Sunteți pe pagina 1din 114

THE EFFECTS OF SOY PROTEIN ISOLATE ADDITION ON THE PHYSICO-CHEMICAL

PROPERTIES OF GUMMI CONFECTIONS

THESIS

Presented in Partial Fulfillment of the Requirements for the Degree Master of Science in
the Graduate School of The Ohio State University

By
Alexander Martin Siegwein
Graduate Program in Food Science and Nutrition

The Ohio State University


2010

Master's Examination Committee:

Dr. Yael Vodovotz, Advisor


Dr. Steven J. Schwartz
Dr. W. James Harper

Copyright by
Alexander Martin Siegwein
2010

ABSTRACT

Soy protein is commonly used to enhance the nutritional quality of food.


However, it also has the ability to act as a functional ingredient, imparting unique
characteristics to foods to which it is added. The effects of soy protein isolate
incorporation on the physico-chemical properties of starch-based gummi confections
were investigated using thermal and rheological analyses, as well as supportive sensory
and spectroscopy studies. The overall objectives of this study were to first characterize
the textural and rheological changes upon varying levels of soy addition to starch
confections and consequent changes to quality parameters and second to assess the
impact of incorporation of soy protein isolate on storage stability.
Texture profile analysis was conducted to simulate mastication and to quantify,
along with a hedonic sensory panel, the effects on gummi acceptability as a function of
texture. Increasing levels of soy protein yielded samples that were progressively less
firm and cohesive, caused either by starch network dilution and/or disruption. The
softening effect of soy protein was observed throughout storage. Addition of soy protein
improved the texture acceptability, presumably by decreasing perceived shortness, a
texture defect associated with overly-firm gelled confections. Rheometric analyses were
conducted to determine the relationship between gummi viscoelastic properties and the
applied stress and rate thereof. Addition of soy protein was found to progressively
ii

suppress the yield increase at each storage interval. The viscoelastic crossover frequency,
an indicator of viscoelastic stability, was also found to decrease as soy protein
concentration increased and to have a lower rate of increase during storage.
Additionally, water-dependent interactions were characterized as they relate to
soy concentration and storage time. Thermogravimetric analysis results indicated that the
primary water population was more easily removed in high-soy protein formulations.
Additionally, as storage time increased, soy protein maintained relative homogeneity of
the primary water population and removal of water required lower temperatures with the
soy formulation, indicating less entrapment of water throughout the storage period
studied.
Results from differential scanning calorimetry did not indicate a significant shift
in the glass transition temperature as soy concentration increased, most likely due to the
low concentration of plasticizing water. Neither gummi system exhibited distinct
endotherms associated with starch melting regardless of treatment; soy protein inclusion
did not prevent complete starch gelatinization during the confection process, as evidenced
by the lack of the typical starch melting endotherm. However, over time, the addition of
soy protein was able to prevent an increase in the T g of the gummi confections, perhaps
by interfering with starch re-association and retrogradation.
The changes in monomeric anthocyanin quantity, expressed as units of cyanidin
3-glucoside, which occurred as a result of confection processing were also analyzed using
ultraviolet spectroscopy. The soy gummi confections yielded higher anthocyanin
recovery compared to the standard formulation. This is probably due to the greater heat
iii

dissipation caused by a lower cooking viscosity from the soy formulation during the
confection process. Sensory analysis, conducted using a hedonic scale of acceptability,
showed that addition of soy protein improved acceptance for both texture and flavor. The
mechanism of improvement is thought to be related to the decreased rigidity (shortness)
of the product which augments mouth-feel and possibly flavor release.
All results indicate that soy protein has potential as a processing aide in gelled
confections by modulating physical properties. Gummi gel rheology was shown to be
dependent on soy concentration and storage time. Texture became less firm with
increasing soy protein concentration, and soy protein was shown to decrease firming over
time. The physical results support those of sensory analysis, in which it was found that
soy protein improved acceptability of the gummi confections for both texture and flavor.
The distribution of water and its dynamics were also shown to be time and concentration
dependent. Finally, addition of soy protein may aide in the processing stability of
anthocyanins by increasing heat dissipation.

iv

DEDICATION

This document is dedicated to my grandmother Stella

ACKNOWLEDGEMENTS

I wish to extend my sincerest gratitude to my academic advisor Dr. Yael


Vodovotz for her guidance and patience during my years at The Ohio State University. It
was my privilege to be mentored by one so dedicated to scientific progression. My
appreciation also extends to Dr. Steven Schwartz, Dr. W. James Harper, and Dr. Mike
Mangino for their instruction and assistance. I am grateful to everyone in the conjoined
office; with their wisdom and support, my experience was all the more enriched. Dr.
Alvarez and the pilot plant staff also have my thanks for granting me access to needed
equipment.

vi

VITA

January 1983 ............................................... Born, Boise Idaho


May 2001 .................................................... Valedictorian, Glenns Ferry High School
May 2005 .................................................... B.S. Food Science, University of Idaho
June 2005 to December 2007 ....................... Graduate Research Associate, Department
of Food Science, The Ohio State University
May 2008 to present .................................... Associate Scientist, Abbott Nutrition

FIELDS OF STUDY

Major Field: Food Science and Nutrition

vii

TABLE OF CONTENTS

ABSTRACT ................................................................................................................... ii
DEDICATION ................................................................................................................v
ACKNOWLEDGEMENTS ........................................................................................... vi
VITA ............................................................................................................................ vii
FIELDS OF STUDY .................................................................................................... vii
TABLE OF CONTENTS............................................................................................. viii
LIST OF TABLES ....................................................................................................... xii
LIST OF FIGURES ..................................................................................................... xiii
1. INTRODUCTION .......................................................................................................1
2. STATEMENT OF PROBLEM ....................................................................................4
3. LITERATURE REVIEW ............................................................................................6
3.1

Gummi Confections...........................................................................................6

3.2

Soy functionality in foods ..................................................................................6

3.3

The role of ingredients in gummy confections ...................................................7

3.4

The role of water in gummy confections ............................................................9

viii

3.5

Characterization of physical properties ............................................................ 17

4. METHODOLOGY .................................................................................................... 23
4.1

Gummi Formulation and Manufacture ............................................................. 23

4.2

Thermogravimetric Analysis (TGA) ................................................................ 26

4.3

Texture Profile Analysis (TPA) ....................................................................... 26

4.4

Rheological Analysis ....................................................................................... 27

4.5

UV-Spec Analysis ........................................................................................... 28

4.6

Sensory Analysis ............................................................................................. 28

4.7

Statistical Analysis .......................................................................................... 28

5. ADDITION OF SOY PROTEIN TO STARCH-BASED CONFECTIONS ALTERS....


TEXTURAL PROPERTIES ....................................................................................... 29
5.1

Abstract ........................................................................................................... 29

5.2

Introduction ..................................................................................................... 30

5.3

Methodology ................................................................................................... 32

5.3.1.

Thermogravimetric Analysis (TGA) .......................................................... 34

5.3.2.

Texture Profile Analysis (TPA) ................................................................. 34

5.3.3.

Rheological Analysis ................................................................................ 35

5.3.4.

UV-Spec Analysis ..................................................................................... 36

5.3.5

Sensory Analysis ....................................................................................... 36


ix

5.3.6.
5.4

Statistical Analysis ................................................................................... 36

Results & Discussion ....................................................................................... 36

5.4.1

Effects on texture and macro-structure ..................................................... 36

5.4.2

Effects on gel structural properties ........................................................... 39

5.4.3

Effect on water properties and distributions ............................................. 48

5.4.4

Sensory Analysis ....................................................................................... 51

5.4.5

UV-Spec Analysis ..................................................................................... 52

5.5

Conclusions ..................................................................................................... 54

6. EFFECTS OF SOY PROTEIN ON THE STORAGE STABILITY OF STARCHBASED GELLED CONFECTIONS ........................................................................... 56
6.1

Abstract ........................................................................................................... 56

6.2

Introduction ..................................................................................................... 56

6.3

Methodology ................................................................................................... 59

6.3.1

Differential Scanning Calorimetry (DSC) ................................................. 61

6.3.2

Thermogravimetric Analysis (TGA) .......................................................... 61

6.3.3

Texture Profile Analysis (TPA) ................................................................. 62

6.3.4

Rheological Analysis ................................................................................ 62

6.3.5.

Statistical Analysis ................................................................................... 63

6.4.

Results & Discussion ....................................................................................... 64


x

6.4.1

Texture Profile Analysis ........................................................................... 64

6.4.2

Effects of soy protein on the dynamic rheological properties over time ..... 70

6.4.3

Effect of soy on water dynamics over time ................................................ 74

6.5

Conclusions ..................................................................................................... 79

7. CONCLUSIONS ....................................................................................................... 80
Future Work............................................................................................................... 82
APPENDIX A ............................................................................................................... 83
APPENDIX B ............................................................................................................... 84
APPENDIX C ............................................................................................................... 85
REFERENCES .............................................................................................................. 86

xi

LIST OF TABLES

Table 1

Formulations for gummi confections used in this investigation..................... 24

Table 2

Formulations for gummi confections used in the concentration study. .......... 33

Table 3

Mean hedonic scores for Standard and 50% SPI gummi confections. ........... 51

Table 4

Formulations for gummi confections used in storage time study. .................. 60

xii

LIST OF FIGURES

Figure 1

Typical DSC thermogram of a gummi confection made with wheat starch.


The highlighted area is a putative glass transition. ........................................ 13

Figure 2

Typical TGA thermogram of wheat starch-based gummi confections. Events


at temperatures greater than 160 oC are considered sample destruction. ....... 16

Figure 3

Typical texture profile analysis for starch-based gummi confections. ........... 18

Figure 4

Example of the yield stress (seen as the onset point of decay) of a material
being subjected to an increasing oscillatory stress sweep. ............................. 20

Figure 5

Viscoelastic crossover (G'-G") frequency obtained from an oscillatory


frequency sweep. ......................................................................................... 22

Figure 6

Batching kettle, refractometer, and stoving oven used in the manufacture of


gummi confections. ...................................................................................... 25

Figure 7

Confection hardness and cohesiveness as a function of SPI concentration. ... 38

Figure 8

Typical yield stress observed in an oscillatory stress sweep. ......................... 40

Figure 9

Yield stresses (Pa) as a function of soy protein isolate (SPI) substitution. ..... 43

xiii

Figure 10

Elastic modulus (G') as a function of soy protein isolate substitution. ....... 44

Figure 11

Viscoelastic moduli crossover point (G-G) as a function of soy protein


concentration. Inset graph depicts a typical crossover point. .................... 47

Figure 12

Peak termination temperatures of derivative weight loss curves of gummi


candies with increasing levels of soy substitutions. Inset graph depicts
typical dTGA weight loss peak for 66% SPI variable versus Standard
control. ..................................................................................................... 50

Figure 13

Monomeric anthocyanins content, expressed as units of cyanidin-3glucoside, in the gummi confections versus raw and boiled grape juice
concentrate. .............................................................................................. 53

Figure 14

Changes in gumminess (N) as a function of storage time. ......................... 65

Figure 15

Changes in hardness (kgF) as a function of storage time. .......................... 67

Figure 16

Changes in cohesiveness (mm) as a function of storage time. ................... 69

Figure 17

Change in yield stress (Pa) as a function of time for both the Standard and
50% SPI treatments. ................................................................................. 71

Figure 18

Change in visco-elastic (G'-G) crossover frequency as a function of


storage time in both the Standard and 50% SPI treatment. ........................ 73

Figure 19

Changes in the derivative weight loss curve of the Standard (19a, no soy)
and 50% SPI (19b) gummi confections. .................................................... 75

Figure 20

Comparison in the Tg onset temperature between the Standard and 50% SPI
gummi confections after 20 days. ............................................................. 78

xiv

CHAPTER 1

INTRODUCTION

Gelled confections and candies, collectively referred to as gummies (Edwards,


2000), are a large, stable, and growing business (Young, 1998); As of 2006, industry
experts have reported as high as 8.4% growth in the gummi candy sector, mostly
attributed to innovative formulations (Anon, 2006). Many of the most common and
recognizable gummi confections are typically made with gelatin as the active agent of
gelation (Lennox 2002). However, gelatin is not the only option for creating gummi-type
candies. Starches may be preferred for being free of animal by-products and the
associated concerns (Lennox 2002). However, starch-based confections are sometimes
described as short an undesirable texture attribute characterized by a firm, brittle gel
(Burg, 1998).
Texture is a primary determinant of quality in gelled confectionery products and,
due to their relatively simple composition, provide an ideal model for composite gel
investigations (DeMars & Ziegler, 2001). Gumminess and chewiness are texture
analysis descriptors that are particularly applicable to gelled confections (Borwankar,
1992). Firmer (short) gummies that are more prone to fracture during mastication are
often described as chewy and softer gummies that dissolve during mastication are
described as gummy (DeMars & Ziegler, 2001). One of the primary uses of soy protein

in foods is as a texture modifier (Lusas & Riaz, 1995). Depending on the polysaccharide
gel system, soy protein has been shown to have variable results, enhancing (Baeza et al,
2002) or disrupting the gel's structural integrity (Ryan & Brewer, 2005). Addition of soy
protein ingredients to confections was suggested to improve handling properties and
stickiness (Rhee, 1994). Therefore, in gelled starch-sugar confections, a relatively model
gel system, soy proteins would perhaps be beneficial. Therefore, our hypothesis was that
soy protein affects the texture of the starch-based confections on a concentrationdependent basis. It is also hypothesized that soy protein will have a lasting effect on the
texture and related properties of starch-based confections throughout the defined storage
period.
To understand the textural changes that were affected upon addition of soy
protein, the rheological behavior was assessed. Oscillatory stress sweeps (OSS) reveal
the viscoelastic properties and behaviors of foods. Soy protein has been shown to have a
pronounced effect on viscoelasticity in gel and food systems (Ryan & Brewer, 2005;
Vittadini & Vodovotz, 2003). Oscillating frequency sweeps (OFS) were used to
characterize the rheological properties of the confections as a function of time. In these
gelled systems, OFS can reveal long-term stability (Rao, 2007), such as tendencies for
phase separation, retrogradation, syneresis, or sedimentation. If the addition of soy
protein creates a systemic instability, the oscillatory stress and frequency sweeps should
reveal differences between the various concentration treatments and time intervals.
In addition to the solid components of the gummi confection, water is another
critical component that affects the overall physical properties. Traditional methods of
describing water behavior (water activity, moisture content, and solvent retention

capacity) do not adequately describe the distribution and mobility of water in food
systems (Slade et al., 1991). Thermogravimetric analysis (TGA) can be used to
characterize the changes in physico-chemical properties of a food system as a function of
temperature. TGA can be used to track rate of moisture loss and differentiate different
water populations within a multi-component food (Fessas and Schiraldi, 2000).
Therefore, TGA will be used to characterize and compare water populations in the
gummi confection system and how those are affected by addition of soy protein isolate.
The aim of this investigation was to assess the effect of soy protein addition on
the physico-chemical properties of a starch confection system. Specifically, the textural,
thermal, and rheological properties of starch-based grape juice confections will be
evaluated upon incorporation of multiple levels of soy protein. The effect of soy protein
on monomeric anthocyanin stability, such as cyanidin-3-glucoside (see Appendix A),
which is known to be compromised in high-temperature food preparation methods (Patras
et al., 2010; Edwards, 2000), will also be assessed in the gummi confections. The
sensory acceptability of gummi confections made with soy protein will be evaluated by
an untrained panel with particular emphasis on texture and flavor. Lastly, the storage
stability of gummi confections formulated with and without soy protein will be assessed
through 10 and 20 days using thermal and rheological characterizations.

CHAPTER 2

STATEMENT OF PROBLEM

In addition to its recognized nutritional value, soy protein has demonstrated


potential in food systems as a functional ingredient. Its use in other food systems has
demonstrated a potent capacity to change water dynamics and rheological properties;
therefore, it may have potential as a texture enhancer. It is possible that formulating
gummi confections with soy protein may improve texture and stability.
How soy protein addition will affect the physico-chemical properties and
processing of confectionery products has not been formally investigated. It is therefore
vital that an understanding of component interactions be attained in order to successfully
produce acceptable soy-formulated foods. The objective of this investigation was to
characterize the effects of soy protein isolate on the physico-chemical properties of a
starch-set, gummi-type confection as a function of storage time and soy protein content.
Hypothesis:

Soy protein concentration and storage time will modulate the physico-

chemical properties of starch-based gummi confections by changing the physical nature


of the gel matrix and water distribution. These changes are proposed to be caused by
altering the interactions between system components (water, sugars, starches, and
proteins). The following aims are proposed to investigate this hypothesis:

Aim 1: To characterize and compare the thermal and rheological properties of starchbased gummies with multiple levels of soy protein incorporation.
Aim 2: To characterize the changes that occur in gummi candies with and without soy
protein during storage.
Aim 3: To determine how soy protein addition affects the sensory acceptability and
processing stability of anthocyanins using UV spectroscopy.

CHAPTER 3

LITERATURE REVIEW

3.1

Gummi Confections
In 2009, non-chocolate chewy confectionery sales increased by 11.2% (Anon,

2010). Chewy sugar-based candies are a stable part of the American confectionery
industry (Young, 1998) with sales of over $1.6 billion (Anon, 2010). However, the soft
candy sector of the sugar confectionery industry had recently suffered losses to
production and value (USCB 2005) that industry professionals have attributed to lack of
innovative products (Anon, 2006b). Development of a highly acceptable soy-containing
gummi confection may represent a key new niche for this industry and a passageway into
the area of functional foods.
3.2

Soy functionality in foods


The use of soy protein in foods for nutritional purposes is an established area of

food augmentation (Green et al., 2006; Sethi et al., 2007; Prasad, 2009; Serventi et al.,
2009) that is becoming more and more accepted by Americans (Nelson, 2008). However,
it also has great potential as a functional ingredient, imparting unique physical properties
when incorporated into the formulation (Vittadini & Vodovotz, 2003; Lusaz & Riaz,
1995). Soy protein has also been implicated improving product stability, such as
impeding product staling (Vittadini & Vodovotz, 2003; Anhong et al., 2006) and freeze6

thaw stability (Akesowan & Taweesakulvatchara, 2009), perhaps by simple moisture


retention or by interacting with amylopectin and retarding retrogradation (Sciarini et al.,
2008). However, the use of soy products in confections is limited and is usually
restricted to a supporting role for other ingredients (Endres et al., 2003) due to a variety
of reasons. For example, adverse texture and sensory attributes (chalky, floury, beany,
etc.) of soy limit its application (Starling, 2005). However, addition of soy protein
ingredients to confections has been suggested to improve handling properties and
stickiness (Rhee, 1994). Soy protein requires more energy and time to completely
disperse and hydrate (Mai, 2004). Maillard browning reactions may also compromise
visual appeal in soy protein fortified confections if the system is rich in reducing sugars
and not sufficiently acidic.
3.3

The role of ingredients in gummy confections


Confectionery gels typically consist of sugars, water, and a gelling agent (Burey

et al., 2009). The choice of gelling agent can have the biggest impact on final product
quality and attributes, and therefore must receive special attention. Gelatin, the
traditional gelling agent used in gummy formulations, is losing favor due to cultural,
dietary, and safety concerns (Lennox, 2002; McHugh, 2003), as well as the desire for
unique textural properties (Poppe, 1995). Specialty starches have been developed to meet
the demand for gelatin-free candies while providing unique sensory and physical
properties (Warnecke, 1991). Starch-set gummies, the largest volume gum/jelly product
in the United States (Warnecke, 1991), have greater acid and heat stability and have
much shorter gelation time than gelatin gummies (Burg, 1998), thereby conferring
functional and pecuniary advantages. The most commonly used starches for gummi and
7

soft jelly applications are acid-thinned because they have relatively low hot viscosity
(Ellis et al., 1998; Singh et al., 2007) and can therefore be agitated and pumped with
minimum energy expenditure. Wheat starches have received renewed attention due to
their unique properties (Sloan, 2006). Gummies are traditionally made in a batch process
(Scuderi, 2002; Burey et al., 2009) which, although not modern, does prevent mechanical
damage upon the swollen starch granule (Zallie, 2006). In a standard starch-based soft
candy, a thin boiling starch is heated in water beyond the gelatinization temperature,
mixed with a corn syrup sucrose solution, and boiled down to the desired soluble solids
content. In contrast to relatively well-understood gelatin gummi formulations, the
interactions that occur between water and starch in a high sugar environment (>65%,
such as in gummy-like confections) have not been studied adequately. Chang et al.,
(2002) demonstrated that certain sugars can, in fact, plasticize starch and reduce firmness,
but at lower concentrations. Kasapsis et al., (2004) showed that high sugar
concentrations reduce the structural order of a model confection system, as opposed to
gelatin which is reinforced by high sugar (Kasapsis et al., 2003).
Composite or mixed gels are becoming more common in foods due to the desire
for unique texture and handling characteristics (DeMars & Ziegler, 2001; Burey et al.,
2009). However, the interactions (or lacks thereof) that create such a novel system are
less well understood. In protein-polysaccharide systems, the level of interaction or
separation depends on many factors, such as system pH, moisture, material history, as
well as polymer properties such as charge, size, and morphology (Burey et al., 2009).
For gelled confectionery systems, which are held together by a continuous network
whose dynamics are retarded by high solutes, an added protein may create a

discontinuous phase if it is non-structure forming and competes for limited solvent


(Burey et al., 2009) which could theoretically render the entire system discontinuous at
some critical concentration.
The sweetness in gummi and similar confections is achieved primarily by high
fructose corn syrup (HFCS), which is functionally preferable over pure sucrose due to the
crystallization-inhibiting abilities of fructose; crystallization being an undesirable defect
in soft, sugar confections (Warnecke, 1991). Similar sweetness and functionality can be
realized by using fruit juice concentrate with the added advantage of naturally
contributing pigments (such as anthocyanins), which may impart additional health
benefits (Prior et al., 2000). Concord grape juice (CGJ) flavonoids are potent
antioxidants that may protect against oxidative stress and reduce the risk of free radical
damage and chronic diseases (OByrne et al., 2002). Grape anthocyanins are over 50%
acylated (Tamborra et al., 2006) and are therefore considered to be more bio-available
(Harada et al., 2004) and their deep purple color is less likely to fade over time (Bassa et
al., 1987). Additionally, concentrating fruit juice further promotes product color stability
(Giusti et al., 2003). Lastly, the low pH of CGJ (2.8) may help inhibit Maillard browning
reactions. Anthocyanin content is often expressed as units of cyanidin-3-glucose
(Miyazawa et al.,1999; structure displayed in Appendix A)
3.4

The role of water in gummy confections


Starch-protein interactions and the behavior of water in a high-solids

environment, like confections, as they pertain to product stability have not been
investigated. Gel-formation capacity and viscosity of soy protein hydrates is related to
soy protein composition (native vs. denatured and the extent of denaturation),
9

temperature, and pH (Arrese et al., 1991). Ideally, soy protein and starch will each
contribute to gel formation and firmness. However, soy protein isolate (SPI) may affect
sugar recrystallization, glass transitions, starch retrogradation, and other physical effects
of soft jelly candy as observed in other food systems (Vittadini et al., 2003; Luo et al.,
2003; Tsai et al., 1998). Therefore, to understand how SPI affects the thermal and
physical properties of starch-based soft jelly confections, it is important to first
characterize the physico-chemical properties of the model system without SPI
incorporation.
A basic soft jelly and gummy formulation is a heterogeneous mixture of water,
sugars, and gel-forming polymers. In such a system, water acts as a plasticizer and
lowers the glass transition temperature (Tg) of food constituents. Fructose has also
demonstrated high plasticization capacity on the amylopectin fraction of starch (Liu et al.,
2004). In boiled sweets, like gummy candies and toffees in which final water content is
low (<20%), the product is in an amorphous rubber (preferable) or an amorphous glass,
depending on the extent and rate of water removal (Roos et al., 1991). The addition of
polymeric compounds, such as starch and protein, to a sugar solution tends to increase the
Tg (Rogers et al., 2006), slowing crystallization of amorphous sugars due to decrease free
volume, molecular mobility, and diffusivity (Roos et al., 1996).
Changes in water mobility, independent of moisture content, may be caused by
continuous chemical interactions in the system, such as recrystallization, starch
retrogradation, and glassy/rubbery equilibrium (Cornillon et al., 2000) affecting, among
other things, the shelf life of the product. For example, recrystallization of amorphous
sugars due to increased molecular mobility is an undesirable defect in soft, gummy-like

10

candies because it creates a coarse texture (Hartel, 1993). Additionally, in a system


where anthocyanins are present, low water mobility results in greater stability of the
pigments (Tsai et al., 2004). Water may also exist in multiple populations because of the
heterogeneity of the system; to what extent water is associated with gelatinized starch and
how that may change upon soy addition in a low-water, high-sugar environment is also
unknown. Characterizing the behavior of water is therefore essential to understanding
the molecular dynamics and estimating stability of gummy-like candies.
Established methods of examining water behavior (water activity, moisture
content, and solvent retention capacity) do not adequately describe the distribution and
mobility of water in food systems (Slade et al., 1991). Thermal analytical techniques,
including differential scanning calorimetry (DSC), thermogravimetric analysis (TGA),
can be used to characterize the changes in physico-chemical properties of a food system
as a function of temperature.
The DSC detects phase changes indirectly by comparing the sample heat flow rate
to an inert control as both are exposed to the same change in temperature (Schenz, 2003.
What is directly measured is the amount of heat required to keep the sample cell at the
same temperature as the reference, which is commonly an empty pan (Schenz, 2003).
This heat amount will increase in an endothermic reaction such as melting or decrease in
an exothermic reaction such as crystallization. Melting and crystallization are both 1st
order transitions, so upon reaching the melting temperature (T m), a samples temperature
will not rise until all the crystals have melted, meaning even more heat must be applied to
keep the temperature rising at a constant rate relative to the reference cell (Lind, 1991).
Most foods are a mixture of solids, liquids, and amorphous materials, which presents

11

challenges when attempting to characterize individual transitions. Some first order


transitions, such as starch gelatinization, amylopectin re-crystallization, and freezable
water melting have all been quantified using DSC in select carbohydrate systems
(Takashi & Yamada, 1998; Vittadini & Vodovotz, 2003).
Additionally, DSC can be used to quantify 2nd order transitions, such as the
glassy-rubbery transition. This unique transition is associated with an increase in
molecular motion of the amorphous phase of polymers as they change from the glassy
(rigid) to the rubbery (flexible) state (Slade et al., 1991; Walstra, 2008). This reversible
change occurs at a specific temperature known as the glass transition temperature (T g).
The glass-to-rubber transition is a 2nd order endothermic reaction, meaning that while
heat is required to complete the transition, latent heat is not a factor (Wunderlich, 1994).
Glass transitions are observed under DSC analysis as a sigmoidal change in the heat
capacity of a sample (Wunderlich, 1994). If the Tg of the system or ingredient is above a
given temperature (t o), the system is in a glassy state (rigid) at t o with limited polymeric
motion in the amorphous phase. Conversely, if the T g of the system or ingredient is
below to, the system is in a rubbery (flexible) state and the polymeric mobility is high.
Figure 3.1 below depicts a typical DSC thermogram of a starch-based gummi confection.

12

Figure 1

Typical DSC thermogram of a gummi confection made with wheat starch. The
highlighted area is a putative glass transition.

13

Thermogravimetric analysis (TGA) is a thermo-analytical technique used to


determine the distribution of water within a heterogeneous system by monitoring weight
loss of a material as a function of time or temperature. Samples may either be held at a
constant temperature to examine time-dependent changes, such as oxidation and
evaporation, or be subjected to linear or step-wise temperature increases to determine
temperature dependent reactions (Fessas & Schiraldi, 2001). Derivative weight loss
curves provide a peak that distinctly plots the temperature range(s) of greatest rate of
weight loss (Figure 3.2). In food systems, this is usually weight loss attributed to water.
Experimental conditions and sample treatment history must be tightly controlled
for reproducibility. To conduct TGA experiments, small samples representative of the
whole food are placed into an appropriate, non-reactive sample pan. This pan is attached
to one end of a balance, the other end holding an empty pan. The sample is then loaded
into a furnace and subjected to the time or temperature treatment. During the experiment,
the interior of the furnace is continuously purged with gas (or a controlled series of
gasses) that may be either non-reactive or reactive with the sample, depending on the aim
of the experiment. Data acquired from TGA experiments are usually plotted as weight or
%weight versus time or temperature. While these can themselves be used for
determining the amount of a certain component within the sample, such as for proximate
analyses, little is revealed about rates of weight change. Applying the first derivatives of
these curves reveal these rates of weight loss and can more accurately describe
composition of the sample and can also reveal the most stable (little or no weight change)
regions of time or temperature.

14

Fessas and Schiraldi (2001) have described the use of TGA in bakery products to
monitor the rate of water release as a function of temperature, assigning ingredient effects
to the different components of the food. This means that water is not uniformly
partitioned within foods. Water that is less strongly associated with food components is
removed at lower temperatures by a simple diffusion mechanism (Fessas & Schiraldi,
2001), whereas strongly associated water is removed at high temperatures by a change in
the physical properties of the system itself. Therefore, using TGA will characterize the
water distribution of a standard starch-based soft jelly candy and the changes arising from
the addition of SPI to the candy formulation.

15

Figure 2

Typical TGA thermogram of wheat starch-based gummi confections.


Events at temperatures greater than 160 oC are considered sample
destruction.

16

3.5

Characterization of physical properties


Texture is a primary determinant of quality in confectionery products and, due to

their relatively simple composition, provide an ideal model for texture-structure


investigations (DeMars & Ziegler, 1998). The Instron Universal Testing Machine can be
used to measure the effects of large deformations on a material and is ideal for gelled
sugar confections. When correlated with sensory analysis of foods, texture analysis may
be used as a quantitative supplement to sensory panels in the evaluation of texture
(Szczesniak, 1986). Texture profiles analysis (TPA) is a common and robust method to
analyze foods in a way meant to compare to mastication (Daubert & Foegeding, 2003).
From a TPA (see example in Figure 3.3), it is possible to determine various physical
properties of the material. Hardness (resistance to deformation) is expressed as the
maximum force applied during the first compression and is usually, but not always, the
maximum compression point. Cohesiveness, a measure of the intermolecular forces
within a material, is measured as a relative resistance to deformation between the two
compressions. Lastly, gumminess is a product of the hardness and the cohesiveness and
is defined as the force required to masticate the material to dissolution (Szczesniak,
1962). Using a mixed component gummi confection, DeMars & Zielgler (2001)
correlated analysis of texture rigidity with sensory analysis of perceived mouth-feel, and
even observed that it impacted flavor perception. Therefore, texture profile analysis can
be used to characterize shortness, a texture defect in gummi confections characterized by
firm, brittle gels (Burg, 1998), the effects it has on sensory acceptability, and how soy
protein addition impacts shortness perception.

17

Figure 3

Typical texture profile analysis for starch-based gummi confections.

18

By definition, rheology is the study of deformation and flow (Rao, 1999), which
would therefore include texture analysis. However, the underlying viscoelastic properties
that manifest as textural changes can be probed by small-deformation rheological
analysis. Using a rheometer, it is possible to observe the viscoelastic behavior of a given
material as a function of applied stress and the rate thereof. Oscillatory rheometry is an
established technique for measuring flow and deformation of food matter under applied
stress or strain (Rao, 2007). Rheological properties are also likely to correlate strongly
with mouth-feel (Sukha et al., 2003). Oscillatory stress sweeps can reveal basic
relationships between a food's elastic and viscous behavior. When G' > G, the sample
behaves more like an elastic solid and has higher stress resistance (Brummer, 2005).
Yield stress, or the stress at which a material is strained beyond its linear viscoelastic
region (LVR) and begins to deform permanently (Rao, 2007), is indicative of a gel's
structural integrity. Figure 3.4 depicts the results of a typical yield stress analysis using
an oscillatory stress sweep.

19

Figure 3.4.
4

Example of the yield stress (seen as the onset point of decay) of a material being
subjected to an increasing oscillatory stress sweep.

20

In addition to stress sweeps, frequency sweeps can also reveal dependency on rate
of applied stress and how that relates to micro-structural stability (Pai & Khan, 2002). In
particular, the viscoelastic crossover (G'-G) cross-over frequency can be predictive of
long-term stability (Ross-Murphy, 1994) and any interactions which may affect it (Franco
et al., 1997). For example, a liquid food material displaying a frequency-independent
elastic modulus (G') with a steadily decreasing viscous modulus (G) at high frequencies
may be experiencing extensive polymeric interactions and behave more like a solid
(Ikeda & Nishinari, 2001; Sans et al., 2005), something with serious implications for an
industrial food manufacturer. Alternately, high frequencies can induce polymer
dissocation and cause a gel material to behave more like a liquid (Gigli et al., 2009).
Figure 3.5 depicts the cross-over frequency of a typical oscillatory frequency sweep
analysis on a viscoelastic gel (frequency dependency presents itself as changing slopes on
the log-log axes). Sivaramakrishnan et al., 2004, used an oscillatory frequency sweep to
characterize how an added hydrocolloid can disrupt the normal structural associations
within a model bread system and at what point the dominant structural component is the
added ingredient.

21

Figure 3.5.
5

Viscoelastic crossover (G'-G") frequency obtained from an oscillatory


frequency sweep.

22

CHAPTER 4

METHODOLOGY
4.1

Gummi Formulation and Manufacture


Confections were prepared using the ingredients listed in Table 1 as follows: all

ingredients were mixed by shaking prior to pouring into the kettle (30 qt., jacketed, plus
scraping mixer; Schweppe, Addison, IL 60101). Water and 65 oBrix concord grape juice
concentrate (Welch's Foods Inc., Concord, MA 01742) amounts were equivalent across
all variables. Acid-thinned wheat starch (Gemstar 1090, Manildra Group USA, Shawnee
Mission, KS 66205) and soy protein isolate (Prolisse, Cargill, Inc., Minnetonka, MN
55391) were added to the juice-water solution and then heated to cooking temperature,
which was was set to and maintained at 100 oC. Confections were subsequently cooked
to a final soluble solids content of 70 oBrix (approximately 25 minutes), verified using a
high-solids refractometer (Fisher Scientific, Pittsburgh, PA 15275). Confections were
then set in Clean Set 0736 cornstarch (Cargill Inc., Cedar Rapids, IA 52406) and
cured/stoved for 24 hours at 38 oC in an Isotemp drying oven (200 series, model 215F,
Fisher Scientific, Pittsburgh, PA 15275). The final total solids content was
approximately 78-80 oBrix for all formulations, where 80 oBrix is considered standard for
gummi-type confections (Koh & Mulvaney, 2004). All equipment used in the gummi
batching process is shown in Figure 6. After curing, candies were cooled to room
temperature, gently brushed free of cornstarch, and then heat-sealed in polyethylene bags.
23

Samples were stored at room temperature (~ 20 oC) until analysis at day 1, 10, and 20
(note: concentration study used day 1 samples only). Samples were analyzed as-is,
without modification; outer skins were not removed for any analysis. The densities of all
samples was approximately 1.4 (mass/volume) and the pH values (Orion Research,
Beverly, MA) were between 3.20 and 4.30 (Table 01).

Formulation
Standard
33% SPI
50% SPI
66% SPI
Table 1

68 oB
Concord
Grape
Juice (%)
69.9
69.9
69.9
69.9

H20
(%)
19.1
19.1
19.1
19.1

Acid-thinned
soy protein
wheat starch
isolate
(%)
(%)
pH
11
0
3.2
7.3
3.7
4.1
5.5
5.5
4.2
3.7
7.3
4.3
*Shaded rows only used for storage time study

Formulations for gummi confections used in this investigation.

24

Figure 6

Batching kettle, refractometer, and stoving oven used in the manufacture


of gummi confections.

25

4.2

Thermogravimetric Analysis (TGA)


Thermogravimetric analysis (TGA) was used to analyze water distribution within

the confections. Three separate batches were tested in quadruplicate for each
formulation. Samples, removed centrally from the interior of the confections were
weighed to 10 15 mg and loaded into platinum TGA pans. A high-resolution
thermogravimetric analyzer (Model Q5000, TA Instruments, New Castle, DE) was used
for all TGA experiments. Nitrogen was used as the purge gas and set to flow rates of 10
and 25 ml/min for the balance and sample pans, respectively. Samples were heated from
~30 oC to 220 oC using a high resolution (Hi-Res) ramp of 20 oC/min resolved to 3
o

C/min whenever the instrument detects a change in weight. Samples were analyzed for

both %weight loss and derivative weight loss (%/oC). Derivative weight loss (dTGA)
describes the rate of weight loss (%) as a function of temperature (Fessas & Schiraldi,
2001) with peaks indicating temperature ranges of accelerated weight loss. Assuming
that all weight loss up to 175 oC is attributed to moisture loss (Fessas & Schiraldi, 2001),
the moisture content (MC) was calculated using equation 1.
Eq. 1
M.C. = initial mass (g) final mass (g)
initial mass (g)

4.3

x 100

Texture Profile Analysis (TPA)


Texture profile analyses were conducted using 35 mm compression geometry on

an Instron 5542 Universal Testing Machine operating with Bluehill software. For the
concentration study, samples were analyzed in quadruplicate for each of three batches for
the standard (0% soy), 33%, and 50% soy gummies. The 66% soy protein gummies
26

could not be evaluated because the samples were destroyed by the texture profiling
method. For the storage time study, samples were analyzed in quadruplicate for each of
three batches for the standard (0% soy) and 50% SPI samples. Two-step, 50%
compression at a rate of 2 mm/sec was performed to simulate mastication (Daubert &
Foegeding, 2003). Gumminess, a function of cohesiveness and hardness (Bourne, 2002)
was used as one of the traits of comparison between samples and, according to
Szczesniak (1962), is the energy required to disintegrate a semisolid food, through
mastication, to a point where it can be swallowed.
4.4

Rheological Analysis
All sample analyses were performed using a stress-controlled AR2000ex

rheometer (TA Instruments, New Castle, DE). Samples were cut with a circular die to a
flat cylindrical shape and then compressed to a thickness of 2 mm with the 20 mm plate
geometry. Oscillatory experiments for stress, frequency, and time were performed to
elucidate rheological characteristics. Oscillatory stress sweeps (OSS) were performed
first at with a range of 0.10 to 4,000 Pa at 1 Hz to determine the linear viscoelastic region
(LVR). The critical (yield) stress, the stress at which the materials deformation was no
longer elastic (Liehr, 2000) was obtained. Oscillatory frequency sweeps (OFS) were also
performed to determine the linear response range at a stress chosen from within the linear
region of the OSS. To determine temperature-dependent behavior, samples were
deformed at 1 Hz and constant stress (from the LVR) and heated from room temperature
to 100 oC. Lastly, oscillatory time sweeps and creep tests were performed under linear
conditions to determine time-dependent behavior. Oscillatory time sweeps (OTS) were
used to determine pre-shear requirements for other oscillatory sweeps regarding testing
27

within the LVR. Results (data not shown) did not indicate any reproducible pre-shear
requirements for other rheometric tests.
4.5

UV-Spec Analysis
The Standard and the 50% SPI treatments were analyzed and compared at day 1

for monomeric anthocyanin content using the haze-corrected pH differential


methodology described by Giusti & Wrolstad (2000). Specifically, cyanidin-3-glucoside
was used as the basis for comparison between treatments. Analyses were conducted
using the HPCORE Chemstation software with the Hewlett-Packard 8453 UV-Vis
spectrophotometer.
4.6

Sensory Analysis
An untrained hedonic sensory panel consisting of students and faculty recruited at

the Parker Food Science building was conducted to determine the acceptability of fresh
gummi confections made with and without soy protein isolate (50% SPI level only).
Panelists (n = 40) were asked to score texture and taste on a scale from 1 to 10 (10 = like
extremely, 1 = dislike extremely). A sample questionnaire and the participatory letter can
be found in Appendices B & C, respectively.
4.7

Statistical Analysis
Data were analyzed using one-way ANOVA with Tukey's multiple comparison

method (p = 0.05) in Minitab 15. Data that failed a test for normality were tested with
the Kruskal-Wallis one-way ANOVA test of median equality using SPSS 18.

28

CHAPTER 5

ADDITION OF SOY PROTEIN TO STARCH-BASED CONFECTIONS ALTERS


TEXTURAL PROPERTIES

5.1

Abstract
The effects of increasing soy protein concentration on the physico-chemical

properties of starch confectionery gels were investigated using thermal and rheological
analyses. Texture analysis revealed that soy protein decreases hardness and
cohesiveness, but with a more dramatic effect on the latter, perhaps demonstrating
potential as a texture modifier. Rheological analysis determined that increasing soy
protein concentration progressively decreased the elastic properties of the starch network,
both by decreasing yield stress and the viscoelastic crossover frequency. High levels of
soy protein also created a more homogeneous water population, one which is lost at
lower temperatures than standard starch gummi confections. Sensory analysis revealed
an improvement in flavor and texture acceptability of gummi confections upon addition
of soy protein. Finally, monomeric anthocyanin recovery was greater in gummi
confections prepared with soy protein, perhaps by improving heat dissipation during
processing.

29

5.2

Introduction
Gelled confections and candies, collectively referred to as gummies (Edwards,

2000), are a large, stable, and growing business (Young, 1998); As of 2006, industry
experts have reported as high as 8.4% growth in the gummi candy sector, mostly
attributed to innovative formulations (Anon, 2006). Many of the most common and
recognizable gummi confections are typically made with gelatin as the active agent of
gelation (Lennox 2002). However, gelatin is not the only option for creating gummi-type
candies. Starches may be preferred for being free of animal by-products and the
associated concerns (Lennox 2002). However, starch-based confections are sometimes
described as short an undesirable texture attribute characterized by a firm, brittle gel
(Burg, 1998).
Texture is a primary determinant of quality in gelled confectionery products and,
due to their relatively simple composition, provide an ideal model for mixed gel
investigations (DeMars & Ziegler, 2001). Gumminess and chewiness are texture analysis
descriptors that are particularly applicable to gelled confections (Borwankar, 1992).
Firmer (short) gummies that are more prone to fracture during mastication are often
described as chewy, while softer gummies that dissolve during mastication are described
as gummy (DeMars & Ziegler, 2001). Shortness is a texture defect associated with
decreased sensory acceptability of gelled confections (Burg, 1998), so quantifying the
firmness (shortness) of the gummi confections as it relates to sensory acceptability is
prudent. One of the primary uses of soy protein in foods is as a texture modifier (Lusas
& Riaz, 1995). Depending on the polysaccharide gel system, soy protein has been shown
to have variable results, enhancing (Baeza et al, 2002) or disrupting the gel's structural

30

integrity (Ryan & Brewer, 2005). Addition of soy protein ingredients to confections was
suggested to improve handling properties and stickiness (Rhee, 1994). In a gelled starchsugar confections, a relatively model gel system, soy proteins would perhaps be
beneficial. Therefore, our hypothesis was that soy protein affects the texture of the
starch-based confections on a concentration-dependent basis.
To understand the textural changes that were affected upon addition of soy
protein, the rheological behavior was assessed. Oscillatory stress sweeps (OSS) reveal
the viscoelastic properties and behaviors of foods. Soy protein has been shown to have a
pronounced effect on viscoelasticity in gel and food systems (Ryan & Brewer, 2005;
Vittadini & Vodovotz, 2003). Oscillating frequency sweeps (OFS) were used to
characterize the rheological properties of the confections as a function of time. In these
gelled systems, OFS can reveal long-term stability (Rao, 2007), such as tendencies for
phase separation, retrogradation, syneresis, or sedimentation. If the addition of soy
protein creates a systemic instability, the oscillatory stress and frequency sweeps should
reveal differences between the various concentration treatments.
In addition to the solid components of the gummi confection, water is another
critical component that affects the overall physical properties. Traditional methods of
describing water behavior (water activity, moisture content, and solvent retention
capacity) do not adequately describe the distribution and mobility of water in food
systems (Slade et al., 1991). Thermogravimetric analysis (TGA) can be used to
characterize the changes in physico-chemical properties of a food system as a function of
temperature. TGA can be used to track rate of moisture loss and differentiate different
water populations within a multi-component food (Fessas and Schiraldi, 2000).

31

Therefore, TGA will be used to characterize and compare water populations in the
gummi confection system and how those are affected by addition of soy protein isolate.
The aim of this investigation was to assess the effect of soy protein addition on
the physico-chemical properties of a starch confection system. Specifically, the textural,
thermal, and rheological properties of starch-based grape juice confections will be
evaluated upon incorporation of multiple levels of soy protein. The effect of soy protein
on the anthocyanin stability, which is known to be compromised in high-temperature
food preparation methods (Patras et al., 2010; Edwards, 2000), will also be assessed in
the gummi confections.
5.3

Methodology
Confections were prepared using the ingredients listed in Table 02 as follows: all

ingredients were mixed by shaking prior to pouring into the kettle (30 qt., jacketed, plus
scraping mixer; Schweppe, Addison, IL 60101). Water and 65 oBrix concord grape juice
concentrate (Welch's Foods Inc., Concord, MA 01742) amounts were equivalent across
all variables. Acid-thinned wheat starch (Gemstar 1090, Manildra Group USA, Shawnee
Mission, KS 66205) and soy protein isolate (Prolisse, Cargill, Inc., Minnetonka, MN
55391) were added to the juice-water solution and then heated to cooking temperature,
which was was set to and maintained at 100 oC. Confections were subsequently cooked
to a final soluble solids content of 70 oBrix (approximately 25 minutes), verified using a
high-solids refractometer (Fisher Scientific, Pittsburgh, PA 15275). Confections were
then set in Clean Set 0736 cornstarch (Cargill Inc., Cedar Rapids, IA 52406) and
cured/stoved for 24 hours at 38 oC in an Isotemp drying oven (200 series, model 215F,
Fisher Scientific, Pittsburgh, PA 15275). The final total solids content was
32

approximately 78-80 oBrix for all formulations, where 80 oBrix is considered standard for
gummi-type confections (Koh & Mulvaney, 2004). After curing, candies were cooled to
room temperature, gently brushed free of cornstarch, and then heat-sealed in polyethylene
bags. Outer skins were not removed for any analysis. The densities of all samples was
approximately 1.4 (mass/volume) and the pH values (Orion Research, Beverly, MA)
were between 3.20 and 4.30 (Table 01).

68 oB
Concord
Acid-thinned soy protein
Grape
H20
wheat starch
isolate
Formulation
Juice (%)
(%)
(%)
(%)
pH
69.9
19.1
11
0
3.2
Standard
69.9
19.1
7.3
3.7
4.1
33% SPI
69.9
19.1
5.5
5.5
4.2
50% SPI
69.9
19.1
3.7
7.3
4.3
66% SPI
Table 2
Formulations for gummi confections used in the concentration study.

33

5.3.1. Thermogravimetric Analysis (TGA)


Thermogravimetric analysis (TGA) was used to analyze water distribution within
the confections. Three separate batches were tested in quadruplicate for each
formulation. Samples, removed centrally from the interior of the confections were
weighed to 10 15 mg and loaded into platinum TGA pans. A high-resolution
thermogravimetric analyzer (Model Q5000, TA Instruments, New Castle, DE) was used
for all TGA experiments. Nitrogen was used as the purge gas and set to flow rates of 10
and 25 mL/min for the balance and sample pans, respectively. Samples were heated from
~30 oC to 220 oC using a high resolution (Hi-Res) ramp of 20 oC/min resolved to 3
o

C/min whenever the instrument detects a change in weight. Samples were analyzed for

both %weight loss and derivative weight loss (%/oC). Derivative weight loss (dTGA)
describes the rate of weight loss (%) as a function of temperature (Fessas & Schiraldi,
2001) with peaks indicating temperature ranges of accelerated weight loss. Assuming
that all weight loss up to 175 oC is attributed to moisture loss (Fessas & Schiraldi, 2001),
the moisture content (MC) was calculated using equation 1.
Eq. 1
M.C. = initial mass (g) final mass (g)
initial mass (g)

x 100

5.3.2. Texture Profile Analysis (TPA)

Texture profile analyses were conducted using 35 mm compression geometry on


an Instron 5542 Universal Testing Machine operating with Bluehill software. Samples
were analyzed in quadruplicate for each of three batches for the standard (0% soy), 33%,
and 50% soy gummies. The 66% soy protein gummies could not be evaluated because
34

the samples were destroyed by the texture profiling method. Two-step, 50% compression
at a rate of 2 mm/sec was performed to simulate mastication (Daubert & Foegeding,
2003). Gumminess, a function of cohesiveness and hardness (Bourne, 2002) was used as
one of the traits of comparison between samples and, according to Szczesniak (1962), is
the energy required to disintegrate a semisolid food, through mastication, to a point
where it can be swallowed.
5.3.3. Rheological Analysis
All sample analyses were performed using a stress-controlled AR2000ex
rheometer (TA Instruments, New Castle, DE). Samples were cut with a circular die to a
flat cylindrical shape and then compressed to a thickness of 2 mm with the 20 mm plate
geometry. Oscillatory experiments for stress, frequency, and time were performed to
elucidate rheological characteristics. Oscillatory stress sweeps (OSS) were performed
first at with a range of 0.10 to 4,000 Pa at 1 Hz to determine the linear viscoelastic region
(LVR). The critical (yield) stress, the stress at which the materials deformation was no
longer elastic (Liehr, 2000) was obtained. Oscillatory frequency sweeps (OFS) were also
performed to determine the linear response range at a stress chosen from within the linear
region of the OSS. To determine temperature-dependent behavior, samples were
deformed at 1 Hz and constant stress (from the LVR) and heated from room temperature
to 100 oC. Lastly, oscillatory time sweeps and creep tests were performed under linear
conditions to determine time-dependent behavior. Oscillatory time sweeps (OTS) were
used to determine pre-shear requirements for other oscillatory sweeps regarding testing
within the LVR. Results (data not shown) did not indicate any reproducible pre-shear
requirements for other rheometric tests.
35

5.3.4. UV-Spec Analysis


The Standard and the 50% SPI treatments were analyzed and compared for
monomeric anthocyanin content using the haze-corrected pH differential methodology
described by Giusti & Wrolstad (2000) with a Hewlett-Packard 8453 UV-Vis
spectrophotometer plus HPCORE Chemstation software. Specifically, cyanidin-3glucoside was used as the basis for comparison between treatments.
5.3.5 Sensory Analysis
An untrained hedonic sensory panel consisting of students and faculty recruited at
the Parker Food Science building was conducted to determine the acceptability of gummi
confections made with and without soy protein isolate (50% SPI level only). Panelists (n
= 40) were asked to score texture and taste on a scale from 1 to 10 (10 = like extremely, 1
= dislike extremely).
5.3.6. Statistical Analysis
Data were analyzed using one-way ANOVA with Tukey's multiple comparison
method (p = 0.05) in Minitab 15. Data that failed a test for normality were tested with
the Kruskal-Wallis one-way ANOVA test of median equality using SPSS 18.
5.4

Results & Discussion

5.4.1 Effects on texture and macro-structure


The confection hardness (kgF) of the soy treatments analyzed was significantly
different (p < 0.05) from the standard (Figure 7). Interestingly, this was not the case with
cohesiveness (the attractive forces between like molecules), where the statistically
significant decrease was not observed until the 50% soy protein-for-starch substitution
36

was used (Figure 7, differing letters above error bars designate statistical significance, p <
0.05). The results indicate that while the lower concentration of soy protein has a great
impact on hardness, the internal structure is most compromised at higher soy protein
concentrations. These results suggest that soy protein can modulate shortness of texture
in gelled confections without sacrificing the structural integrity of the system. A similar
progressive decrease in hardness was observed by Limroongreungrat and Huang (2007)
in pasta upon addition of soy protein. The decrease in hardness could have resulted from
direct disruption of the starch network or, as has been shown in bread systems, the
hydrated soy protein may have acted as a macro-molecular plasticizer and thereby
lowering starch-network stiffness (Vittadini & Vodovotz, 2003). It is important to note
that hardness in gelled confections can be misrepresented by case hardening, a surface
phenomenon caused by moisture loss that manifests as a skin(Ziegler et al., 2003), that
is not necessarily representative of the total physical nature of the sample. However, case
hardening only became more noticeable as soy protein increased and did not mask the
overall softening effect.

37

Figure 7

Confection hardness and cohesiveness as a function of SPI


concentration.

38

5.4.2 Effects on gel structural properties


Yield stress is the stress that causes permanent deformation (plasticity) of a
material. In an oscillatory stress sweep, this is the stress at which a sample is no longer
within the linear viscoelastic region (LVR). A typical yield stress is shown in Figure 8.
As shown in Figure 9, the first level of soy protein-for-starch substitution (33%) did not
exhibit a significantly lower yield stress from the Standard (240 vs. 220 pa, p > 0.05).
This may be partly due to the rigid standard's tendency to fracture during the analyses, an
analytical issue also observed in the oscillatory frequency sweeps. Elastic modulus
decreases significantly upon the addition of soy protein and was low for all soy
treatments (Figure 10). Interestingly, the elastic modulus increased significantly (p <
0.05) at the 66% level, but this was thought to be due to a combination of case hardening
and sample drying during analysis time based on the progressively decreasing yield
stresses as SPI substitution rate increased and the fact that the viscous modulus was
greater than the elastic modulus (G>G'); when this occurs, the material is behaving more
like a liquid (Heng et al., 2005). This is in agreement with the results observed with the
Instron texture analyzer, where the effects of soy protein substitution were most evident
at the higher levels, especially considering the 66% SPI treatment was too liquid-like for
comparable analysis with the TPA.

39

Figure 8

Typical yield stress observed in an oscillatory stress sweep.

40

This also seems to indicate that soy protein is compromising the rigid starch
network rather than forming a gel. Gelation of soy protein isolate is largely dependent on
concentration, pH, ionic strength, temperature, and time (Renkema, 2001). This gummi
system was low pH (3.2 for the control and 4.1 to 4.3 for the soy protein treatments), low
ionic strength (no added salts),relatively low soy protein isolate concentration (3.75% to
7.25%), and prepared at high temperature over a short time period during which water is
progressively evaporated. It is suggested that, in this confection system, there is
insufficient time and too many interfering solutes for soy protein gelation to occur.
Therefore, the soy protein isolate used in this investigation acts as a non-gelling
ingredient, meaning there is no interaction between protein units in such a way that a
three-dimensional structure is formed (Hermansson, 1985). The extent of denaturation is
also not known. Prolisse SPI is made in such a way that there should be minimal protein
denaturation; however, DSC analysis of ingredients revealed no endotherms associated
with SPI denaturation (data not shown). Gelation of denatured soy protein under gummi
manufacturing conditions would require a bridging cation, commonly calcium, more
time at high temperatures, and/or a higher concentration of soy protein (Wagner et al.,
1995) or at least a lower concentration of soluble solutes (Renkam, 2001, Gu et al.,
2009). It is also possible that soy protein may not be able to form a continuous gel
because the confectionery system's pH is near the isoelectric point (4.5) of this soy
protein (Hermannson, 1986; Srejic, 2006; Gennadios et al, 1993). Because soy protein's
impact on a carbohydrate gel's physical properties have been shown to be heavily
impacted by system pH (Ipsen, 1995), this effect is most pronounced at the 66%

41

concentration. The pH should not have affected the starch network itself (Russell &
Oliver, 1989).
Soy protein may alternatively impede starch hydration during confection
preparation and has been observed in other food systems (Ribotta et al., 2005; Molina et
al., 1976) due to soy's strong affinity for water (Yao et al., 2006). However, DSC
analysis revealed no endothermic peaks that would represent starch crystal melting (data
not shown). Alternately or in parallel, the soy protein may simply have interfered with
starch gelation. This is supported by the decreasing hardness values observed in Figure
8. Hua et al., (2004) demonstrated that soy protein can indeed inhibit continuous
formation of a hydrocolloid network in a carbohydrate system.

42

Figure 9

Yield stresses (Pa) as a function of soy protein isolate (SPI) substitution.

43

Figure 10

Elastic modulus (G') as a function of soy protein isolate substitution.

44

Oscillating frequency sweeps revealed an apparent increase in system dynamics


as soy protein concentration increased. The viscous and elastic responses of all
treatments displayed varying degrees of frequency dependency (e.g., change in slope as
shown in the insert of Figure 11), suggesting that the control and all treatments are
subject to time-dependent changes, such as retrogradation. The frequency dependency of
the treatments can be differentiated based on the G-G crossover frequency, which is an
indicator of long-term confection system stability. The insert in Figure 11 represents a
typical cross-over frequency analysis.
The standard did not always exhibit a crossover point even at high frequencies, as
observed previously in more pure starch systems (Byars et al., 2003) although it did
exhibit frequency dependency. The observed increase in storage modulus as a function
of increasing frequency is to be expected when testing frequency dependency in the
linear viscoelastic region (Kohyama & Nishinari, 1993). For comparison purposes, only
analyses including a cross-over point were included. Increasing soy protein
concentration results in a decrease in the frequency of the G-G crossover, with the most
pronounced decrease occurring between the standard and 33% and also between the 50%
and 66% soy protein concentrations suggesting a progressive shift from an elastic to
more viscous system. The differences in crossover points observed between the standard
and the 33% soy substitution were significantly different (P < 0.05). However, G-G
crossover points of the next two substitution levels, 33% and 50%, were not (P > 0.05).
Lastly, the 66% substitution was significantly lower (P < 0.05) than all other treatments.
The standard exhibited highly elastic (solid-like) behavior, which was expected due to the
high concentration (Rosalina & Bhattacharya, 2001). The progressive decrease in the

45

crossover frequency may be due to a destabilizing effect of soy protein, suggesting that,
by causing the G-G crossover to occur at lower frequencies, soy protein was
compromising the starch network (responsible for the elastic properties) of the gel. The
greatest changes to the system stability occurred at the 33% and 66%, the latter being the
sample which was so destabilized by the high soy protein substitution that it could not be
analyzed with the texture analyzer.

46

Figure 11

Viscoelastic moduli crossover point (G-G) as a function of soy protein


concentration. Inset graph depicts a typical crossover point.

47

5.4.3 Effect on water properties and distributions


All confections samples exhibited total weight losses of approximately 50%, of
which 25-30% is attributed to sample pyrolysis that occurs as the samples approach 200
o

C (Aseeva et al., 2009). This is based on the known final total solids (consistently 78-

80%) and the observation that samples removed from the TGA oven at 150 oC were
browned, dry, and brittle while samples tested through to 200 oC were ash. This
temperature range of 150 to 200 oC corresponds to the second dTGA peak observed in all
samples. Weight lost at lower temperatures (less than 100 oC), assumed to be water
removed by simple diffusion (Fessas & Schiraldi, 2001) was low, 3 6%, with very high
sample-to-sample variability. However, it is still possible to differentiate the variables
based on the primary water population, in which the rate of weight loss is greatest from
approximately 100 to 150 oC. Figure 12 shows that this population of water is relatively
uniform except for the 66% SPI treatment, in which the peak terminates at a lower
temperature. Zhang (2004) observed that, upon adding soy protein to bread, the
derivative weight loss peak shifted to lower temperatures, indicating that the soy system
had a weaker water association. Prolisse SPI is considered to be highly soluble (Ohr,
2006; Adams, 2007), but it would seem that its capacity to hold water at high
temperatures is low relative to the other components of the gummi system (starch,
monosaccharides). Because it still has some ability to hold water, it must not have been
rendered fully denatured by the high temperature (100 oC) gummi process (Elmore et al.,
2007; Hermannson, 1986). In the confection system, the differences in peak
temperatures are only significantly different (p < 0.05) at the highest level, 66%, of soy
protein substitution. This indicates that, above 50% SPI, the network in the confection

48

began to lose its ability to bind water, rather than an incremental decrease as soy protein
increased.

49

Figure 12

Peak termination temperatures of derivative weight loss curves of gummi candies with
increasing levels of soy substitutions. Inset graph depicts typical dTGA weight loss peak for
66% SPI variable versus Standard control.

50

5.4.4 Sensory Analysis


Significant (p < 0.05) improvements to both flavor and texture were observed
upon incorporation of soy protein isolate at 50% starch substitution (see Table 3). This
supports the texture profile analysis data that shows that soy protein decreases the
hardness of the standard, which would be perceived as shortness by a sensory panel.
Surprisingly, improved flavor was also reported, which could also be a texture-related
effect. Rigid gels been shown to suppress fundamental flavors more than softer gels
(Costell et al, 2000; Lethuaut et al, 2003), therefore, it is reasonable to conclude that the
soy gummies may have stronger perceived flavor due to lower gel firmness.

n = 40
Standard
50% SPI
Table 3

Code
Texture
Flavor
a
115
7.56
7.30a
698
5.05b
6.23b
*Different letters indicate significant differences (p < 0.05)
Mean hedonic scores for Standard and 50% SPI gummi confections.

51

5.4.5 UV-Spec Analysis


As shown in Figure 13, there was a significant increase in MACN survival in the
50% SPI treatment over the Standard. This is unexpected given that anthocyanins can
form insoluble complexes with proteins (Edwards, 2000) that would precipitate during
sample preparation and also the greater degradation risk associated with higher pH (Laleh
et al., 2006). The improved heat dissipation of the 50% SPI, which was evident as a
lower cooking viscosity, may have mitigated heat loss of MACN.

52

Figure 13

Monomeric anthocyanins content, expressed as units of cyanidin-3-glucoside, in the


gummi confections versus raw and boiled grape juice concentrate.

53

5.5

Conclusions
The results of this investigation have demonstrated that addition of soy protein

isolate was found to dramatically alter the physico-chemical properties of gummy-type


confections. Soy protein was shown to impart an overall softening of gummy candies,
most likely due to partitioning the water populations away from, and possibly directly
disrupting, the starch gel network. This softening effect, which manifested as decreased
hardness, yield stress, and elasticity, became more pronounced as soy protein
concentration increased. At higher concentrations, soy protein may completely disrupt
the starch network, and may even replace starch as the continuous phase, creating a weak
gel held together more by hydrophobic interactions and overall solids concentration.
Despite the destabilizing effects that soy protein appeared to impart on the macrostructural properties of gummi confection, it demonstrated potential as a processing aide
that improved sensory properties and increased color stability. Sensory analysis
determined that addition of soy protein improved hedonic liking scores for texture and
flavor.
Increasing soy concentration also caused an apparent increase in the molecular
dynamics of the gummy system. The long-term stability of a gummy formulated with
soy protein isolate is questionable given the increase in viscous behavior as soy protein
concentration increased. Further research to analyze the potential shelf stability of the
soy gummies is currently underway.
Soy protein may not be a viable nutritional fortifier of gummy candies given the
pronounced effects observed at relatively low concentrations in this investigation.
However, the weakening of the gel network due to the inclusion of soy protein could

54

potentially be used as a processing aide to modulate a specific textural defect, shortness,


associated with starch-gelled sugar confections.

55

CHAPTER 6

EFFECTS OF SOY PROTEIN ON THE STORAGE STABILITY OF STARCH-BASED


GELLED CONFECTIONS

6.1

Abstract
The effects of soy protein addition on the storage stability of starch-based gummi

confections were investigated using thermal and rheological analytical techniques. Soy
protein addition improved the textural stability of the gummy confection by maintaining
lower cohesiveness and decreases the confection's resistance to stress. During storage,
soy gummi confections demonstrated altered water dynamics, as evidenced by a lower T g
and lower loss temperatures suggesting the potential of using soy protein isolate as a
functional ingredient for improved textural profile during storage.
6.2

Introduction
Gelled confections and candies, collectively referred to as gummies (Edwards,

2000), are a large, stable, and growing business (Young, 1998); As of 2006, industry
experts have reported as high as 8.4% growth in the gummi candy sector, mostly
attributed to innovative formulations (Anon, 2006). Many of the most common and
recognizable gummi confections are typically made with gelatin as the active agent of
gelation (Lennox 2002). However, gelatin is not the only option for creating gummi-type
candies. Starches may be preferred for being free of animal by-products and the
associated concerns (Lennox 2002). However, starch-based confections are sometimes
described as short an undesirable texture attribute characterized by a firm, brittle gel
56

(Burg, 1998). Furthermore, starch gels have a tendency to retrograde which can manifest
as further increase in gel firmness or shortness (Leon et al., 1997) which may lower
consumer acceptance.
Texture is a primary determinant of quality in gelled confectionery products and,
due to their relatively simple composition, provide an ideal model for structure-function
investigations (DeMars & Ziegler, 2001). Gumminess and chewiness are texture
analysis descriptors that are particularly applicable to gelled confections (Borwankar,
1992). Firmer (short) gummies that are more prone to fracture during mastication are
often described as chewy, and softer gummies that dissolve during mastication are
described as gummy (DeMars & Ziegler, 2001). One of the many uses of soy protein in
foods is as a texture modifier (Lusas & Riaz, 1995). Depending on the polysaccharide
gel system, soy protein has been shown to have variable results, enhancing (Baeza et al,
2002) or disrupting the gel's structural integrity (Ryan & Brewer, 2005). Addition of soy
protein ingredients to confections was suggested to improve handling properties and
stickiness (Rhee, 1994). In gelled starch-sugar confections, a relatively model gel
system, soy proteins would perhaps be beneficial to long-term storage stability. In a
previous study (Siegwein, 2010), increasing levels of soy protein added to gummi
confections were shown to modulate firmness and increase water mobility using
rheological and thermal analytical techniques, respectively. Therefore, our hypothesis
was that soy protein has a lasting effect on the texture of starch-based confections
throughout the defined storage period.
To understand the textural properties that were affected upon addition of soy
protein, the viscoelastic behavior obtained through oscillatory stress sweeps (OSS) can be

57

assessed. Soy protein has been shown to have a pronounced effect on viscoelasticity in
gel and food systems (Ryan & Brewer, 2005; Vittadini & Vodovotz, 2003). Analysis of
long-term stability of gelled systems utilizing oscillating frequency sweeps (OFS)
revealed tendencies for phase separation, retrogradation, syneresis, or sedimentation
(Rao, 2007), If the addition of soy protein creates a systemic instability, the oscillatory
stress and frequency sweeps should reveal differences in the dynamic rheological
properties between the time intervals.
Water is another critical component that affects the overall physical properties of
gummy confections. Traditional methods of describing water behavior (water activity,
moisture content, and solvent retention capacity) do not adequately describe the
distribution and mobility of water in food systems (Slade et al., 1991), particularly in
regards to long-term stability (Sablani et al., 2007) and, therefore, quality. Thermal
analytical techniques, including differential scanning calorimetry (DSC) and
thermogravimetric analysis (TGA) can be used to characterize the changes in physicochemical properties of a food system as a function of temperature. DSC can be used to
quantify 2nd order transitions, such as the glassy-rubbery transition, which is associated
with an increase in molecular motion of the amorphous phase of materials as they change
from the glassy (rigid) to the rubbery (flexible) state (Slade et al., 1991). This state
change is driven by plasticizing water (Roos & Karel, 1991), meaning the water itself
must be characterized.
The aim of this investigation was to determine how soy protein affects storage
stability of gummi confections. Specifically, the textural, thermal, and rheological
properties of starch-based grape juice confections will be evaluated at 1, 10, and 20 days.

58

6.3

Methodology
Confections were prepared using the ingredients listed in Table 02 as follows: all

ingredients were mixed by shaking prior to pouring into the kettle (30 qt., jacketed, plus
scraping mixer; Schweppe, Addison, IL 60101). Water and 65 oBrix concord grape juice
concentrate (Welch's Foods Inc., Concord, MA 01742) amounts were equivalent across
both treatments. Acid-thinned wheat starch (Gemstar 1090, Manildra Group USA,
Shawnee Mission, KS 66205) and soy protein isolate (Prolisse, Cargill, Inc.,
Minnetonka, MN 55391) were added to the juice-water solution and then heated to
cooking temperature, which was was set to and maintained at 100 oC. Confections were
subsequently cooked to a final soluble solids content of 70 oBrix (approximately 25
minutes), verified using a high-solids refractometer (Fisher Scientific, Pittsburgh, PA
15275). Confections were then set in Clean Set 0736 cornstarch (Cargill Inc., Cedar
Rapids, IA 52406) and cured/stoved for 24 hours at 38 oC in an Isotemp drying oven
(200 series, model 215F, Fisher Scientific, Pittsburgh, PA 15275). The final total solids
content was approximately 78-80 oBrix for all formulations, where 80 oBrix is considered
standard for gummi-type confections (Koh & Mulvaney, 2004). After curing, candies
were cooled to room temperature, gently brushed free of cornstarch, and then heat-sealed
in polyethylene bags. Samples were stored at room temperature (~ 20 oC) until analysis
at day 1 (fresh), 10, and 20. Samples were analyzed as-is, without modification; outer
skins were not removed for any analysis. The densities of all samples was approximately
1.4 (mass/volume) and the pH values (Orion Research, Beverly, MA) were between 3.20
and 4.30 (Table 01).

59

68 oB
Concord
Acid-thinned
soy protein
Grape
H20
wheat starch
isolate
Formulation
Juice (%)
(%)
(%)
(%)
pH
69.9
19.1
11
0
3.2
Standard
69.9
19.1
5.5
5.5
4.2
50% SPI
Table 4
Formulations for gummi confections used in storage time study.

60

6.3.1 Differential Scanning Calorimetry (DSC)


Differential scanning calorimetry (DSC) was used assess the phase transitions in
the confections. Three separate batches were analyzed in quadruplicate for both
formulations at each interval (1, 10, and 20 days). Samples, extracted centrally from the
interior of the gummies, were weighed to 5 10 mg and loaded into high-volume
stainless steel pans (part #0319-1525, PerkinElmer , Wellesly, MA) and complementary
lids (part #0319-1526, PerkinElmer, Wellesly, MA) sealed by an O-ring (Lot #2541102,
PerkinElmer, Wellesly, MA). A Q100 DSC equipped with a Refrigerated Cooling
System (RCS) (TA Instruments, New Castle, DE) was used for DSC experiments.
Nitrogen was used as the purging gas and set to flow rate of 50 mL/min. Indium was
used to calibrate the instrument and an empty pan used as a reference. Initially, samples
were equilibrated at -50 oC, held isothermally for 3 minutes, and increased 5 oC per
minute to 200 oC. All observed thermal events were analyzed using Universal Analysis
software (TA Instruments, New Castle, DE).
6.3.2 Thermogravimetric Analysis (TGA)
Thermogravimetric analysis (TGA) was used to analyze water distribution within
the confections. Three separate batches were tested in quadruplicate for all
concentrations and at each time interval. Samples, removed centrally from the interior of
the confections were weighed to 10 15 mg and loaded into platinum TGA pans. A
high-resolution thermogravimetric analyzer (Model Q5000, TA Instruments, New Castle,
DE) was used for all TGA experiments. Nitrogen was used as the purge gas and set to
flow rates of 10 and 25 mL/min for the balance and sample pans, respectively. Samples
were heated from ~30 oC to 220 oC using a high resolution (Hi-Res) ramp of 20 oC/min
61

resolved to 3 oC/min whenever the instrument detected a change in weight. Samples


were analyzed for both %weight loss and derivative weight loss (%/ oC). Derivative
weight loss (dTGA) describes the rate of weight loss (%) as a function of temperature
(Fessas & Schiraldi, 2001) with peaks indicating temperature ranges of accelerated
weight loss. Assuming that all weight loss up to 175 oC was attributed to moisture loss
(Fessas & Schiraldi, 2001), the moisture content (MC) was calculated using equation
Eq. 1
M.C. = initial mass (g) final mass (g)
initial mass (g)

x 100

6.3.3 Texture Profile Analysis (TPA)


Texture profile analyses were conducted using 35 mm compression geometry on
an Instron 5542 Universal Testing Machine operating with Bluehill software. Samples
were analyzed using texture profile analysis (TPA) in quadruplicate for each of three
batches for the standard (0% soy) and 50% SPI samples. Two-step, 50% compression at
a rate of 2 mm/sec was performed to simulate mastication (Daubert & Foegeding, 2003).
Gumminess, a function of cohesiveness and hardness (Bourne, 2002) was used as the trait
of comparison between samples and, according to Szczesniak (1962), is the energy
required to disintegrate a semisolid food, through mastication, to a point where it can be
swallowed.
6.3.4 Rheological Analysis
All sample analyses were performed using a stress-controlled AR2000ex
rheometer (TA Instruments, New Castle, DE). Samples from each treatment and interval
were cut with a circular die to a flat cylindrical shape and then compressed to a thickness

62

of 2 mm with the 20 mm plate geometry. Oscillatory experiments for stress, frequency,


and time were performed to elucidate rheological characteristics. Oscillatory stress
sweeps (OSS) were performed first at with a range of 0.10 to 4,000 Pa at 1 Hz to
determine the linear viscoelastic region (LVR). The critical (yield) stress, the stress at
which the materials deformation was no longer elastic (Liehr, 2000) was obtained.
Oscillatory frequency sweeps (OFS) were also performed to determine the linear
response range at a stress chosen from within the linear region of the OSS as well as to
determine rate-dependent viscoelastic responses. Lastly, oscillatory time sweeps were
performed under linear conditions to determine time-dependent behavior. Oscillatory
time sweeps (OTS) were used to determine pre-shear requirements for other oscillatory
sweeps regarding testing within the LVR. Results (data not shown) did not indicate any
reproducible pre-shear requirements for other rheometric tests.
6.3.5. Statistical Analysis
All data were analyzed using one-way ANOVA with Tukey's multiple
comparison method (p = 0.05) in Minitab 15. Data that failed a test for normality were
tested with the Kruskal-Wallis one-way ANOVA test of median equality using SPSS
18.

63

6.4.

Results & Discussion

6.4.1 Texture Profile Analysis


Gumminess results are presented in Figure 14. Gumminess in the standard
increased significantly and peaked at day 10, whereas in the 50% SPI the gumminess did
not increase significantly until day 20. Both products, therefore, would require more
mastication to dissolve after storage. However, since gumminess is a function of both
hardness and cohesiveness, the factors contributing to these observations were also
analyzed.

64

Figure 14

Changes in gumminess (N) as a function of storage time.

65

Hardness results are presented in Figure 15. There was no change in hardness
over 20 days for the standard. Hardness increased significantly (p < 0.05) in 50% SPI
samples at each storage interval studied. For each interval, all samples were significantly
(p < 0.05) different from each other. During storage, the 50% SPI treatment was
observed to have developed a tough skin, possibly due to case hardening caused by
surface dehydration of the water associated with soy protein. Soy protein has high water
affinity (Vittadini & Vodovotz, 2003), but may be easily removed depending on the food
matrix (Smith, 2003). A previous study (Siegwein, 2010) also showed that soy protein
could cause the gummi system to more readily lose moisture, although the distribution
across the sample was not investigated. Phase separations are common in composite gels
(Burey et al., 2009) and may be evidenced by the increased hardness observed in Figure
TPA2.

66

Figure 15

Changes in hardness (kgF) as a function of storage time.

67

Cohesiveness results are presented in Figure 16. In the Standard gummi


confections, cohesiveness increased significantly at each storage interval in the standard.
However, there was no change in cohesiveness in the 50% SPI samples over 20 days.
These results suggest that while the standard transformed into an increasingly
interconnected network over time, most likely due to the relatively high starch
concentration (Prokopowich & Biliaderis, 1995), the internal network of the 50% SPI
remained relatively static. So while gumminess increased in both treatments, the
mechanism by which they increased was different.

68

Figure 16

Changes in cohesiveness (mm) as a function of storage time.

69

6.4.2 Effects of soy protein on the dynamic rheological properties over time
Significant differences between standard and 50% SPI samples at all storage
intervals were observed with the standard exhibiting a much greater stress resistance,
manifesting as a higher yield stress value (Figure 17). The standard samples increased
significantly in yield stress at each storage interval, while, in the 50% SPI samples, the
yield stress increased significantly only at the 20 day interval. Higher yield stress values
are indicative of a system that is progressively becoming more resistant to permanent
deformation and perhaps more glassy than rubbery (Ollett et al., 1991), and the relative
magnitudes of increasing yield stresses, approximately 1000% vs. 100% for the Standard
and the 50% SPI, respectively, reveal that the 50% SPI is much more resistant to the
formation of a highly elastic (rigid or short) than the Standard.

70

Figure 17

Change in yield stress (Pa) as a function of time for both the Standard and 50%
SPI treatments.

71

The visco-elastic (G'-G) crossover frequency is the frequency at which the


material changes from behaving more like a liquid than a solid (Rao, 2007). Significant
differences between the Standard and the 50% SPI sample at each time interval were
observed for the cross-over frequency (Figure 18). Within each treatment, significant
increase occurred from Day 1 to Day 10 but not from Day 10 to Day 20. Therefore, both
treatments exhibited a trend towards more elasticity over time, which was expected in a
starch-based gel network (Wang & Sun, 2002). However, the magnitude of crossover
frequency increase in the Standard was 71%, and only 38% in the 50% SPI. This
indicated that the soy protein both impeded and then delayed rigid network formation, a
property observed in soy containing bread (Vodovotz & Vittadini, 2003), which also
contained a starch gel component. This, like the stress sweep, indicated that the SPI has a
lasting ability to enhance the plasticizing effect of water (Vodovotz and Vittadini, 2003),
due to being highly hydrophilic, in the gummi confection matrix (Morales & Kokini,
1999). This may be due to disruption of the starch network, or perhaps by the greater
amount of water retained at lower temperatures in the soy-confection, see below, may act
as plasticizer of macromolecules.

72

Figure 18

Change in visco-elastic (G'-G) crossover frequency as a function of


storage time in both the Standard and 50% SPI treatment.

73

6.4.3 Effect of soy on water dynamics over time


Derivative weight loss curves, presented in Figure 19a, for the Standard suggests
that the water population is progressively more entrapped as storage time increases since
it is more difficult to remove. There was a significant (p < 0.05) increase in the peak
onset temperature at each time interval. The peak weight loss temperature increased
significantly (p < 0.05) from D1 to D20. There was no significant change in total weight
loss (moisture content) or low-temperature (<100 oC) weight loss. There is also a
significant (p < 0.05) increase in the peak termination temperature at D10 that is not
present at D20. This is probably explained by the shoulder on the D10 peak and the
small peak that appears at D20 from the ~150 oC to the 175 oC temperature range. This
distinct, higher-temperature peak may represent a phase-separated water population, one
that appears as a slight shoulder at D1 which becomes more distinct at D10 (Figure 19a).
This could be the beginnings of glucose or fructose decomposition, which has been
shown to occur in this temperature region (Hurtta et al., 2004), although no melting
endotherms were observed with DSC analysis (possibly obscured by the fact that the Orings regularly failed above 150 oC). It could also be retrograded starch, although the
temperature is much higher, 160 to 170 oC, than has been reported in the past for water
associated with retrograded starch (Collison & Dickson, 1971), but this gummi
confection system is uniquely resistant to water loss due to the high soluble solids content
and starch retrogradation rate is concentration dependent (Tolstoguzov, 2003).

74

Figure 19

Changes in the derivative weight loss curve of the Standard (19a, no soy) and 50% SPI (19b) gummi
confections.

75

A similar trend of increasing temperature of weight loss curves during storage


was observed for the 50% SPI treatment (Figure 19b). The peak onset, maximum, and
termination points increased significantly (p < 0.05) at D10 and remained stable through
D20. There was no change in total weight loss or low-temperature (<100 oC) weight loss.
The shift in the peak derivative weight loss at D10 indicates that, like the Standard, the
water dynamics within the 50% SPI stabilized by D10. However, the D10 and D20
samples still display a fairly homogeneous water population, one in which all water is
lost around 150 oC with peak maxima that are approximately 10 oC lower. There is a
high temperature (140 oC) shoulder that begins to resolve at D10, similar to what was
observed more dramatically in the Standard. Mir & Nath (1995) showed that, even when
controlling for moisture content, soy protein increases the water dynamics of a high-sugar
food system at a given temperature.
No melting or crystallization transitions were observed during DSC analysis
indicating no detectable freezable water remaining in the gummi system, even at higher
soy concentrations. There was potentially a glass transition (T g) that occurs in each
variable at about 15 to 35 oC (Figure 20) attributed to the Tg of fructose based on the
temperature range and the high concentration of fructose (Shinyashiki et al., 2008;
Kalichevsky & Blanshard, 1993). There were no significant differences in the C p
(J/gC) of the glass transitions, although the median onset temperature did increase
significantly (p < 0.05) in the Standard after 20 days of storage, an increase that was not
observed in the 50% SPI. A higher Tg in the D20 Standard may indicate a reduction in
molecular mobility, which is supported by the observed increases to cohesiveness during
76

storage suggesting a shift of the standard towards a less rubbery (more elastic or solidlike) state at room temperature. The lack of any melting endotherms in the ~60 C range
indicates that there is either no detectable crystalline starch (full gelatinization) typical of
freshly made starch-based gummi confections (Burey et al., 2009) or that there is not
enough plasticizing water to facilitate the reaction (Zeleznak & Hoseney, 1987).

77

Figure 20

Comparison in the Tg onset temperature between the


Standard and 50% SPI gummi confections after 20 days.

78

6.5

Conclusions
The results of this investigation have demonstrated that addition of soy protein

isolate has a long-term ability to alter the physico-chemical properties of gummy-type


confections. The starch network disrupting effects observed in the previous study
(Siegwein, 2010) continue throughout prolonged storage. During storage, soy gummi
confections demonstrated altered water dynamics and decreased firming, although some
surface drying (and therefore case hardening) occurred. Overall, the soy gummi
confections demonstrated lower overall elasticity or shortness suggesting the potential
use of soy protein isolate as a functional ingredient for long-term texture stability of these
confections.

79

CHAPTER 7

CONCLUSION

It was hypothesized that adding soy protein to a gummi confection system would
have a profound impact on its physico-chemical properties, both as a factor of soy protein
concentration and storage time, due to interactions between the main system components
(water, sugars, starches, and proteins). To that end, the following aims were addressed.
Aim 1: To characterize and compare the thermal and rheological properties of
starch-based gummies with multiple levels of soy protein incorporation.
Addition of soy protein isolate was found to dramatically alter the physicochemical properties of gummi-type confections. Soy protein seems to affect an overall
softening of gummy candies, most likely due to partitioning the water populations away
from, and possibly directly disrupting, the starch gel network. This softening effect,
which manifested as decreased hardness, yield stress, and elasticity, became more
pronounced as soy protein concentration increased. At higher concentrations, soy protein
may completely disrupt the starch network, and may even replace starch as the
continuous phase or render the whole system discontinuous (no ordered associations),
creating a weak gel held together more by hydrophobic interactions and overall solids
concentration. The softening of the gel network due to the inclusion of soy protein could
potentially be used as a processing aide to modulate a specific textural defect, shortness,
associated with starch-gelled sugar confections. Increasing soy concentration also caused
80

an apparent increase in the water dynamics of the gummy system, as evidenced by


thermogravimetric analysis results showing decreased moisture loss temperatures at high
soy protein concentrations.
Aim 2: To characterize the changes that occur in gummi candies with and without
soy protein during storage.
During storage, soy gummi confections demonstrated increased water dynamics,
manifesting as a resistance to Tg increase in the DSC thermogram and continued loss of
water at lower temperatures during thermogravimetric analysis. Gumminess increased in
both treatments, but via different mechanisms; the standard by increasing cohesiveness
indicative of a retrograding gel and the 50% SPI by increasing hardness (thought to be an
artifact of case hardening). At each storage interval, the soy gummi confections
displayed lower overall resistance to stress, lower crossover frequency, and lower relative
increases to both over storage. These softening effects were continuances of the effects
observed in the first study; the mechanism could be starch network disruption or dilution
combined with water partitioning. These results suggest the potential for using soy
protein isolate as a functional ingredient for long-term texture stability.
Aim 3: To determine how soy protein addition affects the sensory acceptability and
processing stability of anthocyanins using UV spectroscopy.
Sensory Acceptability
Addition of soy protein was found to significantly improve mean hedonic scores
for texture and taste. Both improvements are thought to be caused by a decrease in

81

shortness, which is in agreement with the texture and rheology results, yielding a
smoother texture and enhanced flavor perception.
UV Spectroscopy
Despite the disruptive effects that soy protein appears to have on the macrostructural properties of gummi confection, it has demonstrated potential as a processing
aide that improved nutritive pigment (color) stability. Monomeric anthocyanin recovery
was significantly improved (30% more), in the soy gummi formulation. The mechanism
is thought to be based on the effect of viscosity on heat capacity. Addition of soy protein
noticeably decreased cooking viscosity, allowing for greater dissipation of heat.

Future Work
Characterizing the effects of soy protein addition on gummi confections has
yielded a preliminary understanding of how soy protein behaves in a high-solids starch
gel and how that changes over time. As a seminal investigation, there is still much
research needed to fully understand the molecular nature of the interactions and
underlying dynamics that occur upon creation of a mixed/composite gel by addition of
soy protein to a high sugar starch network. Specifically, how soy proteins degree of
denaturation and comprising subunits affect physical properties. The possible thermoprotectant action of soy protein on juice anthoycanins during the preparation gummi
confections deserves more attention as a potential delivery system. Lastly, it may be
worthwhile to map the drivers of sensory acceptability for gummi confections to
determine how best to apply texture modifiers like soy protein isolate.
82

APPENDIX A

83

APPENDIX B

84

APPENDIX C

85

REFERENCES

Adams C. 2007. Soy-volution. Nutraceuticals World 73: 64-74.


Ahmed J, Ramaswamy HS, Inteaz A. 2006. Thermorheological characteristics of soybean
protein isolate. J Food Sci 71(3): E158-E163.
Akesowan A, Taweesakulvatchara S. 2009. Freeze-thaw stability of native starches as
modified by Konjac flour and soy protein isolate. J Applied Sciences Res 5(12): 24772481.
Allen UD, et al. 1994. Cows milk versus soy-based formula in mild and moderate
diarrhea: a randomized, controlled trial. Acta Paedritr 83(2): 183-187.
American Soybean Association. 2005. State soy crop statistics. Soy Stats Online.
http://www.soystats.com/2005/page_17.htm
Anderson GH, et al. 2004. Dietary proteins in the regulation of food intake and body
weight in humans. J Nutr 134:974S-979S.
Anderson JW, Patterson K. 2004. Soy protein diet ideal for kidney function.
JADA 104(7): p. A3.
Anhong X, Changyi M, Xuechang L. 2006. On postponement of steamed bread staling by
adding soybean flour defatted under low temperature. Cereal & Feed Industry 6: 20-22.
Anonymous. 2007. National Confectioners Association Offers a Glimpse of the Industry.
Food Prod Des. http://www.foodproductdesign.com/hotnews/79h20115344.html
Anonymous. 2010. U.S. confectionery sales 2009. Manufacturing Confectioner 90(4): 25.
Anonymous. 2006. Broadening gummi lines. Prof Candy Buyer 7(6): 12-18.
Anonymous. 2005. Rheological analysis of dispersions by frequency sweep testing.
Bohlin application note MRK602-01. http://www.azom.com/details.asp?articleid=2884
Antipova AS, Semenova MG. 1995. Effect of sucrose on the thermodynamic
incompatibility of different biopolymers. Carb Pol 28(4): 359-365.
86

Arjmandi BH, Khalil DA, Lucas EA, Juma S, Payton ME, Munson ME, Arquitt AB,
Wild RA. 2001. Soy protein with its isoflavones improves bone markers in women
particularly those with low estrogen status. J Bone & Mineral Research 16: S533-S533.
Arrese EL, et al. 1991. Electrophoretic, solubility, and functional properties of
commercial soy protein isolates. J Ag Food Chem 39: 1029-1032.
Aseeva RM, Sakharov PA, Sakharov AM. 2009. The physicochemical characteristics of
oxidized polysaccharides. Russian J Phys Chem B, Focus on Physics 3(5): 844-850.
Baeza RI, Carp DJ, Prez OE, Pilosof AMR. 2002. -Carrageenan protein interactions:
effect of proteins on polysaccharide gelling and textural properties. LWT 35: 741-747.
Baskin ML, et al. 2005. Prevalence of obesity in the United States. Obesity Reviews 6(1):
5-7.
Bassa IA, et al. 1987. Stabiliy of anthocyanins from sweet potatoes in a model beverage.
J Food Sci 52: 1753-1754.
Borwankar RP. 1992. Food texture and rheology: a tutorial review. J Food Eng 16(1-2):
1-16.
Bourne JC. 2002. Food texture and viscosity: concept and measurement. 2nd Ed.
Academic Press. New York, NY.
Brummer R. 2005. Vibration or oscillation measurements. In: Rheology essentials of
cosmetic and food emulsions. Springer. New York, NY. 44-48 p.
Burey P, Bhandari BR, Rutgers RPG, Halley PJ, Torley PJ. 2009. Confectionary gels: A
review of formulation, rheological, and structural aspects. Int J Food Prop 12: 176-210
Burg JC. 1998. Generating yummy gummies. Design Elements. Food Prod Design
Online. http://www.foodproductdesign.com/archive/1998/0598DE.html
Byars JA, Fanta GF, Felker FC. 2003. The effect of cooling conditions on jet-cooked
normal corn starch dispersions. Carb Polym 54(3): 321-326.
Chang YH, Lim ST, Yoo B. 2004. Dynamic rheology of corn starch-sugar composites. J
Food Eng 64(4): 521-527.
Collison R, Dickson A. 1971. Heats of dehydration of starch gels. Starch 23(6): 203-205.

87

Cooper R, et al. 2000. Trends and Disparities in Coronary Heart Disease, Stroke, and
Other Cardiovascular Diseases in the United States. Circulation 102(25): 3137-47.
Cornillon P, et al. 2000. Characterization of water mobility and distribution in low- and
intermediate-moisture food systems. Magn Reson Imag 18(3): 335-341.
Costell E, Peyroln M, Durn L. 2000. Influence of texture and type of hydrocolloid on
perception of Basic tastes in carrageenan and gellan gels. Food Sci Tech Int 6(6): 495499.
Daubert CR, Foegeding EA. 2003. Rheological principles for food analysis. IN: Nielson,
S. Food Analysis. 3rd Ed. Kluwer Academmic/Plenum Publishers, New York NY. 513
p.
Dean, John A. The Analytical Chemistry Handbook. New York. McGraw Hill, Inc. 1995.
p 15.1-15.5
DeMars LL, Ziegler GR. 2001. Texture and structure of gelatin/pectin-based gummi
confections. Food Hydrocolloids 15(4-6): 643-653.
Dietz WH, et al. 2005. Overweight children and adolescents. Clinical Practice 352(20):
2100-2109.
Drake MA, Chen XQ, Tamarapu S, Leeanon B. 2000. Soy protein fortification affects
sensory, chemical, and microbiological properties of dairy products. J Food Sci 65(7):
1244-1247.
Dubois DK, Hoover WJ. 1981. Soya protein products in cereal grain foods. J Am Oil
Chem Soc 58: 343-346.
Edwards WP. 2000. Gums, gelled products and liquorice. In: The science of sugar
confectionery. Essex, UK. RSC Paperbacks. p. 74-75, 110-123.
Ellis RP, et al. 1998. Starch production and industrial use. J Sci Food Agric 77: 289-311.
Elmore DL, Smith SA, Lendon CA, Muroski AR. 2007. Probing the structural effects of
pasteurization and spray drying on soy protein isolate in the presence of trehalose using
FT-IR-ATR and FT-Raman spectroscopy. Spectroscopy 22(5): 38-44.
Endres J, et al. 2003. Soy-enhanced lunch acceptance by preschoolers. J Am Diet Assoc
103(3): 346-351.

88

Eyre C. 2007. Confectioners face health and indulgence challenges. <


http://www.confectionerynews.com/news/ng.asp?n=81744-global-business-insightspower-mints-medicated-confectionery > Posted and viewed 11/30/2007
Felsner ML, Cano CB, Matos JR, de Almeida-Muradian LB, Bruns RE. 2004.
Optimization of thermogravimetric analysis of ash content in honey. J Brazilian Chem
Soc 15(6): 797-802.
Fessas D, Schiraldi A. 2001. Water properties in wheat flour dough I: Classical
thermogravimetry approach. Food Chemistry. 72(2): 237-244.
Franco JM, Berjano M, Gallegos C. 1997. Linear viscoelasticity of salad dressing
emulsions. J Ag Food Chem 45(3): 713-719.
Gennadios A, Brandenburg AH, Weller CL, Testin RF. 1993. Effect of pH on Properties
of Wheat Gluten and Soy Protein Isolate Films. J Ag Food Chem 41: 1835-1839.
Gigli J, Garnier C, Piazza L. 2009. Rheological behaviour of low-methoxy pectin gels
over an extended frequency window. Food Hydrocolloids 23 (5): 1406-1412.
Giusti MM, Wrolstad RE. 2003. Acylated anthocyanins from edible sources and their
applications in food systems. Biochem Eng J 14: 217-225.
Giusti MM, Wrolstad. 2000. Characterization and measurement of anthocyanins by UVVisible spectroscopy. In: Current procotols in food analytical chemistry; Wrolstad, R. E.,
Ed.; John
Wiley & Sons: New York; pp F1.2.1-F1.2.13.
Gu X, Campbell LJ, Euston SR. 2009. Influence of sugars on the characteristics of
glucono--lactone-induced soy protein isolate gels. Food Hydrocolloids 23(2): 314-326.
Harada K, et al. 2004. Absorption of acylated anthocyanins in rats and humans after
ingesting an extract of Ipomoea batatas purple sweet potato tuber. Biosci Biotech
Biochem 68(7): 1500-1507.
Hartel RW. 1993. Controlling sugar crystallization in food products. Food Tech 47(11):
99-107.
Hedlund TP, Maroni P, Ferucci P, Dayton R, Barnes S, Jones K, Moore R, Wahala K,
Ogden L, Sackett H, Gray K. 2004. Soy consumption, soy metabolism, and the potential
prevention of prostate cancer in Caucasian men. J Nutrition 134(5): 1259S-1260S.
Hendricks K, et al. 2006. Maternal and child characteristics associated with infant and
toddler feeding practices. J Am Diet Assoc. 106: 135S-148S.
89

Heng PWS, Chan LW, Chow KT. 2005. Development of novel nonaqueous
ethylcellulose gel matrices: rheological and mechanical characterization. Pharm Res
22(4): 676-684.
Hermansson AM. 1978. Physico-chemical aspects of soy proteins structure formation. J
Texture Studies 9: 33-58.
Hermansoon AM. 1986. Soy protein gelation. J Am Oil Chem Soc 63(5): 658-666.
Hertog MGL, Hollman PCH, Putte B. 1993. Content of potentially anticarcinogenic
flavonoids of tea infusions, wines, and fruit juices. J Ag Food Chem 41(8): 1242-1246.
Hou HJ, Chang KC. 2002. Interconversions of isoflavones in soybeans as affected by
storage. J Food Sci 67(6): 2083-2089.
Hrehorova E. 2005. Color consistency of spot-color water-based gravure inks for product
gravure. Gravure (find the rest of the information for a citation!)
Hsu S. 1999. Rheological studies on gelling behavior of soy protein isolates. J Food Sci
64(1): 136-140.
Hua YF, Cui SW, Wang Q. 2004. Gel rigidity and the phase behavior of soy proteinhydrocolloid mixtures. J Wuxi Univ of Light Industry 2004-02.
Hurtta M, Pitkanen I, Knuutinen J. 2004. Melting behaviour of D-sucrose, D-glucose and
D-fructose. Carb Res 339(13): 2267-2273.
Iacono G, et al. 1998. Intolerance of cows milk and chronic constipation in children. N
Eng J Medicine 339: 1100-1104.
Ikeda S, Nishinari K. 2001. On solid-like rheological behaviors of globular protein
solutions. Food Hydrocolloids 15(4-6): 401-406.
Ipsen, R. 1995. Mixed gels made from protein and k-carrageenan. Carbohydrate
Polymers 28(4): 337-339.
Kalichevsky MT, Blanshard JMV. 1993. The effect of fructose and water on the glass
transition of amylopectin. Carb Polymers 20(2): 107-113.
Kasapsis S, Al-Marhoobi IM, Deszczynski M, Mitchell JR, Abeysekera R. 2003. Gelatin
vs polysaccharide in a mixture with sugar. Biomacromolecules 4(5): 1142-1149.

90

Kasapsis S, Mitchell J, Abeysekera R, MacNaughtan W. 2004. Rubber-to-glass


transitions in high sugar/biopolymer mixtures. Trends in Food Sci & Tech 15(6): 298304.
Koh AD, Mulvaney SJ. 2004. Small amplitude oscillatory shear rheology of carrageenan,
pectin, and gelatin high-solids systems. Technical Presentation, IFT Annual Meeting.
Kohyama K, Nishinari K. 1993. Rheological studies on the gelation process of soybean
7S and 11S proteins in the presence of glucono--lactone. J Ag Food Chem 41(1): 8-14.
Laleh GH, Frydoonfar H, Heidary R, Jameei R, Zare S. 2006. The effect of light,
temperature, pH, and species on stability of anthocyanin pigments in four Berberis
species. Pakistan J Nutrition 5(1): 90-92.
Lammers G, Stamhuis EJ, Beenackers AACM. 1993. Kinetics of hydroxypropylation of
potato starch in aqueous solution. Ind Eng Chem Res 32: 835-842.
Lang V, et al. 1998. Satiating effect of proteins in healthy subjects: a comparison of egg
albumin, casein, gelatin, soy protein, pea protein, and wheat gluten. Am J Clin Nutr 67:
1197-1204.
Lee JC, Timasheff SN. 1981. The stabilizations of proteins by sucrose. J Biol Chem 256:
7193-7201.
Lee S, Cornillon P, Kim YR. 2002. Spatial investigation of the non-frozen water
distribution in frozen foods using NMR SPRITE. J Food Sci 67(6): 2251-2255.
Lennox S. 2002. Gelatin alternatives in gummi confections. The Manufacturing
Confectioner 85(2): 65-72.
Leon A, Duran E, de Barber CB. 1997. Firming of starch gels and amylopectin
retrogradation as related to dextrin production by -amylase. Zeitschrift fuer
Lebensmittel-Untersuchung und-Forschung A/Food Research and Technology 205(2):
131-134.
Lethuaut L, Brossard C, Rousseau F, Bousseau B, Genot C. 2003. Sweetness-texture
interactions in model dairy desserts: effect of sucrose concentration and the carrageenan
type. Int Dairy J 13(8): 631-641.
Lewen KS, Paeschke T, Reid J, Molitor P, Schmidt SJ. 2003. Analysis of the
Retrogradation of Low Starch Concentration Gels Using Differential Scanning
Calorimetry, Rheology, and Nuclear Magnetic Resonance Spectroscopy. J Ag Food
Chem 51(8): 2348-2358.
91

Li S, Dickinson LC, Chinachoti P. 1998. Mobility of unfreezable and freezable water


in waxy corn starch by 2H and 1H NMR. J Ag Food Chem 46(1): 62-71.
Liehr M. 2000. Yield point measurements with a modern CS rheometer. In: Gupta R
(Ed). Rheology Bulletin 6(1).
Limroongreungrat K, Huang YW. 2007. Pasta products made from sweetpotato fortified
with soy protein. Food Sci & Tech 40(2): 200-206.
Liu L, et al. 2004. Type 2 diabetes in youth. Current Problems Pediatric Adolescent
Health Care 34(7): 254-272.
Loveday SM, Hindmarsh JP, Creamer LK, Singh H. 2009. Physicochemical changes in a
model protein bar during storage. Food Res Int 42(7) 798-806.
Luo YK, et al. 2003. Gel properties of surimi from bighead carp (Aristichthys nobilis):
Influence of setting and soy protein isolate. J Food Sci 69(8): 374-378.
Lupano CE, Gonzlez S. 1999. Gelation of whey protein concentrate-cassava starch in
acidic conditions. J Ag Food Chem 47: 918-923.
Lusas EW, Riaz MN. 1995. Soy protein products: processing and use. J Nutr. 125(3S):
573S-580S.
Mai J. 2004. The right formula: creating great-tasting soy beverages. Functional Foods &
Nutraceuticals 27.
Mark JE. 1984. Physical Properties of Polymers. ACS Publications. 102 p.
Maurice J. 1995. Improving starch moulding. Candy Industry 160(5): 23.
McHugh DJ. 2003. Carrageenan. IN: A guide to the seaweed industry. FAO Fisheries
Technical Papers T441: 118 p.
Messina M, Ho S, Alekel DL. 2004. Skeletal benefits of soy isoflavones: a review of the
clinical trial and epidemiologic data. Current Opinion in Clinical Nutrition & Metabolic
Care 7(6): 649-658.
Mir MA, Nath N. 1995. Sorption isotherms of fortified mango bars. J Food Eng 25(1):
141-150.
Miyazawa T, Nakagawa K, Kudo M, Muraishi K, Someya K. 1999. Direct intestinal
absorption of red fruit anthoycanins, cyanidin-3-glucoside and cyanidin-3,5-diglucoside,
into rats and humans. J Ag Food Chem 47(3): 1083-1091.
92

Molina MR, Mayorga I, Bressani R. 1976. Production of high-proteinquality pasta


products using a semolina-corn-soy flour mixture. II. Some physicochemical properties
of the untreated and heat-treated corn flour and of the mixtures studied. Cereal Chem
53(1): 134-140.
Morales A, Kokini JL. 1999. State diagrams of soy globulins. J Rheology 43(2): 315325.
Morris ER. 1989. Polysaccharide solution properties: origin, rheological characterization,
and implications for food systems. In: Bemiller J, Karan RC. Frontiers in carbohydrate
research. I. Food applications. Elsevier Applied Science, New York, NY. 132-163 p.
N. Chapman Associates. 2004. Current knowledge on soy and childrens diets. United
Soy Board. 1-20 p.
Narayan KM, et al. 2003. Lifetime risk for diabetes mellitus in the United States. JAMA
290(14): 1884-1890.
Nelson B. 2008. Food enhancements prove palatable. Frontiers MSNBC.com.
<<http://www.msnbc.msn.com/id/24281450/print/1/displaymode/1098/>>
OByrne DJ, et al. 2002. Comparison of the antioxidant effects of Concord grape juice
flavonoids -tocopherol on markers of oxidative stress in healthy adults. Am J Clin Nutr
76(6): 1367-1364.
Ohr LM. 2006. Proteins power up. Food Tech 60(2): 55-56.
Ollett AL, Parker R, Smith AC. 1991. Deformation and fracture behaviour of wheat
starch plasticized with glucose and water. J Materials Sci 26(5): 1351-1356.
Pai VB, Khan SA. 2002. Gelation and rheology of xanthan/enzyme-modified guar
blends. Carbohydrate Polymers 49(2): 207-216.
Patras A, Brunton N, O'Donnell C, Tiwari BK. 2010. Effect of thermal processing on
anthocyanin stability in foods; mechanisms and kinetics of degradation. Trends in Food
Sci & Tech 21(1): 3-11.
Poppe J. 1995. New approaches to gelling agents in confectionery. In: Editoral Staff of
The Manufacturing Confectioner (Eds.), Proceedings of the 49th Annual Production
Conference, Hershey, PA. PMCA, Center Valley, PA. 68-78 p.
Prior RL, et al. 2000. Flavonoids: Diet and health relationships. Nutr Clin Care 3: 279288.
93

Prokopowich D, Biliaderis CG. 1995. A comparative study of the effects of sugars on the
thermal and mechanical properties of concentrated waxy maize, wheat, potato and pea
starch gels. Food Chem 52(3): 255-262.
Pourfarzad A, Khodaparast MHH, Karimi M, Mortazavi SA, Davoodi MG, Sourki AH,
Jahromi SHR. 2009. Effect of polyols on the shelf life and quality of flat bread fortified
with soy flour. J Food Proc Eng EARLY VIEW NOT YET IN PRINT
Prasad K. Protein fortification of mango and banana bar using roasted Bengal gram flour
and skim milk powder. Ag Eng Int 11: Manuscript 1390.
Pszczola DE. 2005. From soup to soynuts: the broadening uses of soy. Food Tech
59(2): 44-55.
Pungor, Erno. A Practical Guide to Instrumental Analysis. Boca Raton, Florida. 1995.
p181-191
Puppo MC, An MC. 1998. Structural Properties of Heat-Induced Soy Protein Gels As
Affected by Ionic Strength and pH. J Ag Food Chem 46(9): 3583-3589.
Quinn GP, Keough MJ. 2002. Specific comparisons of means. In: Experimental design
and data analysis for biologists. New York, NY. Cambridge University Press. 201 p.
Rao MA. 2007. Rheology of fluid and semisolid foods. 2nd ed. Springer
Science+Business Media, LLC. New York, NY. 108-111 p.
Reid M, et al. 1997. Relative effects of carbohydrates and protein on satiety a review of
methodology. Neurosci & Biobehavioral Rev 21(3): 295-308.
Renkama, JMS. 2001. Formation, structure and rheological properties of soy protein gels.
Ph.D. Thesis, Vageningen University, The Netherlands.
Rhee KC. 1994. Functionality of soy proteins. In: Hettiarachchy NS, Ziegler GR. Protein
functionality in food systems. Marcel Dekker, New York, NY. 319 p.
Ribotta PD, Arnulphi SA, Len AE, An MC. 2005. Effects of soybean addition on the
rheological properties and breadmaking quality of wheat flour. J Sci Food Agric 85:
1889-1896.
Roccini AP. 2002. Childhood obesity and a diabetes epidemic. NEJM 346(11): 854-855.

94

Rogers MA, et al. 2006. Structural heterogeneity and its effect on the glass transition in
sucrose solutions containing protein and polysaccharide. Food Hydrocolloids 20: 774779.
Roos Y, et al. 1991. Plasticizing effect of water on thermal behavior and crystallization of
amorphous food models. J Food Sci 56(1): 38-43.
Roos YH, et al. 1996. Glass transitions in low moisture and frozen foods. Food Tech
50(11): 95-108.
Rosaline I, Bhattacharya M. 2001. Dynamic rheological measurements and analysis of
starch gels. Carbohydrate Polymers 48(2): 191-202.
Ross-Murphy SB. 1994. Physical techniques for the study of food biopolymers. Blackie
Academic & Professional, Glasgow, UK. 362-382 p.
Rotter G, et al. 1992. Dynamic mechanical analysis of the glass transition. Curve
resolving applied to polymers. Macromolecules. 25: 2170-2176.
Russell PL, Oliver G. 1989. The effect of pH and NaCl content on starch gel aging. A
study by differential scanning calorimetry and rheology. J Cereal Sci 10(2): 123-138.
Ryan KJ, Homco-Ryan CL, Jenson J, Robbins KL, Prestat C, Brewer MS. 2002. Lipid
extraction process on texturized soy flour and wheat gluten protein-protein interactions in
a dough matrix. Cereal Chem 79: 434-438.
Ryan KJ, Brewer MS. 2005. Purification and identification of interacting components in
a wheat starch-soy protein system. Food Chem 89: 109-124.
Rzungles A, Bayonove CL, Cordonnier RE, Sapis JC. 1988. Grape Carotenoids: Changes
During the Maturation Period and Localization in Mature Berries. Am J Enol Vitic 39(1):
44-48.
Sabanis D, Tzia C. 2009. Effect of rice, corn and soy flour addition on characteristics of
bread produced from different wheat cultivars. Food Bioprocess Tech 2(1): 68-79.
Sablani SS, Kasapsis S, Rahman MS. 2007. Evaluating water activity and glass transition
concepts for food stability. J Food Eng 78(1): 266-271.
Sans T, Fernandez MA, Salvador A, Muoz J, Fiszman SM. 2005. Thermogelation
properties of methylcellulose (MC) and their effect on a batter formula. Food
Hydrocolloids 9(1): 141-147.

95

Saric SP, Schofield RK. 1946. The dissociation constants of the carboxyl and hydroxyl
groups in some insoluble and sol-forming polysaccharides. Proc Royal Soc. 185(1003):
431-447.
Schenz TW. 2003. Thermal analysis. In: Nielson S. Food Analysis, 3rd ed. KA/PP, New
York. 517-528.
Sciarini, LR, Ribotta PD, Leon AE, Perez GT. 2008. Influence of gluten-free flours and
their mixtures on batter properties and bread quality. Food Bioprocess Tech.
doi:10.1007/s11947-008-0098-2
Schiraldi A, Fessas D. 2003. Classical and Knudsen thermogravimetry to check states
and displacements of water in food systems. J Thermal Analysis & Calorimetry 71: 225235.
Schiraldi A, Piazza L, Riva M. 1996. Bread staling: a calorimetric approach. Cereal
Chem 73: 32-39.
Scuderi M. 2002. Characterising the microstructure and rheology of starch-gelatine-sugar
confectionery jellies. Individual Inquiry, CHEE4006, University of Queensland.
Serra-Majem L, et al. 2002. Determinants of nutrient intake among children and
adolescents: Results from the enKid Study. Ann Nutr Metab 46(1): 31S-39S.
Serventi L, Carini E, Curti E, Vittadini E. 2009. Effect of formulation on
physicochemical properties and water status of nutritionally enhanced tortillas. J Sci
Food & Ag 89(1): 73-79.
Sethi V, Vsudeva KR, Sethi S, Singh G. 2007. Diversified use of guava to combat
malnutrition. Acta Hort 735: 609-619.
Shinyashiki N, Shinohara M, Iwata Y, Goto T, Oyama M, Suzuki S, Yamamoto W,
Yagihara S, Inoue T, Oyaizu S, Yamamoto S, Ngai KL, Capaccioli S. 2008. The glass
transition and dielectric secondary relaxation of fructose-water mixtures. J Phys Chem B
112 (48): 1547015477.
Shu XO, et al. 2001. Soyfood intake during adolescence and subsequent risk of breast
cancer among Chinese women. Cancer Epidemiology-Biomarkers & Prevention 10: 483488.
Siegwein AM. 2010. The effects of soy protein isolate addition on the physico-chemical
properties of gummi confections. Unpublished manuscript.

96

Singh KJ, Roos YH. 2005. Frozen state transitions of sucrose-protein-cornstarch


mixtures. J Food Sci 70(3): E198-E204.
Singh J, Kaur L, McCarthy OJ. 2007. Factors influencing the physico-chemical,
morphological, thermal and rheological properties of some chemically modified starches
for food applications a review. Food Hydrocolloids 21(1): 1-22.
Skoog, Douglas A., F. James Holler and Timothy Nieman. Principles of Instrumental
Analysis. Fith Edition. New York. 1998. p905-908.
Slade L, et al. 1991. Beyond water activity: recent advances based on an alternative
approach to the assessment of food quality and safety. Crit Rev Food Sci Nutr 30(23):115-360.
Sloan E. 2006. Top Ten Functional Food Trends, Food Technology 60(4):23-39.
Sloan AE. 2002. Candy is Dandy. Functional Foods & Nutraceuticals. February.
<http://www.functionalingredientsmag.com/fimag/articleDisplay.asp?strArticleId=35&st
rSite=FFNSite>
Srejic E. 2006. The possibilities of protein. Natural Products Insider.
http://www.naturalproductsinsider.com/articles/2006/05/the-possibilities-of-protein.aspx
Starling S. 2005. Finessing soys flavours. Functional Foods & Nutraceuticals 44.
Steiner A, Coronel P, Simmons V. 2001. A food quality application using dynamic shear
rheometry. American Laboratory News Ed. 33(23): 12-14.
Sukha PR, et al. 2003. Rheological and microstructural changes of starch/sugar/gelatine
gels as a function of processing. 3rd Intl Symp Food Rheo & Structure. Conference
Proceedings: 595-596.
Szczesniak AS. 1962. Classification of textural characteristics. J Food Sci 28(4): 385389.
Szczesniak AS. 1986. Correlating sensory with instrumental texture measurements-an
overview of recent developments. J Texture Studies 18(1):1-15.
Takashi A, Yamada T. 1998. Crystalline copolymer approach for melting behaviour of
starch. Starch 50(9): 386-389.
Tamborra P, et al. 2003. Phenolic compounds in red-berry skins of Uva di Troia and
bombino nero grapes (Vitis vinifera L.). Ital J Food Sci 15: 347-357.
97

Tiziani S, Vodovotz Y. 2005. Rheological effects of soy protein addition to tomato juice.
Food Hydrocolloids 19(1): 45-52.
Tolstoguzov V. 2003. Thermodynamic considerations of starch functionality in foods.
Carb Polymers 51(1): 99-111
Tsai PJ, et al. 2004. Effect of sugar on anthocyanin degradation and water mobility in a
Roselle anthocyanin model system using 17O NMR. J Agric Food Chem 52(10): 30973099.
Tsai SJ, et al. 1998. Thermal properties of restructured beef products at different
isothermal temperatures. J Food Sci 63 (3), 481-484.
Turgeon SL, Beaulieu M. 2001. Improvement and modification of whey protein gel
texture using polysaccharides. Food Hydrocolloids 15(4-6): 583-591.
US Dept Commerce. 2004. Nonchocolate confectionery manufacturing. 2002 Economic
Census, Manufacturing.
US Dept of Health & Hum. Ser.. 1988. The surgeon generals report on nutrition and
health.
US Census Bureau. 2005. Confectionery. Current Industrial Reports.
Vittadini E, Vodovotz Y. 2003. Changes in the physicochemical properties of wheat- and
soy-containing breads during storage as studied by thermal analyses. J Food Sci 68(6):
2022-2027.
Wagner JR, Sorgentini DA, An MC. 1992. Effect of physical and chemical factors on
rheological behavior of commercial soy protein isolates: protein concentration, water
imbibing capacity, salt addition, and thermal treatment.. J Ag Food Chem 40(10): 19301937.
Walstra P. 2003. Glass transitions and freezing. In: Physical chemistry of foods. Marcel
Dekker, Inc. New York. 650-681.
Wang FC, Sun XS. 2002. Frequency dependence of viscoelastic properties of bread
crumb and relation to bread staling. Cereal Chem 79(1): 108-114.
Warnecke M. 1991. Gums and jelly products and formulations. PMCA Prod Conf 45:
140-162.
Weinen WJ, Katz RF. 1991. Factors affecting gel strength of gum candies. Discussion
Period, 45th PMCA Production Conference. 146-153.
98

Widhalm K, et al. 1993. Effect of soy protein diet versus standard low fat, low
cholesterol diet on lipid and lipoprotein levels in children with familial or polygenic
Hypercholesterolemia. J Pediatr 123(1): 30-34.
Wilcox S, et al. 1999. Knowledge and perceived risk of major diseases in middle-aged
and older women. Health Psych 18(4): 346-353.
Wu X, Beecher GR, Holden JM, Hayowitz DB, Gebhardt SE, Prior RL. 2006.
Concentrations of anthocyanins in common foods in the United States and estimation of
normal consumption. J Agric Food Chem 54(11): 4069-4075.
Wunderlich B. 1994. The nature of the glass transition and its determination by thermal
analysis. IN: Seyler RJ, Ed. Assignment of the glass transition. ASTM STP 1249. ASTM,
Philadelphia, PA. 17-31 p.
Yao JJ, Wei LS, Steinberg MP. 2006. Water-imbibing capacity and rheological properties
of isolated soy proteins. J Food Sci 53(2): 464-467.
Young J. 1998. Gummies come of age. Prof Candy Buyer. Special Online Report.
http://www.retailmerchandising.net/candy/reports/gummie.asp
Zaleska H, Ring SG, Tomasik P. 2000. Apple pectin complexes with whey protein
isolate. Food Hydrocolloids 14: 277-382.
Zallie J. 1988. Benefits of quick setting starches. Manufacturing Confectioner 66:41-43.
Zametkin AJ, et al. 2004. Psychiatric aspects of child and adolescent obesity: A review of
the past 10 years. J. Amer.Ac. Child & Ad.Psyc. 43(2): 134-150.
Zeleznak KJ, Hoseney RC. 1987. The glass transition in starch. Cereal Chem 64: 121124.
Zhang J, Mungara P, Jane J. 2001. Mechanical and thermal properties of extruded soy
protein sheets. Polymer 42: 2569-22578.
Zhang, YC. 2004. Physicochemical properties and isoflavone content of bread made with
soy. Ph.D. Dissertation, The Ohio State University.
Ziegler GR, MacMillan B, Balcom BJ. 2003. Moisture migration in starch molding
operations as observed by magnetic resonance imaging. Food Res Int 36(4): 331-34.

99

S-ar putea să vă placă și