Anda di halaman 1dari 124

MICROENCAPSULATION OF FISH

OIL/OLIVE OIL BLENDS USING SUGAR


BEET PECTIN AS AN ENCAPSULANT

Sudheera Polavarapu

Submitted in total fulfilment of the requirements


of the degree of Masters by Research (by Thesis Only)

June 2011

Department of Agriculture and Food Systems


Melbourne School of Land and Environment
The University of Melbourne
Victoria 3010, Australia

Abstract
Fish oil (FO) is rich in the polyunsaturated fatty acids, but sensitive to oxidation by many
factors including, heat, light and atmospheric oxygen. Microencapsulation of FO has been
reported to help protect FO from oxidation, especially in the presence of antioxidants. As
extra virgin olive oil (EVOO) is rich in oleic acid, and some antioxidant compounds, this
study aimed to investigate the potential of EVOO to protect FO from oxidation in the
presence of sugar beet pectin as a wall material. Ethylenediaminetetras-acetic acid (EDTA)
was also tested as a chelating agent to retard the lipid oxidation caused by the transition
metal ions present in the sugar beet pectin.
The oil-in-water emulsions were prepared and spray-dried to form microcapsules containing
fish oil alone (control) and a blend of FO/EVOO (1:1 weight ratio), at 25% (pH 3) and 50%
oil (pH 3 and pH 6) loadings. Fish oil and Fish oil- Extra virgin olive oil emulsions with
EDTA were prepared at 50% oil loading only at pH 6. The physicochemical characteristics
and oxidative stability of all FO and FO-EVOO emulsions and microcapsules were assessed.
The microcapsules showed low moisture content (<3% w/w) and water activity (~0.3). The
encapsulation efficiencies were >90% for all the formulations proving sugar beet pectin an
effective encapsulant.
Oxidative stability was assessed by testing propanal and hexanal in the head space using GC
analysis, fatty acid content, and oxidation stability using an Oxipress. It was concluded that
EVOO improved the oxidative stability of microencapsulated FO during accelerated storage
conditions (80C, 5 bar oxygen pressure), irrespective of oil loading. However, no effect
was detected at pH 3 during storage at room temperature (~27C). The addition of EDTA in
the formulation of the microcapsules significantly (P<0.05) increased the oxidative stability
of the microcapsules. It was also observed that microcapsules prepared from emulsions at
pH 6 were better in terms of long term storage when compared to the microcapsules
prepared from emulsions at pH 3.

II

Declaration
This is to certify that
(i) the thesis comprises only my original work towards the Masters
(ii) due acknowledgement has been made in the text to all other material used,
(iii) the thesis is 22,584 words in length as approved by the Graduate School, Faculty or
RHD Committee.
Sudheera Polavarapu
June 2011

III

Acknowledgements
My greatest gratitude to my supervisors Dr. Said Ajlouni, Dr. Maryann Augustin and Dr.
Christine Oliver who were abundantly helpful and offered invaluable assistance, support and
guidance from the very early stages of my Masters Project till the thesis submission. I
thank them for taking time to attend monthly meetings to discuss results and for the
constructive comments to the numerous reports I submitted. Many thanks for providing new
ideas and approaches to the project. Working under their supervision has triggered a passion
for research in me. I attribute my growth as a student to all of them and that I would benefit,
for a long time to come. I am grateful to them in every possible way. Also, my special thanks
to Dr. Christine Oliver for all her help and efforts in preparing the manuscripts for
publication.
I am grateful to all staff and students in the Department of Agriculture and Food Systems,
The University of Melbourne and CFNS (CSIRO Food and Nutritional Sciences) for their
assistance in my lab work, for sharing instruments and for creating a pleasant working
atmosphere. A warm thank you to the Senior Laboratory Manager Michelle Rhee at the
University for her assistance in ordering and supplying of chemicals and for all her support
while working in the lab. Furthermore, I convey my special acknowledgement to Li Jiang
Cheng, Zhiping Shen, Said Ajlouni, Rangika Weerakkody, Helen French, Christine Oliver
and Roderick Williams for their indispensable help and for being patient while teaching me
how to operate various lab instruments. I also appreciate their time and efforts for helping
me interpret the results obtained.
I am immensely thankful to Dr. Simon Crawford, School of Botany, The University of
Melbourne, Parkville for providing his expertise with the scanning electron microscope and
for helping me obtain the SEM images for all my samples.
Lastly, my parents deserve special mention for their endless love, encouragement and
prayers without which none of this would be of any significance. I thank my sister and
mother-in-law for always being a source of undying love, support and comfort. Words fail

IV

me to express my appreciation to my husband Kiran whose love and persistent confidence


have been the foundations for my strength at difficult times and my motivation to complete
my study. I will be forever be grateful to him for understanding my passions, ambitions and
commitment for the success of the project.

Table of Contents
ABSTRACT

II

DECLARATION

III

ACKNOWLEDGEMENTS

IV

TABLE OF CONTENTS

VI

ABBREVIATIONS

IX

LIST OF FIGURES

LIST OF TABLES

XII

Chapter 1- INTRODUCTION

Chapter 2 - LITERATURE REVIEW

2.1 Bioactive compounds in foods

2.1.1 Introduction

2.1.2 Omega 3 fatty acids

2.1.3 Extra Virgin Olive Oil

2.1.4 Interaction of bioactives

10

2.1.5 Bioavailability

11

2.2 Microencapsulation

12

2.2.1 Microencapsulation- Todays approach

12

2.2.2 Encapsulating materials

14

2.2.3 Emulsion preparation

16

2.2.4 Microencapsulation process

17

2.2.5 Incorporation of microencapsulated bioactives into foods

19

2.2.6 Recent developments in Microencapsulation

19

2.3 Sugar beet pectin as encapsulant

21

Chapter 3 MATERIALS AND METHODS

23

3.1 Raw material Composition

23

3.1.1 Sugar beet pectin

23

3.1.2 Fish oil & Extra virgin olive oil

23

3.1.3 Oxidation status of Fish oil and Extra virgin olive oil

24

3.2 Preparation of emulsions & microcapsules


VI

25

3.3 Spray-dried microcapsules

29

3.4 Emulsion Characteristics

30

3.4.1 Particle size

30

3.4.2 Light microscopy

30

3.4.3 Viscosity

30

3.4.4 Oxidative stability of emulsions under accelerated conditions

30

3.4.5 Zeta-potential of emulsions

31

3.5 Characterisation of the spray-dried microcapsules

31

3.5.1 Moisture content

31

3.5.2 Water activity

31

3.5.3 Particle size of reconstituted powders

32

3.5.4 Fatty acid composition analysis

32

3.5.5 Determination of free oil content

32

3.5.6 Determination of total oil content

32

3.5.7 Encapsulation efficiency

32

3.5.8 Propanal and hexanal analysis

33

3.5.9 Oxidative stability of the microcapsules under accelerated

33

storage conditions
3.5.10 Scanning electron microscopy (SEM)
3.6 Statistical analysis

33
34

Chapter 4 RESULTS AND DISCUSSION

35

4.1 Properties of raw material

35

4.1.1 Sugar beet pectin

35

4.1.2 Fish oil & extra virgin olive oil

35

4.1.3 Oxidation status of fish oil and extra virgin olive oil

37

4.2 Effect of partial substitution of fish oil with extra virgin olive oil

38

on emulsion and microcapsule characteristics


4.3 Effect of EDTA on emulsion and microcapsule characteristics

49

4.4 Effect of pH on the emulsion and microcapsule characteristics

57

4.5 Lipid Oxidation

65

VII

4.5.1 Oxidative stability of emulsions and microcapsulated oil during

66

storage under ambient conditions


4.5.2 Oxidative stability of emulsions and microcapsules exposed

79

to accelerated storage conditions


Chapter 5 GENERAL CONCLUSIONS AND FUTURE DIRECTIONS

86

References

88

Appendices

109

VIII

Abbreviations
SBP - Sugar beet pectin
FO - Fish oil
EVOO - Extra virgin olive oil
DGS - Dried glucose syrup
EDTA - Ethylenediaminetetras-acetic acid
EPA - Eicosapentaenoic acid
DHA - Docosahexaenoic acid
ALA- Alpha-linolenic acid
- CD cyclodextrin
PCL- Polycaprolactone
CMC Carboxy methyl cellulose
PEG Polyethylene glycol
PV- Peroxide value
IP- Induction period

IX

LIST OF FIGURES

Figure 1

Illustration of the steps involved and the factors considered in the

14

production of microcapsules
Figure 2

Schematic structure for generalised sugar beet pectin

22

Figure 3

Standard curve obtained after assessment of the oxidative stability

25

using an oxipress
Figure 4

Process flow chart of the experimental plan for the production of spray

27

dried microcapsules
Figure 5

Emulsions and spray-dried powders

29

Figure 6

Particle size distributions of the original (fresh) emulsions, prior to

39

spray drying
Figure 7

Brightfield micrographs of (A) FO-25 (7.5% oil, wet basis) and (B)

40

FO-EVOO-50 (15% oil, wet basis) emulsions at pH 3


Figure 8

Apparent viscosities of fish oil and fish oil-extra virgin olive oil (1:1 wt

41

ratio) emulsions (30% TS; 2% SBP, 20.5% DGS, 7.5% total oil and
30% TS; 2% SBP, 13% DGS, 15% total oil), intended for manufacture
of spray-dried microcapsules containing 25 and 50% oil, as a function
of shear rate
Figure 9

Autotitration curve for FO emulsion with zeta potential measured as a

43

function of pH at 22 C
Figure 10

Autotitration curve for FO-EVOO emulsion with zeta potential

43

measured as a function of pH at 22 C
Figure 11

Particle size distributions of reconstituted spray-dried microcapsules at

47

0, 1, 2 and 3 month (AD) stored under room temperature conditions


(~25C, exposed to light)
Figure 12

Scanning electron microscopy images of the fish oil (A, C) and fish
oil-extra virgin olive oil (B, D) spray-dried powders (25% oil loading),
prior (A, B) and after 3 month storage (C, D) under ambient conditions
(~25C, 0.5 aw)

48

Figure 13

Particle size distributions of the original (fresh) emulsions (15% oil,

50

wet basis), with and without EDTA, at pH 6


Figure 14

Brightfield micrographs of FO-EDTA (A) and FO-EVOO-EDTA (B)

50

emulsions with 15% oil, at pH 6


Figure 15

Apparent viscosities of fish oil and fish oil-extra virgin olive oil (1:1 wt

51

ratio) emulsions (15% oil, wet basis) with and without EDTA at pH 6,
as a function of shear rate
Figure 16

Particle size distribution of reconstituted spray-dried microcapsules

56

(50% oil, dry basis), with and without EDTA at pH 6 after 0, 1, 2 and 3
month (AD) storage at room temperature (~25C)
Figure 17

Scanning electron microscopy images of the fish oil-extra virgin olive

57

oil spray-dried powders (50% oil, dry basis), with (A, C) and without
(B, D) EDTA, at pH 6, prior (A, B) and subsequent (C, D) to storage for
3 months at room temperature (~25C)
Figure 18

Particle size distributions of the original (fresh) fish oil and fish

58

oil-extra virgin olive oil (1:1 wt ratio) emulsions made at pH 3 and 6


Figure 19

Brightfield micrographs of (A) fish oil (pH 3) and (B) fish oil-extra

59

virgin olive oil (1:1 wt ratio) emulsions (pH 6) with 15% oil
Figure 20

Apparent viscosities of fish oil and fish oil-extra virgin olive oil (1:1 wt

60

ratio) emulsions with 15% oil at pH 3 and pH 6, as a function of shear


rate
Figure 21

Particle size distributions of reconstituted spray-dried fish oil and fish

oil-extra virgin olive oil microcapsules prepared from emulsions at pH


3 and 6 with 50% oil loading prior (A), and subsequent to storage for 1
(B), 2 (C) and 3 (D) month under room temperature conditions (~25 C,
aw 0.5, exposed to light)
Figure 22

Scanning electron microscopy images of fish oil spray-dried


microcapsules from emulsions prepared at pH 3 (A, C) and pH 6 (B, D)
with 50% oil loading, prior (A, B) and subsequent (C, D) to storage at
ambient conditions (~25C, aw 0.5, 3 months)

XI

65

LIST OF TABLES

Table 1

List of types of encapsulation processes

18

Table 2

Formulation of the fish oil and fish oil-extra virgin olive oil (1:1 wt ratio)

28

emulsions at pH 3
Table 3

Formulation of the fish oil and fish oil-extra virgin olive oil (1:1 wt ratio)

28

emulsions with and without EDTA at pH 6


Table 4

Fatty acid composition (%) of the initial oils

36

Table 5

Oxidative stability of original oils during accelerated storage (80C,

38

oxygen pressure of 0.5 MPa)


Table 6

Zeta potential values of FO and FO-EVOO emulsions measured at 22 C

42

Table 7

Properties of spray-dried powders containing microencapsulated fish oil

44

and fish oil-extra virgin olive oil (1:1 wt ratio) (25 % & 50 % oil loading)
at pH 3
Table 8

Changes in moisture content and water activity in FO and FO-EVOO

46

microcapsules (25 % & 50 % oil loadings) over storage time


Table 9

Particle size of the reconstituted spray-dried microcapsules containing

46

fish oil and fish oil-extra virgin olive oil (1:1 wt ratio) during storage of
the powders in stoppered flasks at room temperature (~25C)
Table 10

Properties of powders containing microencapsulated fish oil and fish

53

oil-extra virgin olive oil blends (1:1 wt ratio) with and without EDTA at
pH 6, 50% oil (dry basis)
Table 11

Changes in moisture content and water activity of FO and FO-EVOO

55

microcapsules (50% oil, dry basis) (with and without EDTA) at pH 6 on


storage at room temperature (~25C)
Table 12

Particle size values of the reconstituted spray-dried microcapsules (50%

55

oil, dry basis), with and without EDTA at pH 6, on storage at room


temperature (~25C)
Table 13

Properties of powders containing microencapsulated fish oil and fish


oil-olive oil blends (1:1 wt ratio) at pH 3 and pH 6 with 50 % oil loading

XII

61

Table 14

Changes in moisture content and water activity of fish oil and fish

62

oil-extra virgin olive oil (1:1 wt ratio) microcapsules at pH 3 and pH 6


with 50% oil loading over storage time at room temperature (~25C)
Table 15

Particle size values of the reconstituted spray-dried microcapsules

64

containing fish oil and fish oil-extra virgin olive oil (1:1 wt ratio) at pH 3
and pH 6 with 50% oil loading, during storage of the powders in stoppered
flasks at room temperature (~25C)
Table 16

Propanal content in microencapsulated fish oil and fish oil-extra virgin

68

olive oil (1:1 wt ratio) (25 % & 50 % oil loading) at pH 3 during storage at
room temperature (~25C, 03 month)
Table 17

Propanal content in microencapsulated fish oil and fish oil-extra virgin

68

olive oil (1:1 wt ratio) with and without EDTA (50% oil loading) at pH 6
during storage at room temperature (~25 C, 03 month)
Table 18

Propanal content in microencapsulated fish oil and fish oil-extra virgin

70

olive oil (1:1 wt ratio) with 50 % oil loading at pH 3 and pH 6 during


storage at room temperature (~25C, 03 month)
Table 19

Hexanal content in microencapsulated fish oil and fish oil-extra virgin

71

olive oil (1:1 wt ratio) (25 % & 50 % oil loading) at pH 3 during storage at
room temperature (~25C, 03 month)
Table 20

Hexanal content in microencapsulated fish oil and fish oil-extra virgin

72

olive oil (1:1 wt ratio) with and without EDTA (50 % oil loading) at pH 3
and pH 6 during storage at room temperature (~25C, 03 month)
Table 21

Hexanal content in microencapsulated fish oil and fish oil-extra virgin

72

olive oil (1:1 wt ratio) with 50 % oil loading at pH 3 and pH 6 during


storage at room temperature (~25C, 03 month)
Table 22 a

Fatty acid composition of microencapsulated fish oil (25 % and 50 % oil

75

loading) at pH 3 during storage at room temperature (~25C, 03 month)


Table 22 b

Fatty acid composition of microencapsulated fish oil olive oil (25 % and
50 % oil loading) at pH 3 during storage at room temperature (~25C, 03
month)

XIII

76

Table 23 a

Fatty acid composition of microencapsulated fish oil (50 % oil loading)

77

with and without EDTA at pH 6 during storage at room temperature


(~25C, 03 month)
Table 23 b

Fatty acid composition of microencapsulated fish oil olive oil (50 %

78

oil loading ) with and without EDTA at pH 6 during storage at room


temperature (~25C, 03 month)
Table 24

Oxidative stability of bulk oils (FO, EVOO, FO-EVOO (1:1 wt ratio)

80

during accelerated storage (80C, oxygen pressure of 0.5 bar)


Table 25

Oxidative

stability

of

emulsions

and

powders

containing

81

microencapsulated fish oil and fish oil-extra virgin olive oil (1:1 wt ratio)
(25% and 50% oil loading) at pH 3 during accelerated storage (80C,
oxygen pressure of 0.5 MPa)
Table 26

Oxidative

stability

of

emulsions

and

powders

containing

84

microencapsulated fish oil and fish oil-extra virgin olive oil (1:1 wt ratio)
made with and without EDTA (50% oil loading) at pH 3 during
accelerated storage (80C, oxygen pressure of 0.5 MPa)
Table 27

Oxidative stability of emulsions and powders (50% oil loading)


containing fish oil and fish oil-extra virgin olive oil (1:1 wt ratio) at pH 3
and pH 6 during accelerated storage (80C, oxygen pressure of 0.5 bar)

XIV

85

CHAPTER 1 - INTRODUCTION
This chapter provides the background and purpose to the research topic and discusses the
need for further research related to the area. The project hypothesis and specific aims of the
research are outlined.
Microencapsulation technology has been the central concept of this research and the two
other areas discussed in relation to it are the protection of bioactives (omega 3 fatty acids,
and polyphenols) and the application of sugar beet pectin as an encapsulant. Literature
review on the current scientific information/existing research in the above mentioned fields
was reviewed in the second chapter.
The third chapter includes all details about the raw materials (supplier details and
manufacturers specification), chemicals (supplier details), equipments (with the
manufacturers details and model numbers) used in the research project. It also describes
the procedures/method of analysis adopted for the characterization of emulsions and
microcapsules.
Chapter four describes the results obtained after analysis of the raw materials, emulsions
and microcapsules following the analytical methods described in Chapter 3 - Methods
(Section 3.1, 3.4, 3.5, 3.6). Conclusive inferences from the results obtained were also
presented.
Based on the results of the various analyses conducted and the comparison of results
performed between samples, the final conclusions of this study is summarised in fifth
chapter.

1.1 BACKGROUND
Consumer behaviour and attitudes toward healthy foods continue to grow especially in
developed countries. This has sparked a wide interest in foods, such as, functional foods
which contain ingredients with health promoting properties, and can offer specific health
benefits to consumers in addition to the anticipated nutritional value.
Functional foods have the potential to contribute to a healthier society and are important
components of an overall healthful lifestyle that includes a balanced diet and physical
activity. There have been mounting evidence in recent years that food choices are an
important factor in reducing the risk of developing heart disease, cancer, obesity, high
blood pressure, osteoporosis, and other unhealthy conditions (Australian Indigenous
Healthinfonet, 2008).
Fish oils are considered functional foods as they are a rich source of omega-3 fatty acids
such as eicosapentaenoic acid (EPA) and docosahexaenoic acid (DHA) (Erkkila et al.,
2006). There has been mounting evidence in the recent years that omega-3 fatty acids have
positive long-term and current health benefits. Several prospective studies and clinical
trials have shown that omega-3 fatty acids can reduce the risk of developing heart disease
(Erkkila et al., 2006; GISSI-HF investigators, 2008). Various other studies have shown
that consumption of omega-3 fatty acids have been beneficial in treating neurological
disorders such as Alzheimers disease (Morris et al., 2003), in reducing bone loss (Griel et
al., 2007; Weiss, Barrett-Connor & von Muhlen, 2005) in lowering incidence of
depression and other related conditions (Lin & Su, 2007). However, the incorporation of
functional ingredients in a given food system and the processing and handling of such
foods are associated with nutritional challenges for their healthy delivery. The extreme
sensitivity of fish oils to oxidation can easily lead to the development of off-flavours and
cause significant loss of product quality, stability, nutritional value, and bio-availability
and the overall acceptability of the food product (Nawar, 1996; Watkins & German, 1998).
Consequently, microencapsulation has been successfully used to encapsulate omega-3 fatty
acids in order to prevent oxidation and to improve stability and bioavailability (Wakil et

al., 2010; Sanguansri & Augustin, 2006). Microencapsulation can be defined as a process
in which tiny particles or droplets are surrounded by a coating to give small capsules. The
material inside the microcapsule is referred to as the core, internal phase, or fill, whereas
the coating is sometimes called a shell, membrane, or wall material, as referred to
henceforth. Most microcapsules have diameters between a few micrometers and a few
millimetres (Microencapsulation, 2009). Microencapsulation is one example of technology
that has the potential to meet the challenge of successfully incorporating and delivering
functional ingredients into a range of food types.
The most common technology used for encapsulation has been spray drying because it is
efficient, cost effective, has readily available equipment and produces particles of
reasonably good quality (Dobry et al., 2009). Some of the encapsulant materials used in
the microencapsulation of foods are proteins, sugars, starches, gums, lipids and cellulosic
materials. Chitosan plus maltodextrin and liposomes have been reported in encapsulation
of omega-3 rich fish oils (Klaypradit & Huang, 2008; Reineccius, 1995). A considerable
number of research studies have been conducted on the oxidative stability of fish oil,
microencapsulated with different wall materials using one of the many chemical and
physical microencapsulation processes available. Literature review on the oxidative
stability of omega-3 fatty acids showed a variation of results due to the effectiveness of
wall materials used, changes in the process parameters, storage conditions and methods
employed for assessing the oxidative stability.

Thus, further studies were even conducted

with the addition of antioxidants to improve the shelf life of the fish oil microcapsules
(Velasco, Dobarganes & Marquez ruiz, 2000; Heinzelmann et al., 2000).
Several studies suggested that SBP could be a suitable wall material for manufacturing
lipophilic food ingredients due to its excellent emulsifying properties (Drusch, 2007;
Drusch et al., 2007; Leroux et al., 2003; Martinez-Dominguez, de la Puerta, &
Ruiz-Gutierrez, 2001). This functionality associated with SBP showed that it had potential
as a wall material for encapsulating w-3 acids in our study. Recent studies conducted by
Drusch (2007) also confer that SBP could be a promising alternative to the traditional
encapsulating agents like milk proteins and gum Arabic because of its ability to form

stable emulsions with improved encapsulation properties. However, further research is


needed to test the oxidative stability of the microcapsules encapsulated with SBP over
prolonged storage time. The drawback of using SBP as an encapsulant material is the
presence of transition metals in high concentrations i.e. iron (1.91 ppm) and copper (0.08
ppm) (Katsuda et al., 2008). Data reported in the literature established the fact that metals
like iron (Fe) and copper (Cu) act as catalysts for the formation of highly reactive hydroxyl
radicals which initiate a chain reaction for lipid oxidation (Benedt and Shibamoto, 2007;
Goldstein et al., 1993). Hence, in the present research study, ethylenediaminetetra-acetic
acid (EDTA) was used as a chelating agent and as an antioxidant to mask the negative
effects of these metal ions. EDTA is a polyamino carboxylic acid and a colourless,
water-soluble solid. EDTA is capable of forming complexes with metal ions in a solution
and diminishing their reactivity. EDTA and its derivatives are most commonly used
because they are inexpensive and effective even when used in small amounts in the overall
formulation (EDTA, 2010; Kabara & Orth, 1997).

1.2 THESIS HYPOTHESIS


It is hypothesised that the combination of the fatty acid composition and antioxidants
(polyphenols) in EVOO would improve the oxidative stability of the microencapsulated
fish oil powders during storage, and that SBP as the material would provide a good matrix
to maintain the nutritional value in encapsulated fish oil.

1.3 THESIS PURPOSE AND AIMS


The purpose of the research project was to microencapsulate fish oil (FO) mixed with
EVOO at a 1:1 ratio and using SBP as the wall material. EVOO was added as a source of
oleic acid and polyphenols. Polyphenolic components are known to possess potent
antioxidant properties (Olive Oil, 2008). The addition of the chelating agent, EDTA, was
evaluated for its ability to retard the lipid oxidation caused by the transition metal ions
present in SBP.

The aim of this research study was to examine the:

Oxidative stability of the microencapsulated fish oil (FO) in the presence of extra
virgin olive oil (EVOO).

Effect of SBP as the wall material to prevent oxidative stability of


microencapsulated FO.

Influence of EDTA on the oxidative stability of FO and FO-EVOO blend


emulsions and microcapsules.

Influence of pH on the physico-chemical characteristics and oxidative stability of


FO and FO-EVOO emulsions and microcapsules

CHAPTER 2 - LITERATURE REVIEW

2.1. BIOACTIVE COMPOUNDS IN FOODS

2.1.1 INTRODUCTION
Bioactive compounds are biomolecules present in foods that exhibit the capacity to
modulate one or more metabolic processes, which promote better health. They indulge in
multiple metabolic activities and are beneficial for treating several diseases and target
tissues and are usually found in glycosylated, esterified, thiolyated, or hydroxylated forms.
Bioactive food components are found abundantly in most fruits and vegetables. However,
probiotics, conjugated linolenic acid, long-chain omega-3 polyunsaturated fatty acid, and
bioactive peptides are most commonly found in animal products such as milk, fermented
milk products and cold-water fish (Encyclopaedia of Food and Culture, 2003).
Numerous bioactive components from plant sources (e.g. phytosterol, carotenoids,
flavonoids, soluble fibre, polyphenols etc.) marine oils, milk and milk products (e.g.
peptides / proteins, pre and pro-biotics) are currently used in functional food formulations
(McClements et al., 2009; Engel & Schubert, 2006; Murphy, 2001; Lpez-Lpez et al.,
2009). However, a significant lack of understanding of bioactive ingredients availability,
mechanism of action and synergistic effects exists and hence, the acceptable levels of
specific bioactive ingredients to be used in different foods have to established to establish
the appropriate dosage of healthy bioactive components.
Significant research findings stated in the Encyclopaedia of Food and Culture (2003)
indicated that bioactive food components could act simultaneously at different or identical
target sites. For instance, in cardiovascular disease, the bioactive component isoflavone
may reduce circulating oxidized low-density lipoproteins in the plasma, bind cholesterol in
the intestinal tract and thereby reduces absorption of dietary cholesterol. Another example
is the lipid-lowering mechanism of dietary fibre and phytosterol/stanols. It occurs by
sequestering cholesterol in the intestinal tract and reducing cholesterol absorption. The

bioactive food components have thus been proven to have health-promoting characteristics
and reduce the risk of cancer, cardiovascular disease, osteoporosis, inflammation, type 2
diabetes, and other chronic degenerative diseases (Encyclopaedia of Food and Culture,
2003; Ferguson & Philpott, 2007; Pfeuffer & Schrezenmeir, 2000; Hooper & Cassidy,
2006; Shahidi, 2006).
Clearly, bioactive food components can play an important role in better health and disease
prevention. Consequently, there is a need to produce foods with optimal levels of
health-promoting bioactive food components that would assist consumers to make healthy
food choices for their diet.
2.1.2 OMEGA 3 FATTY ACIDS
Omega-3 fatty acids are considered essential fatty acids. They are essential to human
health but cannot be manufactured by the body. For this reason, omega-3 fatty acids must
be obtained from food. Also known as polyunsaturated fatty acids (PUFAs), omega-3 fatty
acids play a crucial role in brain function as well as normal growth and development (drsri,
2009).
Fish, plant, and nut oils are the primary dietary source of omega-3 fatty acids. Best food
sources containing both eicosapentaenoic acid (EPA), C20:5(n-3), and docosahexaenoic
acid (DHA), C22:6(n-3), are cold-water fatty fishes such as, salmon, mackerel, halibut,
sardines, tuna, and herring. Dietary sources of alpha-linolenic acid (ALA), C18:3(n-3) are
flaxseeds, flaxseed oil, canola (rapeseed) oil, soybeans, soybean oil, pumpkin seeds,
pumpkin seed oil, purslane, perilla seed oil, walnuts, and walnut oil. ALA is converted in
the body into EPA and DHA, which are more readily utilised in vivo (Drsri, 2009).
Extensive research indicates that omega-3 fatty acids have antiinflammatory properties and
reduce the risk associated with chronic diseases such as heart disease, cancer, and arthritis
(Geusens et al., 1994; Richardson, 2004; GISSI-prevenzione investigators, 1999). These
essential fatty acids are an important structural component in the brain and appear to be
particularly important for cognitive (brain memory and performance) and behavioural
function (Willat et al., 1998; Mellor, Langharne & Peet, 1995). Symptoms associated with
7

omega-3 fatty acid deficiency include extreme tiredness (fatigue), poor memory, dry skin,
heart problems, mood swings or depression, and poor circulation (Omega-3-fatty acids,
2010).
An appropriate balance of omega-3 and omega-6 in the diet is vital, as they work together
to promote health. Whilst omega-3 fatty acids help reduce inflammation, most omega-6
fatty acids tend to promote inflammation. An inappropriate balance of these essential fatty
acids contributes to the development of disease, while a proper balance helps maintain and
even improve health. A healthy diet should consist of roughly 2 - 4 times more omega-6
fatty acids than omega-3 fatty acids (Omega-3-fatty acids, 2010).
Omega-3 fatty acids are extremely susceptible to oxidative damage from heat, light, and
oxygen. When exposed to the air for long periods, the fatty acids in the oil become
oxidized and rancid. Rancidity not only gives the oil disagreeable off-flavour and smell,
but it also diminishes the nutritional value. More importantly, free radicals are produced in
this process and which are implicated in the development of cancer and other degenerative
diseases (Pham-Huy, He & Pham-Huy, 2008).
Rancidity arises when the oils are removed from their food package or are stored
improperly. Hence, oils rich in polyunsaturated fatty acids should be stored in dark glass,
tightly closed containers in the refrigerator or freezer.
2.1.3

EXTRA VIRGIN OLIVE OIL

Extra virgin olive oil (EVOO) comes from cold pressing of the olives, contains no more
than 0.8% acidity, and is judged to have a superior taste, delicate flavour and most
antioxidant benefits. Olive oil is composed mainly of oleic acid, C18:1(n-9), and palmitic
acid, C16:0, as well as other minor fatty acids, along with traces of squalene (up to 0.7%)
and sterols (about 0.2% phytosterol and tocosterols). The composition varies by cultivar,
region, altitude, time of harvest, and extraction process (Alarcon de la Lastra et al., 2001;
Olive oil, 2008). Olive oil also contains a group of related natural products with potent

antioxidant properties which give extra-virgin unprocessed olive oil its bitter and pungent
taste (Tripoli et al., 2005).
Diets rich in monounsaturated fats have been shown to reduce the risk of coronary heart
disease (Keys et al., 1986; Willet et al., 1995). This is particularly noteworthy, because
olive oil is extremely rich in monounsaturated fats, most notably oleic acid (Olive oil,
2008). Another health benefit of olive oil is its ability to displace omega-6 fatty acids,
while not having any impact on omega-3 fatty acids. This way, olive oil helps to build a
more healthy balance between omega-6 and omega-3 fatty acids (Haban et al., 2004).
Oleic acid is found in various animal and vegetable sources. It comprises 55-80% of olive
oil, and the saturated form of oleic acid is called stearic acid. Oleic acid shares the normal
reactions of fatty acids and alkenes. Oxidation at the double bond can split the molecule,
yielding chain ends with aldehyde, ketone or carboxylic acid groups. This reaction occurs
slowly in air, and it causes the oil in which it is present to go rancid (Oleic acid, 2008).
Polyphenols are a group of chemical substances found in plants, characterized by the
presence of more than one phenol unit or building block per molecule. Polyphenols are
generally divided into hydrolyzable tannins (gallic acid esters of glucose and other sugars)
and phenylpropanoids, such as lignins, flavonoids, and condensed tannins. Notable sources
of polyphenols include berries, tea, beer, grapes/wine, olive oil, chocolate/cocoa, coffee,
walnuts, peanuts, pomegranates, yerba mate, and other fruits and vegetables. High levels of
polyphenols can generally be found in the fruit skins (Polyphenols, 2008).
Based on the epidemiologic studies conducted by Arts & Hollman, (2005), it has been
observed that olive oil phenolic content, rather than its fatty acid profile is responsible for
at least some of its cardio protective benefits. There has been growing evidence that as
antioxidants, polyphenols may protect cell constituents against oxidative damage and,
therefore, limit the risk of various degenerative diseases associated with oxidative stress
(Luqman & Rizvi, 2006; Pandey, Mishra & Rizvi, 2009). Polyphenols also exert protective
effects in treating conditions like cancer, diabetes, aging, asthma and skin damages
(Pandey & Rizvi, 2009).

2.1.4

INTERACTIONS OF BIOACTIVE COMPOUNDS

Bioactive compounds are susceptible to variations due to changes in their structure and
physicochemical properties as a result of processing and the possible interactions with
other food components and the intestinal microflora.
Functional foods rarely contain a single bioactive component and interactions between
bioactive components in the food is expected which could either be positive or negative.
For example, Vitamin C has been reported to reduce seleniums effectiveness against
chemically induced colon cancer (Ip, 1986). Selenium has been shown enhance the ability
of garlic to inhibit chemically induced mammary cancer in experimental animals
(Amagase, Schaffer & Milner, 1996).
There are very few research studies that can confirm or demonstrate the interaction or
effect of polyphenols with omega-3 fatty acids. A study conducted by Ren et al. (2008)
showed that homozygous sickle cell patients (HbSS) had reduced EPA and DHA in red
cells, platelets, and mononuclear cells when compared to their healthy counterparts
(HbAA). This difference in levels of omega-3 fatty acids was not because of reduced
omega 3 intake. Something else besides dietary deficiency was causing lower cell levels of
cellular omega-3 fatty acids in these patients. The study investigated, and concluded that
the lower levels of membrane EPA and DHA in blood cells of the HbSS patients could be
attributed to peroxidation resulting from a compromised antioxidant competence. Omega-3
fatty acids aid in cell membrane function and stability and polyphenols are known to
directly and indirectly prevent peroxidation of fats. In addition, the antioxidant status
affects cell membrane levels of omega-3 fatty acids in humans, and polyphenols were
clearly found to be superior dietary antioxidants to those studied in the above mentioned
research study (beta-carotene, tocopherol). Hence, there is a possibility that polyphenols
could act to protect omega-3 fatty acids in the cell membrane.

10

2.1.5

BIOAVAILABILITY OF NUTRIENTS

From a nutritional point of view nutrient bioavailability can be defined as the amount our
digestive system will be able to extract in a form that can be absorbed into the bloodstream
(Encyclopaedia of Food and Culture, 2003).
It has been established that during fat absorption, unsaturated long chain fatty acids are
esterified at a higher rate than saturated fatty acids of similar chain length. This
phenomenon has been attributed to differences in the binding affinity of fatty acids to a
cytosolic fatty acid-binding protein (Gang et al., 1980). Studies have been conducted to
evaluate the bioavailability of omega-3 fatty acids in microencapsulated fish-oil-enriched
foods compared with an equal amount of omega-3 PUFAs contained in fish oil capsules.
The results indicated that n-3 PUFA from microencapsulated fish-oil-enriched foods were
as bioavailable as n-3 PUFA in a capsule. Therefore, fortification of foods with
microencapsulated fish oil offers an effective way of increasing omega-3 PUFA intakes
and status in line with current dietary recommendations (Wallace et al., 2000; Volker,
Weng, Quaggiotto, 2005).
Singh et al. (2008) indicated that bioavailability and pharmacokinetics of polyphenolics
are governed by a number of factors; their native form (glycosylated/aglycone), the type of
sugar moiety present, and their physiochemical properties. Bioavailability of polyphenols
varies widely from one compound to another. It depends on their chemical structure, which
determines their absorption rate through the gastrointestinal tract, metabolism, and,
therefore, bioactivity (Manach, 2004). In nature, phenolic compounds occur as glucosides.
Phenolic acids like caffeic acid are easily absorbed through the gut barrier, whereas large
molecular weight polyphenols, such as proanthocyanidins are very poorly absorbed. Once
absorbed, polyphenols are conjugated to glucuronide, sulphate and methyl groups in the
gut mucosa and inner tissues. Non-conjugated polyphenols are virtually absent in plasma.
These reactions facilitate their excretion and limit their potential toxicity. The polyphenols
reaching the colon are extensively metabolised by the microflora into a range of low
molecular weight phenolic acids (Scalbert, 2002).

11

Polyphenols present as aglycones can be absorbed from the small intestine. However, most
of them are present in the form of esters, glycosides, or polymers and are not easily
absorbed in their natural form. Glycosylation influences chemical, physical, and biological
properties of the flavonoids and their absorption. It is generally accepted that the
breakdown of these conjugates to aglycones by acid hydrolysis in the stomach and by
microflora in the gut is required to produce the bioactive components that are readily
bioavailable to the body. However, relatively little is known about the ability of these
aglycone polyphenolics to reach the target cells or what the influence of further
metabolism in the body has on their spectra of biological activities. There are numerous
sites important for the metabolism of dietary polyphenols, including the gastrointestinal
tract, the liver, and various other organs such as the skin and brain (Singh et al., 2008).
The food matrix also plays a vital role in the bioaccessibility of a bioactive component in
functional foods. For example, a lipophilic food matrix is needed for bioavailability of
carotenoids whereas when phenolic compounds are in a food matrix the bioavailability of
certain polyhenols may be lowered depending on the matrix components (Parada &
Aguilera, 2007). Another study by Kean, Hamaker, & Ferruzzi, (2008) found that specific
food preparations of maize based food products could influence the bioaccessibility of
certain bioactive carotenoid species.

2.2 MICROENCAPSULATION

2.2.1 MICROENCAPSULATION-TODAYS APPROACH


Microencapsulation has been defined as the technology of packaging solid, liquid and
gaseous materials in small capsules that release their contents at controlled rates over
prolonged periods of time (Champagne & Fustier, 2007).
Microencapsulation technologies have been used extensively in the pharmaceuticals sector
and subsisted for many decades before its application in the food industry. With the

12

identification of exciting and new microencapsulation techniques, new and innovative


approaches have been developed for the protection and targeted delivery of bioactives.
Today, microencapsulation has become a viable means that enables the use of previously
difficult-to-use ingredients (e.g. sensitivity to air/oxygen, heat, light) and the introduction
of a variety of ingredients and properties into foods that were not previously available.
Recent years have seen a widespread popularity of this technology among food ingredients
companies to achieve a variety of foods with different functions. Figure 1 illustrates the
various factors to consider in the production of microcapsules.
Food manufacturers have now become increasingly receptive to the use of encapsulated
ingredients, which have become more economically viable. Indeed, changing consumer
trends and needs have also been the significant factors for essentially driving innovation in
the field of microencapsulation.
With consumers showing a growing preference for functional food - which now accounts
for a substantial amount of the global wellness market - food companies are looking for
different ways to incorporate health-promoting ingredients that deliver some kind of health
benefit to the consumer. While consumers are demanding more nutritious products, they
are unwilling to compromise on taste or quality. Microencapsulation can provide solutions
to these recent demands because of its ability to mask the undesirable tastes, prevent
oxidation of ingredients and also aid in targeted nutrition in a number of processed food
products.

13

Identification of high
quality source of active
ingredient to be
encapsulated

Physical and
chemical properties
of the bioactive
Stability of the
bioactive

Development of
successful formulation
for a targeted
application

Physical and
chemical properties
of the bioactive
Stability of the
bioactive
Selection of
appropriate
encapsulant material
Selection of suitable
microencapsulation
process
Addition of
antioxidants,
chelating agents,
emulsifiers,
stabilizers and salts

Selection of a format for


the delivery of the
microcapsule

Dry product or
liquid format

Figure 1: Illustration of the steps involved and the factors considered in the production of
microcapsules (Sanguansri and Augustin, 2006) with some modifications.

2.2.2 ENCAPSULANTING MATERIALS


Another definition of microencapsulation is a process in which tiny particles or droplets
are surrounded by a coating, sometimes referred to as a shell, wall material, membrane or
an encapsulant material (Microencapsulation, 2010).
A critical step in the production of microcapsules is the selection of appropriate
encapsulating material. The encapsulant material can be selected from a wide variety of
natural or synthetic polymers depending on the stability and release characteristics
expected from the final microcapsule. It has been reported that the composition as well as

14

the physical and chemical properties of the shell material can influence the functionality of
the final microcapsule and the processing technologies to be used for microencapsulation
(Sanguansri & Augustin, 2006). According to Augustin & Sanguansri (2007) a good
encapsulating material should have neutral taste and odour, low viscosity, good film
forming, gelling and barrier properties. It should also protect the core from degradation
during processing and storage and also mask any undesirable taste or odour associated with
the bioactive core when added into foods.
Some of the common encapsulant wall materials have been summarised by Vasishtha
(2003) and includes the following:

Polysaccharides/hydrocolloids, such as starch, algin/alginate, agar/agarose,


pectin/polypectate, carrageenan, and other gums.

Proteins such as gelatin, casein, zein, soy, and albumin.

Fats and fatty acids such as mono-, di- and triglycerides, and lauric, capric, palmitic
and stearic acid and their salts.

Cellulosic derivatives such as methyl- and ethyl-cellulose and CMC (carboxy


methyl cellulose)

Hydrophilic and lipophilic waxes such as shellac, PEG (polyethylene glycol),


carnauba wax and beeswax.

Sugar derivatives.

Conjugates of casein and whey protein have been suggested as effective encapsulating
material for fish oil (Livney, 2010). Polymers such as -cyclodextrin (-CD) and
polycaprolactone (PCL) were used to encapsulate fish oil using aggregation and emulsion
diffusion methods and it was found that PCL protected FO better than (-CD) (Choia et al.,
2010). The use of complex coacervates composed of gelatin or -lactoglobulin, gum
arabic, starch, and crosslinked with glutaraldehyde have been reported to decrease
oxidation in fish oil (Lumsdon, Friedmann & Green, 2005). Other encapsulant materials
used for the encapsulation of omega-3 fatty acids were Maillard reaction products. Fish oil
here is emulsified with heated aqueous mixtures comprising of a protein source (sodium
caseinate, whey protein isolate, soy protein isolate, or skim milk powder) and
15

carbohydrates (glucose, dried glucose syrup and oligosaccharide) (Augustin, Sanguansri


and Bode, 2006).
Different types of n-octenylsuccinate-derivatised starch have also been used as encapsulant
materials for fish oils and were found to be effective encapsulant materials (Drusch &
Schwarz, 2006). Chitosan plus maltodextrin could also offer an alternative for the
encapsulation of omega-3 rich fish oils (Daniells, 2007). Liposomes have also been
employed to extend the storage stability and shelf life of encapsulated omega 3 fatty acids
(Haynes et al., 1992). Recent studies conducted by Drusch (2007) also showed a
possibility of using sugar beet pectin as an encapsulant material for omega-3 fatty acids
2.2.3 EMULSION PREPARATION
An emulsion is a term used to describe a mixture containing two or more immiscible
liquids where one liquid (the dispersed phase) is dispersed in the other (the continuous
phase). Emulsions are thermodynamically unstable and normally do not form
spontaneously but require some form of mechanical energy input (e.g.shaking, stirring,
homogenizing or spray processes) to form an emulsion (Emulsion, 2009).

An emulsion

is said to be stable when there is no perceptible change in the size distribution of droplets,
its state of aggregation, or its spatial arrangement over the time-scale of observation, which
may vary from hours to months depending on the material (Vega & Roos, 2006;
Dickinson, 1994). Surfactants (e.g. polysorbates, phospholipids), proteins (e.g. milk
proteins) and/or thickening agents (gums, gelatin) can be used to increase the kinetic
stability of the emulsion (Dickinson & Woskett, 1989).
The particle size of emulsions is also a highly relevant factor as it influences the stability
of an emulsion prior to manufacture into spray-dried powder as well as the encapsulation
efficiency after spray drying. Results have shown that larger particles retained more oil
than smaller ones, but at the same time there was more unencapsulated oil at the surface of
big particles than the surface of small particles (Jafari et al., 2008; Jafari, He & Bhandari,
2007; Soottitantawat et al., 2005). The presence of unencapsulated oil on the surface of the

16

powder particles is the most undesirable property of encapsulated powders because of the
detrimental effects it has on the powder quality and stability (Vega et al., 2005).

2.2.4 MICROENCAPSULATION PROCESSES

The selection of a microencapsulation process depends on the properties of the core and
the wall materials, the release mechanisms desired, process type, and desired capsule
morphology and particle size. Most of the microencapsulation processes have been adapted
from the pharmaceutical and chemical industries (Augustin & Sanguansri, 2007). A range
of physical and chemical processes are available to encapsulate bioactive ingredients as
outlined in Table 1.
A process often used for encapsulation of oil and oil-soluble cores involves the use of
biopolymer gels as entrapment matrices. Examples include oil and oil-soluble cores
encapsulated in alginate beads by extrusion to form particles, followed by drying. For
efficient particle formation of this type methods like centrifugal extrusion and microject
methods have been developed (Benita, 1996).
Amongst the encapsulation processes mentioned above spray drying is the most commonly
used technique for microencapsulation. It is one of the oldest methods of encapsulation.
The process of spray drying is efficient, cost effective, has readily available equipment and
produces particles of reasonably good quality (Gharsallaoui, 2007). Food ingredients
microencapsulated by this method include fish oils, essential oils, vitamins, colourants,
flavours and other oil-soluble bioactives.

17

Table 1 List of Types of Encapsulation Processes (Thies, 1996)


Chemical Processes
Complex coacervation

Mechanical Processes
Spray drying

Polymer-polymer incompatibility

Spray chilling

Interfacial polymerization in liquid


media

Fluidized bed
Electrostatic deposition

In situ polymerization

Centrifugal extrusion

In-liquid drying
suspension separation

Spinning disk or rotational

Thermal and ionic gelation in liquid


media

Polymerization at liquid-gas
or solid-gas interface

Desolvation in liquid media


into solvent extraction bath

Pressure extrusion or
spraying

The general process of spray drying involves the dispersion and homogenization of the
substance to be encapsulated and the encapsulating material to form a suspension in water
(slurry). The suspension is then fed into a spray drier. The solution is sprayed into the
drying chamber with the help of an atomiser or spray nozzle. There are different types of
nozzles available like rotary nozzles, single fluid pressure nozzles and ultrasonic nozzles.
Appropriate choice of the kind of nozzle to be used is based on the particle size required
for a specific application. The most common applications are in the 100 to 200 micrometre
diameter range. Hot drying gases are passed as either in the co-current or counter current
direction to the atomiser in the drying chamber. This hot air converts the suspension into
solid and the solvent into vapour. The dried particles are then collected in the particle
separator which is usually a cyclone separator (Spray drying, 2010). The dry powder is
often free flowing and the microcapsules are soluble in water.
Fish oil has been encapsulated by using various formulations, processing conditions and
encapsulation technologies (Gan, Cheng & Easa, 2008; Blatt et al., 2006; Heinzelmann &
Franke, 1999; Lumsdon, Friedmann & Green, 2005; Klinkesorn et al., 2005). According to

18

Sanguansri & Augustin (2006) the formulation and processing steps prior to spray drying
are considered important in the production of stable powder microcapsules. Processing
conditions such as, temperature, pH of the emulsion, total solids, drying rate and the order
of adding each of the components used during the manufacture of fish oil powders also
influence the microcapsule properties and stability.

2.2.5 INCORPORATION OF MICROENCAPSULATED BIOACTIVES INTO


FOODS
The market for functional ingredients and foods has experienced growth in recent years
due to the increased consumer awareness and promotion of healthy eating and lifestyle.
Challenges remain to ensure that functional ingredients survive and remain active and
bioavailable after food processing and storage. Food can be used as a vehicle for the
delivery of bioactives and micronutrients at suitable levels that provide health benefits for
consumers (Day, 2009).
Rapid developments in microencapsulation technologies and delivery strategies have
resulted in increasing numbers of successful omega-3 fortified products in the market. The
omega-3 enriched products available to consumers worldwide are dietary supplements, pet
foods, dairy products, processed fish, meat and egg products, snacks, meals, infant formula
and baby foods, soups, ice tea drinks, cakes, biscuits, bakery products and beverages
(Sanguansri & Augustin, 2006).

2.2.6 RECENT DEVELOPMENTS IN MICROENCAPSULATION (NANOENCA-PSULATION)


Nanoscience is an emerging area of science that has the potential to generate radical new
products and processes. Concepts in nanoscience provide a sound framework for
developing an understanding of the interactions and assembly behaviour of food
components into microstructure, which influence food structure, rheology and functional
properties at the macroscopic scale. Advances in processes for producing nanostructured

19

materials coupled with appropriate formulation strategies have made possible the
production and stabilisation of nanoparticles that have potential applications in the food
and related industries (Sanguansri & Augustin, 2006).
Nanoencapsulation technology can be used to address some extraordinary needs of many
applications by producing nanosized particles and capsules. Nanocapsules in combination
with other microencapsulation methods can lead to new release characteristics (Persyn &
Oxely, 2008).
Nanotechnology and nanoscience involve the study of phenomena and materials, and the
manipulation of structures, devices and systems that exist at the nanoscale (<100 nm in
size). The properties of nanoparticles are not governed by the same physical laws as larger
particles, but by quantum mechanics. The physical and chemical properties of
nanoparticles for example, colour, solubility, strength, chemical reactivity and toxicity can therefore be quite different from those of larger particles of the same substance (Miller,
2008).
The altered properties of nanoparticles have created the possibility for many new profitable
products and applications. Engineered nanoparticles have been used to develop hundreds
of products and are available on supermarket shelves including transparent sunscreens,
light-diffracting cosmetics, penetration enhanced moisturisers, stain and odour repellent
fabrics, dirt repellent coatings, long lasting paints and furniture varnishes, and some food
products (Miller & Kinnear, 2007). It has been speculated that nanoencapsulation may gain
importance in the near future to develop designer probiotic bacterial preparations that
could be delivered to certain parts of the gastro-intestinal tract where they interact with
specific receptors (Paques & van Rijn, 2007). Biodegradable nano/microparticles of poly
(d,l-lactide-co-glycolide) (PLGA) and PLGA-based polymers have been widely explored
as carriers for controlled delivery of macromolecular therapeutics such as proteins,
peptides, vaccines, genes, antigens and some growth factors, etc (Raghavendra et al.,
2008).

20

Nanoencapsulation just like microencapsulation has many applications like protein, DNA
and RNA stabilization, small molecule delivery, extending circulatory half-life, modifying
drug transport, clear liquid formulations, stable colloid dispersions, controlled release,
targeted delivery, triggered release (Persyn & Oxely, 2008).
There has been a rise in concern in the use of nanotechnology as it involves the
manipulation of matter at the scale of atoms and molecules and hence, potentially
introduces some serious risks to human and environmental health. Toxicological literature
suggests that nanoparticles are more reactive, more mobile, and more likely to be toxic to
humans and the environment than larger particles. Preliminary scientific research has
shown that many types of nanoparticles can result in increased oxidative stress which can
result in the formation of free radicals that can lead to cancer, DNA mutation and even cell
death. Additionally, fullerenes, carbon nanoparticles, have been found to cause brain
damage in largemouth bass, a species accepted by regulatory agencies as a model for
defining ecotoxicological effects (Miller & Kinnear, 2007) thus, limiting the use of
nanotechnology extensively.
2.3 SUGAR BEET PECTIN AS WALL MATERIAL
Pectins are a family of plant-derived heteropolysaccharides comprised predominantly of a
linear chain of (14)-linked -D-galacturonic acid residues. These residues may be
partially esterified with a small percentage of rhamnose units to yield branches consisting
of neutral sugars, notably galactose and arabinose (Cho, 2001). Unlike citrus pectins, beet
pectins have ferulic acid groups esterified to some of the neutral sugars in the side-chains
of the so-called hairy regions (Guillon and Thibault, 1990; Micard and Thibault, 1999;
Saulnier and Thibault, 1999). Apples, guavas, quince, plums, gooseberries, oranges and
other citrus fruits, contain large amounts of pectin, while soft fruits like cherries, grapes
and strawberries contain small amounts of pectin (Cho, 2001).
The amount, structure and chemical composition of pectin differs between plants, within a
plant over time and in different parts of a plant (Cho, 2001). Pectins are generally
classified as high methyl ester (HM) pectins and low methyl ester (LM) pectins based on
21

the percentage of ester groups they possess or the degree of esterification (DE). If the DE
of the pectin is greater than 50% it is called as HM pectin, while if it is less than 50% it is
called a LM pectin (IPPA International Pectin Producers Association, 2001).
Pectins are produced commercially as a white to light brown powder, mainly extracted
from citrus fruits, and is used in food as a gelling agent particularly in jams and jellies. It is
also used in fillings, sweets, as a stabilizer in fruit juices and milk drinks and as a source of
dietary fiber (Pectin, 2010). In contrast to other pectins, sugar beet pectin (SBP) is more
interesting because of its unusual emulsifying properties. The emulsifying properties of
SBP have been attributed to its high protein content (up to 11%) and high level of
acetylation (up to 5%) (Leroux et al., 2003; Funami et al., 2007). SBP also contains the
known antioxidant, ferulic acid, which occurs esterified to the oxygen on C-2 of arabinose
or C-6 of galactose (Fry, 1983) (Figure 2). However, Drusch et al., (2007) and Katsuda
et al., (2008) showed that commercial SBP contains significant amounts of transition metal
ions, namely copper and iron, which promote lipid oxidation. Hence, antioxidants were
used in this research study to improve the oxidative stability of microencapsulated fish oil
with SBP as the wall material.

Rhamnose

Methyl groups

Galacturonic acid
Galactose

Arabinose

Figure 2 Schematic structure for generalised sugar beet pectin. Adapted from Morris et al.,
(2010) with some modifications made.

22

CHAPTER 3 - MATERIALS AND METHODS


Fish oil (FO) (Hi-DHA 25N, EPA = 5.3%, DHA =26%, Omega-3 fatty acids =35%, PV =
1.4 meq O2/kg) was obtained from Nu-Mega Ingredients, Melbourne, Australia) and
extra-virgin olive oil (EVOO) (PV= < 20 meq O2/kg) from Boundary Bend Marketing Pty
Ltd, Lara, Australia. Sugar beet pectin (SBP, GENU pectin type BETA, of 60 kDa and
degree of acetylation of 23.8% according to manufacturers specifications) was kindly
provided by CPKelco, Melbourne, Australia. Dried glucose syrup (DGS) (with dextrose
equivalent of 2630 according to manufacturers specifications) was obtained from
Manildra, NSW, Australia. FO and EVOO were stored under nitrogen in dark bottles at
4C (up to 6 month) or at -20C (6-24 month).

3.1. RAW MATERIAL COMPOSITION

3.1.1 SUGAR BEET PECTIN

3.1.1.1PROTEIN CONTENT
The protein content of SBP was determined in triplicate using LECO (FP-2000 LECO
Corp., Michigan, USA) method, and the protein content was calculated using 6.25 as a
conversion factor.

3.1.1.2 IRON AND COPPER CONTENT


The iron and copper content in SBP was analyzed at DTS Laboratories, Melbourne.

3.1.2 FISH OIL AND EXTRA-VIRGIN OLIVE OIL

3.1.2.1 FATTY ACID COMPOSITION

23

The fatty acid composition of the oils prior to encapsulation was determined according to
Christie (2003) with some modification (Shen, 2010). The original and extracted oils were
derivatized using transmethylesterification and subsequently analyzed using a Varian 3400
gas chromatograph (Varian Associates Inc., USA) equipped with a BPX70 column (30 m

0.25 mm i.d., 0.25 m film thickness).

3.1.2.2 TOTAL PHENOLICS CONTENT IN EXTRA VIRGIN OLIVE OIL


The total phenolics content was determined using the method of Hrncirik and Fritsche
(2004). Extra virgin olive oil (2.5 g) was dissolved in 5 ml hexane (EMD Chemicals, USA)
and the phenolics were extracted with 5 ml methanol (Merck Chemicals, Germany)/water
(60:40 vol/vol) for 2 min under nitrogen atmosphere using an automatic shaker. The
mixture was then centrifuged (3500 rpm,10min) for phase separation and an aliquot (0.2
ml) of the methanolic phase was diluted with water to a total volume of 5 ml, followed by
addition of 0.5 ml Folin-Ciocalteu reagent (Sigma Chemicals Co., Australia). After 3 min,
1 ml sodium carbonate solution (35%, wt/vol) was added to the reagent mixture, which
was finally mixed and diluted with water to 10 ml. The absorbance of the solution was
measured after 2 hr against a blank sample using a UV-vis spectrophotometer (UV-1700,
Shimadzu, Suzhou manufacturing co. ltd., China) at a wavelength of 725 nm. The
calibration curve was constructed using standard solutions of caffeic acid (Sigma
Chemicals Co., Australia) within the range of 0.050.5 mg/ml.

3.1.3 OXIDATION STATUS OF FISH OIL AND EXTRA VIRGIN OLIVE OIL

3.1.3.1 ASSESSMENT OF OXIDATIVE STABILITY USING OXIPRESS


The oxidative stability of microcapsules containing 4g oil was assessed under oxygen
pressure (0.5 MPa) at 80C in an ML Oxipres apparatus (DK-8270, Mikrolab Aarhus
A/S, Hjbjerg, Denmark) equiped with Paralog software. The induction period (IP) was
defined as the time (h) required for the oxidation process to be initiated. The IP was
obtained directly from the Oxipres curve and represented the starting point of pressure

24

decline. The slope (-mbar hour-1) indicates the rate at which the sample is oxidized after
the initiation of oxidation. The slope was calculated by drawing a line at the IP point
parallel to the part of the curve recording the change in pressure (Figure 3).

P
t1

Pressure
(MPa)

FIGURE 3 Illustration of oxidation stability measurements using ML OXIPRES as


reported by Trojakova, Reblova & Pokorny, 2001).
P- Oxygen pressure (MPa); t- Reaction time (h); t1- Tangent drawn on the curve
corresponding to the induction period; t2- Tangent drawn on the curve corresponding to the
change in oxygen pressure; p- Cross-section of the two tangents; x- end of induction
period.

3.2 PREPARATION OF EMULSIONS


Encapsulant materials were primarily selected from a range of biopolymers of various
sources such as carbohydrates( Maltodextrins, Dried Glucose syrups, Alginates, Chitosan,
Starches

etc),

proteins

(Sodium

Caseinate,

Whey

Proteins,

Gelatins

etc),

lipids (Hydrogenated fats, Vegetable oils, Palm Stearin etc) and waxes (Shellac or
Carnauba Wax etc) (Smith & Charter, 2010). In our study, DGS was used as a bulking
agent along with SBP to improve stability and microencapsulation of fish oil. Sugars in
association with emulsifying bioploymers have been shown to improve encapsulation
properties. Drusch (2007) successfully produced stable fish oil microcapsules using sugar

25

beet pectin as a wall material and glucose syrup as a bulk one. The same author reported
also that a pectin content of 1- 2% was sufficient for the formation of stable emulsions for
spray drying and proposed that the maximum oil loading in microcapsules containing FO
as the core may be limited to 50 % because of the amount of non-encapsulated fat obtained
in them. Hence, formulations were developed with 2% SBP along with DGS in the
preparation of FO and FO-EVOO emulsions with 25% and 50% oil loading at 30% TS.
SBP (2 g) was dispersed in distilled water at 70C and held at that temperature for a further
15 min before the addition of DGS. For samples with EDTA in the formulation, SBP (2 g)
was dispersed in 85 % of the total distilled water required at 70C and EDTA was added to
chelate the metal ions before the addition of DGS. The solution was subsequently cooled to
60C. The pH was then adjusted by adding 1M NaOH solution to samples whose pH had
to be altered to 6. The amount of NaOH solution was recorded to determine the remaining
water to be added to the pre-heated solution. The solution was subsequently cooled to
60C. FO and EVOO were pre-heated to 60C. A FO-EVOO blend was prepared by
combining the oils at a 1:1 weight ratio. The FO and the FO-EVOO blend were dispersed
into the pre-heated mixtures of SBP and DGS at 60C for 1-2 min using a high shear mixer
(Silverson, London, UK) at maximum speed to obtain a pre-emulsion. The pre-emulsions
were then homogenized at 60C using a high pressure two-stage homogenizer (APV
Rannie AS Homogenizer, Denmark) at 35 and 10 MPa. All emulsions were prepared at
30% total solids. The composition of the pre-emulsion with the percentage of each
component stated in terms of ingredient weight is outlined in Table 2 & 3. The process
flow chart of the experimental plan for the production of spray dried powders has been
presented in Figure 4. FO and FO-EVOO emulsions were prepared at 25% (pH 3) and 50%
(pH 3 & pH 6 ) oil loading in powder and emulsions with EDTA were prepared at 50% oil
loading in powder only at pH 6 yielding a total of 8 different emulsions (FO-25,
FO-EVOO-25, FO-50 (pH 3), FO-EVOO-50 (pH 3), FO-EDTA-50,FO-EVOO-EDTA-50,
FO-50 (pH 6), FO-EVOO-50 (pH 6)) (Figure 5 (A)). All emulsions were prepared in two
runs with duplicate samples in each run.

26

Dispersion of SBP (2%) in distilled water/ buffer solution (pH adjusted to 6)


at 70C
Addition of EDTA then DGS
SBP-DGS mixture cooled to 60C
Adjusted to pH 6 & Addition of FO/FO
EVOO pre-heated to 60C
SBP-DGS-Oil mixture

Pre-emulsion

Homogenization
(35+10 MPa)

Stable Oil Emulsion


(30 % total solids)

Spray Drying

Powder Microcapsules
Figure 4 Process flow chart of the experimental plan for the production of spray-dried
microcapsules.

27

Table 2 Formulation of the fish oil and fish oil-extra-virgin olive oil (1:1 wt ratio)
emulsions at pH 3.

Percentage contents in each emulsion (Ingredient weight)


Constituent
FO-25

FO-50

FO-EVOO-25

FO-EVOO-50

SBP

DGS

20.5

13

20.5

13

FO

7.5

15

3.75

7.5

EVOO

3.75

7.5

Water

70

70

70

70

Note: Moisture content of SBP: 9.92 %, DGS: 6.99 %


Table 3 Formulation of the fish oil and fish oil-extra-virgin olive oil (1:1 wt ratio)
emulsions with and without EDTA at pH 6.

Percentage contents in each emulsion (Ingredient weight)


Constituent

FO-EDTA-50

FO-EVOO-EDTA-50

FO-50

FO-EVOO-50

SBP

EDTA(Na

0.05

0.05

DGS

12.95

12.95

13

13

FO

15

7.5

15

7.5

EVOO

7.5

7.5

Water

70

70

70

70

salt)

Note: Moisture content of SBP: 9.92 %, DGS: 6.99 %

28

(A) Emulsions prior to spray drying

(B) Spray-dried powders

Figure 5 Emulsions (A) and spray-dried powders (B)

3.3 SPRAY-DRIED MICROCAPSULES


The homogenized emulsions were converted to spray-dried microcapsules/powders (Figure
4 (B)) using a Drytec laboratory spray dryer (Tonbridge, UK). The dryer had a water
evaporation rate of 1 L h-1 with a twin fluid nozzle at 2.5 bar atomizing pressure. The feed
was heated to 60C before atomization. The inlet and the outlet air temperatures were
180C and 80C, respectively. Two independent processing runs were carried out. The
process flow chart for the production of microcapsules has been illustrated in Figure 4.

3.3.1 STORAGE CONDITIONS


Aliquots (20 g) of each of the microcapsule formulations were stored in duplicate in
transparent, stoppered, oxygen-permeable 100 ml plastic containers at room temperature
(~25C) for 0, 1, 2 and 3 months. Zero month refers to samples analysed immediately after
the production of the microcapsules without storage. At the end of each storage time the
plastic containers were covered with aluminium foil and transferred to frozen storage
(-18C) prior to further analysis.

29

3.4 EMULSION CHARATERISTICS

3.4.1 PARTICLE SIZE


The particle size distribution of the emulsions was determined by laser light scattering
(Mastersizer 2000G, Malvern Instruments, Worcestershire, UK) using standard optical
parameters, and a refractive index of 1.456. The emulsion was dispersed in recirculating
water within the measuring cell (Hydro SM, Worcestershire, UK) until an obscuration rate
of 10-20% was reached. The particle size measurements were reported as the
surface-volume mean diameter, d3,2.
3.4.2 LIGHT MICROSCOPY
The emulsions were visualized by light microscopy (Olympus BH-2, Anax Pty. Ltd, Japan)
equipped with an attached camera. A drop of Oil-Red-Oil dye (BDH laboratory supplies,
England) was added to each emulsion (~1 ml) and visualized using a 10x objective
magnification. The images were processed using image acquisition software (anaLYSIS
getIT).

3.4.3 VISCOSITY
The viscosity of the emulsions was measured using a cup and bob geometry (Paar Physica
Rheometer, MCR 300, Anton Paar, Austria) at 25C within 2 h of preparation. The shear
rate () was increased in 30 steps from 0.2 to 291 s1. The duration of measurement at each
shear rate was 10 s.

3.4.4 OXIDATIVE STABILITY OF EMULSIONS UNDER ACCELERATED


CONDITIONS
The oxidative stability of the emulsions (equivalent to 4 g oil in sample) was assessed by
following the procedure described in Section 3.1.3.1.

30

3.4.5 ZETA-POTENTIAL OF EMULSIONS


Measurement of zeta potential of FO and FO-EVOO emulsions as a function of pH was
performed at 22C with a Zetasizer Nanoseries (Nano ZS), Malvern Instruments,
Worcestershire,

United

Kindom).

Titrations

were

performed

by

Autotitrator-MPT-2(Malvern Instruments, Worcestershire, United Kindom). Samples were


prepared by diluting (7l emulsion) in buffer solution (20 ml) (the pH of distilled water
was adjusted to 6 by the addition of NaOH and HCl). Diluted emulsions (15ml) were then
filled in sample tube with a magnetic stirrer. A liquid filled pH probe was then inserted
into the sample tube. Three different titrants (0.5M HCL, 0.2 M HCL and 0.5M NaOH)
were connected simultaneously to the sample tube. The system automatically selected the
appropriate concentrations of the titrant to be mixed into the sample to adjust to the
required pH. The diluted emulsion was pumped into a capillary cell by the integral
peristaltic pump and the cell was then inserted into the measurement chamber of Zetasizer.
The program was set to measure up to 100 points between the pH range 6 1 and to titrate
with either acid or base.

3.5 CHARACTERIZATION OF THE SPRAY-DRIED MICROCAPSULES

3.5.1 MOISTURE CONTENT


The moisture content of the spray-dried powders was determined at 80C using a Sartorius
infrared balance (MA30, Sartorius Mechatronics, Gottingen, Germany).

3.5.2 WATER ACTIVITY (AW)


The aw of the spray-dried powders was determined at 25C using a Aqua-Lab Water
Activity Meter (Series 3, Decagon Devices Inc, USA).

3.5.3 PARTICLE SIZE OF THE RECONSTITUTED POWDERS

31

The spray-dried powders (10% w/v) were reconstituted at 70C for 3 h under constant
stirring, and then rested for 1 h at room temperature (~25C). Particle size was determined
as described above (Section 3.4.1).

3.5.4 FATTY ACID COMPOSITION


Fatty acid composition of the oils extracted from the spray-dried microcapsules was
determined as described in Section 3.1.2.1.

3.5.5 DETERMINATION OF FREE OIL CONTENTS


The free (solvent-extractable) oil content of each spray-dried powder was determined
gravimetrically according to Pisecky (1997), except that petroleum ether was used in place
of carbon tetrachloride. In this method, petroleum ether (50 ml) was added to 10 g powder.
The mixture was agitated in a stoppered tube for 15 min. The mixture was filtered
(Whatman No.541) and the solvent evaporated at 60C using a rotary evaporator. The
remaining fat residue was then dried in an oven at 105C for 1 h.

3.5.6 DETERMINATION OF TOTAL OIL CONTENT


The total oil content of the spray-dried microcapsules was determined gravimetrically
according to the Schmid-Bondzyndki-Ratzlaff method (1988) (AS 2300.1.3).

3.5.7 ENCAPSULATION EFFICIENCY


Encapsulation efficiency was calculated after estimating the free and total fat contents in
spray-dried microcapsules using the following formula:
Encapsulation efficiency (%) = 100- (% solvent-extractable fat / % total fat) 100
[equation1]

32

3.5.8 PROPANAL AND HEXANAL ANALYSIS


Aliquots (2 g) of the spray-dried powders were placed in 10 ml headspace vials, sealed and

equilibrated at 60 C for 15 min in a water bath. An aliquot (1 ml) of the headspace was
then analyzed using a Varian 3400 gas chromatograph (Varian Associates,Inc., USA)
equipped with a BPX-70 fused silica capillary column (30 m

0.25 mm i.d., 0.25 m

film thickness) and a FID (flame ionisation detector). The injector and detector

temperatures were 250 C and 275 C, respectively. The oven temperature was

programmed at 60 C and held for 5 minutes. The temperature was then increased to 175 C

at 10 C min-1. The temperature was then further increased to 220

C at 4C min .
-1

Propanal and hexanal were identified using external standards.

3.5.9 OXIDATIVE STABILITY OF THE SPRAY-DRIED MICROCAPSULES


UNDER ACCELERATED STORAGE CONDITIONS
The oxidative stability of the spray-dried microcapsules (equivalent to 4 g oil in sample)
was assessed was assessed by following the procedure described under Section 3.1.3.1.

3.5.10 SCANNING ELECTRON MICROSCOPY (SEM)


The morphology of the spray-dried microcapsules was examined using the SEM (Philips
XL30FEG scanning electron microscope, Eindhoven, Netherlands) in the departments of
Zoology and Biology, The University of Melbourne, Australia.

3.6 STATISTICS
All analytical determinations were carried out at least in duplicate. The results were
reported as the mean standard deviation (SD) of these measurements. The analysis of
variance (One way ANOVA) was performed at 95% confidence level using SPSS 18
(SPSS Inc, Chicago, USA) and LSD (least significance) was used to separate the means.

33

CHAPTER 4 - RESULTS AND DISCUSSION

4.1 PROPERTIES OF RAW MATERIALS

4.1.1 SUGAR BEET PECTIN


The protein content in SBP was 4.9 0.05 (%). This value is similar to that previously
reported by Drusch (2007), and Siew & Williams (2008). However, other reported values
for protein content of SBP ranged from as low as 1.95 % (Leroux, 2003) to as high as ~
10.4 % (Thibault, 1988). This large variation in protein content reflects the sensitivity of
the protein moiety in SBP to extraction and purification processes (Kirby, Dougall &
Morris, 2006).
The iron and copper contents in the SBP were determined to be 310 ppm and 10 ppm,
respectively. This corresponds to a total iron and copper content in the emulsions of
0.000114 moles/L (i.e. 6.4 ppm, comprising 6.2 ppm Fe and 0.2 ppm Cu). Katsuda et al.,
(2008) cited that to maximize the oxidative stability of commercial oils, iron and copper
contents in oil should be as low as possible, with most specifications recommending <0.1
ppm and 0.02 ppm, respectively. The high levels of iron and copper in the SBP used in this
study are likely to promote lipid oxidation, particularly if they are not complexed by
chelating agents. To investigate the effect of the metal ions on lipid oxidation, the
emulsions (at pH 6) were prepared in the presence and absence of the metal chelator,
EDTA. EDTA was added to the emulsions at a level of 0.05% w/v (i.e. 0.000171 moles
EDTA/L emulsion). This equated to an excess of EDTA to metal ions, and thus ensured
that all the metal ions were present in the emulsion as EDTA-metal ion complexes,
preventing their interaction with lipid hydroperoxides.

4.1.2 FISH OIL AND EXTRA VIRGIN OLIVE OIL

4.1.2.1 FATTY ACID COMPOSITION

34

Results of fatty acid composition of FO and EVOO was quantified using GC are shown in
Table 4. The major fatty acids found in FO were palmitic acid (C16:0) (21%), oleic acid
(C18:1) (~14%), eicosapentaenoic acid, EPA (C20:5n-3) (6.6%), and docosahexaenoic
acid, DHA (C22:6n-3) (26.5%).The fatty acid composition of FO is highly dependent on
the age, sex, spawning cycle, season, species and origin of the fish (Donmez, 2009). In this
study, the FO was derived from tuna. The EPA and DHA content of the FO were found to
be very close to that of the manufacturers specification. Gotoh et al., (2006) quantified
fatty acids present in big eye tuna oil, and the reported EPA and DHA percentages were
6.5% and 24% respectively.
Results of fatty acids analysis in EVOO were in agreement with those reported in the
literature. The major fatty acid in EVOO is oleic acid, which can vary from 5583%. It
was present at a level of 80% in the EVOO used in this work. The other major fatty acid
present is palmitic acid and normally its content ranges between 7.520% (Olive oil,
2008). It was 7% in the EVOO used in this work.
Table 4 Fatty acid composition (%) of the initial oils

Fatty Acida
Palmitic acid, C16:0

Fatty acid (%)


Extra Virgin
Fish Oil
Olive Oil
21.09 1.94
6.99 0.13

Oleic acid, C18:1

13.51 1.79

79.54 0.12

Eicosapentaenoic acid, C20:5n-3

6.64 0.92

ND

Docosahexaenoic acid, C22:6n-3

26.54 0.53

ND

Others

32.21 4.12

13.47 0.01

Values are average of duplicate measurements SD. ND = not detected

35

4.1.2.2 TOTAL PHENOLICS CONTENT IN EXTRA VIRGIN OLIVE OIL


The total phenolics content in EVOO was 98.11 0.86 (ppm). This value falls in the
expected range specified by the manufacturer of 70120 ppm).The phenolic content in
EVOO varies largely because of a number of factors like the olive variety, time of harvest,
environmental factors, storage conditions, and refining process, for example (Olive oil
source, 2010). EVOO with a phenolic content as high as 1500 ppm has been reported
(Oliveras-Lpez et al., 2008).

4.1.3 OXIDATION STATUS OF FISH OIL AND EXTRA VIRGIN OLIVE OIL
4.1.3.1 ASSESSMENT OF OXIDATIVE STABILITY USING THE OXIPRES
The Oxipres results of the original FO, EVOO and their 1:1 blend are given in Table 5.
The IP for these oils under the accelerated conditions (80C, 0.5 bar oxygen pressure) used
in Oxipres assessments were 11.85 0.07 h, >100 h and 16.3 0.14 h, respectively, for
FO, EVOO and the 1:1 mixture of FO-EVOO. The uptake of oxygen after the IP was -210
9.19 mbar h-1 and -119 4.24 mbar h-1, respectively, for FO and a 1:1 mixture of
FO-EVOO. EVOO was monitored for up to 105 h, during which time no detectable
oxidation was observed. Hence no oxidation rate (slope) for this sample was provided.
This confirmed that olive oil was more stable to oxidation than FO. That might be
expected based on the fatty acid composition alone as the omega-3 fatty acids in FO are
very prone to oxidation. In addition, olive oil contains phenolic compounds, which are
known to be antioxidative (Psomiadou & Tsimidou, 2002).

36

Table 5 Oxidative stability of original oils during accelerated storage (80C, oxygen
pressure of 0.5 bar)
a

Induction period (h)

Slope (mbar h-1)

FO

11.85 0.07

-210 9.19

EVOO

>100 0.00

FO-EVOO (1:1 wt ratio)

16.3 0.14

-119 4.24

Sample

Values are average of duplicate measurements SD

4.2 EFFECT OF PARTIAL SUBSTITUTION OF FISH OIL WITH EXTRA


VIRGIN

OLIVE

OIL

ON

EMULSION

AND

MICROCAPSULE

CHARACTERISTICS
The average particle size (d3,2) of the different emulsions ranged between 0.410.43 m
(Figure 6). These values were comparable to those reported for oil-in-water emulsions
made by emulsifying 20% w/w orange oil with 2% SBP (Leroux et al., 2003) and
homogenised with three passes at 200 bars. SBP dissolved in water alone had a (d3,2) of
2.38 0.31 m, and contributed to the particle size of the emulsions. This observation also
accounts for the minor peak present in the particle size distribution graph of the emulsions.
There was no significant difference (P > 0.05) in the average particle sizes between the
emulsions containing 25% and 50% oil loading. This indicated that the presence of 2%
SBP was sufficient emulsifier to produce physically stable fish oil-in-water emulsions
containing 15% oil (wet basis). Similarly, other studies (Drusch, 2007; Drusch et al., 2007;
Leroux et al., 2003; Nakauma et al., 2008; Siew & Williams, 2008) have reported that
1.52% SBP was sufficient to produce oil-in-water emulsions containing up to 18% oil
(wet basis). For example, Nakauma et al., (2008) made stable oil-in-water emulsions with
1.5% SBP and 15% w/w oil (medium chain triglyceride) homogenised with two passes at
50 MPa.

37

Solubility is an important criterion in choosing a wall material. If the wall material is


highly insoluble it may precipitate. When this occurs the wall material (pectin) is not able
to move to the oil/water interface and the emulsifying power of the pectin will be reduced.
The particle size distribution of the emulsions revealed that SBP was fully dispersed and
that there were no large aggregates of undissolved SBP (Figure 6). Visual examination of
the emulsions by light microscopy also showed the emulsions were homogeneous and no
visual particle aggregation or free (nonencapsulated) oil was present (Figure 7).

7
6

Volume (%)

5
4
3
2
1
0
0.01

0.1

10

100

1000

Particle size (m)

Figure 6 Particle size distributions of the original (fresh) emulsions, prior to spray drying.
Values are the averages of triplicate measurements SD;
(FO-EVOO-50);

(FO-50);

(FO-25);

(FO-EVOO-25).

Emulsion stability is an important property to be considered in the microencapsulation


process. Emulsion stability arises from steric repulsion of aggregates by a hydrated layer
(Willats, Knox & Mikkelsenc, 2006). Pectin stabilises an emulsion by creating a layer
around the particles at the interface. It has also been reported that stable emulsions have
been formed when using low molecular weight pectins (70 kDa) (Akhtar et al., 2002) and
also because of the link between charged molecules (pectin) and protein (Leroux et al.,
2003).

38

Figure 7 Brightfield micrographs of (A) FO-25 (7.5% oil, wet basis) and (B)
FO-EVOO-50 (15% oil, wet basis) emulsions at pH 3.
The apparent viscosities of the emulsions as a function of shear rate are shown in Figure 8.
All the formulations displayed shear thinning or pseudo-elastic behaviour, which is a
common feature of oil-in-water emulsions (McClements, 1999).
Emulsions containing 15% oil (wet basis) had greater apparent viscosities than
corresponding emulsions containing 7% oil, as expected. However, emulsions containing
either FO or a 1:1 FO-EVOO mixture with the same gross formulation had different
viscosities. It is possible that the factor responsible for these differences was the presence
of phenolic compounds in the olive oil. For example, polyphenols are known to complex
with protein (Spencer et al., 1988). An interaction between polyphenols and residual
protein in SBP can alter the interfacial properties of the oil droplet or the properties of the
bulk phase and also affect viscosity of the emulsion. Other studies showed that
polyphenol--casein complexes alter the properties of an air/water interface and also exist
as complexes in the bulk solution (Aguie-Beghin et al., 2008).

39

1800
1600

Viscosity (mPa/s)

1400
1200
1000
800
600
400
200
0
0

50

100

150

200

250

300

Shear Rate (1/s)

Figure 8 Apparent viscosities of fish oil and fish oil-extra virgin olive oil (1:1 wt ratio)
emulsions (30% TS; 2% SBP, 20.5% DGS, 7.5% total oil and 30% TS; 2% SBP, 13%
DGS, 15% total oil), intended for manufacture of spray-dried microcapsules containing 25
and 50% oil, as a function of shear rate. Values are the average of duplicate measurements
SD;

(SBP-DGS-FO-50);

(SBP-DGS-FO-25);

(SBP-DGS-FO-EVOO-50);

(SBP-DGS-FO-EVOO-25).
Zeta potential measurements of the FO and FO-EVOO emulsions (2% SBP, 13% DGS,
15% total oil) intended for the manufacture of spray-dried microcapsules with 50% oil
loading, as a function of pH, were determined using an autotitrator to predict the stability
of the emulsions (Table 6). The emulsions were measured at pH values between 6 and 1.
The zeta potential of a particle can be defined as the overall charge that the particle
acquires in a particular medium. Depending on the zeta potential of the particle, the
stability of an emulsion can be explained. If the particles in a suspension have a large
negative or positive zeta potential then they will tend to repel each other and resist the
formation of aggregates. Conversely, particles of opposite charge tend to associate. The
dividing line between stable and unstable suspensions is generally taken at either +30 mV

40

or -30 mV. Particles with zeta potentials more positive than +30 mV or more negative than
-30 mV are normally considered stable (Zeta Potential Measurement Using an Autotitrator
from Malvern Instruments and the Effect Of pH, 2005). The autotitration plots revealed
that both FO and FO-EVOO emulsions were most stable at pH greater than 5 (Figure 9 &
10). Both emulsions would be less stable as pH is reduces below 5, and would be least
stable at pH ~1.5 as the iso-electric point range for the emulsions was between 1.2 and 1.5.
It has been observed that at pH less than 5 the zeta potential of the emulsions decreases
rapidly and eventually reached an iso-electric point after pH 1.5.
Table 6 Zeta potential values of FO and FO-EVOO emulsions measured at 22 C.
a

Sample

Zeta potential (mV)

FO-50 (pH 3)

-12.12 0.39

FO-50 (pH 6)

-39.8 0.51

FO-EVOO-50 (pH 3)

-13.56 0.47

FO-EVOO-50 (pH 6)

-41.4 0.66

Values are average of 5 measurements SD

After manufacture, the moisture content of the spray-dried microcapsules made with SBP
and containing 25% and 50% oil loadings ranged from 1.61.8%, with a water activity of
~0.27 (Table 7).The maximum moisture specification for most spray-dried powders in the
food industry was between 34 % (Masters, 1991). The encapsulation efficiency of the
spray-dried microcapsules made with SBP containing 25% and 50% of FO and FO-EVOO
blend was >90%. A significant difference (P < 0.05) in the encapsulation efficiencies of
the microcapsule powders containing different oil loadings was recorded (Table 7). The
microcapsule powders with 50% oil loading had lower encapsulation efficiency (~90%)
than those with 25% oil loading (encapsulation efficiency ~98%). In concurrence with the
solvent-extractable fat contents, encapsulation efficiency was not affected by the
composition of the microencapsulated oil.

41

10

Zeta Potential (m V)

0
0

-10
-20
-30
-40
-50
pH

Figure 9 Autotitration curve for FO emulsion with zeta potential measured as a function of
pH at 22 C.

5
0

Zeta Potential (m V )

-5 0

-10
-15
-20
-25
-30
-35
-40
-45
pH

Figure 10 Autotitration curve for FO-EVOO emulsion with zeta potential measured as a
function of pH at 22 C.

42

Table 7 Properties of spray-dried powders containing microencapsulated fish oil and fish
oil-extra virgin olive oil (1:1 wt ratio) (25% and 50% oil loading) at pH 3.
a

Property

Microcapsule

FO-25

FO-EVOO-25

FO-50

FO-EVOO-50

Moisture content

1.83 0.02

1.73 0.01

1.66 0.01

1.62 0.03

Water activity (aw)

0.27 0.02

0.26 0.01

0.27 0.02

0.27 0.00

Total oil (%)

24.39 0.06

24.38 0.06

49.09 0.14

49.22 0.12

Solvent-extractable oil
(% of free oil in
powder)
Encapsulation
efficiency (%)

0.53 0.01

0.52 0.00

4.99 0.01

5.00 0.01

97.85 0.04a

97.87 0.04a

90.43 0.09b

90.42 0.08b

0.42 0.00

0.43 0.00

0.41 0.00

0.43 0.00

Mean particle
diameter, d32, prior to
spray drying
a

Values are averages of triplicate analysis SD of 2 individual runs

Means within the row followed by different superscript letters differ significantly at P = 0.05

The amount of solvent-extractable fat in the spray-dried powders with the same oil loading
but different oil composition was comparable (Table 7). The FO and FO-EVOO powders
gave ~2% solvent-extractable fat as a percentage of total fat in powder at 25% oil loading,
and ~10% solvent-extractable fat at 50% oil loading. The amount of free oil in the powders
can generally be described as the amount of oil that can be found on the surface of the
powder particles and can include oil that is within cracks in the powder particle that is
readily accessible to solvent (Buma, 1971). Several studies had been conducted to evaluate
the factors affecting the oxidative stability of spray-dried powders containing encapsulated
oil (Velasco, Dobarganes & Mrquez-Ruiz, 2003; Drusch et al., 2007; Mrquez-Ruiz,
Velasco & Dobarganes, 2003). However, due to the different matrix and core materials
used, and the various methods and conditions employed to monitor oil oxidation, no
consensus has yet been reached. Velasco, Dobarganes and Marquez- Ruiz (2003) stated
that lipid distribution in a microcapsule is an important factor influencing oxidation in
encapsulated oil. They discussed various studies conducted to evaluate the oxidation of
43

microencapsulated oils during storage by separate extractions of surface and encapsulated


oil fractions. They demonstrated that it could not be hypothesised that surface oil oxidation
occurs at a faster rate than encapsulated oil even though it is more exposed to oxygen and
had no protection by the matrix. Interestingly, Drusch & Berg (2008) investigated the
oxidative storage stability of microencapsulated FO prepared under various spray drying
conditions and concluded that the surface oil protects other solvent-extractable oil fractions
(e.g. fat within capillary pores of the microcapsules) against oxidation. Furthermore, these
authors revealed that solvent-extractable oil cannot be used to predict shelf-life of
microencapsulated oils.
The moisture content and aw of the spray-dried microcapsules containing 25% and 50% oil
loadings after 3 months storage at ambient conditions were measured (Table 8). A general
increase in moisture content and aw, approximating that of the external environment (aw
~0.5), was observed with increased storage duration. The highest percentage increase in
moisture content was recorded for microcapsules with 50% oil loading.
The particle size distributions and average particle size values (d3,2) of the reconstituted
spray-dried microcapsules at 0, 1, 2 and 3 month storage under room temperature
conditions were determined (Table 9, Figure 11). The particle size distribution for the
fresh powders reconstituted before storage showed that they had particle sizes similar to
those of the respective emulsions prior to spray drying. However, there was a general
increment in the particle size distribution of the reconstituted microcapsules during storage
at room temperature. The data revealed a positive relationship between the rate of
increment in particle size and storage time. Additionally, microcapsules containing 25% oil
loading generally had smaller particle sizes than those containing 50% oil loading under
same storage conditions.

44

Table 8 Changes in moisture content and water activity in FO and FO-EVOO


microcapsules (25% and 50% oil loadings) over storage time.

Water activity (aw)

Moisture content (%)


a

Microcapsule

0 month

3 month

0 month

3 month

FO-25

1.83 0.02

2.18 0.05

0.27 0.02

0.49 0.00

FO-50

1.66 0.01

2.06 0.03

0.27 0.02

0.50 0.01

FO-EVOO-25

1.73 0.01

1.95 0.01

0.26 0.01

0.47 0.01

FO-EVOO-50

1.62 0.03

1.93 0.02

0.27 0.00

0.48 0.00

Values are averages of duplicate analysis SD of 2 individual runs

Table 9 Particle size of the reconstituted spray-dried microcapsules containing fish oil
and fish oil-extra virgin olive oil (1:1 wt ratio) during storage of the powders in stoppered
flasks at room temperature (~25C).
Mean particle diameter, d3,2 (m) a
Microcapsule

0 month

1 month
a

2 month
b

3 month
a

2.91 0.10b

FO-25

0.50 0.01

FO-50

0.70 0.00 c

0.78 0.00c

2.06 0.08c

3.06 0.05c

FO-EVOO-25

0.50 0.00 a

0.50 0.00a

0.90 0.01a

1.62 0.01a

FO-EVOO-50

0.69 0.00 b

0.79 0.01d

1.69 0.03b

1.69 0.02a

0.50 0.00

0.92 0.00

Values are averages of triplicate analysis SD of 2 individual runs

Means within the row followed by different superscript letters differ significantly at P = 0.05

45

5
4
3
2
0
0.01

10

100

3
2

0.1

Particle size (m)

III

0.01

1000

6
Volume (%)

Volume (%)

0.1

4
3
2
1

10

100

1000

Particle size (m)

IV

5
4
3
2
1

0
0.01

II

6
Volume (%)

Volume (%)

0
0.1

10

100

1000

0.01

Particle size (m)

0.1

10

100

1000

Particle size (m)

Figure 11 Particle size distributions of reconstituted spray-dried microcapsules at 0, 1, 2

and 3 month (AD) stored under room temperature conditions (~25C, exposed to light).
Values are the averages of triplicate measurements SD;
(FO-EVOO-50);

(FO-50);

(FO-25);

(FO-EVOO-25).

Visual inspection of the microcapsule powders revealed that the samples stored for 2 and 3
months contained clumps that were not evident in those of the fresh microcapsule powders
(0 month), or those stored for 1 month. These observations suggested the development of
cohesive interactions between particles over the period of storage time. A number of
variables are known to contribute to the time consolidation effects on food powders.
Storage temperature, exposure of powders to moisture content in air, physical and chemical
changes in the powder sample during storage time, and variations in powder bulk density
have all been found to contribute significantly to the time-consolidation effects of powders
(Teunou, 2000; Fitzpatrick 2004; Onwulata 2005).

46

The scanning electron micrographs of the spray-dried microcapsules revealed a


polydisperse particle size distribution between ~125 m in diameter (Figure 12). The
microcapsules were irregular in shape with wrinkled surfaces. The extent of surface
wrinkling has been previously associated with the concentration of SBP in the feed
emulsion (Drusch, 2007). The wrinkled surface may also be attributed to the mechanical
stresses by uneven drying at different parts of the liquid emulsion droplets at the early
stages of drying (Moreau & Rosenberg, 1993). Considerable agglomeration of the particles
was observed in powders after 3 months of storage, irrespective of the oil loading or oil
composition. Small pores were also noticed on the surfaces of a portion of particularly in
the microcapsule powders that had been stored for 3 months. The large number of pores
evident in the stored microcapsule powders indicates degradation of the matrix on storage
and consequent release of the microencapsulated oil.

Figure 12 Scanning electron microscopy images of the fish oil (A, C) and fish oil-extra

virgin olive oil (B, D) spray-dried powders (25% oil loading), prior (A, B) and after 3
month storage (C, D) under ambient conditions (~25C, 0.5 aw).

47

4.3 EFFECT OF EDTA ON EMULSION AND MICROCAPSULE


CHARACTERISTICS

The average particle size (d3,2) of the FO and FO-EVOO emulsions, with and without
EDTA, ranged between 0.350.36 m. The particle size distributions of the emulsions
revealed a multimodal distribution of particles, with the majority of particles ranging in
size from ~0.21 m in diameter, and a minor proportion ranging from 110 m (Figure
13). As mentioned in (section 4.2), SBP dissolved in water had an average particle size

(d3,2) of ~ 2.5 m and contributed to the particle size of the emulsions. A slight difference
in mean particle size distributions was observed in FO and FO-EVOO-EDTA when
compared to FO-EVOO and FO-EDTA emulsions. That indicated that the addition of
EDTA as chelating agent had very little influence on the mean particle size of emulsions.
The average particle size (d3,2) of all the emulsions were determined after 1 month of
storage at 22C, and ranged between 0.35-0.36 m. That indicated that all emulsions were
stable to droplet aggregation, as evidenced by the lack of change in the average particle
size values, (d3,2) of the fresh emulsions and corresponding emulsions after 1 month
storage. The absence of droplet aggregation could be attributed to the strong repulsive
forces between droplets which were large enough to overcome the various interactions
(mainly Van der Waals forces and hydrophobic interactions) between droplets. It was also
suggested that as pH was increased from the natural pH of pectin solution (pH 3) to pH 6,
the ratio of protonated and deprotonated carboxyl groups present on pectin molecule would
decrease, thereby increasing the negative charge of the pectin molecules. These changes in
pH values increased repulsion between the emulsion droplets and stabilises the droplets
against aggregation.
Visual examination of the emulsions by brightfield microscopy showed that all emulsions
were homogeneous without any visible particle aggregation or any marked presence of
free (nonencapsulated) oil (Figure 14). Based on the results and literature findings
discussed in section 4.2, those results again established that 2% SBP was sufficient to
produce fine and stable oil-in-water emulsions containing at least 15% w/w oil.

48

10
9
8

Volume (%)

7
6
5
4
3
2
1
0
0.01

0.1

10

100

1000

Particle size (m)

Figure 13 Particle size distributions of the original (fresh) emulsions (15% oil, wet basis),

with and without EDTA, at pH 6. Values are the averages of triplicate measurements SD;
(FO-EDTA-50);

(FO-EVOO-EDTA-50);

(FO-50);

(FO-EVOO-50).

Figure 14 Brightfield micrographs of FO-EDTA (A) and FO-EVOO-EDTA (B) emulsions

with 15% oil, at pH 6.


All the emulsions exhibited shear thinning behaviour with the viscosities decreasing with
increase in shear rate (Figure 15). There were some slight differences in the viscosities
noticed between the emulsions. It was observed that FO and FO-EVOO-EDTA emulsions
had slightly lower viscosities when compared to FO-EVOO and FO-EDTA over the shear

49

rate range measured. The presence of high concentrations of metal ions in SBP (Drusch et
al., 2007; Katsuda et al., 2008) and phenolic compounds in EVOO (Cicerale et al., 2009;
Frankel, 2010) could be the underlying cause for the differences in the flow properties of
FO emulsions containing both EVOO and EDTA when compared to FO emulsions
containing either EVOO or EDTA. Also, the differences in interfacial properties of FO and
FO-EVOO emulsions stabilised with SBP have already been established and discussed in
relation to polyphenol-protein complexation (section 4.2).
1000
900

Viscosity (mPa/s)

800
700
600
500
400
300
200
100
0
0

50

100

150

200

250

300

Shear rate (1/s)

Figure 15 Apparent viscosities of fish oil and fish oil-extra virgin olive oil (1:1 wt ratio)

emulsions (15% oil, wet basis) with and without EDTA at pH 6, as a function of shear rate.
Values are the average of duplicate measurements SD;
(FO-EVOO-EDTA-50);

(FO-50);

(FO-EDTA-50);

(FO-EVOO-50)

The moisture content of the spray-dried powders made with SBP (with and without EDTA)
at pH 6 with 50% oil loadings ranged from 2.832.96 %, with an aw of ~0.3 (Table 10).
The moisture content values were typical of that for most spray-dried powders. The
maximum moisture specification for most spray-dried powders in the food industry is
between 34 % (Masters 1991). This is in order to achieve microbiological stability. The
moisture content and aw of the spray-dried powders made with SBP (with and without

50

EDTA) at pH 6 with 50% oil loadings was also determined after 3 months of storage
(Table 11). An increase in moisture content and aw was observed in all microcapsules and

the percent increase in the moisture content and aw was found to be highest in
microcapsules which did not contain EDTA. That increment in moisture content was
expected as the sample containers in which the powders were stored were permeable to
moisture and the aw of the atmosphere (~0.5) was greater than that of the freshly
manufactured powders.
The spray-dried powders with and without EDTA at pH 6 and 50% oil loading gave ~10%
solvent-extractable fat of total fat in powder. The effect of solvent-extractable fat on lipid
oxidation has been investigated and it has been concluded that it cannot be used to predict
the shelf-life of microencapsulated oils (Drusch & Berg, 2008; Velasco, Dobarganes &
Mrquez-Ruiz, 2003). Furthermore, since the amount of solvent-extractable fat was
considerably small, it was unlikely to correlate with the oxidative stability of the
microencapsulated oils. Also, the spray-dried powders were found to be free-flowing,
indicating that the powders did not have high free fat on the surface.
The encapsulation efficiency of the spray-dried microcapsules with and without EDTA at
pH 6 with 50% oil loading was ~90%. There was no difference in the encapsulation
efficiencies of microcapsules without EDTA and microcapsules with EDTA. Thus, it could
be concluded that neither the addition of EDTA nor the oil composition in the core of the
microcapsule had any effect on either solvent-extractable fat or encapsulation efficiency.
The particle size distribution and d3,2 values of the reconstituted spray-dried microcapsules
at 03 month storage under room temperature conditions was determined (Figure 16 &
Table 12). A general increase in particle size was observed in all microcapsules over

storage time at room temperature. The particle size values for FO and FO-EVOO
microcapsules over storage time were similar indicating no influence of type of oil/oils in
the core of the microcapsule on particle size.

51

Table 10 Properties of powders containing microencapsulated fish oil and fish oil-extra

virgin olive oil blends (1:1 wt ratio) with and without EDTA at pH 6, 50% oil (dry basis).
a

Property

Microcapsule

FO-EDTA

FO-EVOO-

-50

EDTA-50

2.83 0.01

2.89 0.02

2.98 0.03

2.96 0.04

Water activity (aw)

0.30 0.01

0.30 0.00

0.29 0.00

0.29 0.00

Total oil (%)

49.27 0.12

49.22 0.13

49.29 0.17

49.29 0.03

Solvent-extractable

4.73 0.00

4.73 0.00

4.72 0.00

4.72 0.00

90.41 0.08

90.40 0.09

90.41 0.12

90.42 0.02

0.36 0.00

0.35 0.00

0.35 0.00

0.35 0.00

Moisture content

FO-50

FO-EVOO
-50

(%)

oil
(% of free oil in
powder)
Encapsulation
efficiency (%)
Mean particle
diameter, d3,2,
prior to spray
drying
a

Values are averages of duplicate analysis SD of 2 individual runs

The average oil droplet size (d3,2) of all the reconstituted microcapsules at zero month was
~0.80.9 m. This showed that the particle size of zero month powders were slightly
greater than the particle size of the corresponding emulsions prior to drying. This increase
in particle size may be due to changes to the interfacial structure of SBP, causing
coalescence of oil droplets and also, some pectin aggregation during the spray drying
process. According to Rees (1969), the pectin chain interactions are promoted under low
aw conditions. Kirby, MacDougall, & Morris (2006) used atomic force microscopy (AFM)

52

to assess the structure of pectin molecules and also, to characterise protein-polysaccharide


interactions in SBP. They reported that when SBP was deposited on a mica surface it
formed small aggregates. This aggregation process was believed to be because of the
adsorbed protein-protein complexes at the air-water interface as the solvent is removed
during drying. Particle size (d3,2) values for FO and FO-EVOO microcapsules with and
without EDTA over storage time were similar indicating no influence of EDTA. The
average oil droplet size range increased from 0.840.90 0.01 m for microcapsules
stored for 1 month, to 1.271.35 0.04 m after 2 months and 1.441.46 0.01 m after
3 months. Visual inspection of all the stored microcapsules revealed that there were a
number of lumps in 2nd and 3rd month stored powders and absolutely no lumps in either the
0 or 1st month of storage. Also, microcapsules stored for 2 months and 3 months were
increasingly difficult to reconstitute. The increase in moisture content and aw during
storage time in the microcapsules (Table 11) and the increase in particle size / aggregation
of particles could be attributed to the change in aw over storage time. When moisture is
taken up by the powders, plasticisation of the encapsulant material occurs, leading to
increased particle-particle interaction and consequent agglomeration of powder particles
into lumps. The phenomenon of stickiness, caking or crystallisation that effect the quality
of spray-dried powders have been shown to be influenced by time, temperature and
moisture and have been numerously reported to have been encountered during the
production and storage of food powders (Levine & Slade, 1986; Bhandari et al., 1997;
Ozkan, Withy & Chen, 2003).

SEM of the microcapsules indicated that the stored powders had an increase in
agglomeration in the powder particles when compared to the SEM images of the fresh
powders (Figure 17). The microcapsules were regular in shape with wrinkled surface and
no cracks or pores. This surface wrinkling has been associated to be caused by the
mechanical stress caused during spray drying process and has been discussed in section
4.2. Most images had microcapsules as discrete units with no incomplete or damaged shell.

53

Table 11 Changes in moisture content and water activity of FO and FO-EVOO

microcapsules (50% oil, dry basis) (with and without EDTA) at pH 6 on storage at room
temperature (~25C).
a

Microcapsule

Moisture content (%)

Water activity (aw)

0 month

3 month

0 month

3 month

FO-EDTA-50

2.83 0.01

3.25 0.01

0.30 0.01

0.49 0.01

FO-EVOO-EDTA-50

2.89 0.02

3.01 0.06

0.30 0.00

0.48 0.00

FO-50

2.98 0.03

3.53 0.06

0.29 0.00

0.49 0.01

FO-EVOO-50

2.96 0.04

3.19 0.12

0.29 0.00

0.49 0.00

Values are averages of duplicate analysis SD of 2 individual runs

Table 12 Particle size values of the reconstituted spray-dried microcapsules (50% oil, dry

basis), with and without EDTA at pH 6, on storage at room temperature (~25C).


Mean particle diameter, d3,2, (m)a
Microcapsule

0 month

1 month

2 month

3 month

FO-EDTA-50

0.87 0.01b

0.88 0.01a

1.30 0.03ab

1.46 0.03a

FO-EVOO-EDTA-50

0.84 0.00a

0.89 0.00a

1.35 0.04b

1.44 0.03a

FO-50

0.89 0.00b

0.89 0.00a

1.28 0.06a

1.44 0.04a

FO-EVOO-50

0.87 0.01c

0.90 0.02a

1.27 0.04a

1.46 0.03a

Values are averages of triplicate analysis SD of 2 individual runs

Means within the row followed by different superscript letters differ significantly at P = 0.05

54

5
4
3
2
1
0
0.01

5
4
3
2
1

0.1

10

100

0
0.01

1000

10

100

1000

6
5
4
3
2

6
Volume (%)

Volume (%)

0.1

Particle size (m)

Particle size (m)

5
4
3
2
1

1
0
0.01

6
Volume (%)

Vo lum e (% )

0.1

10

Particle size (m)

100

1000

0
0.01

0.1

10

100

1000

Particle size (m)

Figure 16 Particle size distribution of reconstituted spray-dried microcapsules (50% oil,

dry basis), with and without EDTA at pH 6 after 0, 1, 2 and 3 month (AD) storage at
room temperature (~25C). Values are the averages of triplicate measurements SD;
(FO-EDTA-50);

(FO-EVOO-EDTA-50);

(FO-50);

55

(FO-EVOO-50).

Figure 17 Scanning electron microscopy images of the fish oil-extra virgin olive oil

spray-dried powders (50% oil, dry basis), with (A, C) and without (B, D) EDTA, at pH 6,
prior (A, B) and subsequent (C, D) to storage for 3 months at room temperature (~25C).

4.4

EFFECT

OF

pH

ON

EMULSION

AND

MICROCAPSULE

CHARACTERISTICS

The mean particle diameters (d3,2) of FO and FO-EVOO emulsions at pH 3 with 50% oil
loading were 0.41 0.00 m and 0.43 0.00 m, respectively. In comparison, both the FO
and FO-EVOO emulsions at pH 6 with 50% oil loading had a slightly smaller mean
particle diameter of 0.35 0.0 m. The difference in particle size observed for emulsions
at pH 3 and pH 6 indicated a slight influence of pH on the emulsion particle size. It has
already been discussed in section 4.2 that 2% SBP was sufficient to prepare stable FO
emulsions with 50% oil loading. The small difference in mean particle diameters of the

56

emulsions made at pH 3 and pH 6 suggests that there was sufficient emulsifier present to
prepare finely-dispersed emulsions and any change in conformation / charge of the SBP as
pH is raised from pH 3 to 6 did not have a marked effect on the emulsifying capacity of
SBP.
The particle size distributions of FO and FO-EVOO emulsions at pH 3 and pH 6 showed a
non-Gaussian, largely uni-modal distribution (Figure 18). It was observed that there was a
shoulder on the major peak in the distribution curve in all the 4 emulsions, corresponding
to particles having a mean diameter between 110 m, and this was more prominent in the
emulsions prepared at pH 3. This shoulder could be because of the presence of small
aggregates caused by the interaction between particles. When the pH of a solution is
altered the proteins present in the mixture may experience some conformational changes
and as they approach their isoelectric point they precipitate (Vaclavik & Christian, 2008).
This precipitation may cause the proteins to loose their emulsifying capacity as the
precipitated particles collide, stick and break apart until a stable mean particle size is
reached and in the due process causes some aggregation.
9
8

V o l u m e (% )

7
6
5
4
3
2
1
0
0.01

0.1

10

100

1000

Particle size (m)

Figure 18 Particle size distributions of the original (fresh) fish oil and fish oil-extra virgin

olive oil (1:1 wt ratio) emulsions made at pH 3 and 6. Values are the averages of triplicate
measurements SD;

(FO-50 (pH 3));

(FO-EVOO-50 (pH 3));

(FO-EVOO-50 (pH 6)).

57

(FO-50 (pH 6));

The brightfield micrographs of the emulsions are shown in Figure 19. The micrographs of
pH 3 and pH 6 emulsions were comparable and appeared to have no large aggregates of
undissolved SBP. Additionally, all the emulsions were homogeneous with no indication of
any unencapsulated oil.

Figure 19 Brightfield micrographs of (A) fish oil (pH 3) and (B) fish oil-extra virgin olive

oil (1:1 wt ratio) emulsions (pH 6) with 15% oil.


In general, all the emulsions exhibited shear thinning behaviour with the viscosities
decreasing with increase in shear rate (Figure 20). FO and FO-EVOO emulsions

having

15% oil (wet basis) that were adjusted to pH 6 were less viscous than the FO and
FO-EVOO emulsions at pH 3, indicating the influence of pH. The zeta potential data
showed that the SBP-stabilised oil droplets had an isoelectric point of ~!1.5. Thus the
droplets were more negatively charged at pH 6 (-potential = -41.5 2.12 mV) than pH 3
(-potential = -14.0 1.41 mV). As the pH is increased to pH 6, the carboxyl groups on the
pectin molecule (as well as protein) lose H+ and therefore the negative charge on the SBP
will increase, resulting in increased electrostatic repulsion. At pH 3, there is less charge
(smaller zeta potential) and therefore the tendency for the particles to aggregate is
increased. The increased interaction between particles lead to an increase in viscosity of
the emulsions made at pH 3.

58

1800
1600
Viscosity (mPa/s)

1400
1200
1000
800
600
400
200
0
0

50

100

150

200

250

300

Shear rate (1/s)

Figure 20 Apparent viscosities of fish oil and fish oil-extra virgin olive oil (1:1 wt ratio)

emulsions with 15% oil at pH 3 and pH 6, as a function of shear rate. Values are the
average of duplicate measurements SD;
(FO-50 (pH 6));

(FO-50 (pH 3));

(FO-EVOO-50 (pH 3));

(FO-EVOO-50 (pH 6)).

The moisture content of the FO and FO-EVOO powders at pH 6 with 50% oil loading was
2.98 0.03% and 2.96 0.04%, respectively at 0 month storage (Table 13). The moisture
content of the FO and FO-EVOO powders at pH 3 with 50% oil loading was 1.66
0.01% and 1.62 0.03%, respectively. A water activity of ~0.3 was found for all the
powders. The higher moisture content in pH 6 microcapsules was directly related to the
smaller particle size diameter. Smaller particle size provided the larger surface area and
facilitated entrapping more water molecules between particles.
Moisture content and aw of FO and FO-EVOO powders made at pH 3 and pH 6 with 50%
oil loading was determined after 3 months of storage under ambient conditions to assess
their physical stability. Both properties increased with increased storage duration. It was
observed that the percent increase in moisture content and aw was greater in microcapsules
made at pH 3 (Table 14) and as pointed out previously (section 4.3), that was probably due
to absorption of moisture from the surrounding environment (aw ~0.5) as the samples were
stored in (plastic) containers that were permeable to moisture. The moisture content

59

measurements were in agreement with the results of aw, with the aw increasing with the
increase in moisture content.
FO and FO-EVOO microcapsules prepared at pH 3 and pH 6

with 5 % oil loading both

had ~10% solvent-extractable fat and encapsulation efficiencies of ~90% (Table 13). Thus,
it can be concluded that neither pH nor the type of oil/oils in the core of microcapsule had
any effect on either solvent-extractable fat or encapsulation efficiency.
It can be collectively concluded that there were no differences among the microcapsules at
pH 3 and pH 6 when total oil percentage, extractable oil percentage and encapsulation
efficiency was compared.
Table 13 Properties of powders containing microencapsulated fish oil and fish oil-olive oil

(1:1 wt ratio) at pH 3 and pH 6 with 50% oil loading.


a

Property

Microcapsule

Moisture content (%)

FO-50
(pH 3)
1.66 0.01

FO-EVOO-50
(pH 3)
1.62 0.03

FO-50
(pH 6)
2.98 0.03

FO-EVOO-50
(pH 6)
2.96 0.04

Water activity (aw)

0.27 0.02

0.27 0.00

0.29 0.00

0.29 0.00

Total oil (%)

49.09 0.14

49.22 0.12

49.29

49.29 0.03

0.17
Solvent-extractable oil (%)

4.99 0.01

5.00 0.01

4.72 0.00

4.72 0.00

Encapsulation efficiency

90.43 0.09

90.42 0.08

90.41

90.42 0.02

(%)
Mean particle diameter,
d3,2, prior to spray drying
(m)
a

0.12
0.41 0.00

0.43 0.00

Values are averages of duplicate analysis SD of 2 individual runs

60

0.35 0.00

0.35 0.00

Table 14 Changes in moisture content and water activity of fish oil and fish oil-extra

virgin olive oil (1:1 wt ratio) microcapsules at pH 3 and pH 6 with 50% oil loading over
storage time at room temperature (~25C).

Moisture content (%)


a

Microcapsule

Water activity (aw)

0 month

3 month

0 month

3 month

FO-50 (pH 3)

1.66 0.01

2.06 0.03

0.27 0.02

0.5 0.01

FO-EVOO-50 (pH 3)

1.62 0.03

1.93 0.02

0.27 0.0

0.48 0.0

FO-50 (pH 6)

2.98 0.03

3.53 0.06

0.29 0.0

0.49 0.01

FO-EVOO-50 (pH 6)

2.96 0.04

3.19 0.12

0.29 0.0

0.49 0.0

Values are averages of duplicate analysis SD of 2 individual runs

The particle size distribution of the reconstituted spray-dried microcapsules at 0, 1, 2 and 3


month storage under room temperature conditions was determined (Figure 21 & Table
15). A general increase in particle size was observed in all microcapsules over storage time

at room temperature. The increase in particle size for FO and FO-EVOO microcapsules
over storage time were similar, indicating only a minor influence of the type of oil/oils in
the core of the microcapsule. The particle size of all zero month powders at pH 3 and pH 6
with 50% oil loading were slightly greater than the particle size of the corresponding
emulsions prior to spray drying. As stated previously in section 4.3, this could be attributed
to the destabilisation of emulsion droplets during the spray drying process as well the
formation of pectin aggregates.
A significant (P < 0.05) increase in particle size of all the microcapsules at pH 6 with 50%
oil loading was observed after spray drying. However, the increase in particle size of the
microcapsules over storage was comparatively small in comparison with the particle size
of the FO and FO-EVOO microcapsules at pH 3 with 50% oil loading during storage.

61

4
3
2

5
4
3
2
1

1
0
0.01

Volum e (% )

V olum e (% )

0.1

10

100

0
0.01

1000

0.1

100

1000

Volum e (% )

V olu m e (% )

10

Particle size (m)

Particle size (m)

5
4
3
2
1
0
0.01

5
4
3
2
1

0.1

10

100

1000

0
0.01

0.1

10

100

1000

Particle size (m)

Particle size (m)

Figure 21 Particle size distributions of reconstituted spray-dried fish oil and fish oil-extra

virgin olive oil microcapsules prepared from emulsions at pH 3 and 6 with 50% oil loading
prior (A), and subsequent to storage for 1 (B), 2 (C) and 3 (D) month under room
temperature conditions (~25 C, aw 0.5, exposed to light). Values are the averages of
triplicate measurements SD;
(pH 6));

(FO-50 (pH 3));

(FO-EVOO-50 (pH 6))

62

(FO-EVOO-50 (pH 3));

(FO-50

Table 15 Particle size values of the reconstituted spray-dried microcapsules containing

fish oil and fish oil-extra virgin olive oil (1:1 wt ratio) at pH 3 and pH 6 with 50% oil
loading, during storage of the powders in stoppered flasks at room temperature (~25C).
Mean particle diameter, d3,2, (m)
a

Microcapsule

0 month

1 month

FO-50 (pH 3)

0.70 0.00c

0.78 0.00c

2.06 0.08c

3.06 0.05c

FO-EVOO-50 (pH 3)

0.69 0.00b

0.79 0.01d

1.69 0.03b

1.69 0.02a

FO-50 (pH 6)

0.89 0.00b

0.89 0.00a

1.28 0.06a

1.44 0.04a

FO-EVOO-50 (pH 6)

0.87 0.01c

0.90 0.02a

1.27 0.04a

1.46 0.03a

2 month

3 month

Values are averages of triplicate analysis SD of 2 individual runs

Means within the row followed by different superscript letters differ significantly at P = 0.05

SEM of FO and FO-EVOO microcapsules at pH 3 and pH 6 revealed comparable surface


topology. The microcapsules had a wrinkled surface (Figure 22). As discussed previously
in section 4.2 and 4.3 the wrinkles have been associated with the SBP (Drusch, 2007) but
may also be due to particle shrinkage during the early stages of the drying process (Moreau
& Rosenberg, 1993). Given the surface of the microcapsules is continuous and devoid of
any cracks or pores, it is likely that the microcapsules would protect the encapsulated oil
against rapid oxidation.
The SEM images of the FO and FO-EVOO microcapsules after storage at room
temperature for 3 months show an increased agglomeration of the powder particles.

This

phenomenon of lumping can be attributed to the change in moisture content and aw that
occurred in the microcapsules over storage time (Table 14).

63

A
A
C

C
Figure 22 Scanning electron microscopy images of fish oil spray-dried microcapsules

from emulsions prepared at pH 3 (A, C) and pH 6 (B, D) with 50% oil loading, prior (A,
B) and subsequent (C, D) to storage at ambient conditions (~25C, aw 0.5, 3 months).

4.5 LIPID OXIDATION

Lipid oxidation is a major issue when dealing with oils and powders encapsulated with oils
because of the thermal and oxidative reactions that take place during storage. The primary
products of lipid oxidation are the hydroperoxides which are colourless, tasteless and
odourless. These hydroperoxides are unstable and can further degrade into various low
molecular weight compounds with distinctive colour, odour and flavour characteristics.
These low molecular weight compounds include alkanes, alkenes, alcohols, ketones,
aldehydes, acids and esters. Some of these compounds impart off-flavor/taste to the
powders and drastically decrease the shelf-life of the product (Tamime, 2009).

64

Many secondary products have been identified from the radical and photosensitized
oxidations of polyunsaturated lipids. These secondary products mainly consist of
oxygenated monomeric materials including epoxy-hydroperoxides, oxo-hydroperoxides,
hydroperoxy

epidioxides,

dihydroperoxides,

hydroperoxy

bis-epidioxides,

and

hydroperoxy bicycloendoperoxides. Some higher molecular weight dimeric compounds


have also been identified from autoxidized methyl linoleate and linolenate. Decomposition
of these oxidation products form a wide range of carbonyl compounds, hydrocarbons and
furans, for example, that contribute to the flavour deterioration of foods (Frankel, 1987).

4.5.1 OXIDATIVE STABILITY OF EMULSIONS AND MICROCAPSULATED


OIL DURING STORAGE UNDER AMBIENT CONDITIONS

4.5.1.1 PROPANAL HEADSPACE ANALYSIS

Propanal and hexanal are the main volatiles formed by oxidative decomposition of
omega-3 fatty acids and omega-6 fatty acids, respectively (Frankel et al., 1994). Hence, the
propanal and hexanal contents measured at different storage times in the microcapsule
powders can be used as indicators of oxidative stability.
Propanal was detected in all the fresh (0 month) FO and FO-EVOO microcapsules made
with 25% and 50% oil loading at pH 3 (Table 16) indicating a possibility of lipid oxidation
taking place either during the emulsion preparation or spray drying process. An increase in
the propanal content in all powders was positively correlated with the increase in storage
time. The amount of propanal detected in the headspace of all the fresh (0 month)
microcapsules were similar to amounts detected by Drusch et al., 2007; Serfert, Drusch &
Schwarz, 2009 for SBP-, caseinate-, n- octenylsuccinate starch and gum arabic stabilised
FO-in-water emulsions, but only in the absence of added antioxidants. The highest content
of propanal in the microcapsule powders was recorded after 3 months of storage in FO-50
(32.35 0.92 g/g powder) and FO-EVOO-50 (28.70 4.53 g/g powder) powders made
at pH 3. Significant differences in propanal content (P < 0.05) amongst all the samples
stored for up to 2 months were noted. The propanal content in microcapsule powders

65

stored 02 months was higher in microcapsule powders containing 50% oil loading than
powders containing 25% oil loading. This is expected as a more robust interface is
anticipated for the 25% microcapsules, resulting in reduced porosity causing less oxygen
diffusion and hence, less oxidation on storage. The results also showed that the FO-EVOO
powders generally had slightly higher propanal concentrations compared to FO powders
alone, irrespective of oil loading, at most times during storage. However, there were no
significant differences (P > 0.05) in the propanal content in the FO and FO-EVOO samples
at 3 months. The continuous increment of propanal in all powders during storage indicated
the steady oxidation and degradation of omega-3 fatty acids. These observations revealed
that the addition of EVOO was not able to protect FO from oxidation. One of the reasons
for EVOO not being able to increase the oxidative stability of microencapsulated FO could
be the presence of transition metal ions (e.g. Fe, Cu) present in SBP and also, the low pH
environment that led to rapid oxidation.
No propanal was detected in all the fresh (0 month) FO and FO-EVOO powders made with
and without EDTA (50% oil loading) at pH 6 (Table 17). However, a general increase in
propanal content was observed with increase in storage time. The microcapsules with
EDTA had less propanal content compared to the microcapsules without EDTA indicating
that EDTA exerted a significant (P < 0.05) protective effect on the microcapsules with
regard to their oxidative stability.
The results also showed that the FO-EVOO powders had propanal contents significantly (P
< 0.05) lower than FO powders alone indicating that the antioxidant compounds in EVOO
were effective at pH 6. For example, addition of EVOO at 50% oil loading decreased
propanal content by 2.13 fold compared to microcapsules without EDTA after 3 months of
storage (Table 17).

66

Table 16 Propanal content in microencapsulated fish oil and fish oil-extra virgin olive oil

(1:1 wt ratio) (25% and 50% oil loading) at pH 3 during storage at room temperature (~25
C, 03 month) .
Propanal content (g g-1 powder)
a

Microcapsule

0 month

1 month

2 month

3 month

FO-25

1.49 0.09ab

2.63 0.25a

8.50 0.14a

25.85 2.62 a

FO-50

1.66 0.03b

4.55 0.21b

19.15 1.63 b

28.70 4.53 a

FO-EVOO-25

1.41 0.02a

3.97 0.24b

9.28 0.16 a

24.60 3.25 a

FO-EVOO-50

1.58 0.03ab

6.35 0.12c

20.30 0.14 b

32.35 0.92 a

Values are average of duplicate analysis SD of 2 individual runs

Means within columns followed by different superscript letters differ significantly at 5% level

Table 17 Propanal content in microencapsulated fish oil and fish oil-extra virgin olive oil

(1:1 wt ratio) with and without EDTA (50% oil loading) at pH 6 during storage at room
temperature (~25 C, 03 month).
Propanal content (g g-1 powder)
a

Microcapsule

0 month

1 month

2 month

3 month

FO-EDTA-50

0a

1.90 0.42a

1.95 1.01a

18.90 0.00 b

FO-EVOO-EDTA-50

0a

1.06 0.08a

1.19 0.08a

7.68 0.40 a

FO-50

0a

2.37 0.01a

9.15 0.92 b

27.55 0.07 c

FO-EVOO-50

0a

1.88 0.21a

7.50 0.85 b

8.80 0.71 a

Values are average of duplicate analysis SD of 2 individual runs

Means within columns followed by different superscript letters differ significantly at 5% level

Data reported in literature established the fact that metals like Fe and Cu act as catalysts for
the formation of highly reactive hydroxyl radicals which initiate a chain reaction for lipid
oxidation (Benedt & Shibamoto 2007; Goldstein, Meyerstein & Czapski 1993). The
significant (P < 0.05) differences in the propanal content between microcapsules with and
without EDTA confirmed EDTAs efficiency in inhibiting oxidation and also that metal

67

ion-induced oxidation reactions was one of the major reasons for the oxidative
deterioration of the FO and FO-EVOO microcapsules during storage. Jacobsen (2010)
performed some studies to retard lipid oxidation in foods enriched with omega-3-fatty
acids. It was found that the addition of EDTA (5 mg/kg milk) to skimmed milk
significantly retarded the oxidation when the emulsions were enriched with 1.5% FO.
Data gained from those studies also indicated that EDTA was effective in preventing
oxidation in omega-3 fatty acid-enriched mayonnaises and salad dressings when compared
to fitness bars. The article also illustrated that the effectiveness of EDTA might also
depend on the initial level of lipid hydroperoxides in the FO or the emulsion.
At pH 6, improved oxidative stability can be expected as the solubility of iron in water
increases with decreasing pH (Donnelly et al., 1998; Graf et al., 1984; Mancuso et
al.,1999). Therefore, the high propanal contents of FO and FO-EVOO microcapsules at pH
3 as compared to FO and FO-EVOO microcapsules at pH 6 could be attributed to the
increased solubility of iron. Indeed, some amount of propanal was found in FO and
FO-EVOO microcapsules at pH 3 prior to storage (zero month), whereas no detectable
propanal was observed in FO and FO-EVOO microcapsules at pH 6 at zero month. The
propanal content of FO and FO-EVOO with 50% oil loading at pH 6 were significantly
lower than the FO and FO-EVOO samples with 50% oil loading at pH 3 (P < 0.05),
irrespective of the storage duration (Table 18). This also showed that there was an effect of
pH on the oxidative stability of the powders. The results revealed that the FO-EVOO
powders at pH 6 had propanal contents (8.8 0.71 g/g powder) significantly (P < 0.05)
lower than FO powders (27.55 0.07 g/g powder) alone after 3 months of storage.
In addition to its influence on the solubilisation of transition metals, pH also affects oxygen
solubility and mobility, and the rate of non-enzymatic browning reaction in foods . There
have been studies where FO-enriched mayonnaise showed increased oxidation with
decreasing pH during storage and also that metal ions significantly promoted oxidation
(Tong et al., 2000; Jacobsen, Timm & Meyer, 2001). In another study, polyoxyethylene 10
lauryl ether was used as an emulsifier in making model emulsions and it was found that
iron showed a highly increased pro-oxidative effect at pH 3 compared to that at pH 7

68

because of the increased solubility of iron at low pH (Cho et al., 2003). However, in the
current study the data indicated that the effect of pH was not significant (P > 0.05) in the
case of FO without added EVOO. Consequently, it could be concluded that the antioxidant
effect of EVOO was readily available at pH 6.
Table 18 Propanal content in microencapsulated fish oil and fish oil-extra virgin olive oil

(1:1 wt ratio) with 50% oil loading at pH 3 and pH 6 during storage at room temperature
(~25C, 03 month) .
Propanal content (g g-1 powder)
a

Microcapsule

0 month

1 month

FO-50 (pH 3)

1.66 0.03

FO-EVOO-50 (pH 3)

2 month
b

3 month

4.55 0.21

19.15 1.63

28.70 4.53bc

1.58 0.03ab

6.35 0.12c

20.30 0.14b

32.35 0.92c

FO-50 (pH 6)

0a

2.37 0.01a

9.15 0.92a

27.55 0.07b

FO-EVOO-50 (pH 6)

0a

1.88 0.21a

7.50 0.85a

8.80 0.71a

Values are average of duplicate analysis SD of 2 individual runs

Means within columns followed by different superscript letters differ significantly at 5% level

4.5.1.2 HEXANAL HEADSPACE ANALYSIS

There was no detectable hexanal for 02 months storage in either FO or FO-EVOO


microcapsules (25% and 50% oil loading) at pH 3 (Table 19) and FO or

FO-EVOO

microcapsules (50% oil loading) with EDTA at pH 6 (Table 20). FO and FO-EVOO
microcapsules (50% oil loading) without EDTA at pH 6 had some amount of detectable
hexanal in the 2nd and 3rd month and no detectable hexanal prior to 2 months (Table 21).
The later development of hexanal compared to propanal may be expected as hexanal and
propanal are secondary oxidation products of omega-6 (e.g. C18:2) and omega-3 fatty
acids (e.g. DHA, EPA, C18:3), respectively. It is well known that omega-3 fatty acids are
more prone to oxidation than omega-6 fatty acids. It has, however, been reported that
hexanal cannot be used as an adequate marker for the beginning of oxidation in case of

69

virgin olive oil, although it has been successful with refined vegetable oils (Snyder et al.,
1988; Warner et al., 1988).
Table 19 Hexanal content in microencapsulated fish oil and fish oil-extra virgin olive oil

(1:1 wt ratio) at pH 3 during storage at room temperature (~25C, 03 month)


a

Microcapsule

Hexanal content (g g-1 powder) after 3 months storage

FO-25

1.4 0.21a

FO-50

1.6 0.36a

FO-EVOO-25

2.1 0.42a

FO-EVOO-50

1.8 0.71a

Values are average of duplicate analysis SD of 2 individual runs

Means within columns followed by different superscript letters differ significantly at 5% level

No significant difference (P > 0.05) in the hexanal content between FO and FO-EVOO
microcapsules (25% and 50% oil loading) at pH 3 microcapsules after 3 months of storage
was observed. FO and FO-EVOO microcapsules with EDTA had significantly less hexanal
content compared to the samples without EDTA (P < 0.05). It was also observed that
FO-EVOO microcapsules with and without EDTA had hexanal contents significantly
lower than FO microcapsules with and without EDTA (P < 0.05). Also, the hexanal
content of FO and FO-EVOO with 50% oil loading at pH 6 were significantly lower than
the FO and FO-EVOO samples with 50% oil loading at pH 3 (P < 0.05).. That was in
agreement with the propanal results and confirmed that pH was also an influencing factor
in the oxidation of the microencapsulated oil.
Objectionable rancid odour was noted in all microcapsule powders on sniffing the samples
stored in containers for 3 months at ambient conditions .There are a wide range of volatile
compounds present in FO and EVOO and these volatiles bring out different tastes and
odours. Each of these compounds have different thresholds of perception. It is anticipated
that propanal and hexanal give rise to a rancid odour.

70

Table 20 Hexanal content in microencapsulated fish oil and fish oil-extra virgin olive oil

(1:1 wt ratio) with and without EDTA (50% oil loading) at pH 6 during storage at room
temperature (~25C, 03 month) .

Hexanal content (g g-1 powder)


a

Microcapsule

0 month

1 month

2 month

3 month

FO-EDTA-50

0.00 0.00a

0.00 0.00a

0.00 0.00a

1.26 0.02c

FO-EVOO-EDTA-50

0.00 0.00a

0.00 0.00a

0 .00 0.00a

0.18 0.04a

FO-50

0.00 0.00a

0.00 0.00a

0.12 0.01a

0.51 0.01b

FO-EVOO-50

0.00 0.00a

0.00 0.00a

0.41 0.28a

0.44 0.06b

Values are average of duplicate analysis SD of 2 individual runs

Means within columns followed by different superscript letters differ significantly at 5% level

Table 21 Hexanal content in microencapsulated fish oil and fish oil-extra virgin olive oil

(1:1 wt ratio) with 50% oil loading at pH 3 and pH 6 during storage at room temperature
(~25C, 03 month).
Hexanal content (g g-1 powder)
a

Microcapsule

0 month

1 month

2 month

3 month

FO-50 (pH 3)

0.00 0.00a

0.00 0.00a

0.00 0.00a

1.60 0.36a

FO-EVOO-50 (pH 3)

0.00 0.00a

0.00 0.00a

0 .00 0.00a

1.80 0.71a

FO-50 (pH 6)

0.00 0.00a

0.00 0.00a

0.12 0.01a

0.51 0.01b

FO-EVOO-50 (pH 6)

0.00 0.00a

0.00 0.00a

0.41 0.28a

0.44 0.06b

Values are average of duplicate analysis SD of 2 individual runs

Means within columns followed by different superscript letters differ significantly at 5% level

4.5.1.3 EVALUATION OF LIPID OXIDATION BY DETERMINING THE


CHANGES IN THE FATTY ACIDS COMPOSITION OVER STORAGE TIME

The major fatty acids present in the oils (FO and EVOO) and all encapsulated oils were
palmitic acid (C16:0), oleic acid (C18:1), EPA (C20:5) and DHA (C22:6). Hence, the
71

percentage composition of each of the major fatty acids mentioned above was determined
[Table 22 (a & b), Table 23 (a & b) and Table 24 (a & b)] after extraction of the oil

from the microcapsules. The ratio of EPA to C16:0 and DHA to C16:0 were also provided.
Other minor fatty acids present in each of the oils were calculated as a percentage and
represented as others in the tables. The data has been expressed as the unsaturated fatty
acids to C16:0 because the absolute value of C16:0 did not change, only its % relative to
the unsaturated fatty acid changed as oxidation proceeded.
As expected, EPA and DHA, being polyunsaturated fatty acids had significant changes in
their composition over the storage period (P > 0.05). The amounts of EPA (C20:5n-3) and
DHA (C22:6n-3) in the oil extracted from all fresh powders containing microencapsulated
FO were slightly lower than that of the pure FO. The extreme susceptibility of
polyunsaturated fatty acids to oxygen, light and temperature is widely known (Garg et al.,
2006; Kolanowski & Laufenberg, 2006) and hence, the oxidation of the omega-3-fatty
acids in the oil must have occurred during the preparation of emulsions and/or the spray
drying process. The monounsaturated fatty acid composition was comparable. As
expected, the polyunsaturated omega-3 fatty acids EPA and DHA in all microcapsules
decreased over the storage period. In contrast, there was no significant change (P > 0.05)
in the oleic acid content with storage time. This result might be explained in terms of the
structure of the fatty acids. Generally, the greater the number of double bonds in the fatty
acid, the more unstable the fatty acid and the greater the ease of its oxidation (e.g. by
oxygen, light) (Olive oil source, 2010).
The percentage decrease in EPA and DHA was found to be greater in microcapsules
having 50% oil loading when compared to microcapsules with 25% oil loading, at pH 3
(Tables 22, a & b). The highest stability of EPA and DHA was found to be in FO

microcapsules (25% oil loading) with the EPA decreasing from 6.4% at 0 month to 6.2% at
the 3rd month, and DHA from 23.6% at 0 month to 21.8% at the 3rd month. FO-EVOO-50
microcapsules and FO-50 microcapsules were found to be the least stable after 3 months of
storage. The EPA content decreased from 3.3% to 2.7% in FO-EVOO-50 and from 6.7%
to 5.6% in FO-50, and DHA from 12% to 9.3% in FO-EVOO-50 and from 24% to 18% in

72

FO-50 microcapsules. A percentage decrease of 16.31% and 22.84% in EPA and DHA
was observed in FO-EVOO-50 microcapsules over 3 months of storage whereas, a
percentage decrease of 16.49% EPA and 25.55% DHA was recorded in FO-50
microcapsules indicating that EVOO had a slight antioxidative effect on controlling lipid
oxidation of microencapsulated FO. On the contrary, FO-EVOO-25 had a greater
percentage decrease of EPA and DHA, 12.04% and 17.75%, respectively, when compared
to 2.82% EPA and 7.83% DHA in FO-25. That indicated that the concentration of
antioxidative components present in EVOO was not sufficient to protect the FO from
oxidation or other mechanisms come into play which offset or (negated) the antioxidant
effects of EVOO components.
FO and FO-EVOO microcapsules without EDTA at pH 6 in the formulation were found to
be least stable when compared to the microcapsules with EDTA (Table 23, a & b). That
suggested that the metal ions present in SBP were one of the underlying factors responsible
for lipid oxidation in FO and FO-EVOO microcapsules prepared at pH 3. FO-EVOO
microcapsules were less oxidised than FO microcapsules made with 50% oil loading at pH
6. EPA and DHA contents decreased by 0.33% and 1.99%, respectively, in FO-EVOO-50
microcapsules and 0.74% for EPA and 4.46% for DHA in FO-50 microcapsules.
Consequently, it could be concluded that EVOO was as effective in preventing lipid
oxidation in FO microcapsules as EDTA during long term storage. That might be
attributed to the high antioxidant content and radical scavenging activity of EVOO.
FO and FO-EVOO microcapsules (50% oil loading) at pH 3 were found to be less stable
when compared to FO and FO-EVOO microcapsules (50% oil loading) at pH 6 [Table 22
(a & b) and Table 23 (a & b)]. EPA and DHA contents decreased by 16.49% and

25.55%, respectively, in FO-50 microcapsules and EPA by 16.31% and DHA by 22.84%
in FO-EVOO-50 microcapsules at pH 3. EPA and DHA contents decreased by 12.71% and
19.16%, respectively in FO-50 microcapsules and 10.41% and 15.89%, respectively, in
FO-EVOO-50 microcapsules at pH 6. This shows that FO microcapsules were less stable
than FO-EVOO microcapsule at both pH 3 and pH 6.

73

Table 22 a Fatty acid composition (%) of microencapsulated fish oil (25% and 50% oil loading) at pH 3 during storage at room
temperature (~25C, 03 months).

Microcapsule
FO-25

Fatty acid

FO-50

0 month

1 month

2 month

3 month

0 month

1 month

2 month

3 month

C16:0

19.78 1.25

21.14 0.13

22.16 0.01

23.05 0.29

20.78 0.15

21.53 0.49

23.58 0.08

25.34 0.08

C18:1

14.07 0.86

15.07 0.39

14.72 0.13

16.11 0.17

14.55 0.12

14.690.65

16.03 0.05

15.31 0.04

C20:5n-3

6.39 0.14

6.36 0.17

6.42 0.05

6.21 0.07

6.67 0.29

6.14 0.12

5.40 0.01

5.57 0.11

C22:6n-3

23.63 0.17

22.55 0.92

21.66 0.22

21.78 0.08

24.23 1.24

21.07 0.26

18.35 0.11

18.04 0.12

Others

37.13 0.67

35.05 0.57

33.77 0.42

36.56 0.61

33.77 0.57

36.56 0.21

24 0.24

24.46 0.19

3.5 0.01

3.63 0.03

3.44 0.02

3.34 0.01

3.24 0.04

DHA:EPA

3.7 0.11

3.54 0.05

3.4 0.00

EPA:C16

0.32 0.01

0.30 0.01

0.29 0.00

0.27 0.00

0.32 0.01

0.29 0.01

0.23 0.00

0.22 0.00

DHA:C16

1.20 0.08

1.07 0.05

0.98 0.01

0.94 0.01

1.17 0.05

0.98 0.01

0.78 0.01

0.71 0.01

Values are average of duplicate measurements SD

74

Table 22 b Fatty acid composition (%) of microencapsulated fish oilextra virgin olive oil (1:1 wt ratio) (25% and 50% oil loading) at
pH 3 during storage at room temperature (~25C, 03 months).

Microcapsule
FO-EVOO-25

Fatty acid

FO-EVOO-50

0 month

1 month

2 month

3 month

0 month

1 month

2 month

3 month

C16:0

14.71 0.08

15.02 0.04

14.87 0.21

15.72 0.01

14.57 0.17

15.01 0.05

14.94 0.05

15.52 0.16

C18:1

46.05 0.27

46.89 0.09

46.59 0.37

48.52 0.07

45.41 0.15

46.18 0.10

46.98 0.06

48.62 0.30

C20:5n-3

3.24 0.03

3.15 0.01

3.07 0.03

2.85 0.10

3.25 0.06

3.09 0.07

2.92 0.01

2.72 0.03

C22:6n-3

12.00 0.23

11.30 0.13

11.02 0.03

9.87 0.03

12.04 4.4

10.74 0.32

10.26 0.03

9.29 0.17

Others

24.00 0.61

24.4 0.26

24.73 0.11

24.91 0.01

24.73 0.45

24.91 0.34

24.91 0.15

23.84 0.34

DHA:EPA

3.70 0.04

3.61 0.01

3.6 0.01

3.46 0.13

3.71 0.06

3.48 0.04

3.50 0.00

3.42 0.03

EPA:C16

0.22 0.00

0.21 0.00

0.21 0.00

0.18 0.00

0.22 0.01

0.21 0.00

0.19 0.00

0.18 0.00

DHA:C16

0.82 0.01

0.75 0.01

0.74 0.00

0.63 0.00

0.81 0.05

0.72 0.02

0.69 0.00

0.60 0.02

Values are average of duplicate measurements SD

75

Table 23 a Fatty acid composition (%) of microencapsulated fish oil (50% oil loading) with and without EDTA at pH 6, during
storage at room temperature (~25C, 03 months).

Microcapsule
FO-EDTA-50

Fatty acid

FO-50

0 month

1 month

2 month

3 month

0 month

1 month

2 month

3 month

C16:0

19.50 0.66

19.66 0.35

20.13 0.50

19.71 0.44

19.10 0.57

19.48 0.65

20.37 0.11

20.16 0.38

C18:1

14.71 0.72

13.56 0.28

15.06 0.85

15.36 1.93

17.38 2.29

16.95 2.88

18.16 2.70

18.74 2.29

C20:5n-3

6.06 0.24

6.44 0.33

5.89 0.21

5.48 0.35

5.82 0.06

5.64 0.02

5.36 0.05

5.08 0.06

C22:6n-3

24.13 0.18

25.05 0.33

22.52 0.16

20.57 1.33

23.28 0.15

22.30 0.37

20.43 0.60

18.82 0.18

Others

35.52 0.36

35.30 0.27

36.65 0.29

38.88 1.90

34.42 1.66

34.98 1.85

35.84 1.94

37.20 2.00

DHA:EPA

3.99 0.13

3.89 0.04

3.83 0.16

3.76 0.02

4.00 0.06

3.95 0.06

3.78 0.08

3.71 0.03

EPA:C16

0.31 0.00

0.33 0.01

0.29 0.00

0.28 0.02

0.31 0.01

0.29 0.01

0.26 0.00

0.25 0.00

DHA:C16

1.22 0.03

1.29 0.04

1.12 0.04

1.04 0.06

1.22 0.03

1.16 0.02

1.00 0.02

0.93 0.01

Values are average of duplicate measurements SD

76

Table 23 b Fatty acid composition (%) of microencapsulated fish oilextra virgin olive oil (1:1 wt ratio) (50% oil loading) with and
without EDTA at pH 6, during storage at room temperature (~25C, 03 months).

Microcapsule
FO-EVOO-EDTA-50

Fatty acid

FO-EVOO-50

0 month

1 month

2 month

3 month

0 month

1 month

2 month

3 month

C16:0

15.30 1.02

16.31 0.25

16.39 0.55

16.56 0.78

15.76 0.23

15.86 0.13

16.21 0.21

16.25 0.08

C18:1

35.88 2.09

40.40 0.68

39.80 0.37

40.70 2.01

40.26 1.21

41.18 1.54

42.19 0.99

42.10 0.90

C20:5n-3

3.69 0.08

3.22 0.24

3.22 0.14

3.19 0.04

3.17 0.28

3.04 0.26

2.84 0.15

2.84 0.12

C22:6n-3

14.39 0.25

12.66 0.68

12.60 0.03

12.11 0.11

12.52 0.68

11.88 0.61

10.77 0.08

10.53 0.38

Others

30.74 2.78

27.99 0.51

27.99 0.33

27.44 2.86

28.30 0.09

28.04 0.54

27.99 0.55

28.28 0.96

DHA:EPA

3.90 0.02

3.90 0.09

3.92 0.17

3.79 0.06

3.96 0.14

3.91 0.14

3.80 0.17

3.71 0.06

EPA:C16

0.24 0.02

0.24 0.01

0.20 0.00

0.19 0.01

0.18 0.02

0.18 0.01

0.16 0.01

0.16 0.01

DHA:C16

0.92 0.08

0.77 0.03

0.77 0.03

0.73 0.03

0.79 0.03

0.75 0.03

0.66 0.00

0.65 0.02

Values are average of duplicate measurements SD

77

4.5.2 OXIDATIVE STABILITY OF EMULSIONS AND MICROCAPSULES


EXPOSED TO ACCELERATED STORAGE CONDITIONS
Accelerated oxidation tests are generally performed to compare the oxidative stability of
samples containing different ingredients or samples held at varying storage conditions.
Such studies are used also to understand the effectiveness of various antioxidants used in
preventing oxidative deterioration in samples based on values of the induction period (IP).
Accelerated storage tests artificially hasten or speed up the oxidation process by exposing
samples to heat, oxygen, light, metal catalysts, or enzymes. Shelf-life analysis of samples
under ambient conditions should also be conducted for comparison. Accelerated oxidation
tests should be used basically as a rapid screening test. The main difficulty with
accelerated storage test results is that they are conducted under artificial conditions and we
assume that the reactions occurring under these conditions are similar to those reactions
that occur at ambient conditions (Nielsen, 2010).
Results from accelerated storage revealed that the IP of the pure oils and FO-EVOO blend
was greatest for EVOO followed by FO-EVOO and then FO (Table 24). A similar trend
was observed in their rate of oxidation. Oxipres results for microencapsulated FO and
FO-EVOO at 25% and 50% oil loadings, with and without inclusion of EDTA, at pH 3 and
pH 6 are shown in Tables 25
27. The emulsions and microencapsulated oils had a lower
IP than the pure oils (Table 24 & 25). One reason for this is the larger surface area of the
emulsified oil droplets and microencapsulated oil compared to that of the bulk oil/air
interface. The presence of small oil droplets in the emulsions and microencapsulated oil
allows rapid access of oxygen to the oil.

78

Table 24 Oxidative stability of bulk oils (FO, EVOO, FO-EVOO (1:1 wt ratio) during
accelerated storage (80C, oxygen pressure of 0.5 bar).
a

Induction period (h)

Slope (mbar h-1)

FO

11.85 0.07

-210 9.19

EVOO

>100 0.00

FO:EVOO (1:1 wt ratio)

16.3 0.14

-119 4.24

Sample

Values are averages of duplicate analysis SD

There was no significant (P > 0.05) difference in the IP values between emulsions made
for manufacture of powders with 25% and 50% oil loading at pH 3 (Table 25). That also
suggested that there was no effect of oil composition or oil loading on the IP values of
emulsions. Similarly, FO microcapsules (25% and 50% oil loading) revealed the same IP
values (6.95 0.07 h). However, the IP value increased from 7.90 h to 11.90 h when the
oil load was increased from 25% to 50% in FO-EVOO microcapsules. The results
indicated that spray-dried FO-EVOO microcapsules had greater oxidative stability than
spray-dried FO microcapsules under the accelerated storage conditions.
The enhanced oxidative stability of the FO-EVOO microcapsules could be due to the
presence of olive oil, which contains large amounts of oleic acid (5583%) and much
smaller amounts of saturated fatty acids (14%). That interpretation was in agreement with
the findings of Parkanyiova et al., (2000) who reported that triacylglycerols containing
bound linoleic acid were oxidised several times faster than triacylglycerols containing only
oleic and saturated fatty acids. However, the Oxipres data were in direct contrast to those
obtained for powders prepared at pH 3 and stored under ambient conditions, as determined
by propanal and hexanal analysis and fatty acid composition data. This could be attributed
to a variety of auto oxidative mechanisms that take place with respect to the change in
temperature of oxidation. It has been widely reported that the antioxidative potential of
virgin olive oil is affected by both storage and thermal treatment due to modifications in

79

oxidative/hydrolytic reactions. Other studies have shown that thermal treatment of virgin
olive oil resulted in an increase in the formation of hydroxytyrosol and tyrosol from virgin
olive oil secoiridoids (Beter et al., 2008; Brenes et al., 2002; Lerma-Garcia et al., 2009;
Sacchi et al., 2002). Another study by Carrasco-pancorbo et al., (2005) demonstrated that
hydroxytyrosol contributed to the oxidative stability of virgin olive oil.
Table 25 Oxidative stability of emulsions and powders (25% and 50% oil loading)
containing fish oil and fish oil-extra virgin olive oil (1:1 wt ratio) at pH 3 during
accelerated storage (80C, oxygen pressure of 0.5 bar).

Induction period (h)

Slope (mbar h-1)

FO-25

4.90 0.14a

-28.00 4.24b

FO-50

4.60 0.00a

-38.00 1.41a

FO-EVOO-25

5.00 0.14a

-21.50 0.71b

FO-EVOO-50

4.65 0.07a

-27.00 1.41b

Sample
(A) Emulsion

(B) Microcapsule
FO-25
FO-50
FO-EVOO-25
FO-EVOO-50

0.07
6.95 0.07
7.90 0.00
6.95

-225.00 14.14d

-2350.50 7.78a

11.95 0.35c

-612.50 3.54c
-792.00 1.41b

Values are averages of duplicate analysis SD of 2 individual runs


Means within columns followed by different superscript letters differ significantly at 5% level

As expected, the addition of antioxidant i.e EDTA, in the formulation of the microcapsules
significantly increased the IP of FO and FO-EVOO microcapsules and emulsions with
EDTA, offering better oxidative stability. The FO-EVOO emulsions and microcapsules
showed greater oxidative stability with greater IPs as compared to FO microcapsules at
pH 6 with 50% oil loading (Table 26B). The enhanced stability of FO-EVOO
microcapsules could be attributed to the absence of polyunsaturated fatty acids.
Monounsaturated fatty acids possess only one single double bond making them more
80

stable than polyunsaturated fatty acids yet not as stable as saturated fats (Unsaturated fats
2010; Extra virgin olive oil 2010). All emulsions (FO and FO-EVOO) made with and
without EDTA at pH 6 had relatively small slope values indicating a slow rate of oxidation
in the emulsions, whereas these values were much larger in all the microcapsules (Table
26). This difference in oxidative stability between emulsions and microcapsules may be
due to the orientation of the EDTA and the phenolic compounds present in EVOO in an
aqueous phase and in a dried state. Transition metal ions, such as iron and copper, would
be located in the aqueous phase of oil-in-water emulsions and oriented in the oil/water
interface (Frankel et al., 1994; Jacobsen et al., 2008). EDTA, omega-3 hydroperoxides and
the polar phenolics would also be present in the aqueous phase of the liquid emulsions.
Therefore, it is possible that the shared location and high level of contact between the proand anti- oxidants in the aqueous phase of the liquid emulsions was more effective in
preventing lipid oxidation than in dried state, due to the water restricted environment in the
latter. Other factors that could possibly cause differences in lipid oxidation in emulsions
and microcapsules would be factors affecting the antioxidant activity and rate of oxidation
in both systems. The effectiveness of antioxidants in oil-in-water emulsions is dependent
on several parameters, such as their structure, polarity, location, radical scavenging and
metal-chelating attributes (Mattia et al., 2010; Paiva-Martins & Gordon, 2002,
Paiva-Martins et al., 2006). The rate and extent of lipid oxidation may also be influenced
by numerous variables like pH, type of emulsifier, oxygen availability and structure,
thickness and composition of the interface (Coupland & McClements, 1996; McClements
& Decker, 2000; Velasco et al., 2003, 2006).
The FO and FO-EVOO emulsions had greater IP values at pH 3 than FO and FO-EVOO
emulsions at pH 6 (Table 27). No significant difference (P > 0.05) in IP values was
observed between FO and FO-EVOO emulsions made with 50% oil loading at pH 3. The
FO-EVOO emulsion had a similar IP value to FO emulsion made with 50% oil loading at
pH 6. All emulsions made with 50% oil loading at pH 3 and pH 6 had lower slope values
than the microcapsulated oil, indicating a slow rate of oxidation. No significant difference
(P > 0.05) between the spray-dried FO and FO-EVOO microcapsules at pH 3 with 50% oil
loading when compared to the corresponding spray-dried FO and FO-EVOO

81

microcapsules at pH 6 (Table 27). Data revealed that the FO-EVOO microcapsules were
better in terms of oxidative stability when compared to FO microcapsules both at pH 3 and
pH 6. In case of FO-EVOO and FO emulsions, the IP values were very small and did not
vary significantly (P > 0.05) from each other irrespective of oil composition and pH. This
result indicated that microencapsulation was significantly (P < 0.05) effective in
preventing lipid oxidation than in liquid emulsions. It has been previously discussed
(section 4.5.2) as to how the effectiveness of the antioxidants may be affected by numerous
variables like structure, polarity, location (e.g. water, oil, interface) and levels of metals
(Fe & Cu) availability. It is possible that the efficacy of antioxidants present in EVOO has
been affected by one of these factors and that could have caused these differences in
microcapsules and emulsions lipid oxidation.

82

Table 26 Oxidative stability of emulsions and powders (50% oil loading) containing fish
oil and fish oil-extra virgin olive oil (1:1 wt ratio) made with and without EDTA (50% oil
loading) at pH 6 during accelerated storage (80C, oxygen pressure of 0.5 bar).

Induction period (h)

Slope (mbar h-1)

FO-EDTA-50

13.90 0.14b

-16.50 2.12ab

FO-EVOO-EDTA-50

23.65 0.49c

-11.50 4.95c

FO-50

3.20 0.28a

-24.00 7.07ab

FO-EVOO-50

3.90 0.28a

-31.50 3.54a

Sample

(A) Emulsion

(B) Microcapsule
FO-EDTA-50

10.00

0.28

-421.50 6.36b

FO-EVOO-EDTA-50

19.65 1.06d

-131.50 3.54d

FO-50

7.35

0.21

-480.50 4.95a

FO-EVOO-50

12.70

0.14

-97.00 16.97c

Values are averages of duplicate analysis SD of 2 individual runs


Means within columns followed by different superscript letters differ significantly at 5% level

83

Table 27 Oxidative stability of emulsions and powders (50% oil loading) containing fish
oil and fish oil-extra virgin olive oil (1:1 wt ratio) at pH 3 and pH 6 during accelerated
storage (80C, oxygen pressure of 0.5 bar).

Induction period (h)

Slope (mbar h-1)

FO-50 (pH 3)

4.60 0.00a

-38.00 1.41a

FO-EVOO-50 (pH 3)

4.65 0.07a

-27.00 1.41b

FO-50 (pH 6)

3.20 0.28a

-24.00 7.07ab

FO-EVOO-50 (pH 6)

3.90 0.28a

-31.50 3.54a

Sample

(A) Emulsion

(B) Microcapsule

0.07

-2350.50 7.78a

FO-50 (pH 3)

6.95

FO-EVOO-50 (pH 3)

11.95 0.35c

-792.00 1.41b

FO-50 (pH 6)

7.35

0.21

-480.50 4.95a

FO-EVOO-50 (pH 6)

12.70

0.14

-97.00 16.97c

Values are averages of duplicate analysis SD of 2 individual runs


Means within columns followed by different superscript letters differ significantly at 5% level

84

CHAPTER 5 - GENERAL CONCLUSIONS AND FUTURE DIRECTIONS


The experimental results obtained in the study confirmed some previous conclusions
drawn by other researchers and generated some new findings supporting initial research
objectives. The following were the main conclusions made:

SBP was an effective encapsulant only in terms of microencapsulation efficiency of


the microcapsules and required the addition of a chelating agent, such as EDTA, to
neutralize the negative effects of transition metals that were associated with the
polysaccharide to improve the oxidative stability of the microcapsules.

At pH 3
- EVOO did not improve the oxidative stability of microencapsulated FO during
storage at room temperature.
- In contrast, EVOO improved the oxidative stability of microencapsulated FO
during accelerated storage conditions.

At pH 6
- FO-EVOO microcapsules had greater oxidative stability than FO microcapsules
both during storage at room temperature and under accelerated storage conditions.

Addition of EDTA as a chelating agent significantly increased the oxidative


stability of the microcapsules.

Microcapsules prepared from emulsions at pH 6 were better in terms of long term


oxidative storage stability than microcapsules prepared from emulsions at pH 3.

The main focus of this work was to produce physically stable powders containing
microencapsulated FO with EVOO added for its natural antioxidants and additional health
benefits. Results demonstrated that EVOO did contribute to some antioxidative effect
when microcapsules were prepared at pH 3 and 6 under accelerated storage conditions.

85

However, the same could not be said about the protective effect of EVOO on FO-EVOO
microcapsules prepared at pH 3 when stored under ambient conditions. A significant and
positive effect on oxidative stability of FO microcapsules was observed with the addition
of EDTA in the microcapsule formulation. Similarly, SBP was effective as a wall material
but, for long term stability of FO microcapsules it required EDTA (a metal-chelating
additive) to counteract the pro-oxidative effects caused by the metal ions (iron and copper)
associated with it.
The results obtained through this research project helped to improve understanding of
o the protection requirements to minimise oxidation of FO during processing of
FO microcapsules.
o how and where to add the microencapsulated omega 3 fatty acids in a product.
o the possible interactions of omega-3-fatty acids with other ingredients.
o the physical and shelf-life properties of the FO-EVOO powders.

FUTURE DIRECTIONS
This study identified the potential application of SBP as a food matrix, and tried to
protect FO with the natural antioxidants present in EVOO by optimising the process
parameters to stabilize microcapsules. However, it could be suggested that testing SBP
after depolymerising could yield good results. It is recommended also for further
studies that concentrate on natural food ingredients rich in antioxidants such as
phytosterols could be useful to protect FO from oxidation. Furthermore, the in vitro
analysis of FO-EVOO powders could be examined to examine the susceptibility of the
microcapsule to simulated gastric and intestinal digestion.
Finally, testing the delivery of omega-3 fatty acids in the form of FO-EVOO powder
into various food products (cereal products e.g. breakfast cereals, bakery products e.g.
bread or into cake mixes and fruit nut bars) could have significance nutritional value
and possible industrial applications.

86

REFERENCES
Aguie-Beghin, V., Sausse, P., Meudec, E., Cheynier, V. & Douillard, R. (2008).
Polyphenol- beta-casein complexes at the air/water interface and in solution: effects of
polyphenol structure. Journal of Agricultural & Food Chemistry. 56(20): 9600-9611.
Akhtar, M., Dickinson, E., Mazoyer, J., Langendorff, V., Valle, G.D. & Popineau, Y.
(2002). Emulsion-stabilizing properties of depolymerised pectin. Food Hydrocolloids. 16:
249-256.
Alarcon de la Lastra, C, Barranco, M.D., Motilva, V., & Herrerias J.M. (2001).
Mediterranean diet and health: biological importance of olive oil. Curr Pharm. 7(10):
933-950.
Amagase, H., Schaffer, E.M., & Milner, J.A. (1996). "Dietary components modify the
ability of garlic to suppress 7, 12-dimethylbenz (a) anthracene-induced mammary DNA
adducts. J Nutr.126 (4): 817-824.
American Oil Chemists Society, A. (2005). Official methods and recommended practices
of the American Oil Chemists Society. AOCS Official method No. Cd 8b-90.Peroxide
Value. Champaigne, American Oil Chemists Society.
Arts, I. C., & Hollman, P.C. (2005). Polyphenols and disease risk in epidemiologic studies.
American Journal Clinical Nutrition. 81(1): 317S-325S.
Augustin, M. A., Sanguansri, L., & Bode, O. (2006). Maillard Reaction Products as
Encapsulants for Fish Oil Powders. Journal of Food Science. 71(2): E25 - E32.
Augustin, M. A., & Sanguansri, L. (2007). Encapsulation of bioactives. Food materials
science: principles and practice. Aguilera, J.M., Lillford, P.J. New York, springer: 577601.

87

Australian indigenous healthinfonet. (2008). [online]. Australian indigenous health infonet:


Reviews.

Available

at:

http://www.healthinfonet.ecu.edu.au/health-risks/

nutrition/

reviews/background-information [Retrieved 18 December, 2008].


Benedt, J., & Shibamoto, T. (2008). Role of transition metals, Fe(II), Cr(II), Pb(II) and
Cd(II) in lipid peroxidation. Food Chemistry. 107(1): 165- 168.
Benita, S. (1996). MICROENCAPSULATION methods and industrial applications.
NewYork, Marcel Dekker Inc.
Betera, E., Butinarb, B., Buar-Miklavia, M., & Golobc, T. (2008). "Chemical changes
in extra virgin olive oils from Slovenian Istra after thermal treatment " Food Chemistry.
108(2): 446-454.
Bhandari, B.R., Datta, N., & Howes, T. (1997). " Problems associated with spray drying of
sugar-rich foods". Drying Technology. 15(2): 671-684.
Bioactive Food Components. (2003). Encyclopaedia of Food & Culture. Ed. Katz. S.H.
Volume

1.

enotes.com.

The

Gale

http://www.enotes.com/food-encyclopedia/bioactiv

Group,

Inc.

Available

at:

e-food-components [Retrieved 17

November 2010].
Blatt, Y., Pinto, R., Safronchik, O., Sedlov, T., & Zelkha, M. (2006). Stable coated
microcapsules. freepatentsonline, Bio-Dar Ltd. (Yavne, IL). United States Patent
7097868
Brenes, M., Garca, A., Dobarganes, M.C., Velasco, J. & Romero, R. (2002). "Influence of
Thermal Treatments Simulating Cooking Processes on the Polyphenol Content in Virgin
Olive Oil". J. Agric. Food Chem. 50(21): 59625967.

88

Buma, T. J. (1971). "Free fat in spray-dried whole milk.1.General introduction and brief
review of literature." Netherlands Milk and Dairy Journal. 25: 33-41.
Carrasco-Pancarbo, A., Cerretani, L., Bendini, A., Segura-Carretero, A., del Carlo, M.,
Gallina-Toschi, T., Lecrcker, G., Compagnone, D., & Fernanadez-Gutierrez, A. (2005).
"Evaluation of antioxidant capacity of individual phenolic compounds in virgin olive oil."
J Agric Food Chem. 53: 8913-8925.
Champagne, C.P., & Fustier, P.

(2007). "Micro encapsulation for the improved delivery

of bioactive compounds into foods." Current opinion in biotechnology. 18: 184-190.


Cho, S., & Dreher, M. L. (2001). Handbook of dietary fiber. New York, Marcel Dekker,
Inc.
Cho, Y.J., Alamed, J., McClements, D.J., & Decker, E.A. (2003). "Ability of Chelators to
Alter the Physical Location and Prooxidant Activity of Iron in Oil-in-Water Emulsions."
Journal of Food Science. 68(6): 19521957.
Choia, M.J., Ruktanonchaia, U., Minb, S.G., Chunb, J.Y., & Soottitantawat, A. (2010).
"Physical characteristics of fish oil encapsulated by -cyclodextrin using an aggregation
method or polycaprolactone using an emulsiondiffusion method " Food Chemistry. 119
(4): 1694-1703.
Christie, W. W. (2003). Lipid analysis: Separation, Identification and structural analysis
of Lipids. Bridgwater, The Oily Press, UK.
Cicerale, S., Conlan, X.A., Sinclair, A.J., & Keast, R.S. (2009). "Chemistry and health of
olive oil phenolics." Crit Rev Food Sci Nutr. 49(3): 218-236.
Coupland, J.N., and McClements, D.J. (1996). "Lipid oxidation in food emulsions." Trends
in Food Science & Technology 7(3): 83-91.

89

Daniells, S. (2007)."Omega-3-encapsulation-in-chitosan-gets-study-boost." Foodnavigator


Available at: http://www.foodnavigator.com/Science-Nutrition/Omega-3-encapsulation-inchitosan-gets-study-boost. [Retrieved 10 December, 2008]
Day, L., Seymour, R.B., Pitts, K.F., Konczak, I. and Lundin, L. (2009). "Incorporation of
functional ingredients into foods." Trends In Food Science and Technology. 20(9):
388-395.
Dickinson, E., & Woskett, C. C. (1989). "Competitive adsorption between proteins and
small-molecule surfactants in food emulsions." In the proceedings of the Royal Society of
Chemistry, London, UK. Bee, R.D., Richmond,P., & Mingins, J. (Ed.). Food colloids.
74-96.
Dickinson, E. (1994). "Colloidal aspects of beverages." Food Chem. 51: 343347.
Dobry, D. E., Settell, D.M., Baumann, J. M., Ray, R.J., Graham, L. J., & Beyerinck, R. A.
(2009). "A Model-Based Methodology for Spray-Drying Process Development" Journal of
Pharmaceutical Innovation 4(3): 133-142.
Donmez, M. (2009). "Determination of fatty acid compositions and cholesterol levels of
some freshwater fish living in Porsuk Dam, Turkey." Chemistry of Natural Compounds
45(1): 14-17.
Donnelly, J. L., Decker, E.A., & McClements, D.J. (1998). "Iron-Catalyzed Oxidation of
Menhaden Oil as Affected by Emulsifiers." Journal of Food Science 63(6): 9971000.
Drsri. ( 1 July 2009). "Health Benefits of Omega-3 Fatty Acids!!!!" Bukisa. Available
at:http://www.bukisa.com/articles/115057_health-benefits-of-omega-3-fatty-acids.
[Retrieved 21 December, 2010]

90

Drusch, S. (2007). "Sugar beet pectin: A novel emulsifying wall component for
microencapsulation of lipophilic food ingredients by spray-drying." Food Hydrocolloids
21(7): 1223 - 1228.
Drusch, S., Yvonne, S., Scampicchio, M., Schmidt-Hansberg, B & Schwarz, K. (2007).
"Impact of Physicochemical Characteristics on the Oxidative Stability of Fish Oil
Microencapsulated by Spray-Drying." J. Agric. Food Chem. 55(26): 1104411051.
Drusch, S., & Berg, S. (2008). "Extractable oil in microcapsules prepared by spray-drying:
localisation, determination and impact on oxidative stability." Food Chemistry 109: 17-24.
Drusch, S., & Schwarz, K. (2006). "Microencapsulation properties of two different types
of n -octenylsuccinate-derivatised starch." European Food Research and Technology. 222:
1-2.
EDTA. [online]. Wikipedia. Available at: http://en.wikipedia.org /wiki/Ethylenediamine
tetraacetic_acid [Retrieved 12 August 2010].
Emulsion. Wikipedia. [online]. Available at: http://en.wikipedia.org/wiki/Emulsion
[Retrieved 18 April 2009].
Engel, R. & Schubert, H. (2005). "Formulation of phytosterols in emulsions for increased
dose response in functional foods." Innovative Food Science &amp; Emerging
Technologies. 6(2): 233-237.
Erkkila, A.T., Matthan, N.R., Herrington, D.M., & Lichtenstein, A.H., (2006). "Higher
plasma docosahexaenoic acid is associated with reduced progression of coronary
atherosclerosis in women with CAD." Journal of Lipid Research. 47: 2814-2819.
Extra virgin olive oil. [online]. Wikipedia. Available at: http://en.wikipedia.org/wiki/Oli
ve_oil#Unique_Extra_Virgin_Olive_Oils [Retrieved 25 July 2010].

91

Ferguson, L. R., Philpott, M. (2007). "Cancer Prevention by Dietary Bioactive


Components that Target the Immune Response" Current Cancer Drug Targets. 7(5):
459-464.
Fitzpatrick, J. J., Delaney, T. I. C., Twomey, T., & Keogh, M.K. (2004). "Effect of powder
properties and storage conditions on the flowability of milk powders with different fat
contents." Journal of Food Engineering. 64: 435-444.
Frankel, E. N. (1987). "Secondary products of lipid oxidation." Chem Phys Lipids. 44(2-4):
73-85.
Frankel, E. N., Parks, E.J., Xu, R., Schneeman, B.O., Davis, R.A., & German, J.B.
(1994). "Effect of n-3 Fatty Acid-Rich Fish Oil Supplementation on the Oxidation of Low
Density Lipoproteins." Lipids. 29(4): 233-236.
Frankel, E. N. (2010). "Chemistry of Extra Virgin Olive Oil: Adulteration, Oxidative
Stability, and Antioxidants." J. Agric. Food Chem. 58(10): 59916006.
Fry, S. C. (1983). "Feruloylated pectins from the primary cell wall: their structure and
possible functions." Planta. 157: 111123.
Funami, T., Zhang, G., Hiroe, M., Noda, S., Nakaumaa, M., Asai, I., Cowman, M. K.,
Al-Assaf, S., & Phillips, G.O.

(2007). "Effects of the proteinaceous moiety on the

emulsifying properties of sugar beet pectin." Food Hydrocolloids. 21(13191329).


Gana, C.Y., Chenga, L.H., & Easa, A.M. (2008). "Evaluation of microbial
transglutaminase and ribose cross-linked soy protein isolate-based microcapsules
containing fish oil " Innovative Food Science & Emerging Technologies 9(4): 563-569
Gang, A., Kornauth,W., Mlczoch, J., Sulm, O., & Klose, B. (1980). "Different metabolism
of saturated and unsaturated long chain plasma free fatty acids by intestinal mucosa of
rats." Lipids 15(2): 75-79.

92

Garg, M. L., Wood, L.G., Singh, H., & Moughan, P.J. (2006). "Means of Delivering
Recommended Levels of Long Chain n-3 Polyunsaturated Fatty Acids in Human Diets."
Journal of Food Science 71(5): R66R71.
Geusens, P., Wouters, C., Nijs, J., Jiang, Y., & Dequeker, J. (1994). "Long-term effect of
omega-3 fatty acid supplementation in active rheumatoid arthritis". Arthritis &
Rheumatism. 37(6): 824829.
Gharsallaoui, A., Roudauta, G., Chambina, O., Voilleya, A., & Saurela, R. (2007).
"Applications of spray-drying in microencapsulation of food ingredients: An overview."
Food Research International. 40(9): 1107-1121.
GISSI-HF investigators. (2008). "Effect of n-3 polyunsaturated fatty acids in patients with
chronic heart failure (the GISSI-HF trial): a randomised, double-blind, placebo-controlled
trial." The Lancet. 372(9645): 1223-1230.
GISSI-PrevenzioneInvestigators.

(1999).

"Dietary

supplementation

with

n-3

polyunsaturated fatty acids and vitamin E after myocardial infarction: results of the GISSI
-prevenzione trial." Lancet. 354:447-455.
Goldstein, S., Meyerstein, D., & Czapski, G. (1993). "The Fenton reagents." Free Radical
Biology and Medicine. 15: 435-445.
Gotoh, N., Aoki, T., Nakayasu, K., Tokairin, S., Noguchi, N., & Wada, S. (2006).
"Quantification Method for Triglyceride Molecular Species in Fish Oil with High
Performance Liquid Chromatography - Ultraviolet Detector - Evaporative Light Scattering
Detector." Journal of Oleo Science. 55(9): 457-463.
Graf, E., Mahoney, J.R., Bryant, R.G., & Eaton, J.W. (1984). "Iron-catalyzed hydroxyl
radical formation." The Journal of Biological Chemistry. 259: 3620-3624.

93

Griel, A.E., Kris-Etherton, P.M., Hilpert, K.F., Zhao, G., West, S.G., & Corwin, R.L.
(2007). "An increase in dietary n-3 fatty acids decreases a marker of bone resorption in
humans." Nutrition Journal. 6: 2-10.
Guillon, F., & Thibault, J.F. (1990). "Oxidative cross-linking of chemically and
enzymatically modified sugarbeet pectin." Carbohydr Polym. 12(4): 353374.
Haban, P., Klvanova, J., Zidekova, E., & Nagyova, A. (2004). "Dietary supplementation
with olive oil leads to improved lipoprotein spectrum and lower n-6 PUFAs in elderly
subjects." Med Sci Monit. 10(4): I49-54.
Haynes, L.C., Finley, J.W. & Levine, H. 1992. Method and liposome composition for the
stabilization of oxidizable substances, Patentgenius 5139803.
Heinzelmann,

K.,

Franke,

K.,

Velasco,

J.,

&

Mrquez-Ruiz,

G.

(2000).

"Microencapsulation of fish oil by freeze-drying techniques and influence of process


parameters on oxidative stability during storage." European Food Research and
Technology. 211: 234-239.
Heinzelmann, K., & Franke, K. (1999). "Using freezing and drying techniques of
emulsions for the microencapsulation of fish oil to improve oxidation stability " Colloids
and Surfaces B: Biointerfaces. 12(3-6): 223-229.
Hooper, L. & Cassidy, A. (2006). "A review of the health care potential of bioactive
compounds." J Sci Food Agric.86: 18051813.
Hrncirik, K., & Fritsche, S.

(2004). "Comparability and reliability of different techniques

for the determination of phenolic compounds in virgin olive oil." European Journal of
Lipid Science and Technology. 106: 540549.

94

Ip, C. (1986). "Interaction of vitamin C and selenium supplementation in the modification


of mammary carcinogenesis in rats." J Natl Cancer Inst. 77(1): 299-303.
IPPA International Pectin Producers Association. (2001). [online]. Available at:
http://www.ippa.info/types_of_pectin.htm [Retrieved 21 January 2009].
Jacobsen, C., Timm, M., & Meyer, A.S. (2001). "Oxidation in Fish Oil Enriched
Mayonnaise: Ascorbic Acid and Low pH Increase Oxidative Deterioration." J. Agric. Food
Chem. 49(8): 39473956.
Jacobsen, C. (2010). "Enrichment of foods with omega-3 fatty acids: a multidisciplinary
challenge." Annals of the New York Academy of Sciences. 1190: 141150.
Jacobsen C., L., M.B., Nielsen, N.S., & Meyer, A.S. (2008). "Antioxidant strategies for
preventing oxidative flavour deterioration of foods enriched with n-3 polyunsaturated
lipids : a comparative evaluation." Trends in food science & technology. 19(2): 76-93.
Jafari, S. M., He, Y. H., & Bhandari, B. (2007). "Role of powder particle size on the
encapsulation efficiency of oils during spray drying." Drying Technology. 25(4-6):
1081-1089.
Jafari, S. M., Assadpoor, E., Bhandari, B., & He, Y (2008). "Nano-particle encapsulation
of fish oil by spray drying." Food Research International. 41(2): 172-183.
Kabara.J. J., & Orth, D. S., Ed. (1997). Preservative-free and self-preserving cosmetics
and drugs: principles and practices, CRC Press.
Katsuda, M. S., McClements, D. J.,

Miglioranza, L. H. S., & Decker, E. A.

(2008).

"Physical and Oxidative Stability of Fish Oil-in-Water Emulsions Stabilized with


-Lactoglobulin and Pectin " J. Agric. Food Chem. 56(14): 59265931.

95

Kean, E. G., Hamaker, B.R., & Ferruzzi, M.G. (2008). "Carotenoid Bioacessability from
Whole Grain and Degermed Maize Meal Products." J Agric Food Chem. 56: 9918- 9926.
Kirby, A. R., MacDougall, A.J., & Morris, V.J. (2006). "Sugar Beet PectinProtein
Complexes " Food Biophysics. 1(1): 51-56.
Klaypradit, W., & Huang, Y.W.

(2008). "Fish oil encapsulation with chitosan using

ultrasonic atomizer." LWT-Food Sci. Technol. 41: 1133-1139.


Klinkesorn, U., Sophanodora, P., Chinachoti, P., Decker, E.A., & McClements, D.J.
(2005a). "Encapsulation of emulsified tuna oil in two-layered interfacial membranes
prepared using electrostatic layer-by-layer deposition." Food Hydrocolloids 19(6):
10441053.
Kolanowski, W., & Laufenberg, G. (2006). "Enrichment of food products with
polyunsaturated fatty acids by fish oil addition " European Food Research and
Technology. 222(3-4): 472-477.
Lerma-Garcia, M. J., Sim-Alfonso, E.F., Chiavaro, E., Bendini, A., Lercker, G., &
Cerretani, L. (2009). "Study of Chemical Changes Produced in Virgin Olive Oils with
Different Phenolic Contents during an Accelerated Storage Treatment." J. Agric. Food
Chem. 57(17): 78347840.
Leroux, J., Langendorff, V., Schick, G., Vaishnav, V., & Mazoyer, J. (2003). "Emulsion
stabilizing properties of pectin." Food Hydrocolloids. 17: 455462.
Levine, H., & Slade, L. (1986). "A polymer physiochemical approach

to the study of

commercial starch hydrolysis products (SHPs)." Carbohydr. Polym. 6: 213-244.

96

Lin, P.-Y., & Su, K-P. (2007). "A Meta-Analytic Review of Double-Blind,
Placebo-Controlled Trials of Antidepressant Efficacy of Omega-3 Fatty Acids." J Clin
Psychiatry. 68(7): 10561061.
Livney, Y. D. (2010). "Milk proteins as vehicles for bioactives " Current Opinion in
Colloid & Interface Science. 15: 73-83.
Lpez-Lpez, I., Cofrades, S., Ruiz-Capillas, C., & Jimnez-Colmenero, F. (2009).
"Design and nutritional properties of potential functional frankfurters based on lipid
formulation, added seaweed and low salt content." Meat Science. 83(2): 255-262.
Lumsdon, S. O., Friedmann, T.E., & Green, J.H. (2005). Encapsulation of oils by
coacervation. WO 2005/105290.
Luqman S, & Rizvi, S.I. (2006). "Protection of lipid peroxidation and carbonyl formation
in proteins by capsaicin in human erythrocytes subjected to oxidative stress." Phytother
Res. 20: 303306.
Manach, C., Scalbert, A., Morand, C., Rmsy, C., & Jimnez L. (2004). "Polyphenols Food sources and bioavailability." Am. J. Clin. Nutr. 79: 727-747.
Mrquez-Ruiz, G., Velasco, J., & Dobarganes, C. (2003). Oxidation in Dried
Microencapsulated Oils. Lipid Oxidation Pathways. A. Kamal-Eldin, AOCS Publishing
Martinez-Dominguez, E., de la Puerta, R., & Ruiz-Gutierrez, V. (2001). "Protective effects
upon experimental inflammation models of a polyphenol-supplemented virgin olive oil
diet." Inflammatory Research. 50(2): 102-106.
Masters, K. (1991). Spray drying handbook. New York, John Wiley & Sons.

97

Mattia, C. D. d., Sacchettia, G., Mastrocolaa, D., & Pittiaa, P. (2009). "Effect of phenolic
antioxidants on the dispersion state and chemical stability of olive oil O/W emulsions "
Food Research International. 42(8): 1163-1170
McClements, D. J. (1999). Food emulsions-principles, practice and techniques. Boca
Raton,FL, CRC Press LLC.
McClements, D. J., & Decker, E.A. (2000). "Lipid Oxidation in Oil-in-Water Emulsions:
Impact of Molecular Environment on Chemical Reactions in Heterogeneous Food
Systems." Journal of Food Science. 65(8): 12701282.
McClements, D. J. (2005). Food emulsions: Principles, Practices & Techniques. Boca
Raton, Fl, CRC Press
McClements, D. J., Decker, E.A., Park, Y. & Weiss, J. (2009). "Structural Design
Principles for Delivery of Bioactive Components in Nutraceuticals and Functional Foods."
Critical Reviews in Food Science and Nutrition.49(6): 577-606.
McNamee, B. F., O'Riordan, E.D., & O'Sullivan, M. ( 1998). "Emulsification and
Microencapsulation Properties of Gum Arabic." J. Agric. Food Chem. 46(11): 45514555.
Mellor, J.E., Laugharne, J., & Peet, M. (1995). "Schizophrenic symptoms and dietary
intake of n-3 fatty acids." Schizophrenia Research. 18(1): 85-86.
Micard, V., & Thibault, J.F. (1999). "Oxidative gelation of sugarbeet pectins: use of
laccases and hydration properties of the cross-linked pectins." Carbohydr Polym 39(3):
265273.
Micro-encapsulation. [online]. Wikipedia. Available at: http://en.wikipedia.org/wiki/Mic
roencapsulation [Retrieved 9 December 2008].

98

Miller, G. (2008). "Nanotechnology the new threat to food." Available at:


http://www.globalresearch.ca/index.php?context=va&aid=10755.[Retrieved 22 November,
2010]
Miller, G., & Kinnear, S. (2007). Nanotechnology the new threat to food. Clean Food
Organic. 4.
Moreau, D. L., & Rosenberg, M. (1993). "Microstructure and fat extractability in
microcapsules based on whey proteins or mixtures of whey proteins and lactose." Food
Structure. 12: 457468.
Morris, G. A., Ralet, M.C, Bonnin, E., Thibault, J.F., & Harding, S.E. (2010). "Physical
characterisation of the rhamnogalacturonan and homogalacturonan fractions of sugar beet
(Beta vulgaris) pectin." Carbohydrate Polymers 82(4): 1161-1167.
Morris, M.C., Evans, D.A., Bienias, J.L., Tangney, C.C., Bennett, D.A, Wilson, R.S.,
Aggarwal N, & Schneider, J.(2003). "Consumption of fish and n-3 fatty acids and risk of
incident Alzheimer disease." Arch Neurol. 60(7): 940-946.
Murphy, O. (2001). "Non-polyol low-digestible carbohydrates: food applications and
functional benefits." British Journal of Nutrition. 85(1): S47-S53.
Nakauma, M., Funami, T., Noda, S., Ishihara, S., Al-Assaf, S., Nishinari, K. & Phillips,
G.O. (2008). "Comparison of sugar beet pectin, soybean soluble polysaccharide, and gum
arabic as food emulsifiers. 1. Effect of concentration, pH, and salts on the emulsifying
properties." Food Hydrocolloids. 22(7): 1254-1267.
Nawar, W. W., Ed. (1996). Lipids. Food chemistry. New York, Marcel Dekker.
Nielsen, S. S., Ed. (2010). Food Analysis. Technology & Engineering, Springer.

99

Nutrient Bioavailability. (2003). Encyclopaedia of Food and Culture, Answers.com. The


Gale Group, Inc, Avaialble at: http://www.answers.com/topic/nutrient-bioavailability
[Retrieved 22 November 2010].
Omega-3 fatty acids. [online]. Available at: http://www.umm.edu/altmed/articles/omega-3000316.htm [Retrieved 21 December 2010].
Oleic acid. Wikipedia. [online]. Available at: http://en.wikipedia.org/wiki/Oleic acid
[Retrieved 15 November 2008].
Oliveras-Lpez, M. J., Bern, G., Carneiro, E.M., de la Serrana, L.G.H, Martn, F. &
Lpez, M.C. (2008). "An Extra-Virgin Olive Oil Rich in Polyphenolic Compounds Has
Antioxidant Effects in Of1 Mice1,2." J. Nutr. 138(6): 1074-1078.
Olive oil. [online]. Wikipedia. Available at: http://en.wikipedia.org/wiki/ Olive_oil
[Retrieved 25 July 2010].

Olive oil source. [online]. Available at: http://www.oliveoilsource.com/page/chemical-char


acteristics. [Retrieved 23 October 2010].
Onwulata, C., Ed. (2005). Encapsulated and powdered foods, CRC Press.
zkan, N., Withy, B., & Chen, X.D. (2003). "Effects of time, temperature, and pressure on
the cake formation of milk powders " Journal of Food Engineering. 58(4): 355-361.
Paiva-Martins, F., & Gordon, M.H. (2002). "Effects of pH and ferric ions on the
antioxidant activity of olive polyphenols in oil-in-water emulsions." Journal of the
American Oil Chemists' Society. 79(6): 571-576.

100

Paiva-Martins, F., Santos, V., Mangerico, H., & Gordon, M.H. (2006). "Effects of copper
on the antioxidant activity of olive polyphenols in bulk oil and oil-in-water emulsions." J
Agric Food Chem. 54(10): 3738-3743.
Pandey, K.B, Mishra. N., & Rizvi, S.I. (2009). "Protective role of myricetin on markers of
oxidative stress in human erythrocytes subjected to oxidative stress." Nat Prod Commun.
4: 221-226.
Pandey, K. B., & Rizvi, S.I. (2009). "Plant polyphenols as dietary antioxidants in human
health and disease". Oxid Med Cell Longev. 2(5): 270278.
Paques, J. P., & van Rijn, C.J.M. (2007). Nanoencapsulation of probiotics. Micro Nano
conferentie 2007.
Parada, J., & Aguilera, J.M.

(2007). "Food Microstructure Affects the Bioavailability of

Several Nutrients." J Food Sci. 72(2): 21-31.


Parkanyiova L., T., L., Reblova, Z., Zainuddin, A., Nguyen H.T.T., Sakurai, H., Miyahara,
M., Pokorny, J. (2000). "Resistance of high-oleic peanut oil against autoxidation under
storage and deep frying conditions." Czech J. Food Sci. 18 (Special): 125-126.
Pectin. [online]. Wikipedia. Available at: http://en.wikipedia.org/wiki/pectin [Retrieved 21
January 2010].
Persyn,

J.,

&

Oxely,

J.

(2008).

"Micro/Nano

Encapsulation."

Available

at:

http://www.swri.org/4org/d01/microenc/microen/nanoencap.htm. [Retrieved 12 December


2008].
Pfeuffer, M. & Schrezenmeir, J. (2000). "Bioactive substances in milk with properties
decreasing risk of cardiovascular diseases." British Journal of Nutrition. 84(1): S155-S159.

101

Pham-Huy, L.I., He, H., Pham-Huy, C. (2008). "Free radicals, antioxidants in disease and
health." Int J Biomed Sci. 4(2): 89-96.
Polyphenols. Wikipedia. [online]. Available at: http://en.wikipedia.org/wiki/polyphenols
[Retrieved 26 December 2008].
Protein precipitation. [online]. Wikipedia. Available at: http://en.wikipedia.org/w
iki/Protein_precipitation [Retrieved 5 May 2011].
Psomiadou, E., & Tsimidou, M. (2002). "Stability of virgin olive oil 1:Autoxidation
studies." J. Agric. Food Chem. 50: 716721.
Raghavendra, C. M., Babu, V.R., Rangaswamy, V., Patel, P. & Aminabhavi, T.M. (2008).
"Nano/micro

technologies

for

delivering

macromolecular

therapeutics

using

poly(d,l-lactide-co-glycolide) and its derivatives." Journal of Controlled Release. 125(3):


193-209.
Rees, D. (1969). "Structure, conformation, and mechanism in the formation of
polysaccharide gels and networks." Adv Carbohydr Chem Biochem. 24: 267-332.
Reineccius, G. A. (1995). Liposomes for controlled release in the food industries. ACS
Symposium Series, Encapsulation and Controlled Release of Food Ingredients.
Ren, H., Ghebremeskel, K., Okpala, I., Lee, A., Ibegbulam, O., Crawford, M. (2008).
"Patients with sickle cell disease have reduced blood antioxidant protection." Int J Vitam
Nutr Res. 78(3): 139-147.
Richardson, A. J. (2004). "Clinical trials of fatty acid treatment in ADHD, dyslexia,
dyspraxia and the autistic spectrum." Prostaglandins Leukot Essent Fatty Acids 70(4):
383-390.

102

Sacchi, R., Paduano, A., Fiore,F., Medaglia,D.D., Ambrosino, M.L., & Medina, I. (2002).
"Partition Behavior of Virgin Olive Oil Phenolic Compounds in OilBrine Mixtures
during Thermal Processing for Fish Canning." J. Agric. Food Chem. 50(10): 28302835.
Sanguansri, L., & Augustin, M.A. (2006). Microencapsulation and Delivery of Omega-3
Fatty Acids. Functional Food Ingredients and Nutraceuticals, CRC Press.
Saulnier, L., & Thibault, J.F. (1999). "Ferulic acid and diferulic acids as components of
sugarbeet pectins and maize bran heteroxylans." J Sci Food Agric. 79(3): 396402.
Scalbert, A., Morand, C., Manach, C., & Rmsy, C. (2002). "Absorption and metabolism
of polyphenols in the gut and impact on health." Biomedecine & Pharmacotherapy. 56(6):
276-282.
Schmid-Bondzyndki-Ratzlaff method. (1988). Standards Australia. Australian Standard
AS 2300.1.3 Determination of fat: Gravimetric method. Australian Standard Methods of
Chemical and Physical testing for the Dairying Industry: General methods and Principles;
SAI GLOBAL. Publisher of Australian Standards:Parramatta,NSW,Australia.
Serfert, Y., Drusch, S. & Schwarz, K. (2009). "Chemical stabilisation of oils rich in
long-chain polyunsaturated fatty acids during homogenisation, microencapsulation and
storage." Food Chemistry. 113(4): 1106-1112
Shahidi, F. (2004). "Functional Foods: Their Role in Health Promotion and Disease
Prevention." Journal of Food Science. 69: R146R149.
Shen, Z. (2010). Personal communication.
Siew, C. K., & Williams, P. A. (2008). "Role of protein and ferulic acid in the
emulsification properties of sugar beet pectin." J. Agric and Food Chemistry. 56:
4164-4171.

103

Singh, M., Arseneault, M., Sanderson, T., Murthy, V., & Ramassamy, C. (2008).
"Challenges for Research on Polyphenols from Foods in Alzheimers Disease:
Bioavailability, Metabolism and Cellular and Molecular Mechanisms." J. Agric. Food
Chem. 56(13): 4855-4873.
Smith, J., & Charter, E., Ed. (2010). Functional Food Product Development. Blackwell
Publishing Ltd.
Snyder, J. M., Frankel, E.N., Selke, E., & Warner, K. (1988). "Comparison of gas
chromatographic methods for volatile lipid compounds in soybean oil." J. Am. Oil Chem.
Soc. 65: 1617-1620.
Soottitantawat, A., Bigeard, F., Yoshii, H, Furuta, T., Ohkawara, M., & Linko, P. (2005).
"Influence of emulsion and powder size on the stability of encapsulated d-limonene by
spray drying." Innovative Food Science & Emerging Technologies. 6(1): 107-111.
Spencer, C. M., Cai, Y., Martin, R., Gaffney, S.H., Goulding, P.N., Magnolato, D., Lilley,
T.H. & Haslam, E. (1988). "Polyphenol complexationsome thoughts and observations "
Phytochemistry. 27(8): 2397-2409.
Spray drying. [online]. Wikipedia. Available at: http://en.wikipedia.org/wiki/ spray_drying
[Retrieved 22 December 2010].
Tamime, A. Y., Ed. (2009). Dairy powders and concentrated milk products. Technology &
Engineering, ohn Wiley and Sons.
Teunou, E., & Fitzpatrick, J. J. (2000). "Effect of storage time and consolidation on food
powder flowability " Journal of Food Engineering. 43: 97-101.

104

Thibault, J. F. (1988). "Characterisation and oxidative crosslinking of sugar-beet pectins


extracted from cossettes and pulps under different conditions." Carbohydr Polym. 8:
209-223.
Thies, C. (1996). A Survey of Microencapsulation Processes. Microencapsulation:
Methods and Industrial Applications. S. Benita. New York, Marcel Dekker: 1-20.
Tong, L. M., Sasaki, S., McClements, D.J., & Decker, E.A. (2000). "Antioxidant Activity
of Whey in a Salmon Oil Emulsion." Journal of Food Science. 65(8): 13251329.
Tripoli, E., Giammanco, M., Tabacchi, G., Di Majo, D., Giammanco, S., & Maurizio La
Guardia (2005). "The phenolic compounds of olive oil: structure, biological activity and
beneficial effects on human health." Nutrition Research Reviews. 18: 98112.
Trojakova, L., Reblova, Z. & Pokorny, J. (2001). "Determination of oxidative stability in
mixtures of edible oil with nonlipidic substances." Czech J. Food Sci. 19: 19-23.
Unsaturated fats. [online]. Wikipedia. Available at: http://en.wikipedia.org/wiki/Unsatu
rated_fat [Retrieved 16 June 2010].
Vaclavik, V.A., & Christian, E.W. (2008). Proteins in food : An introduction. Essentials of
Food Science. New York, Kluwer Academic/ Plenum Publishers: 131-145.
Vasishtha, N. (2003). "Microencapsulation: Delivering a market advantage " Prepared
foods, Available at: http://www.preparedfoods.com/Articles/Cover_Story/7ae5dfd391788
010VgnVCM100000f932a8c0. [Retrieved 12 December 2010]
Vega, C., Kim, E. H. J., Chen, X. D., & Roos, Y. H. (2005). "Solid-state characterization
of spray-dried ice cream mixes." Colloids and SurfacesB: Biointerfaces. 45: 66-75.

105

Vega, C., & Roos, Y. H.

(2006). "Invited Review: Spray-Dried Dairy and Dairy-Like

EmulsionsCompositional Considerations." J. Dairy Sci. 89: 383-401.


Velasco, J., Dobarganes, C., & Mrquez-Ruiz, G. (2003). "Variables affecting lipid
oxidation in dried microencapsulated oils." Grasas y Aceites. 54(3): 304-314.
Velasco.J., Dobarganes. M.C., & Marquezruiz .G. (2000). "Oxidation of free and
encapsulated oil fractions in dried microencapsulated fish oils." Grasas y Aceites. 51(6):
439-446.
Volker, D. H., Weng, X., & Quaggiotto, P. (2005). "Bioavailability of long-chain omega-3
polyunsaturated fatty acid enriched luncheon meats." Nutrition & Dietetics. 62(4):
130-137.
Wakil, A., Mackenzie, G., Diego-Taboada, A., Bell .J.G., & Atkin, S. L.(2010). "Enhanced
Bioavailability of Eicosapentaenoic Acid from Fish Oil After Encapsulation Within Plant
Spore Exines as Microcapsules." Lipids. 45(7): 645-649.
Wallace, J.M., McCabe, A.J., Robson, P.J., Keogh, M.K., Murray, C.A., Kelly, P.M.,
Mrquez-Ruiz, G., McGlynn, H., Gilmore, W.S., & Strain, J.J. (2000). "Bioavailability of
n-3 polyunsaturated fatty acids (PUFA) in foods enriched with microencapsulated fish oil."
Ann Nutr Metab. 44(4): 157-162.
Warner, K., Frankel, E.N., & Moulton, K.J. (1988). "Flavour evaluation of crude oil to
predict the quality of soybean oil." J. Am. Oil Chem. Soc 65: 386-391.
Watkins, S.M., & German, J.B., Ed. (1998).Omega fatty acids. Lipid chemistry. New
York, Marcel Dekker.

106

Weiss, L.A., Barrett-Connor, E., & von Muhlen, D. (2005). "Ratio of n-6 to n-3 fatty acids
and bone mineral density in older adults: the Rancho Bernardo Study." Am J Clin Nutr.
81(4): 934-938.
Willats, W. G. T., Knoxb, J.P. & Mikkelsenc, J.D. (2006). "Pectin: new insights into an old
polymer are starting to gel." Trends in Food Science & Technology. 17(3): 97-104.
Willatts, P., Forsyth, J.S., DiModugno, M.K., Varma, S., & Colvin, M. (1998). " Influence
of Long-Chain Polyunsaturated Fatty Acids on Infant Cognitive Function." Lipids. 33:
973-980.
Willet, W. C., Sack, F., Trichopoulou, A., Drescher, G., Ferro-Luzzi, A., Helsing, E., &
Trichopoulos, D. (1995). "Mediterranean diet pyramid: a cultural model for healthy
eating." American Journal of Clinical Nutrition. 61: 1402S-1406S.

Zeta Potential Measurement Using An Autotitrator From Malvern Instruments And The
Effect Of pH. (2005). [online]. Available at: http://www.azom. com/details .asp? Article
ID=2840 [Retrieved 9 December 2010].

107

APPENDICES
1. Percentage protein present in the SBP sample.
The nitrogen content in SBP sample (triplicate analysis) found by LECO was:
SBP replicate 1 = 3.887 mg
SBP replicate 2 = 3.871 mg
SBP replicate 3 = 3.950 mg
Average of replicate 1 +

2 + 3= 3.9 mg

N % = N in mg/( Mass of sample (g) * 10)


= 3.9/ (0.5*10)
= 0.78
% protein = N% * factor
(Factor for general samples= 6.25
Factor for samples containing milk= 6.38)
Therefore, % protein = 0.78 *6.25
= 4.9

2. Propanal content calculation


std 0.1 l = 1987073 peak area counts
Sample (2 g) gave 80137 peak area counts
Therefore amount of propanal equivalents is;
X l propanal = 80137
so 80137 x 0.1/1987073
=0.00403 l/2 g
=0.0020 l/g sample

108

Density of propanal is 0.81 g/ml, therefore we need to convert the propanal concentration
to ml/g to have the same units, so have to multiply by 10-3
so 0.0020 x 10-3 ml (0.81 g/ml)
is 1.62 * 10-6 g
which is 1.62 g/g of powder.

3. Hexanal content calculation


std 0.1 l = 1996252 peak area counts
Sample (2 g) gave 75559 peak area counts
Therefore amount of hexanal equivalents is
X l hexanal = 75559
So 75559 x 0.1/1996252
=0.0038 l/2 g
=0.00189 l/g sample
Density of hexanal is 0.81 g/ml, therefore we need to convert the hexanal concentration to
ml/g to have the same units, so have to multiply by 10-3
so 0.00189 x 10-3 ml (0.81 g/ml)
is 0.00000153 g
which is 1.53 g/g of powder.

4. Total phenolic content in EVOO


The equation obtained from the calibration curve plotted for the caffeic acid is as follows:
y= 3.8617x+ 0.0759
109

Substituting the value of y (absorbance value of the extract at 725 nm) in the above
equation,
x = 0.257- 0.0759 / 3.8617
= 0.047 mg/ ml caffeic acid in 0.2 ml
Therefore, in 5.2 ml of the extract there is 0.244 mg caffeic acid.
0.244 mg caffeic acid is present in 2.5 g oil and hence, there is 97.5 mg/kg oil.

110

Anda mungkin juga menyukai