Nutrition And Cancer, 45(2), 226–235.

Page 109: Nanoencapsulation of Tea Catechins in Casein Micelles

100

Hofmann T., Glabasnia A., Schwarz B., Wisman K. N., Gangwer K. A. and Hagerman A.E.,

(2006). Protein Binding and Astringent Taste of a Polymeric Procyanidin, 1,2,3,4,6-Penta-O-

galloyl-β-D-glucopyranose, Castalagin and Grandinin. J Agric Food Chem. 54(25), 9503–9509.

Holt C., Sawyer L.,(1993). Caseins as rheomorphic proteins - interpretation of primary and

secondary structures of the αs1-caseins, β-caseins and κ-caseins. J. Chem. Soc. Faraday Trans.

89, 2683-2692.

Holt C., Horne D.S., (1996). The hairy casein micelle: evolution of the concept and its

implications for dairy technology. Neth. Milk Dairy J. 50, 85-111.

Hong J., Lu H., Meng X., Ryu J.H, Hara Y., Yang C.S., (2002). Stability, Cellular Uptake,

Biotransformation, and Efflux of Tea Polyphenol Epigallocatechin-3-Gallate in HT-29 Human

Colon Adenocarcinoma Cells. Cancer Research 62, 7241–7246.

Horne D.S., (2006).Casein micelle structure: models and muddles. Current Opinion in Colloid

and Interface Science.11:11 (2-3), 148-153.

Horne D.S., (2009). Casein micelle structure and stability. in Milk Proteins-From Expression to

Food. ed. Thompson, A., Boland, M. and Singh, H., Academic Press, New York. pp.442-473.

Hsu S, Bollag W.B., Lewis J., Huang Q., Singh B., Sharawy M., Yamamoto T., and Schuster

G.S., (2003) Tea polyphenols induce differentiation and proliferation in epidermal

keratinocytes.J Pharmacol Exp Ther 306, 29–34.

Hwang J.T., Ha J., Park I.J., Lee S.K., Baik H.W, Woon H., Min Kim Y., Park O.J., ( 2007).

Apoptotic effect of EGCG in HT-29 colon cancer cells via AMPK signal pathway. Cancer

Letters 247, 115–121.

Inglingstad R.A., Devold T.G., Eriksen E.K., Holm H., Jacobsen M., Liland K.H., Rukke E.O.,

Vegarud G.E., (2010). Comparison of the digestion of caseins and wheyproteins in equine,

bovine, caprine and human milks by human gastrointestinal enzymes. Dairy Sci. Technol., 90,

549–563.

Jobstyl E., O'Connell J., Fairclough J. P. A., and Williamson M. P., (2004). Molecular Model for

Astringency Produced by Polyphenol/Protein Interactions. Biomacromolecules, 5 (3), 942-949.

Jobstyl E., Howse J. R., Fairclough J. P. A., and Williamson M.P., (2006). Noncovalent Cross-

Linking of Casein by Epigallocatechin Gallate Characterized by Single Molecule Force

Microscopy. J. Agric. Food Chem., 54 (12), 4077-4081.

Jung Y.D., Kim M.S., Shin B.A. , Chay K.O., Ahn B.W., Liu W., Bucana C.D., et al.,(2001).

EGCG, a major component of green tea, inhibits tumour growth by inhibiting VEGF induction in

human colon carcinoma cell. British Journal of Cancer 84(6), 844–850.

Page 110: Nanoencapsulation of Tea Catechins in Casein Micelles

101

Kalkwarf HJ., Khoury JC., and Lanphear B.P., (2003). Milk intake during childhood and

adolescence, adult bone density, and osteoporotic features in US women. The American journal

of clinical nutrition. 77, 257-265.

Karak T., Bhagat R.M., (2010). Review:Trace elements in tea leaves, made tea and tea infusion.

Food Research International 43, 2234–2252.

Kartsova L. A. and Alekseeva A. V., (2008). Effect of Milk Caseins on the Concentration of

Polyphenolic Compounds in Tea. Journal of Analytical Chemistry, 63(11), 1107–1111.

Keogh J.B., Mcinerney J., and Clifton P.M., (2007). The Effect of Milk Protein on the

Bioavailability of Cocoa Polyphenols . J. Food Science. 72 ( 3).

Kilmartin P. A., Hsu C. F., (2003). Characterisation of polyphenols in green, oolong,and black

teas, and in coffee, using cyclic voltammetry. Food Chemistry, 82, 501–512.

Kim M., Murakami A., Kawabata K., Ohigashi H., (2005). Epigallocatechin-3-gallate promotes

pro-matrix metalloproteinase-7 production via activation of the JNK1/2 pathway in HT-29

human colorectal cancer cells. Carcinogenesis 26, 1553-1562.

Kim Y., Goodner K.L., Park J.D., Choi J., Talcott S.T., (2011) Changes in antioxidant

phytochemicals and volatile composition of Camellia sinensis by oxidation during tea

fermentation. Food Chem. 129, 1331–1342.

Korhonen H., Pihlanto A., (2006). Bioactive peptides: Production and functionality.

International Dairy Journal 16, 945–960.

Kyle J. A., Morrice P. C, Mcneill G., and Duthie G. G., (2007). Effects of Infusion Time and

Addition of Milk on Content and Absorption of Polyphenols from Black Tea. J. Agric. Food

Chem. 55, 4889−4894.

Kumamoto M., Sonda T., Nagayama K., Tabata M., (2001). Effects of pH and metal ions on

antioxidative activities of catechins. Biosciences Biotechnology Biochemistry 65 (1), 126–132.

Lakowicz J.R., (2010). Quenching Of Fluorescence. Principles of Fluorescence Spectroscopy

Third Edition. 227-331.

Lambert J.D., Lee M.J., Diamond L., Ju J., Hong J., Bose M., Newmark H.,Yang C. S., (2007a).

Dose-dependent levels of epigallocatechin-3-gallate in human colon cancer cells and mouse

plasma and tissues. Drug metabolism and disposition. 34 (1), 8-11.

Lambert J.D., Sang S., Yang C.S., (2007b). Possible Controversy over Dietary Polyphenols:

Benefits vs Risks. Chem. Res. Toxicol., 20, 583 585.

Langley-Evans S.C., (2000).Consumption of black tea elicits an increase in plasma antioxidant

potential in humans. International Journal of Food Sciences and Nutrition (2000) 51, 309–

315.

Page 111: Nanoencapsulation of Tea Catechins in Casein Micelles

102

Leenan R., Roodenburg A., Tijburg L., Wiseman S., (2000). A single dose of tea with or without

milk increases plasma antioxidant activity in humans. European Journal of Clinical Nutrition, 54,

87–92.

Lesschaeve I., Noble A.C., (2005). Polyphenols: factors influencing their sensory properties and

their effects on food and beverage preferences. Am J Clin Nutr.,81.

Li L., Hang Chan T., (2001). Enantioselective Synthesis of Epigallocatechin-3-gallate (EGCG),

the Active Polyphenol Component from Green Tea. Organic Letters. Vol. 3 (5), 739 -741.

Li G.H., ChenY.k., Hou Z., Xiao H., Jin H., Lu G., Lee M.J., Liu B., Guan F., et. al., (2010).

Pro-oxidative activities and dose–response relationship of (2)-epigallocatechin-3-gallate in the

inhibition of lung cancer cell growth: a comparative study in vivo and in vitro. Carcinogenesis 31

(5), 902–910.

Livney Y. D., Verespej E. and Dalgleish D. G., (2003) .Steric effects governing disulfide

bond interchange during thermal aggregation in solutions of beta-lactoglobulin B and

alpha-lactalbumin.J.Agric.Food Chem.51,8098-8106.

Livney Y.D., Dalgleish D.G. (2009). Casein micelles for nanoencapsulation of hydrophobic

compounds. U.S. Patent US2009311329 (A1).

Livney Y. D. (2010). Milk proteins as vehicles for bioactives. Current Opinion in Colloid &

Interface Science 15, 73–83.

López-Fandiñoa1 R., Olanoa1 A., San Joséa C. and Ramos M., (1993). Application of reversed-

phase HPLC to the study of proteolysis in UHT milk. Journal of Dairy Research, 60: 111-116

Luck G., Liao H., Murray N. J., Grimmer H. R., Warminski E., Williamson M. P., Lilley T. H.

and Haslams E., (1994). Polyphenols, astringency and proline-rich proteins. Phytochemistry,

37(2), 357-371.

Maiti T.K,. Sundar Ghosh K, and Dasgupta S., (2006). Interaction of Epigallocatechin-3-Gallate

With Human Serum Albumin: Fluorescence, Fourier Transform Infrared,Circular Dichroism, and

Docking Studies. Proteins: Structure, Function, and Bioinformatics 64, 355–362.

Malaki Nik A., Wright A. J., Corredig M., (2010). Surface adsorption alters the susceptibility of

whey proteins to pepsin-digestion. Journal of Colloid and Interface Science 344, 372–381.

Malaki Nik A., Wright A. J., Corredig M., (2011). Impact of interfacial composition on emulsion

digestion and rate of lipid hydrolysis using different in vitro digestion models. Colloids and

Surfaces B: Biointerfaces 83, 321–330

Page 112: Nanoencapsulation of Tea Catechins in Casein Micelles

103

Manach C., Scalbert A., Morand C., Rémésy C., and Jimenez L., (2004). Polyphenols: food

sources and bioavailability. Am J Clin Nutr. 79, 727–747.

Manach C., Scalbert A., Williamson G., Morand C., Rémésy C., (2005). Bioavailability and

bioefficacy of polyphenols in humans. I. Review of 97 bioavailability studies. Am J Clin

Nutr2005;81(suppl):230S–42S.

McDougall G.J., Stewart D., (2005). The inhibitory effects of berry polyphenols on digestive

enzymes. BioFactors 23, 189–195.

McDougall G.J., Kulkarni N.N., Stewart D., (2008). Current developments on the inhibitory

effects of berry polyphenols on digestive enzymes. BioFactors 33,1–8.

McLoughlin P., Roengvoraphoj M., Gissel C., Hescheler J., Certa U., Sachinidis A., (2004).

Transcriptional responses to epigallocatechin-3 gallate in HT 29 colon carcinoma spheroids.

Genes to Cells. 9, 661–669.

Na H.k., Surh Y.J., (2006).Review: Intracellular signaling network as a prime chemopreventive

target of (–)-epigallocatechin gallate. Mol. Nutr. Food Res., 50, 152 – 159.

Nagle D.G., Ferreira D., Zhou Y.D., ( 2006).Epigallocatechin-3-gallate (EGCG): Chemical and

biomedical perspectives. Phytochemistry 67,1849-1855.

O’connel J.E., Fox P.D., Tan-Kintia R.and Fox P.F., (1998). Effects of tea, coffee and cocoa

extracts on the colloidal stability of milk and concentrated milk. Int. Dairy Journal.689-693.

O’Connell J.E., Fox P.F., (2001). Significance and applications of phenolic compounds in the

production and quality of milkand dairy products: a review. International Dairy Journal 11,103–

120.

Papadopoulou A. and Frazier R. A., (2004). Characterization of protein–polyphenol interactions.

Trends in Food Science & Technology 15,186–190.

Pascal C., Poncet-Legrand C., Cabane B., Vernhet A., (2008). Aggregation of a Proline-Rich

Protein Induced by Epigallocatechin Gallate and Condensed Tannins: Effect of Protein

Glycosylation. J. Agric. Food Chem., 56 (15), 6724-6732.

Picariello G., Ferranti P., Fierro O., Mamone G., Cairaa S., et. al., (2010). Peptides surviving the

simulated gastrointestinal digestion of milk proteins: Biological and toxicological implications.

Journal of Chromatography B, 878, 295–308.

Poncet-Legrand C., Edelmann A., Putaux J.L., Cartalade D., Sarni-Manchadoa P., Vernhet A.,

(2006). Poly(L-proline) interactions with flavan-3-ols units: Influence of the molecular structure

and the polyphenol/protein ratio. Food Hydrocolloids 20, 687–697.

Page 113: Nanoencapsulation of Tea Catechins in Casein Micelles

104

Rahimi Yazdi S., Corredig M., (2012). Heating of milk alters the binding of curcumin to casein

micelles. A fluorescence spectroscopy study. Food Chemistry 132, 1143–1149.

Rawel H.M., Frey S. K., Meidtner K., Kroll J., Schweigert F.J., (2006). Determining the binding

affinities of phenolic compounds to proteins by quenching of the intrinsic tryptophan

fluorescence. Mol. Nutr. Food Res., 50, 705 – 713.

Reis C.P., Neufeld R.J., Ribeiro, A.J. and Veiga F., (2006). Nanoencapsulation I. Methods for

preparation of drug-loaded polymeric nanoparticles. Nanomed. Nanotech. Biol. Med..2, 8-21.

Ribadeau B, Mercier J.C.,Grosclau F., (1973). Amino-acid composition and sequence

of bovine αs1-caseins and β-caseins. Neth. Milk Dairy J. 27, 304-312.

Richard T., Lefeuvre D., Descendit A., Quideau S., Monti J.P., (2006). Recognition characters in

peptide–polyphenol complex formation. Biochimica et Biophysica Acta 1760, 951–958.

Roura E., Lacueva C. A., Estruch R., Lourdes Mata Bilbao M., Izquierdo-Pulido M. and.

Lamuela-Raventos R. M., (2008). The effects of milk as a food matrix for polyphenols on the

excretion profile of cocoa epicatechin metabolites in healthy human subjects. British Journal of

Nutrition, 100, 846–85.

Sahu A., Kasoju N., and Bora U., (2008). Fluorescence Study of the Curcumin Casein Micelle

Complexation and Its Application as a Drug Nanocarrier to Cancer Cells. Biomacromolecules 9,

2905–2912.

Salucci M., Stivala L. A., Maiani G., Bugianesi R. and Vannini V., (2002). Flavonoids uptake

and their effect on cell cycle of human adenocarcinoma cells. British J.Cancer 86, 1645-1651.

Sandra S., Alexander M., Dalgleish D. G., (2007). The rennet coagulation mechanism of skim

milk as observed by transmission diffusing wave spectroscopy. Journal of Colloid and Interface

Science. 308, 364–373.

Sang S., Lambert J.D., Ho C.T., Yang C.S., (2011).The chemistry and biotransformation of tea

constituents. Pharmacol Res., 64(2), 87-99.

Saura-Calixto F., Serrano J., Gon I., (2007). Intake and bioaccessibility of total polyphenols in a

whole diet. Food Chemistry. 101, 492–501.

Scalbert A., Manach C., Morand C., Remesy C., (2005). Dietary Polyphenols and the Prevention

of Diseases. Critical Reviews in Food Science and Nutrition, 45, 287–306.

Schwarz B. and Hofmann T., (2008). I s there a direct relationship between oral astringency and

human salivary protein binding. Eur Food Res Technol 227, 1693–1698.

Semo E., Kesselman E., Danino D., Livney Y.D., (2007). Casein micelle as a natural nano-

capsular vehicle for nutraceuticals. Food Hydrocolloids 21, 936-942.

Page 114: Nanoencapsulation of Tea Catechins in Casein Micelles

105

Sharma R., (2005). Market trends and opportunities for functional dairy beverages. Australian

Journal of Dairy Technology. Melbourne: 60 (2).

Sharma V.and Rao L.J.M., (2009). A Thought on the Biological Activities of Black Tea. Critical

Reviews in Food Science and Nutrition, 49:, 379–404.

Skehan P., Storeng R., Scudiero D., Monks A. , McMahon J., Vistica, D., et al.,( 1990). New

Colorimetric Cytotoxicity Assay for Anticancer-Drug Screening. Articles. 82, 13, 1107-1112.

Shpigelman A., Israeli G., Livney Y. D., (2010). Thermally-induced proteinepolyphenol co–

assemblies: beta lactoglobulin-based nanocomplexes as protective nanovehicles for EGCG. Food

Hydrocolloids 24, 735-743.

Shukla A., Narayanan T. Zanchi D., (2009). Structure of casein micelles and their complexation

with tannins. Soft Matter 5, 2884–2888.

Soares S., Mateus N., de Freitas V., (2007). Interaction of Different Polyphenols with Bovine

Serum Albumin (BSA) and Human Salivary -Amylase (HSA) by Fluorescence Quenching .J.

Agric. Food Chem., 16, 6726-6735.

Spencer J.P.,(2003). Metabolism of Tea Flavonoids in the Gastrointestinal Tract. American

Society for Nutritional Sciences. 3255- 3261.

Stahl W., van den Berg H., Arthur J., Bast A., Dainty J., Faulks R.M., et al. (2002).

Bioavailability and metabolism. Molecular Aspects of Medicine. 23, 39-100.

Stevenson D. E., Hurst R. D., (2007). Review: Polyphenolic phytochemicals – just antioxidants

or much more? Cell. Mol. Life Sci., 64, 2900 – 2916.

Tagliazucchi D., Verzelloni E., Conte A.,(2005). Effect of Some Phenolic Compounds and

Beverages on Pepsin Activity during Simulated Gastric Digestion. J. Agric. Food Chem., 53,

8706−8713.

Vaidyanathan J.B.and Walle T., (2003). Cellular Uptake and Efflux of the Tea Flavonoid

Epicatechin-3-gallate in the Human Intestinal Cell Line Caco-2. Journal Pharmacology and

Experimental Therapeutics 307, 745–752.

Valeille K., Ferezou J., Parquet M., Amsler G., and Gripois D., Quignard-Boulanque A., and

Martin J.C., (2006). The natural concentration of the conjugated linoleic acid, cis-9,trans-11 in

milk fat has antiatherogenic effects in hyperlipidemic hamsters. Journal of Nutrition, The. 136,

1305-1310.

van der Burg-Koorevaar M. C. D., Miret S., and Duchateau G.,(2011). Effect of Milk and

Brewing Method on Black Tea Catechin Bioaccessibility. J. Agric. Food Chem., 59, 7752–7758.

Page 115: Nanoencapsulation of Tea Catechins in Casein Micelles

106

Vichai V.,Kirtikara K., (2006).Sulforhodamine B colorimetric assay for cytotoxicity screening.

Nature protocols 1 (3), 1112.

Visioli F., De La Lastra C.A., Andres-Lacueva C., Aviram M., Calhau C. et al., (2011).

Polyphenols and Human Health:A Prospectus. Critical Reviews in Food Science and Nutrition,

51.

Vuong U.V. Stathopoulos C.E., Nguyen M.H., Golding J.B., Roach P.D., (2011).Isolation of

Green Tea Catechins and Their Utilization in the Food Industry. Food Reviews International,

27,227–247.

Walstra P., Wouters J.T.M., Geurts T. J., (2006). Dairy science and technology. Taylor &

Francis, New York.

Weinstein I. B., Joe A., (2008). Oncogene addiction. Cancer Res. 68, 3077–3080; discussion

3080

Willig J.A.,(2009). Thesis: The effect of anthocyanin acylation on the inhibition of ht-29 colon

cancer cell proliferation. Graduate School of The Ohio State University.

Yamamoto T., Hsu S., Lewis J., Wataha J., Dickinson D., Singh B., Bollag W. B. , Lockwood P.,

Osaki T., and Schuster G., (2003). Green Tea Polyphenol Causes Differential Oxidative

Environments in Tumor versus Normal Epithelial Cells. J. Pharmacology and Experimental

Therapeutics. 307 (1).

Yan Y., Hu J., and Yao P., (2009). Effects of Casein, Ovalbumin, and Dextran on the

Astringency of Tea Polyphenols Determined by Quartz Crystal Microbalance with Dissipation.

Langmuir 25, 397- 402.

Yang G., Liao J., Kim K., Yurkow E.J., Yang C.H.,(1998). Inhibition of growth and induction of

apoptosis in human cancer cell lines by tea polyphenols. Carcinogenesis. 19 (4 ), 611–616.

Yang C. S., Sang S., Lambert J.D., Hou Z., Ju J., Lu G., (2006). Possible mechanisms of the

cancer‑preventive activities of green tea. Mol. Nutr. Food Res.50, 170–175.

Yang C. S., Sang S., Lambert J.D., Lee M.J., (2008). Review: Bioavailability issues in studying

the health effects of plant polyphenolic compounds. Mol. Nutr. Food Res., 52, 139 –151.

Yang C. S., Wang X., Lu G. and Picinich S. C., (2009). Cancer prevention by tea: animal studies,

molecular mechanisms and human relevance. Nature reviews:cancer 9, 429-439.

Yang C. S., Pan E., (2012). Review: The effects of green tea polyphenols on drug metabolism.

Expert Opin. Drug Metab. Toxicol. 8(6), 677-689.

Yuan J.M., Sunb C., Butlerc L.M., (2011). Review: Tea and cancer prevention: Epidemiological

studies. Pharmacological Research 64,123– 135.

Page 116: Nanoencapsulation of Tea Catechins in Casein Micelles

107

Yuksel, Z., Avci, E., & Erdem, Y. K., (2010). Characterization of binding interactions between

green tea flavanoids and milk proteins. Food Chemistry, 121(2), 450-456.

Zaveri N.T., (2006). Green tea and its polyphenolic catechins: Medicinal uses in cancer and

noncancer applications. Life Sciences 78, 2073–2080.