Download as pdf or txt
Download as pdf or txt
You are on page 1of 37

This article was downloaded by: [University of Central Florida]

On: 17 October 2014, At: 09:01


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

International Journal of Food Properties


Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/ljfp20

Confectionery Gels: A Review on


Formulation, Rheological and Structural
Aspects
a b c d
P. Burey , B.R. Bhandari , R.P.G. Rutgers , P.J. Halley & P.J.
d
Torley
a
Centre for Nutrition and Food Sciences , The University of
Queensland , Brisbane , Australia
b
School of Land, Crop and Food Sciences , The University of
Queensland , Brisbane , Australia
c
Plantic Technologies Ltd , Victoria , Australia
d
Centre for High Performance Polymers , The University of
Queensland , Brisbane , Australia
Published online: 13 Jun 2012.

To cite this article: P. Burey , B.R. Bhandari , R.P.G. Rutgers , P.J. Halley & P.J. Torley (2009)
Confectionery Gels: A Review on Formulation, Rheological and Structural Aspects, International
Journal of Food Properties, 12:1, 176-210, DOI: 10.1080/10942910802223404

To link to this article: http://dx.doi.org/10.1080/10942910802223404

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the
“Content”) contained in the publications on our platform. However, Taylor & Francis,
our agents, and our licensors make no representations or warranties whatsoever as to
the accuracy, completeness, or suitability for any purpose of the Content. Any opinions
and views expressed in this publication are the opinions and views of the authors,
and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content
should not be relied upon and should be independently verified with primary sources
of information. Taylor and Francis shall not be liable for any losses, actions, claims,
proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or
howsoever caused arising directly or indirectly in connection with, in relation to or arising
out of the use of the Content.

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &
Conditions of access and use can be found at http://www.tandfonline.com/page/terms-
and-conditions
Downloaded by [University of Central Florida] at 09:01 17 October 2014
International Journal of Food Properties, 12: 176–210, 2009
Copyright © Taylor & Francis Group, LLC
ISSN: 1094-2912 print / 1532-2386 online
DOI: 10.1080/10942910802223404

CONFECTIONERY GELS: A REVIEW ON FORMULATION,


RHEOLOGICAL AND STRUCTURAL ASPECTS

P. Burey1, B.R. Bhandari2, R.P.G. Rutgers3, P.J. Halley4,


and P.J. Torley4
1
Centre for Nutrition and Food Sciences, The University of Queensland,
Downloaded by [University of Central Florida] at 09:01 17 October 2014

Brisbane, Australia
2
School of Land, Crop and Food Sciences, The University of Queensland,
Brisbane, Australia
3
Plantic Technologies Ltd. Victoria, Australia
4
Centre for High Performance Polymers, The University of Queensland,
Brisbane, Australia

A confectionery gel (CG) consists of high sugar components of sucrose and glucose syrup,
combined with gelling components such as starch, gelatin, or pectin, along with food acid,
flavourings and colourings. Common CG products include “jelly snakes,” “jelly babies,”
“jelly beans,” and form a portion of the lucrative confectionery market; however, there are
continual consumer demands for more interesting and innovative products that have new
and exciting textures, flavors and appearances. Improving or modifying CG textures can
meet these demands, but first an understanding of how the behaviour and structure of the
gel is developed must be achieved. Companies that will gain a competitive advantage in the
confectionery market will be those able to actively manipulate and control sensory properties
to meet exacting customer demands. This paper is a review of literature available on con-
fectionery gels, their components, and factors that may affect their microstructure, texture,
and rheology.

Keywords: Confectionery gels, Confectionery manufacture, Confectionery texture,


Confectionery rheology.

INTRODUCTION
Gels exist in many material systems including diverse media such as polymers, plant
and animal tissues, and food. The majority of foods are, or consist of gels[1] and gels found
in foods provide easily recognized, although widely varying textures.[2] National tastes
can often dictate which are the gelling ingredients of choice to provide pleasing textures to
consumer markets.[3] There are several definitions of a gel within literature. Hermans[4]
stated that a gel must be: (1) a coherent two component system formed by a solid sub-
stance finely dispersed, or dissolved in, a liquid phase that also (2) exhibits solid-like
behaviour under mechanical forces. Another view put forward is that all gels are solid, as

Received 26 April 2007; accepted 23 May 2008.


Address correspondence to P. Burey, Centre for Nutrition and Food Sciences, The University of Queensland,
St. Lucia, Brisbane QLD 4072, Australia. E-mail: polly.burey@ihug.com.au

176
CONFECTIONERY GELS FORMULATION 177

they are self-supporting and can recover elastically after deformation, however some gels
may have some unrecovered stresses that can be present after deformation, and gels may
deform in a brittle manner, such as in agar or kappa-carrageenan gels.[5] Flory[6] suggested
that gels consist of polymeric molecules crosslinked to form a tangled interconnected net-
work immersed in a liquid medium. Djabourov and LeBlond[7] mentioned that in a gel
both the dispersed component and the solvent should extend continuously throughout the
whole system, each phase being interconnected, which is similar to the theory proposed by
Flory.[6] A gel has also been described as a giant polymer made up of molecules that are
branched in three dimensions and form a lattice.[8]
Obviously there are several types of gels and they are commonly classified into two
main groups – chemical and physical gels, which are distinguished by the differences in
nature of the bonds linking the gels’ molecules together.[8] Chemical gels usually describe
Downloaded by [University of Central Florida] at 09:01 17 October 2014

covalently crosslinked gels, whilst physical gels consist of chains which are “physically”
(non-covalently) crosslinked into networks.[9,10]
The crosslinks in physical gels are of small but finite energy and/or finite lifetime,
and can be found in biological and synthetic polymers. In biopolymer gels the crosslinks
are formed by physical gel mechanisms such as Coulombic, dipole-dipole, van der
Waal’s, charge transfer, hydrophobic, and hydrogen bonding interactions. The term phys-
ical gel is often assumed to imply thermoreversibility in the gel, such as in the case of gel-
atin, although this is not the case in every physical gel system (e.g. casein gels).[9]

COMPOSITE GELS
Most food gel products in the market today are in fact composite gels, containing two
or more gelling components. Most components contribute to structure and physical proper-
ties of food, with the two main structural materials being proteins and polysaccharides.[11]
Biopolymer mixtures are used in many industries, including food, to impart specific flow
behaviours, textures, appearances, and where required, tactile and mouthfeel properties, to
products.[12,13] As a consequence of the wide use of biopolymer mixtures in food, these
types of systems have been greatly researched, but it is only recently that biopolymer
phase separation in composite gels and the effects of this on composite gel properties have
been studied.[12–15] Most food components have limited miscibility on a molecular level
and tend to form multicomponent heterophase and non-equilibrium dispersed systems.[11]
For polymers dissimilar in shape or structure, segregation leads to a reduction of the
polymer concentration near the other type of polymer particle. Once a critical polymer
concentration is exceeded, this leads to phase separation.[12,13] Interactions of proteins and
polysaccharides range from complete segregation to complexation (Fig. 1). These interac-
tions can greatly affect mechanical gel properties whilst chain structure, molecular weight
and composition are important for structure.[16] Often, linear polysaccharides are more
incompatible with proteins than branched polysaccharides[17] due to rigidity and the pres-
ence of less independently moving macromolecular segments and lower mixing entropy
than in flexible branched polymers.[11] The differences in hydrophobicity in proteins and
polysaccharides are of great importance for phase equilibrium in protein-polysaccharide-water
systems.[17] The level of phase separation, and shape and appearance of phase compo-
nents, in these systems depends on polymer concentration and the compatibility of the
biopolymers.[12,13]
Phase separation in composite gels is a result of thermodynamic incompatibility of
gel components[16,17]. Phase separation is entropically unfavourable, but enthalpically
178 BUREY ET AL.
Downloaded by [University of Central Florida] at 09:01 17 October 2014

Figure 1 Main trends in the behaviour of protein/polysaccharide mixtures (modified from de Kruif and
Tuinier[13]).

advantageous as molecules prefer like molecules.[12,17] Temperature changes can also


affect biopolymer phase separation, as well as pH and shear.[12,16,18–20] Interfacial tension
between gel phases is important in determining biopolymer composite behaviour, whilst
thermodynamic incompatibility and complex formation by food hydrocolloids can greatly
affect mechanical and other physicochemical gel properties.[12,16] It has been possible to
apply synthetic polymer blending laws to food gel composites, in order to predict and
quantify mechanical behaviour of such composites.[21] This theory is described by the
Isostrain and Isostress models, where the Isostrain model sets an upper bound for com-
posite modulus, whilst the Isostress model sets a lower bound. The equations for each are
shown below where equation 1 is applicable for the Isostrain condition and equation 2
applies to the Isostress situation:

Gc′ = qX GX ′ + qY GY ′ , and (1)

qX qY
Gc′ = + , (2)
GX ′ GY ′

where G′C is the elastic modulus of the composite; θX is the fraction present component X;
θY is the fraction present of component Y; G′X is the elastic modulus of component X and
CONFECTIONERY GELS FORMULATION 179

G′Y is the elastic modulus of component Y. For systems in which one component forms a
continuous matrix with the other dispersed through it as a discontinuous filler, the modu-
lus of the composite is expected to lie close to the upper bound if the matrix is more rigid
(higher modulus) than the filler and close to the lower bound if the filler is more rigid than
the matrix.[22]

CONFECTIONERY GELS
Confectionery gels (CGs) are often high sugar systems, with one or more gelling
components, which are chosen for their textural attributes to give firmer or softer textures
to the CGs. The ingredients are formed into a molten material that can be moulded into
many different shapes (Fig. 2). The aforementioned definitions of gels and composite gels
Downloaded by [University of Central Florida] at 09:01 17 October 2014

refer more to the conventional ideas of a biopolymer gel system made in an aqueous envi-
ronment. Recent studies show that on the addition of sugars to such a system, the mor-
phology of the high solid network is distinctly different from that of the aqueous system.
This means that the standard concepts of gelation, which apply in aqueous systems are no
longer valid in CG systems.[23] However, the aqueous gel system model could be valid if a
non-nutritive sweetener is used in a sugar-free CG formulation. This type of product could
have a viable market as obesity in children is an increasing problem worldwide.
In the gelling of aqueous biopolymer systems, network formation is governed by
some important characteristics of the system which are: 1. the minimum critical gelling
concentration (C0), which is the minimum concentration of biopolymer that allows gel net-
works to form; and 2. the concentration of coil overlap and entanglement of biopolymer
chains (C*), which affects density of the network. These characteristics are no longer the
main influence on biopolymer gel network formation once sugars are introduced into the
system, as is the case in confectionery gels. Sugars change morphology of single biopoly-
mer gel systems, as well as affecting phase separation characteristics in mixed biopolymer
gel systems. The major textural properties of CGs are imparted by the gelling components
used, such as starch or gelatin. The sugar co-solutes are not part of the polymer network in
the confectionery gels but can greatly contribute to formation and behaviour of CGs.[24]

Figure 2 Examples of jelly products a) Jelly snakes, b) Jelly babies, c) Jelly frogs.
180 BUREY ET AL.

Studies of single biopolymer systems containing sugar have shown that in polysac-
charides, such as starch that as sugar concentration approaches levels found in confectionery
gels, chain-chain association is reduced. This leads to a less aggregated structure than in
the case of the aqueous system, meaning that C0 must be increased as the sugar is inhibiting
gel network aggregation. Gelatin gels behave quite differently to starch gel systems, with
chain association being increased in the presence of sugar, rather than reduced. Sugars
tend to destabilize polysaccharide gel networks at levels of 40–60% sugars, but increase
gelatin gel networks at these levels of sugar.[25]
In aqueous systems, gelatin gels appear featureless, but upon addition of sugar they
separate into sugar-rich and gelatin-rich phases, showing high biopolymer aggregation. In
aqueous systems, polysaccharides show fibrillar network structures, but upon addition of
sugar some of these features are lost. This is due to the threshold of thermodynamic stability
Downloaded by [University of Central Florida] at 09:01 17 October 2014

of structure formation for polysaccharides being exceeded at these levels, and consider-
able parts of the network “dissolving” in the saturated sugar environment. Thermody-
namic stability of gelatin gel network formation is increased at 40–60% sugar and
continues to increase at levels above this.[23]
In mixed biopolymer systems phase separation is governed by critical polymer con-
centration, polymer shape, and compatibility of the biopolymers present. The extent of
mixing is governed by the “Flory-Huggins interaction parameter,” which is dependent on
solvent quality and volume as well as polymer concentration and size.[26,27] Altering the
quality of the solvent (e.g., adding sugar to water), results in changes to the extent of mixing
in the mixed biopolymer system. This leads to mixed biopolymer systems that contain
sugars having a different morphology to those formed in an aqueous environment.[25]
Some work has been carried out into investigating the structure of confectionery gel
systems [28,29], as well as thermomechanical behaviour of such systems.[25,29,30] Moisture
content and water activity can also have an effect on the behaviour of these gels, with the
equilibrium relative humidity of confectionery gels being 65–75%. Due to the complexity
of these systems, and the interactions of confectionery gel system components, limited
information is available on quantifiable structure-property interactions of these gels.

CONFECTIONERY GEL FORMULATIONS


A range of gelling agents can be used in the manufacture of confectionery gels,
which include: agar, starch, pectin, alginates, and gums.[31] However, the majority of gel
confectionery consists primarily of sucrose, glucose syrup, starch, gelatin and water, with
a number of minor components including food acids, flavourings, and colourings. Pectin
may also be used as a gelling agent.

Sucrose and Glucose Syrup


Sucrose used in food is often in granule form and is used as a sweetener. It is often
used in concert with glucose syrup, as the syrup can enhance sucrose solubility and retard
sucrose crystallization in food products.[3,32] It serves to contribute to the texture and sen-
sory properties of the gels and may be used to increase product bulk or weight, giving
body or mouthfeel to the product. The form of sucrose used in jelly and gum confectionery
manufacture is non-crystalline.[33,34]
Glucose syrups refer to products with a dextrose equivalent (DE) of between 20 and
80 (where 100 indicates pure glucose and 0 indicates no glucose). Products with a DE of
CONFECTIONERY GELS FORMULATION 181

less than 20 are termed maltodextrins, and those with a DE of greater than 80 are called
hydrolysates or hydrols. Glucose syrup is specified by its DE, carbohydrate composition,
solids, and sulphur dioxide content in order to allow manufacturers to produce a product
exacting to customer requirements.[31] Traditionally, 42DE glucose syrup is used in con-
fectionery to prevent sucrose crystallization [31,35] to meet the requirement of confectionery
that it must not undergo any change in physical properties during storage. Use of higher
viscosity glucose syrup (lower moisture or higher DE) will slow sucrose molecule migra-
tion and inhibit graining.[31,36]
Glucose syrup is essentially shelf-stable and no preservatives need to be added to
prevent microbial growth, due to the high dissolved solids content which reduces the
water activity to below the level required for microbial growth.[31] Because confectionery
must not undergo fermentation, mould growth or other microbiological spoilage during a
Downloaded by [University of Central Florida] at 09:01 17 October 2014

long storage life, glucose syrup also aids in promoting a lower water activity in confec-
tionery gels, thus preventing microbial growth.[35] Glucose syrup is used in the production
of jelly and gum confectioneries, where its principal role is as a sweetener, although it also
contributes to texture and microbial stability, as well as stabilizing other ingredients, e.g.
sucrose or gelatin.[32,37] The gelling agents (e.g., starch, gelatin) in CGs usually make up
approximately 10% of the food gel, and the rest consists of glucose syrup and sucrose.
Confectionery products can often contain more glucose syrup than sucrose.[3]

Water
Water is a major constituent in many foods, supporting chemical reactions and acting
as a reactant in hydrolytic processes,[38] so removing water or binding it by increasing the
concentration of common salt or sugar in foodstuffs inhibits many reactions and retards
micro-organism growth, aiding to improve shelf life of the food product.[32] The physical
interaction of water with proteins, polysaccharides, lipids, and salts can contribute signifi-
cantly to food texture, as foods tend to become plastic when their hydrophilic components
are hydrated. The water content can affect glass transition temperature of a food (the tem-
perature where the material goes from glassy to rubbery behaviour).[32] In confectionery
gels, water often acts as a plasticizer to aid gel formation.
Studies by Cornillon, Andrieu, Duplan, and Laurent[39] showed bound water values
in sucrose/starch/gelatin systems which were: sucrose, 0.05 g/gDM; starch, 0.26 g/gDM;
gelatin, 0.44 g/gDM, where g/gDM denotes grams of water per gram of dry material. The
study shows that gelatin has the highest affinity for water when combined with sucrose
and starch, and hence requires it to hold its integrity.[32] The physical state of metastable
foods can depend on composition, temperature, and storage time. Water can affect their
properties, causing them to be glassy, rubbery, or viscous by altering glass transition
temperature. Variations in moisture content can lead to quality variations such as prema-
ture crystallization, stickiness, accelerated rancidity, lack of body, difference of chew,
hardness, poor handling on forming and cutting machines, and product flaws in surface
texture.[31]

Starch
Starch is present in most plant tissues, laid down in granular form in defined
cells,[35] as a storage carbohydrate. The major botanical sources of industrial starch are
tubers, grains and pulses,[40] with the composition and place of storage of the granules, as
182 BUREY ET AL.

well as granule shape and size being specific to the plant source.[41] Starches from various
origins have characteristic properties that are related back to granule size distribution.[32]
In addition to its role in plant physiology, starch is the most important source of car-
bohydrates in human nutrition[42] and is widely used, alone or in conjunction with other
gelling agents, to thicken and bind foods,[32] and to provide a wide range of jelly and gum
products. Native and modified starches are important ingredients of many fabricated foods [40]
and are often added to semisolid food products to contribute to their structure, and thus
improve fat and water-holding properties.[43]
Composition and structure of starch. Starch granules consist of starch, mois-
ture and small amounts of lipids and protein. The starch consists of two main components:
amylose and amylopectin.[32,41] Amylose is a predominantly linear α(1–4)-glucan
(Fig. 3a), whereas amylopectin is a highly-branched molecule consisting of an α(1–4)-
Downloaded by [University of Central Florida] at 09:01 17 October 2014

glucan chains with α(1–6) branch points (Fig. 3b).[44] The proportions of amylose and
amylopectin in the starch granules, as well as structure of the molecules (e.g. average
molecular weight, frequency of branching in amylopectin, naturally occurring level of
phosphorylation) also depends on plant source.[41]
The role of starch in confectionery gels is to provide the base of the gel structure,
and hence many of the gels’ textural characteristics. Confectionery gels often contain
“thin boiling starches.”[32,33] These starches are formed by adding a small amount of acid
to a starch suspension and heating at a temperature below the starch gelatinization temper-
ature producing starch hydrolysis.[31] When the required degree of chain concision is
achieved, the mix is neutralized and the starch filtered off and dried.
The acid hydrolyses some bonds within the starch, leading to a less linked structure
and a reduction in molecular weight of some of the chains. This causes the starch granules
to be readily soluble in boiling water [32] and disintegrate when cooked to give a lower hot
paste viscosity and higher gel viscosity than non-acid modified starches.[43–45] This type of
chemical modification is done to alter the nature of the interactions between polysaccharide
chains in the starch.[43,44] The differences between native and acid modified starches are
varied depending on the acid concentration used as well as time for acid hydrolysis.[46,47]
Acid-thinning can “roughen” the surface of starch granules, as well as decrease granule
amylose content, as amylose is more easily cleaved by acid hydrolysis than amylopectin.[48]
This causes a drop in crystallinity of the starch, and hence a more amorphous material,
which causes starch processing for food applications to be easier.

Figure 3 Chemical Structure of a) amylose and b) amylopectin.


CONFECTIONERY GELS FORMULATION 183

Figure 4 The starch gelatinization mechanism. a) Starch granules; b) Swelling of granules upon application of
heat and moisture; c) Leaching of amylose; and d) Creation of starch gel matrix (modified from Remsen and
Clark[52]).
Downloaded by [University of Central Florida] at 09:01 17 October 2014

Starch gelatinization. Starch gelatinization is often a major step in starch


processing and is loosely termed as a loss of starch crystallinity, coupled with granule
swelling.[43,49,50] Starch gelatinization occurs when aqueous suspensions of starch granules
are heated to above their gelatinization temperature Tgel, then cooled to form a rubbery gel
(Fig. 4).[51,52] When suspended in cold water, air-dried starch granules swell increasing in
diameter by 30–40%. When heating is applied to these swollen granules, irreversible
changes occur at Tgel,[32] for native starches, often 60–70°C at 30% moisture.[53] At Tgel
starch granules swell further, and amylose begins to leach out of the granules. On cooling,
the amylose forms a gel network, whilst amylopectin remains in granules during moderate
cooking.[35,54] Starch gels can be composite gels as the matrix and granules have distinct,
different properties. Swollen granules are remarkably elastic and give textural body to the
gel.[55] Depending on severity of processing and the starch source, starch gels can have
varying structures, including: 1. amylose with water entrapped in a three dimensional net-
work; 2. highly swollen and fragmented granules present in a gel matrix; 3. highly swollen
and intact granules in a gel matrix; and 4. unswollen and intact granules.[56]
On cooling and re-melting of starch gels, they have been found to consist of a crys-
talline phase and two immiscible liquid phases.[57] However, this may be dependent on
moisture content; Yuryev et al.[1] stated that at high water contents (> 70%) starch gels are
biphasic, whilst at low water contents (15–40%) they are monophasic. Tolstoguzov[58]
found that amylose and amylopectin can phase separate in starch gels. In acid-thinned
starch the gelatinization behaviour is altered depending on the native starch source. In the
case of barley and maize starches, acid modification can weaken gel formation if heating
is to 90°C, with the weakness due to increased amounts of amylopectin in the continuous
phase of hydrolyzed starch pastes [47]; heating hydrolyzed barley starch to 98°C can over-
come this problem as the amylose in the granules is liberated more easily.
The gelatinization temperature and the breadth of the gelatinization endotherm
(range of temperatures over which starch is gelatinizing) have been shown to increase on
acid hydrolysis of starch.[59] Acid modification has been found to increase solubility and
gel strength and decrease viscosity of starches.[60, 61] The viscoelastic properties of
starches are also affected by acid hydrolysis. Virtanen et al.[47] reported that the gel of acid
modified oat starch, although less rigid, was more elastic than the corresponding native
starch gel.
Composition and structure of starch gelatinization in high sugar systems.
Many factors affect the process of starch gelatinization and gel formation and so starch gela-
tinization must be carefully controlled to give a product with preferred consumer
184 BUREY ET AL.

attributes, such as texture.[62] Water concentration affects the heat treatment required to gela-
tinize starch. Starch requires approximately 30% moisture to fully gelatinize, any lower and
the gelatinization extent reduces whilst the gelatinization temperature, Tgel, will rise.[53]
Some studies have shown that the temperature profile of the starch gelatinization process,
and the moisture present, can be used to control the degree of gelatinization, with little
effect on final mechanical properties.[63] The presence of sugar can also affect starch gela-
tinization. Sugar type and concentration can have several effects, with the most important
being to raise starch Tgel. In limited water (<30%), starch Tgel increases with decreasing
moisture content. When sugar is dissolved in the water being used for gelatinization, part
of the water is bound by the sugar, effectively decreasing the moisture content, leading to
an increased starch Tgel.[53,64]
It has been proposed that upon reduction of available water, a point is reached at
Downloaded by [University of Central Florida] at 09:01 17 October 2014

which the limited extent of starch granule swelling is insufficient to disrupt the starch
granule completely.[65] Adding 12% or 24% sucrose has been found to delay swelling
of wheat starch granules at up to 50°C; above 50°C amylose leakage and granule
fragmentation is still delayed, but swelling is accelerated.[55] There are several theo-
ries as to why sugars inhibit starch gelatinization, and how and where the sugar acts to
inhibit starch gelatinization. One theory is that sugar molecules interact with starch
granule amorphous regions and increases energy required to melt them, while another
theory is that sugar displaces water inside the starch granules, therefore inhibiting one
of the promoters of starch gelatinization. Sugars are also thought to stabilize crystal-
line regions in the starch and hence immobilize water molecules, hindering starch
gelatinization.[53]
Starch retrogradation. Upon completing gelatinization, the starch now has a
gel structure; however this is a metastable structure, and is subject to changes over time,
depending on its molecular configuration. Starch molecules are known to associate on
ageing, and crystallites form.[40] This is a form of retrogradation, and syneresis (separation
of liquid and solid) can occur in conjunction.[66] While acid-thinned starches have shown
slower retrogradation rates than native starch [32], they can display an increase in retrogra-
dation rate with increased hydrolysis.[46]
Retrogradation affects quality, acceptability, and shelf life of starch gelled products.[66]
It is often detected during storage where the starch gels become firmer, which is associated to
the retrogradation and syneresis,[35] often in the form of amylopectin recrystallization.[67,68]
Retrogradation can usually be quantified through differential scanning calorimetry (DSC)
studies where the movement of the glass transition temperature Tg (the temperature where
a material changes from an amorphous to a crystalline structure or vice versa) can indicate
the growth of crystalline areas in the starch gel.[21,69]
In gelled starch, sugars tend to help stabilize the structure, inhibiting chain reorgani-
zation (also known as recrystallization or retrogradation), where starch molecules reorga-
nize into their crystalline form, though they do not achieve the structure found in the
native starch granule.[70] Waxy maize and wheat starches have been found to be the more
sensitive to this effect. The order of stabilizing effect provided by different sugars is
sucrose>glucose>fructose.[71]

Gelatin
Gelatin is a thermoreversible gel formed in aqueous solvents through lowering
the temperature.[72] It is derived from collagen via controlled acid or alkaline hydrolysis.
CONFECTIONERY GELS FORMULATION 185

Collagen may come from hide, bone, or other collagenous material.[32] Commonly, the
collagen used is of bovine (cow), porcine (pig), or piscine (fish) origins.[73] The properties
of the gelatins derived are affected by the source, age and type of collagen.[45] Gelatin has
a role as a gelling agent, providing texture and water binding properties to food materials.
Like starch, it is one of the most widely studied functional biopolymers.[74] The functional
properties of gelatin depend on the collagen source, and the extraction process used, both
of which affect molecular weight of the gelatin product.[32]
Composition and structure of gelatin. A typical gelatin consists of 14%
moisture, 84% protein and 2% ash.[45] The protein portion consists of several different
amino acids and an approximate composition of gelatin is shown in Fig. 5. The major
amino acids present are glycine, proline and hydroxyproline.[18,73]
These amino acids are arranged in gelatin gels to form long molecular chains, similar
Downloaded by [University of Central Florida] at 09:01 17 October 2014

to the collagen source from whence they came. These chains then form structures, which
then interact to lead to the overall gelatin gel network structure. Gelatin molecules contain
repeating sequences of glycine-X-Y triplets, where X and Y are frequently proline and
hydroxyproline amino acids. The glycine is said to be responsible for chain flexibility.[75]
These sequences lead to a triple helical structure in gelatin and its ability to form gels where
helical regions form in the gelatin protein chains immobilizing water. This triple helix struc-
ture is synonymous with many proteins. The basic macromolecular unit of collagen in which
the individual chains are wound in a gentle superhelix around a common molecular axis,[72]
which to some extent carries through to the gelatin structure. These helices are stabilized by
hydrogen bonds perpendicular to their axes.[76] The triple helices interact via secondary
forces to create a three-dimensional network. The arrangement and appearance of this net-
work is dependent on processing methods by which the gelatin becomes a gel.
Gelation of gelatin. Gelatin can undergo thermoreversible gelation at protein
concentrations > 2–3%. The sol-gel transition of gelatin has been studied extensively and
has been shown to be affected by many different factors.[77–80] A gelatin solution, which
can eventually produce a gelatin gel, is normally formed through either an indirect solu-
tion or a direct solution. To form an indirect solution, gelatin particles are added to cold
water just ensuring that particle wetting and that the granules swell until a soft friable
mass is formed. A gelatin solution is formed when this mass is heated to 50–60°C. Con-
stant stirring of the solution aids dissolution.[73]

Glycine
10% 9%
6% Proline and
Hydroxyproline
Glutamic acid
25% 8%
Alanine

Aspartic acid
15%
Arginine

Other amino acids


27%

Figure 5 A typical gelatin amino acid composition.


186 BUREY ET AL.

Forming a direct solution is the more common method, since it eliminates the cold
soaking stage in the indirect solution method, thereby reducing production time. This
method requires higher temperatures (60–80°C) and high speed agitation to prevent
clumping when gelatin is added to the water. Water is heated, and stirred vigorously
enough to create a vortex. Gelatin powder or granules are then slowly added, with stop-
ping and stirring to ensure even dispersion and dissolution. After all the requisite gelatin is
added, extra stirring is carried out to ensure complete dispersion of the gelatin in the
water.[73]
There are several theories about how formation of gelatin gels occurs. Many of these
may be valid for the types and concentrations of gelatin, and processing methods, used in
confectionery gels. Above 40°C gelatin in solution behaves like a typical synthetic poly-
mer with the individual macromolecules each assuming random-coil configurations [79]
Downloaded by [University of Central Florida] at 09:01 17 October 2014

with typical molecular weights of 2 × 105 Daltons.[77] These random coils consist of single
polypeptide chains, termed α–chains that may be entangled. Upon cooling these coils
undergo a random coil-to-helix transition, leading to gelation.[79] These solutions may
then be cooled to form a gelatin gel. This occurs through a sol-gel transition.
The two important steps in gelatin gelation which are often termed setting and age-
ing. Setting involves the linking together of irregular regions on the triple helices to form a
network throughout the whole gelatin solution. Ageing is the step where gel strength is
developed -it appears to go on indefinitely, however the rate of gelation decreases with
time at a constant temperature (Fig. 6).[79] Ageing involves two mechanisms: first, there is
continuous adjustment of the molecular network, through motions of the chains between
links; and second, by the dynamic nature of hydrogen bonding. The original interchain
links formed in the setting stage are strengthened in this way, as only the strongest bonds
of these can survive at constant temperature. These links are extended by incorporating
adjacent parts of chains.[76] Occurring simultaneously is the linking together of adjacent
collagen-like regions, the network thickens and becomes more fibrillar. The setting pro-
cess of gelatin is naturally slow and problems of too rapid aggregation are not normally
encountered.
The ageing step shown in Fig. 6 takes a long period of time, and in confectionery
gels, it is believed that interactions between gelatin and other components would slow this
process — as sucrose and glucose syrup tend to stabilize gelatin structure. Sugars are good
at enhancing the stability of conformationally ordered junctions in gelatin gels.[23] Finer,
Franks, Phillips, and Sugget[79] studied pure α-chain gelation and found that random coils
in the gelatin sol would take some time to become nucleated random coils, but after this

Polymer
chains Setting Ageing

Figure 6 Sol-gel transition in a gelatin system.


CONFECTIONERY GELS FORMULATION 187

stage would rapidly gel into a rigid state molecule. The gelation process is connected to
conformation transitions of protein chains at <36°C. The triple helix structure of the col-
lagen can be partially recovered, but can depend on local thermal history.[7]
It has been suggested that the most suitable technique to investigate the sol-gel
transition of gelatin is rheological analysis, particularly small angle oscillatory measurements,
as viscosity and elasticity can be measured without disturbing the gelation process if car-
ried out carefully. Some rheological results on gelatin gelation have shown various phases
or steps in the gelatin gelation process and displayed that with the growth of G′ (storage or
elastic modulus) the gelatin system has become more solid-like.[74]
Factors affecting gelation of gelatin. Many of the factors affecting starch
gelatinization also affect gelatin gelation. With regards to temperature, since gelatin is
a thermoreversible gel, gelling temperature, and temperature gradient during cooling
Downloaded by [University of Central Florida] at 09:01 17 October 2014

from the sol-state and temperature fluctuations during the gelation process all affect
product properties.[81] The lower the ageing temperature, the faster the helix content
within the gel will increase, but the lower the stability of the helices will be.[80] A small
increase in temperature leads to melting of some junction zones, along with formation
at new ones.
The presence of water aids gelatin gelation, however the higher the solids (which
could be gelatin granules, powder or leaf) content of the gelatin solution precursor, the
faster the gelatin gel will form. The initial change in modulus with time tends to become
steeper as gelatin concentration increases, and the modulus plateaus sooner with higher
concentrations.[82] Shear can also have an effect on gelatin gelation, and some studies have
shown that viscosity of the sol-gel can depend on a combination of shear rate and the
probability of bonding occurring.[78,81]
The presence of other ingredients during gelatin gel formation can affect the gelation
process as well as final properties of the gel. When mixing gelatin with gelatinized starch,
there is the possibility that amylose leaches into the gelatin phase — perhaps forming a
composite and affecting crosslink formation in the gelatin gel. Polymer blending laws indi-
cate that if the concentration of leached amylose is significant in a starch-gelatin-composite
system, then the strength of the system may be weaker than a pure gelatin gel due to
crosslink inhibition in the gelatin phase.[21] Many factors can affect gelatin gel structure
and strength, which in turn affects gel stability.
The final gel, after gelation is complete, is a clear, orange-tinged gel with elastic
properties. The specific properties of the gel should indicate what use it is good for. One
of the measured properties of a gelatin gel is Bloom. Bloom is determined by measuring the
mass required to press a 4 mm diameter probe to a depth of 12.5 mm into a 6.666 w/w%
gelatin gel at 10°C. It is the weight that mass, expressed in grams, that is the Bloom num-
ber. The Bloom number will determine what application the gelatin is used for, and is set
by the method used for extraction of gelatin from collagen, and the conditions under
which this extraction is carried out.[73]
As with starch, sugars also serve to stabilize the gelatin gel configuration. Sucrose
can help with gelatin dissolution, and stabilize the final product.[23] This is because
sucrose/glucose syrup blends can establish a continuous liquid phase with gelatin.[23,83]
The presence of sugar co-solutes can increase the strength of gelatin gels up to a peak co-
solute concentration, above which the gel weakens due to a lack of water available to
maintain gel integrity.[23,83] When combined with typical confectionery gel ingredients of
starch, sucrose, and water, gelatin tends to retain bound water more easily than the other
ingredients.[39]
188 BUREY ET AL.

Ionic force and pH can also have an effect on gelatin gel properties,[84] where pH
can affect turbidity (transparency) of gelatin, with the transmittance of gelatins dependent
on the isoelectric point of the gelatin.[73] The isoelectric point is the pH at which no net
migration occurs within the gelatin when placed in an electric field, and it is dependent on
the process for forming the gelatin from collagen, as well as collagen source. A low pH
can lower viscosity of gelatin solution during processing due to degradation effects. This
pH degradation effect is amplified at high temperature and hence food acids are usually
the last ingredient to be added to confectionery gel products during manufacture, usually
during cooling.[45,73]

Pectin
Downloaded by [University of Central Florida] at 09:01 17 October 2014

Pectin is found in virtually all land-based plants and is a structural polymer, the
intercellular “glue” that helps to reinforce the basic cellulose structure of plant cell
walls.[85,86] Commercial pectin is extracted under mildly acidic conditions from citrus peel
or apple pomace (dried pulp) and sometimes from sugar beet residues or sunflower
heads.[86,87] The chemical structure of pectin consists of a linear chain of galacturonic acid
units and can have a molecular weight of up to 150,000 (Fig. 7).[85,86]
While still in the fruit, there is, on average, one free acid group (COO-) to every
three to four methyl esters of galacturonic acid; although, there is no repeating sequence
within the polymer chain. This corresponds to a degree of esterification of 70–80%. Ester-
ification can be controlled during the extraction process, so that the degree of esterifica-
tion of the final pectin product can range from 0–75%. It is the degree of esterification and
the arrangement of methyl esters along the pectin molecule that controls how the pectin
behaves as a gelling agent or protein-stabilizing agent.[87]
Pectin with a degree of esterification of < 50% is termed low ester or low methoxy
pectin and does not require either sugar or acid to gel,[31] but does require calcium to aid
gel formation.[87] Pectin with a degree of esterification of > 50% is termed high ester or
high methoxy pectin, and requires the presence of sugar and acid to set to a firm gel.
Grade strength of this type of pectin is determined by the ratio of sucrose:pectin required
to form a gel of a particular strength.[31] It is the high methoxy form of pectin that is
widely used in the manufacture of jam and gel confectionery.[28,88]

O
O

OH

COOCH3

OH
(1 4)-α−galacturonic acid (1 4)-α−galacturonic acid
with carboxyl acid group partially methylated with methyl ester

Figure 7 Pectin chemical structure.


CONFECTIONERY GELS FORMULATION 189

Commercial pectin used in food production often has sugar added to standardize
viscosity and to prevent clumping when the pectin is added to water. Gel formation is car-
ried out by addition of the pectin, with added sugar, in a thin stream to well-agitated water.
As pectin absorbs water, the mixture thickens and gelation depends on acidification to the
correct level of pH, as well as on the concentration of dissolved sugar. Gelation of pectin
occurs very rapidly, so pectin gel confectionery is able to be produced in a continuous
manner.[31,33] Recent advances in the study of pectin have uncovered a wide range of
physical pectin structures that may be used for a variety of applications, including confec-
tionery, and this is discussed in greater detail in an excellent recent review by Willats,
Knox, and Mikkelsen.[89]

Other Confectionery Gelling Agents


Downloaded by [University of Central Florida] at 09:01 17 October 2014

Recently there has been growing interest in the use of bacterial polysaccharides for
gel formation as procedures for forming these polysaccharides on a larger scale are devel-
oped. Gellan and curdlan are examples of such materials and they have been proposed as
useful gelling agents in confectionery.[3,35,90–92] Gellan is a linear, anionic heteropolysac-
charide, with a molecular weigh of approximately 5 × 105 Da. The structure contains of
tetrasaccharide repeat units consisting of 1,3 – β-D-glucose, 1,4-β-D-glucuronic acid,
1,4 – β-D-glucose and 1,4,-α-L-rhamnose, along with acyl groups on the 3-linked glu-
cose. If left acylated, gellan forms soft, elastic, transparent and flexible gels but once de-
acylated it forms hard, non-elastic brittle gels.[3,91] Gels formed from gellan are usually
very stable over a wide pH range, while an increase in sugar or ion concentration can
greatly increase gellan gel strength.[3,35,91]
Curdlan is a moderate molecular weight unbranched linear 1,3-β-D glucan with a
molecular weight of approximately 100,000 with no side-chains. An advantage of using
curdlan as a gelling agent is that gels of differing strength are formed depending on the
heating temperature, time of heat-treatment and curdlan concentration.[90] It is commonly
used in Japan to improve texture of tofu, bean jelly and fish paste and may have other
applications in confectionery jelly.[92]

Food Acids
Food acids are added to confectionery gels primarily to give a tart tangy taste, and in
the case of pectin, to aid gel formation. The choice of food acid is related to the desirable
level of sharpness and the likely effect of the acid on rather raw materials present. Common
organic food acids used are citric acid, malic acid, tartaric acid, adipic acid and fumaric
acid. Lactic acid may also be used, particularly when gelatin is added to fermented
milk.[73] The most commonly used acid is citric acid as it causes the least degradation in
other food materials used in food products, and it is the food acid being utilized in the gels
being studied. The other acids are used in foods to achieve a different flavour profile to
citric acid.
Citric acid occurs naturally in lemon juice. It is produced by fermentation through
the action of certain moulds on sugar syrups. It is commercially available in anhydrous
form and as a monohydrate, is odourless and colourless and dissolves readily in water. In
confectionery manufacture, a 50% citric acid solution is commonly added near the end of
confectionery gel processing to avoid the harsh combination of acid and high temperature
affecting the other ingredients.[36, 45] All fruit acids have the ability to break down sucrose
190 BUREY ET AL.

to invert sugar, which is a mixture of dextrose (glucose) and laevulose (fructose).[31] Food
acids also lower viscosity in gelatin,[73] as well as cause coacervation of gelatin in mixed
gels. The food acids alter the pH of the confectionery gels, and this acidity has some pre-
servative action which helps shelf stability of the gels.[36]

Colourings
Colourings make CGs more attractive to the consumer. There are three main catego-
ries of food colourings.
• Synthetic — no similar natural colour.
• Nature identical — synthetic material but identical to a natural colour.
• Natural — obtained from plants or animals.[36]
Downloaded by [University of Central Florida] at 09:01 17 October 2014

It is assumed that colourings have little effect on the mouthfeel/texture of the gels, only
affecting appearance properties and perhaps flavour perception.[3] Most colourings are
water soluble and are added in very small amounts <1%) during confectionery gel pro-
cessing. The choice of colour depends on many factors, which can include the effect of
highly acid conditions on colour stability, and the presence of protein and various other
ingredients which can affect colour intensity.[31]
Food standards of the relevant jurisdiction will affect the choice and concentration
of colouring agents.[3] Other factors that influence colour choice include those that affect
colour stability and appearance, including: excessive heat and light exposure; use of
strong acid; reducing ingredients such as sugars; proteins that bind some types of colours;
strong light, preservatives, background colour of raw materials, trace metallic contami-
nants, and micro-organisms. Colour defects that can arise from non-conducive conditions
include discolouration, fading, loss of sparkle, which can be affected by supply variations,
inadequate mixing, and poor storage.[31]

Flavourings
Chemically, flavourings are very complex substances, and like colourings they are
divided into the three main groups of synthetic, nature-identical, and natural.[3] Flavour-
ings can be essential oils obtained from basic fruit or spice, or a wholly synthetic mixture
produced by blending approved chemical materials. Flavouring agents are added at the
last moment of production, as they are volatile and can be affected by high temperature.[31]
Flavourings must be dosed into the product accurately and blended well to ensure even
flavour dispersion. The flavour compounds can be intense, and so are diluted before incor-
porating into the food product. Natural flavour compounds are more volatile, less stable,
than synthetic flavour compounds, but are perceived as better, healthier, and safer.[3]

CONFECTIONERY GEL FORMING PROCESSES


Currently most CG products are prepared in the food industry via a multi-step batch
process. This process consists of several mixing and heating stages, and a prolonged dry-
ing period (Fig. 8).[3,93] Confectionery gels processes serve to gelatinize the starch in the
product and blend the sugars, starch gel and gelatin gel together. The ‘cooking’ step can
be attained through a range of manufacturing methods using different apparatus.
CONFECTIONERY GELS FORMULATION 191

RAW
MATERIALS Optional mixing with
other gelling agent
for composite gels

1.
Dissolving of sucrose, 4.
2. 3. Stoving
glucose syrup and Cooking Depositing
gum or gelling agent /Drying

5.
Sweet
finishing
Downloaded by [University of Central Florida] at 09:01 17 October 2014

FINISHED
PRODUCT

Figure 8 Schematic of confectionery gel process (modifed from Edwards[3]).

Batch Processes
If starch is being used as a gelling agent, this step is often used to form the starch
gel, as well as dissolving and concentrating the sugar ingredients. Common cooking methods
used in industry include jet cooking,[3,94] coiled heater cooking,[95] and the more tradi-
tional, and now less common, open-pan boiling.[36] Jet cooking involves the injection of
high temperature steam at pressure to cook the confectionery gel ingredients. The mix of
ungelatinized starch suspended in the sugar solution is often preheated to 70–80°C, prior
to being fed into the jet cooker. The jet cooker is often a small vessel, in the order of litres,
with a short residence time. Alteration of pump rates and steam valve back pressures can
alter the amount of starch gelatinized, and hence, the degree of cooking. This process can
be very efficient as both sensible heat and latent heat of the steam can be used to cook the
gel ingredients. The main disadvantage is that the steam becomes an ingredient and so
care must be taken with boiler water, so it can be of sufficient food grade.[3]
Coil cooking involves passing the carbohydrate slurry through a coiled length of
pipe, which is often heated by high temperature steam. Studies have shown that heating
and the amount of starch gelatinization is highest in tightly wound coils versus being lowest
in straight tubes. At low coil diameters, the increase in relative viscosity of the carbohy-
drate solution is the highest.[95] The increase in starch gelatinization could be due to the
introduction of increased shear, via a more convoluted flow path, and hence higher
destruction of starch granules.
These processes can stand alone as continuous processes, but because of the
presence of the stoving step and limited resources (ovens) and time associated with this
step, the overall process must be carried out in batches. For the Australian confectionery
gels mentioned here, there is a “blending” step prior to the ‘depositinG′ step, whereby a
gelatin solution is mixed with the cooked carbohydrate ingredients, which can be done via
large mixing tanks or complex inline mixing apparatus. Additives are also added prior to
depositing.
192 BUREY ET AL.

Depositing often involves moulding of the cooked gel into starch moulds. The stov-
ing step is necessary to reduce the moisture content of the final gel products down to an
acceptable level, whereby microbial activity will not occur. Stoving may involve drying at
room temperature for 4–6 days, or oven drying for 2–3 days. This step within the process
is time and resource-intensive, with high wastage, and there is a need for reducing the time
taken to carry this out. The rate of stoving is determined by: size of the jelly product; vis-
cosity of the deposited gel; degradation temperature; and moisture diffusion limitations.[3]
Moisture content of the moulding starch will also have an effect, as the moisture gradient
between the gel and the starch will determine diffusion rate, as well as moisture diffusivi-
ties in these materials.

Continuous Processes
Downloaded by [University of Central Florida] at 09:01 17 October 2014

A trend in recent times has been to develop continuous processes for making food
products. Gel confectionery can be formed using a scraped-surface heat exchanger,[33]
however, in recent times research into confectionery gel formation has turned to some of
the more traditional synthetic polymer processing techniques for inspiration. Two of these
techniques are injection moulding and extrusion processing.
Extrusion of confectionery gels. Extrusion processing is becoming increas-
ingly important in food processing and is characterized by diverse applications. The tech-
nology is more available than it was in the past, but the science base is difficult to generate
because of complex applications.[96] Extrusion is a series of unit operations [97] including:
solids and liquid conveying, melting, pumping (mixing), forming, and cooling. Extrusion
processing is not completely new to the food industry and has been utilized to produce
pasta for more than 60 years.[98] It is widely used in the food industry for products such as
pastas and breakfast cereals, due to its versatility, high productivity, low cost and energy
efficiency, resulting in the potential for improved cost-effectiveness when compared to
traditional processing methods.[99,100]
A screw extruder consists of several main features-a barrel, a screw or screws, heat-
ing source, feed hopper and feed ports, a drive, and a die where the extruded product
emerges (Fig. 9). The process begins with feeding, whereby granules or powder are placed

Raw
material in

Screw Barrel Die


Hopper

Product out
Screen
Heaters

Figure 9 Schematic of an extruder.


CONFECTIONERY GELS FORMULATION 193

in the feed hopper and metered down into the screw or screws to be conveyed; liquid may
also be injected to promote plasticization. The extruder screw/s has three main zones
which are the feed zone, the transition zone and the metering zone. The appearance
and physical configuration in each of these zones may differ, to carry out the relevant
operations.[101]
The screws can be shaped as a solid piece, or they can be modular with interchange-
able screw elements. Modular screws are particularly common in twin screw extrusion.
Depending on what type of material manipulation is required a screw configuration can be
“built” to achieve this. Elements that just have a basic screw profile are used for convey-
ing, whilst more intricate screw elements such the angled paddles, are used for mixing,
with reverse elements providing some of the most intense mixing.[102]
Extrusion cooking allows starchy and proteinaceous food materials to be cooked
Downloaded by [University of Central Florida] at 09:01 17 October 2014

and formed in a single continuous high temperature short time (HTST) process.[103] It has been
studied as a processing technique to process starch,[98,104–106] starch-gelatin composites,[107]
pectin-starch composites,[63] fruit-starch gels,[108] and due to its flexible nature as a pro-
cessing technique, it has the potential to extrude other systems containing biopolymers.
Injection moulding of confectionery gels. Injection moulding involves a
similar setup to the extrusion process, but with the use of a die capable of being closed and
pressurized to form shapes, rather than a continuous extrudate. It has been studied as a via-
ble process for moulding 85% solids gelatin fruit gums containing 40–60% gelatin. In the
study, the parameters of die cylinder temperature, mould time, and plunger dwell time
were optimized to give suitable gum products. The main difficulty in the process seemed
to be the slow gelling of gelatin, and its low susceptibility to frozen in stress. The gums
had strong relaxation on demoulding, indicating that processing parameters had a lesser
effect on product properties and all the products formed under different circumstances had
similar characteristics.[109] Current processing methods require controlled time and
energy. For injection moulded gums, one minute in the mould seemed appropriate to give
the requisite gum textural characteristics. However, this amount of time would also be a
hindrance unless several hundred products could be moulded with each die cycle. For this
reason, extrusion processing may be a better option as a continuous process for confec-
tionery gels containing gelatin. The product continuously emerges in extrusion processing,
and there is no waiting for process cycle times.[109]
The rational development of food extrusion processing is dependent on under-
standing the relationship between microstructural changes of the cooking material,
equipment design, and processing conditions. Knowledge of melt rheological charac-
teristics is a prerequisite to performing design calculations on extruder screws and
dies and modeling the overall extrusion process.[103] Parameters such as barrel tem-
perature, screw speed, water content and feed rate can be optimized to obtain the
required product.[98,108]
In extruders, starch material is modified all along the screw shown by an intrinsic
viscosity decrease, and increase in degree of gelatinization. In single screw extrusion this
happens progressively, but in twin screw extrusion there is a characteristic very short
melting zone,[105] powder is compacted and dissolved, from which a polymer + melt mix
is evolved, then a homogenous molten phase is obtained in a very short time (3–10 s) at
high shear rate (approximately 100/s).[110] Transformations occurring here are difficult to
analyse experimentally with respect to instrumentation and sampling, as melting is initi-
ated as soon as the first restrictive screw element (die, reverse screw, kneading block)
encountered after feeding.[110]
194 BUREY ET AL.

GEL CHARACTERIZATION TECHNIQUES


In order to understand gels; how they form; what properties they have; and what
causes the behaviour of these properties; it is necessary to characterize gel properties. The
relevant properties of jellies for processing and eating quality are: rheology which can
affect pumping and transport during processing, molecular structure, molecular weight,
and texture, which gives an indication of gel eating quality.[3]

MICROSTRUCTURE
Foods have microstructure that is imparted to them by nature or through processing.
Structure is important, as almost all relevant food properties are structure sensitive,
Downloaded by [University of Central Florida] at 09:01 17 October 2014

reflecting interactions of components and structural elements occurring at different levels


in food.[111] Microscopy can be used as a technique to better explain published physical
properties, such as rheological or texture measurements.[112–114] Microstructure has been
practically ignored by food engineers in the past for several reasons, particularly:
• The emphasis of the food industry in the last century has been on designing safe and
reliable processes for large volumes of traditional products.
• Innovation and quality has not been a major driving force in the food market until recently.
• Food engineers knew little about the underlying sciences linking food structure to product
properties.
• The techniques aiding the study of microstructure were not developed to the point that
they could provide useful data (e.g. quantitative information) to engineers.[111]
Because of this, attempts to observe food microstructure and link food properties to this
have been made only recently. Advances in computer technology and development of
more sophisticated structure analysis programs have allowed the rapid analysis of micro-
structural images, greatly expanding their use. Now it is possible to perform several mea-
surements on the structures, which include:
• feature specific measurements (area, length, perimeter, volume, surface area, grey level,
etc.);
• shapes (form factor, aspect ratio, fractal dimension, number of holes, convexity, and
solidity); and
• global measurements (area fraction, line length and curvature etc).[115]
Light microscopy (LM) may be useful in observing phases in CGs; however, it does not
have the resolution of scanning electron microscopy (SEM) and transmission electron
microscopy (TEM). It is, nevertheless, a quick and simple method to check what is
observable within the gel microstructure before a more complex method is utilized to
examine it further. Light microscopy has proved useful in observing the microstructure of
agar gels,[116] as well as confectionery gel systems (Fig. 10), which helped in determining
the structural composition of the gels.[28]
Bright field and fluorescence LM are frequently used because they allow selective
staining of different chemical components. Cryo-sectioning involves freezing, and
ice crystals may damage the structure. Plastic embedding involving dehydration can cause
shrinkage. Confocal Scanning Laser Microscopy (CSLM) offers more advantages in that
there is minimal sample processing, and it can yield three-dimensional images. Electron
microscopy is required to investigate fine details.[117]
CONFECTIONERY GELS FORMULATION 195

Pastille 1 Pastille 2 Pastille 3


Downloaded by [University of Central Florida] at 09:01 17 October 2014

Continuous Protein Sugar syrup Starch


Phase

Dispersed Starch/Sugar Protein/Starch Protein/Sugar


Phase syrup syrup

Texture Chewy, rubbery, Soft, moist, Soft, jelly, moist


elastic, firm breaks down

Figure 10 Structure models of three jelly pastille products (modified from Groves 2003 [28]).

Scanning Electron Microscopy (SEM) is a useful tool for examining the surface of
microstructural components.[118] Samples are mounted on stubs, sometimes fractured, and
coated with a metal compound.[119] Environmental Scanning Electron Microscopy
(ESEM) allows viewing of the surface of a hydrated sample without the need for coating;
this is possible by observing the samples under vacuum.[118]
Transmission Electron Microscopy (TEM) is a method that may be used to view the
microstructure of a thin sample by passing electrons through it. In the case of food gels,
the section must be fixed and stained with a heavy metal compound to provide contrast
between the various components.[115] A successful method of carbohydrate/gelatin food
gel sample preparation for microscopic examination involves fixing in glutaraldehyde to
stabilize the gel by crosslinking the gelatin protein.[115,120] Following this step, the sample
is then post-fixed in osmium tetraoxide in to assist the staining contrast.[121] Finally the
samples are stained with uranyl acetate, dehydrated in a gradient series of acetone, and
embedded in a hard medium, often resin.[14]
TEM studies of gelatin-pectin-sugar gummy confection ultrastructure showed
phase-separation. Most of the combinations of pectin and gelatin gave gels that formed a
196 BUREY ET AL.

pectin-rich continuous phase with a gelatin-rich dispersed phase.[14] Due to the sample
preparation method gelatin-rich phases showed up as dark, while pectin-rich phases
showed up lightly coloured. Phase separation also occurred when sucrose and glucose
syrup were added to a pectin-gelatin mixture.[83]
DeMars and Ziegler[14] found that in gelatin-pectin-sugar confections that a gela-
tin:pectin ratio of 3:1, an extensively coalesced dispersed phase was found while in con-
fections with a gelatin:pectin ratio of 4.5:1 a more homogeneous structure was observed.
This was explained by the competing effects of coalescence (affected by phase viscosity)
and rate of gelation (affected by gel concentration). Lower total polymer (gelatin and pec-
tin) concentrations caused increased rate of coalescence and decreased gelation rates,
leading to larger dispersed phase microstructures, whereas the opposite is true of higher
total polymer concentrations. Recent research into the structure of gelatin-starch-sugar
Downloaded by [University of Central Florida] at 09:01 17 October 2014

confections has shown a range of CG microstructures ranging from a gelatin-rich matrix


enclosing starch/sugar-rich inclusions to bicontinuous structures, to a starch/sugar-rich
matrix enclosing gelatin-rich inclusions (Fig. 11).[122]
Confocal Laser Scanning Microscopy (CLSM) and TEM combined with stereological
image analysis have been used to investigate gelatin/maltodextrin gel systems. For gel
compositions ranging from 4% gelatin/2.25% maltodextrin to 5% gelatin/5% maltodextrin,

Figure 11 TEM micrographs of four different commercial confectionery gels showing a) a starch/sugar-rich
matrix enclosing gelatin-rich inclusions; b-c) a bicontinuous structure d) a gelatin-rich matrix enclosing starch/
sugar-rich inclusions (scale bars all 2μm) from Burey[122].
CONFECTIONERY GELS FORMULATION 197

there was a continuous gelatin phase with maltodextrin inclusions. The size of the inclu-
sions increased with maltodextrin content. At high temperature the gel appeared relatively
homogeneous and upon cooling over time, phase separation occurred.[19] The system goes
from a fine dispersion at 60°C to a greatly phase-separated gel at 30°C.[123]
In gelatin/maltodextrin gels, it is possible to cause an inversion from gelatin-continuous
to maltodextrin-continuous gels by decreasing the gelatin chain length and/or increasing
maltodextrin chain length.[124] This allows the buildup of long chain links throughout the
gel by the polysaccharide component. Gelatin/maltodextrin is one of the widely studied
composite gel systems and Butler[15] found when looking at gelatin/maltodextrin systems
that in the system there was a loss of molecular mobility caused by gelation, although over
time the gelled system continued to evolve. He found that in this system droplet coales-
cence was the dominant structure coarsening mechanism in the liquid system, which
Downloaded by [University of Central Florida] at 09:01 17 October 2014

agrees with Lorén, Altskär, and Hermansson.[123] Nucleation also occurs, leading to phase
separation.
In addition to observation of food gel phases, it has been possible to observe the net-
work structure of food gels using a surface replication technique and then viewing under
an electron microscope with a goniometric (moveable) stage. The images of gelatin
showed a dense filamentous network, the features of which were dependent on the type of
gelatin being investigated and method for forming it.[125] The addition of sugar to gelatin
or polysaccharide systems can result in changes in gel network morphology, when com-
pared to aqueous gels, and it has been shown by phase separation in gelatin gels, and in a
less fibrillar network in gellan gels upon the addition of sugar.[23] These changes show that
polysaccharide gel networks are weakened by the addition of sugar, whilst gelatin gel net-
works are reinforced by self-attraction of the gelatin chains and repulsion from the sugar
phase. These changes in morphology will obviously affect other gel properties such as
mechanical properties and sensory properties.
Atomic Force Microscopy (AFM) is a very recent technique which can observe
structures at much higher resolution than other techniques such as LM, SEM, TEM, and
CLSM. Typically, a cantilever with a sharp tip is scanned over the surface of a sample; a
laser is aimed at the top of the cantilever above the tip, and the deflections of the cantilever
are measured by a detector which converts the signal received into a topographical image
of the sample. Traditionally AFM has been used to observe surface structures of materials,
and is an attractive technique to utilize for this purpose as it has the ability of performing
three-dimensional measurements with a resolution on the order of nanometres.[126] Some
other advantages of this technique is the very simple sample preparation, hence, avoiding
damaging or altering the sample prior to measurement,[127] and also the ability to observe
samples in air or under solution. The use of AFM for observing fine gel network structures
has developed only recently and hence the capabilities are still being explored. It has been
used to investigate structure in pectin, gelatin and starch structures.[128–130] AFM also has
the capability of carrying out mechanical measurements, providing further information on
sample characteristics.[131]

MECHANICAL BEHAVIOUR
There are several empirical mechanical tests that can be carried out in order to char-
acterize a gel, depending on how rigid it is. Some tests that depend on linear motion
include: shearing, puncture, extrusion, bending, compression, tension, and cutting.[132]
The results obtained using these methods are strongly dependent on sample mass and
198 BUREY ET AL.

Mode 1
or

Mode 2
Downloaded by [University of Central Florida] at 09:01 17 October 2014

Mode 3

Stage 1. Start 2. At maximum 3. After test


deformation

Figure 12 The three stages during deformation testing. Stage 1. Deformation during small deformation testing;
Stage 2. deformation during intermediate deformation testing; Stage 3. Deformation during large deformation
testing (modified from Olkku and Sherman[137]).

geometry, sample surface, cell geometry, test speed, and soon.[133,134] The irregular shape
of CG product samples can also cause problems in texture analysis, so it is important to
prepare uniform samples.[135] Large deformation mechanical testing is used to investigate
failure and texture, whilst small deformation testing is used to investigate network gel
structure.[136] The tests differ on how their loading affects the sample dimensions during
and after the test (Fig. 12).[137]

Large Deformation Testing — Texture Profile Analysis


In an industrial context, the ultimate judge of food texture is the consumer.[138] The
term texture is often misunderstood, and hence, there is a requirement for a standard and
consistent use of terminology when investigating texture.[139] Texture as applied to foods
in general has been defined in many different ways. One official definition provided by
the International Organization for Standardization is that texture encompasses all rheolog-
ical and structural (geometrical and surface) attributes of a food product, which are per-
ceptible by means of mechanical, tactile and when appropriate, visual and auditory
receptors.[138] Often these attributes are evaluated in relation to deformation under an
applied force and measured objectively in terms of force applied, distance the force is
applied through and time of deformation.[139] The first implication of this definition is that
texture is related to sensory perception of food product attributes. Examples of sensory
descriptors in the case of CGs are chewiness and gumminess.[138] Jellies can be termed
“chewy,” and this behaviour is a result of their ductile nature.[140]
Texture arises from the arrangement of various chemical species by physical forces
into distinct macro-and microstructures. Texture is an external manifestation of these
structures.[111] Texture awareness is often at a subconscious level, if texture is as people
have come to expect it to be, it may go un-noticed; however if the texture is not as
CONFECTIONERY GELS FORMULATION 199

expected, it may become a point of criticism.[138,141] For this reason, food manufacturers
are often reluctant to move from an existing food production process to a more economic
one, for fear that they will lose the very textural attributes that attract consumers in the
first place.
The relationship between objective texture measurements and sensory properties are
often poorly characterized or misunderstood. Only through application of a number of
objective methods, based on different principles (e.g., compression, shear) and property
measurement on differing scales (e.g., large versus small deformation) can a food material’s
texture be fully characterized. However, complete texture is rather ambitious, therefore,
only a few highly relevant tests are often carried out.[139] Large deformation studies are rel-
evant to food texture as large deformations occur in the mouth.[138] The selection of instru-
mental method for texture analysis is important, and it is necessary to consider the purpose
Downloaded by [University of Central Florida] at 09:01 17 October 2014

of the test, ensuring that it accurately measures quality characteristic of choice, with instru-
ment choice being influenced by ruggedness and versatility within cost constraints.[132]
A puncture test measures the force required to push a probe or punch into food to a
depth that causes irreversible crushing of flow of food. This is the simplest and most
widely used texture testing technique, and it is very useful, as the test can measure
compression or shear or a combination of the two.[142] The choice of testing speed is impor-
tant, and it is necessary to select one that correlates well with sensory perception.[143] In
compression testing, it is possible for the finish of the compression surfaces to influence
the results obtained under certain conditions. This is because surface finish has a great
bearing on contact stresses developed. Samples may be tested under constant area condi-
tions (bonded) or lubricated (sample is free to move).[144]
The instrumental texture profile analysis (TPA) method was developed at General
Foods in the early 1960s [145] and involves imitating the first two bites of the food mastication
process with an instrument, and relating the data obtained to textural descriptors.[138,146]
The test involves a two-cycle penetration test into the food sample, with the force devel-
oped observed over time. The output obtained shows a two-peak force curve, and calcula-
tions are carried out to determine the magnitude of textural parameters.[146] The
magnitudes of the textural parameters are dependent on testing geometries, so it is impor-
tant to be consistent with sample size when testing products.
Gellan gum is a gel-forming polysaccharide and gellan gels have been studied using
a texture analyzer.[147] The samples were 21 mm in diameter and height and compressed at
a rate of 0.3 mm/s until failure. Failure strains ranged from 0.5–1.6% and failure stresses
ranged from 10–70 kPa based on polymer concentration and calcium presence.[147] Studies
were also carried out on the effects of gellan gum, citric acid and sweetener on lemon jellies.
The main findings were that jellies were harder with low citric acid and high gum pres-
ence and that the influence of citric acid varied with gum contents.[148]
Texture of gelatin-pectin confectionery gels has also been investigated. The gels
were made up of 33.4% sucrose, 29.8% 42DE corn syrup solids and varying gelatin and
high-methoxyl pectin contents. Using a tensile test on a texture analyzer it was observed that
these gels had a fracture strain on order of 0.5–1.2 and a fracture stress of 290–300 kPa
based on gelatin and pectin concentrations.[14] TPA is a technique commonly used in
industry for the evaluation of food textural behaviour,[146,149,150] as it can give an indica-
tion of sensory eating properties.[151] An interesting engineering approach is to use TPA as
a quality control tool, with engineering specifications of textural parameters required to be
met, in order for food materials to be deemed suitable for the consumer market.[140] This
approach can aid in setting up a model for texture-sensory relationships.
200 BUREY ET AL.

Small Deformation Testing — Rheological Behaviour


Small deformation studies allow the study of effects of composition and processing
variables to lead to a greater understanding of mechanical properties of food.[132] Rheology
is the study of deformation (solid) and flow (liquid) of matter. The rheological properties
of interest are elasticity (solid behaviour) and viscosity (liquid behaviour) and foods
generally tend to have both.[138] Rheological information of a food gel can give an indica-
tion of some of the sensory and textural properties it will have when it reaches the
consumer,[152] as well as provide useful information on sol-gel transitions and gel charac-
teristics.[153]
Constant strain dynamic (oscillatory) rheological tests can provide valuable rheo-
logical information on the viscoelastic nature of foods. In dynamic oscillatory rheometry a
Downloaded by [University of Central Florida] at 09:01 17 October 2014

specimen is subjected to a small amplitude strain where the strain is applied at a given
frequency (Hz or angular frequency). If the material is viscoelastic, it will exhibit solid-
like and liquid-like behaviour.[154] If the material is linearly viscoelastic, stress will vary
sinusoidally with strain, but will be out of phase with it. In an ideal elastic solid, stress and
strain are in phase, while in a true Newtonian fluid, stress is 90° out of phase with strain.
Thus when a sinusoidal deformation is applied and the resultant stresses observed, real
materials are found to be intermediate between the two elastic and viscous extremes.[21,155]
Modulus is defined as stress/strain in any constant deformation experiment. For a dynamic
sinusoidal experiment it follows that two moduli can be defined: 1. in-phase stress/strain
or storage modulus (G′); and 2. out-of-phase stress/strain or loss modulus (G″) G is the
nomenclature used for shear tests, and E is used when strain (tension or compression) tests
are performed. G′ is a measure of the energy stored and recovered from a material during
each cycle whilst G″ is a measure of energy dissipated or lost (as heat) during each cycle
of sinusoidal deformation.[21,55] A complex modulus, G*, which takes into account both
viscous and elastic behaviour, can be developed from the two moduli G′ and G″, where:

G* = G ′ 2 + G ′′ 2 . (3)

There are several dynamic oscillatory tests that can be carried out. Strain sweeps need to be
carried out to determine appropriate conditions for non-destructive testing and to determine
the linear viscoelastic range — where moduli are virtually independent of strain.[21,155]
Following on from the dynamic strain sweep test there are three other rheological tests
that can be carried out which are:
• frequency sweeps, where frequency of oscillation is varied, and the temperature and
strain are kept constant. The viscosity profile, moduli, and other parameters are mea-
sured as a function of frequency;
• temperature sweeps where temperature is varied, and the frequency of oscillation and
strain are kept constant. Viscometric parameters, moduli, and other parameters are mea-
sured as a function of temperature; and
• time sweeps where temperature, frequency and strain are kept constant and the test is
run for a period of time. The viscometric and moduli time profile is determined.[153]
Because gels are viscoelastic, dynamic tests are well suited for studying characteristics of
gels as well as of gelation and melting.[153] Starch pastes or gels can have viscous and elastic
properties.[40] Rheological tests have been carried out on several food gels and provided
CONFECTIONERY GELS FORMULATION 201

information of interest. Investigations carried out on maize starch of 20% moisture at


110°C found that viscosity was 1000–2000 Pa.s.[103] An 11% wheat starch dispersion has
shown a complex modulus G* on cooling over time of 2000–3500 Pa.[43]
Time sweeps carried out on wheat starch solutions containing up to 40% starch,
showed that the 40% solution took the longest time to plateau when tests were carried out
at 0.2 Hz and 2.0% strain for 24 hours.[40] Other time sweep investigations carried out on
starch gels showed a plateau for most concentrations after pasting for 45 minutes, however
the heating method used was not stated.[156] Jet-cooked starch has shown a crossover from
solid behaviour dominance to viscous behaviour dominance during frequency response
tests.[94] This behaviour is dependent on the presence of stirring during gel formation, and
time.
Penetration tests may be utilized to obtain rheological information. Gelatin gels
Downloaded by [University of Central Florida] at 09:01 17 October 2014

show a shear modulus which is independent of penetration depth, whilst polysaccharide


gels are initially independent of depth as long as the ratio of sample height:sample radius
is > 0.2.[157] It was found that results for gelatin were similar to those obtained in a rota-
tional rheometer, whereas this was not so for polysaccharide gels.
Studies of the effects of sucrose/glucose syrup on gelatin gels have yielded some
interesting rheological results. A combination of 5% gelatin and 66% sucrose/glucose
syrup shows predominantly elastic behaviour over the range of frequencies 0.01–10Hz
while 3% gelatin in 70% co-solute exhibits a cross-over frequency of 3–5 Hz.[83] Temper-
ature sweeps carried out on amylopectin-gelatin mixtures show two different thermal tran-
sitions occurring when carried out at 1 and 5 Hz at a rate of 2°C/min.[158] This provides
some inkling of how confectionery gel systems may behave. Investigating rheology of
gels online during processing could prove useful in providing a control measure.[159]

SIGNIFICANCE OF GLASS TRANSITION


Glass transition of an amorphous polymer system is the change in system behaviour
from rubbery to hard and relatively brittle. In recent times the importance of this transition
has gained wider appreciation of its use in understanding and quality control of such sys-
tems. Confectionery gels are amorphous polymer systems and hence the rubber to glass
transition is a significant factor in analysis of the gels. A common method to find when
glass transition occurs is to determine glass transition temperature (Tg), but other tech-
niques such as dynamic mechanical analysis may be used to observe where the changes in
behaviour from rubber to glass occur. Fig. 13 shows a master curve detailing the four
regions of mechanical behaviour for amorphous polymer systems.[25,29]
Section I of the curve denotes where molecular flow is predominant in protein or
polysaccharide solutions and concentrated preparations of glucose syrup, leading to G″
being greater than G′. In section II of the curves, the samples form elastic networks of sta-
ble physical associations (G′ > G″) which can be seen in gummy candies at room temper-
ature.[123] In the absence of crystallization, cooling of rubbery foods ushers in the glass
transition region shown in section III of the curve at which the viscous response becomes
again dominant (G″ > G′). Finally, the moduli crossover for a third time and the system
enters the glassy state in section IV at which the rates of chemical, enzymatic and micro-
bial processes decelerate dramatically.[25]
The appearance of the master curve gives some indication of molecular movements
within the material during changes in frequency or temperature. In the glass transition
region (section III) backbone adjustments of polymer chains (elastic contribution) is
202 BUREY ET AL.

Decreasing temperature

G’
Log (modulus)

G”

Tan δ
1
Downloaded by [University of Central Florida] at 09:01 17 October 2014

I II III IV

Increasing frequency, molecular weight and concentration

Figure 13 Variation of G′, G″, and tan δ as a function of temperature, frequency, molecular weight, and concen-
tration for amorphous polymers (modified from Kasapis, Mitchell, Abeysekera and MacNaughtan[23]).

limited, but short-range pendant group, or molecular movements capable of dissipating


energy are present. Upon entering the glassy zone (section IV), extremely high frequencies
prevent configurational rearrangement of chains; only stretching and bending of bonds are
permitted leading to an elastic, rather than a viscous, character.[29]
The appearance of this curve, and the temperatures and frequencies at which the
boundaries of the curve sections occur are important for understanding of confectionery gel
behaviour. Frequency dependence curves obtained from dynamic mechanical thermal analy-
sis (DMTA) can be compared to such a master curve, and the structural behaviour analyzed.

STRUCTURE-PROPERTY RELATIONSHIPS IN FOOD GELS


In an attempt to investigate how properties of food gels are developed, a few recent
studies have looked at the relationships between microstructure and properties of food gel
type systems. Many studies have begun to look at composite gels and related properties,
while taking into account the effects of microstructural characteristics on these proper-
ties.[113,114,160–165] Composite gels often consist of protein and polysaccharide components,
and hence have a tendency to phase separate.[11–13,166,167] The studies state some reasons
why the molecular arrangement of the composite gel systems affects such properties. New
techniques have even looked at combining microstructural analysis with rheological analysis,
in one machine measurement,[168–170] whilst models are now being developed to predict
biopolymer gel properties based on their structural characteristics.[141,171]
In some studies, polymer blending models have been used to explain the properties
of composite food gels. Studies by Abdulmola, Hember, Richardson, and Morris[21] were
done on gelatin-starch composite gels where X in equations 1 and 2 was considered as gel-
atin, the “stronger” gelling component. In the study, there was little change in composite
moduli with starch concentration when starch granules were present in a gelatin matrix,
CONFECTIONERY GELS FORMULATION 203

however at higher gelatin concentrations the composite modulus went closer to the upper
bound at lower temperature (5°C). At low gelatin concentration, the lower bound fitted.
The focus in the past has been on the effects of additives or processing changes on
individual gel characteristics, without considering the relationships of gel characteristics
to each other.[172–174] Very few studies have looked at complete confectionery gel systems,
containing carbohydrate, protein and high sugar co-solutes, and the relationships between
gel properties and structure.[14,28,175] Some studies have been carried out looking at gel
composites with co-solutes of sucrose and glucose syrup, as these systems are representa-
tive of true food gel products.[25,67,83]
It has been simpler in the past to investigate structure-property relationships in gels
containing only one biopolymer, as the system is not so complex.[94,176] Now that a greater
understanding is required of complex food systems, in order to develop models for pro-
Downloaded by [University of Central Florida] at 09:01 17 October 2014

cessing-property relationships and quality control, there is a need for further study into
structure-property-processing interactions.

SUMMARY
Whilst much is known about confectionery gel ingredients, and interactions between
some of these ingredients, there is limited knowledge available about how the entire sys-
tem is developed, due to the complexity of ingredient interactions and processing-induced
changes. Although there is a lack of knowledge about the development of confectionery
gel microstructure, and texture development in typical industrial processes, gel formation
of individual biopolymers (e.g., starch, gelatin) has been studied separately; and, the fun-
damental phenomenon of phase separation in similar biopolymer mixtures has also been
investigated. Previous food gel studies have focused on looking at individual gel charac-
teristics, such as microstructure, texture and rheological behaviour, as well as behaviour of
single biopolymer gel systems. It is proposed by further studies that textural behaviour and
rheological behaviour of food systems is dependent on the structural characteristics of the
food, and this principle can be applied to a confectionery gel system. It is necessary to
acknowledge the importance of creating consumer-acceptable textures, and understanding
how these textures come about, through processing and structure manipulation.

ACKNOWLEDGMENTS
The authors would like to acknowledge the Australian Research Council (ARC Linkage projects LP
LP0211915 and LP0453529), and the industrial partners on these projects. PB would also like to
thank The University of Queensland for providing her scholarship.

REFERENCES
1. Yuryev, V.P.; Nemirovskaya, I.E.; Maslova, T.D. Phase state of starch gels at different water
contents. Carbohyd. Polym. 1995, 26 (1), 43–46.
2. Edwards, S.; Lillford, P.; Blanshard, J. Gels networks in practice and theory. In Food Structure
and Behaviour; Blanshard, J.; Lillford, P.; Eds.; Academic Press Ltd.: London, 1987; 1–12.
3. Edwards, W.P. The Science of Sugar Confectionery; The Royal Society of Chemistry:
Cambridge, 2000; 166 pp.
4. Hermans, P.H. Gels. In Colloid Science, Volume II; Kruyt, H.; Ed.; Elsevier Science: Amster-
dam, 1949; 483–651.
204 BUREY ET AL.

5. Stainsby, G. In Gelation and Gelling Agents, Proceedings of the UK Symposium on Gelation


and Gelling Agents, London, UK, Oct 20th-, 1971, The British Food Manufacturing Industries
Research Association: London, 1972.
6. Flory, P.J. Principles of Polymer Chemistry; Cornell University Press: Ithaca, 1953; 672 pp.
7. Djabourov, M.; LeBlond, J. Thermally reversible gelation of the gelatin-water system. ACS
Sym. Ser. 1987, 350, 211–223.
8. Peyrelasse, J.; Lamarque, M.; Habas, J.P.; El Bounia, N. Rheology of gelatin solutions at the
sol-gel transition. Phys. Rev. E. 1996, 53 (6), 6126–6133.
9. Kavanagh, G.M.; Ross-Murphy, S.B. Rheological characterisation of polymer gels. Prog.
Polym. Sci. 1998, 23 (3), 533–562.
10. Burchard, W.; Ross-Murphy, S.B. Introduction: Physical gels from synthetic and biological
macromolecules. In Physical Networks — Polymers and Gels; Burchard, W.; Ross-Murphy, S.;
Eds.; Elsevier Science: Essex., 1990; 1–14.
Downloaded by [University of Central Florida] at 09:01 17 October 2014

11. Tolstoguzov, V. Phase behaviour of macromolecular components in biological and food systems.
Food. 2000, 44 (5), 299–308.
12. Norton, I.T.; Frith, W.J. Microstructure design in mixed biopolymer composites. Food Hydro-
colloid. 2001, 15 (4–6), 543–553.
13. de Kruif, D.G.; Tuinier, R. Polysaccharide protein interactions. Food Hydrocolloid. 2001, 15
(4–6), 555–563.
14. DeMars, L.L.; Ziegler, G.R. Texture and structure of gelatin/pectin-based gummy confections.
Food Hydrocolloid. 2001, 15 (4–6), 643–653.
15. Butler, M.F. Mechanism and Kinetics of Phase Separation in a Gelatin/Maltodextrin Mixture
Studied by Small-Angle Light Scattering. Biomacromol. 2002, 3 (4), 676–683.
16. Zasypkin, D.V.; Braudo, E.E.; Tolstoguzov, V.B. Multicomponent biopolymer gels. Food
Hydrocolloid. 1997, 11 (2), 159–170.
17. Grinberg, V.Y.; Tolstoguzov, V.B. Thermodynamic incompatibility of proteins and polysaccha-
rides in solutions. Food Hydrocolloid. 1997, 11 (2), 145–158.
18. Cuppo, F.; Venuti, M.; Cesàro, A. Kinetics of gelatin transitions with phase separation: T-jump
and step-wise DSC study. Int. J. Bio. Macromolecules. 2001, 28 (4), 331–341.
19. Lorén, N.; Hermansson, A.-M. Phase separation and gel formation in kinetically trapped gelatin/
maltodextrin gels. Int. J. Bio. Macromol. 2000, 27 (4), 249–262.
20. Wolf, B.; Scirocco, R.; Frith, W.J.; Norton, I.T. Shear-induced anisotropic microstructure in
phase-separated biopolymer mixtures. Food Hydrocolloid. 2000, 14 (3), 217–225.
21. Abdulmola, N.A.; Hember, M.W.N.; Richardson, R.K.; Morris, E.R. Application of polymer
blending laws to starch-gelatin composites. Carbohyd. Polym. 1996, 31 (1–2), 53–63.
22. Clark, A.H. The application of network theory to food systems. In Food Structure and Behav-
iour; Blanshard, J.; Lillford, P., Eds.; Food Science and Technology, Academic Press Limited:
London, 1987; 13–34.
23. Kasapis, S.; Al-Marhoobi, I.M.; Deszczynski, M.; Mitchell, J.R.; Abeysekera, R. Gelatin vs
Polysaccharide in Mixture with Sugar Biomacromolecules. 2003, 4 (1), 1142–1149
24. Morris, V.J. Food gels – roles played by polysaccharides. Chem. Ind. (London). 1985, (5), 159–164.
25. Kasapis, S.; Mitchell, J.; Abeysekera, R.; MacNaughtan, W. Rubber-to-glass transitions in high
sugar-biopolymer mixtures. Trends Food Sci. Tech. 2004, 15 (6), 293–304.
26. Clark, A. Kinetics of demixing. In Biopolymer Mixtures; Harding, S.; Hill, S.; Mitchell, J.; Eds.;
Nottingham University Press: Nottingham, 1995; 37–64.
27. Danner, R. Handbook of polymer solution thermodynamics; American Institute of Chemical
Engineers: New York, 1993; 184 pp.
28. Groves, K. Microscopy Techniques for Confectionery Gels. Manuf. Confect. 2003, 83 (9),
110–114.
29. Ong, M.H.; Whitehouse, A.S.; Abeysekera, R.; Al-Ruqaie, I.M.; Kasapis, S. Glass transition-
related or crystalline forms in the structural properties of gelatin/oxidised starch/glucose syrup
mixtures. Food Hydrocoll. 1998, 12 (3), 273–281.
CONFECTIONERY GELS FORMULATION 205

30. Nickerson, M.; Paulson, A.; Speers, R. Time-temperature studies of gellan polysaccharide gela-
tion in the presence of low, intermediate and hgh levels of co-solutes. Food Hydrocolloid. 2004,
18 (5), 783–794.
31. Lees, R. Faults, causes and remedies: in sweet and chocolate manufacture; Specialised Publica-
tions: Surbiton, England, 1980; 384 pp.
32. Belitz, H.-D.; Grosch, W. Food Chemistry; Springer Verlag: Berlin, New York, 1999; 774 pp.
33. Cakebread, S. Sugar and Chocolate Confectionery; Oxford University Press: London, 1975; 60 pp.
34. Richards, G.N. Initial steps in thermal degradation of sucrose. Int. Sugar J. 1986, 88 (1052),
145–145.
35. Jackson, E.B., ed. Sugar Confectionery Manufacture. 2nd ed., Blackie Academic and Profes-
sional: London. 1995; 400 pp.
36. Minifie, B. Chocolate, Cocoa and Confectionery; Van Nostrand Reinhold: New York, 1989;
904 pp.
Downloaded by [University of Central Florida] at 09:01 17 October 2014

37. Gillies, M.T. Candies and Other Confections; Park Ridge: New Jersey, 1979; 354 pp.
38. Slade, L.; Levine, H. Beyond Water Activity — Recent Advances Based on an Alternative Approach
to the Assessment of Food Quality and Safety. Crit. Rev. Food Sci. 1991, 30 (2–3), 115–360.
39. Cornillon, P.; Andrieu, J.; Duplan, J.-C.; Laurent, M. Use of Nuclear Magnetic Resonance to
Model Thermophysical Properties of Frozen and Unfrozen Model Food Gels. J. Food Eng.
1995, 25 (1), 1–19.
40. Abd Karim, A.; Norziah, M.H.; Seow, C.C. Methods for the study of starch retrogradation. Food
Chem. 2000, 71 (1), 9–36.
41. Kearsley, M.W.; Dziedzic, S.Z., eds. Handbook of Starch Hydrolysis Products and their Deriv-
atives. 1st ed., Blackie Academic and Professional: Glasgow. 1995; 275 pp.
42. Veiga, V.; Ryan, D.H.; Sourty, E.; Llanes, F.; Marchessault, R.H. Formation and characteriza-
tion of superparamagnetic cross-linked high amylose starch. Carbohyd. Polym. 2000, 42 (4),
353–357.
43. Hermansson, A.-M.; Svegmark, K. Developments in the understanding of starch functionality.
Trends Food Sci. Tech. 1996, 7 (11), 343–353.
44. Gaillard, T. Starch: Properties and Potential; Wiley: Chichester, 1987; 151 pp.
45. Rix, A. Gelling and Whipping Agents. In Sugar Confectionery Manufacture; Jackson, E.B., Ed.
Van Nostrand Reinhold: Glasgow; Blackie; New York, 1990; 57–76.
46. Wang, Y.-J.; Truong, V.-D.; Wang, L. Structures and rheological properties of corn starch as
affected by acid hydrolysis. Carbohyd. Polym. 2003, 52 (3), 327–333.
47. Virtanen, T.; Autio, K.; Suortti, T.; Poutanen, K. Heat-induced changes in native and acid-modified
oat starch pastes. J. Cereal Sci. 1993, 17 (2), 137–145.
48. Atichokudomchai, N.; Shobsngobb, S.; Varavinita, S. Morphological properties of acid-modified
tapioca starch. Starch. 2000, 52 (8–9), 283–289.
49. Tester, R.F.; Morrison, W.R. Swelling and gelatinization of Cereal Starches. I. Effects of
Amylopectin, Amylose and Lipids. Cereal Chem. 1990, 67 (6), 551–557.
50. Atkin, N.J.; Abeysekera, R.M.; Cheng, S.L.; Robards, A.W. An experimentally-based predictive
model for the separation of amylopectin subunits during starch gelatinization. Carbohyd. Polym.
1998, 36 (2–3), 173–192.
51. Carnali, J.O.; Zhou, Y. An examination of the composite model for starch gels. J. Rheol. 1996,
40 (2), 221–234.
52. Remsen, C.H.; Clark, J.P. A viscosity model for cooking dough. J. Food Process Eng. 1978, 2
(1), 39–64.
53. Beleia, A.; Miller, R.A.; R.C., H. Starch Gelatinization in Sugar Solutions. Starch. 1996, 48 (7–8),
259–262.
54. Miles, M.; Morris, V.; Orford, P.; Ring, S. The roles of amylose and amylopectin in the gelation
and retrogradation of starch. Carbohyd. Res. 1985, 135 (2), 247–269.
55. Richardson, G.; Langton, M.; Bark, A.; Hermansson, A.-M. Wheat starch gelatinization – the effects
of sucrose, emulsifier and the physical state of the emulsifier. Starch. 2003, 55 (3–4), 150–161.
206 BUREY ET AL.

56. Appelqvist, I.A.M.; Debet, M.R.M. Starch-biopolymer interactions – a review. Food Rev. Int.
1997, 13 (2), 163–224.
57. Takahashi, A.; Yamada, T. Crystalline Copolymer Approach for Melting Behaviour of Starch.
Starch. 1998, 50 (9), 386–389.
58. Tolstoguzov, V. Thermodynamic considerations of starch functionality in foods. Carbohyd.
Polym. 2003, 51 (1), 99–111.
59. Shi, Y.-C.; Seib, P.A. The structure of four waxy starches related to gelatinization and retrogra-
dation. Carbohyd. Res. 1992, 227 131–145.
60. Osunsami, A.T.; Akingbala, J.O.; Oguntimein, G.B. Effect of storage on starch content and
modification of cassava starch. Starch. 1989, 41 (2), 54–57.
61. Kim, R.E.; Ahn, S.Y. Gelling properties of acid-modified red bean starch gels. Agric. Chem.
Biotech. 1996, 39, 49–53.
62. Iizuka, K.; Aishima, T. Starch Gelation Process Observed by FT-IR/ATR Spectrometry with
Downloaded by [University of Central Florida] at 09:01 17 October 2014

Multivariate Data Analysis. J. Food Sci. 1999, 64 (4), 653–658.


63. Fishman, M.L.; Coffin, D.R.; Konstance, R.P.; Onwulata, C.I. Extrusion of pectin/starch blends
plasticized with glycerol. Carbohyd. Polym. 2000, 41 (4), 317–325.
64. Perry, P.A.; Donald, A.M. The effect of sugars on the gelatinisation of starch. Carbohyd. Polym.
2002, 49 (2), 155–165.
65. Donovan, J.W. Phase transitions of the starch water system. Biopolymers. 1979, 18 (2), 263–275.
66. Biliaderis, C. The structure and interactions of starch with food constituents. Can J. Physiol.
Pharm. 1991, 69 (1), 60–78.
67. Farhat, I.A.; Mousia, Z.; Mitchell, J.R. Water redistribution during the recrystallisation of amy-
lopectin in amylopectin/gelatin blends. Polymer. 2001, 42 (10), 4763–4766.
68. Keetels, C.; Oostergetel, G.; van Vliet, T. Recrystallization of amylopectin in concentrated
starch gels. Carbohyd. Polym. 1996, 30 (1), 61–64.
69. Jang, J.K.; Pyun, Y.R. Effect of Moisture Content on the Melting Point of Wheat Starch. Starch.
1996, 48 (2), 48–51.
70. Evageliou, V.; Richardson, R.K.; Morris, E.R. Effect of sucrose, glucose and fructose on gela-
tion of oxidised starch. Carbohyd. Polym. 2000, 42 (3), 261–272.
71. Prokopowich, D.J.; Biliaderis, C.G. A comparative study of the effect of sugars on the thermal
and mechanical properties of concentrated waxy maize, wheat, potato and pea starch gels. Food
Chem. 1995, 52 (3), 255–262.
72. Djabourov, M. Gelation – A Review. Polym. Int. 1991, 25 (3), 135–143.
73. Johnston-Banks, F.A. Gelatine. In Food Gels; Harris, P., Ed. Elsevier Science Publishers Ltd:
1990; 233–289.
74. Normand, V.; Muller, S.; Ravey, J.-C.; Parker, A. Gelation Kinetics of Gelatin: A Master Curve
and Network Modeling. Macromolecules. 2000, 33 (3), 1063–1071.
75. Clark, A.; Ross-Murphy, S. Shear modulus – concentration relationships for biopolymer gels. com-
parison of independent and cooperative crosslink descriptions. In Physical Networks — Polymers
and Gels; Burchard, W.; Ross-Murphy, S.; Eds.; Elsevier Science: London, 1990; 209–230.
76. Veis, A. The Macromolecular Chemistry of Gelatin; Academic Press Ltd.: New York, 1964;
433 pp.
77. Bohidar, H.B.; Jena, S.S. Kinetics of sol-gel transition in thermoreversible gelation of gelatin. J.
Chem. Phys. 1993, 98 (11), 8970–8977.
78. de Carvalho, W.; Djabourov, M. Physical gelation under shear for gelatin gels. Rheol. Acta.
1997, 36 (6), 591–609.
79. Finer, E.G.; Franks, F.; Phillips, M.C.; Sugget, A. Gel formation from solutions of single chain
gelatin. Biopolymers. 1975, 14 (10), 1995–2005.
80. Michon, C.; Cuvelier, G.; Relkin, P.; Launay, B. Influence of thermal history on the stability of
gelatin gels. Int. J. Bio. Macromolecules. 1997, 20 (4), 259–264.
81. Huang, H.; Sorenson, C.M. Shear effects during the gelation of aqueous gelatin. Phys. Rev. E.
1996, 53 (5), 5075–5078.
CONFECTIONERY GELS FORMULATION 207

82. Tosh, S.; Marangoni, A.; Hallet, F.; Britt, I. Aging dynamics in gelatin gel microstructure.
Food Hydrocolloid. 2003, 17 (4), 503–513.
83. Al-Ruqaie, I.M.; Kasapis, S.; Abeysekera, R. Structural properties of pectin-gelatin gels Part
II: Effect of sucrose/glucose syrup. Carbohyd. Polym. 1997, 34 (4), 309–321.
84. Djabourov, M.; Maquet, J.; Theveneau, H.; Leblond, J.; Papon, P. Kinetics of Gelation of
Aqueous Gelatin Solutions. Brit. Polym. J. 1985, 17 (2), 169–174.
85. Christensen, S.H. Pectins. In Food Hydrocolloids; Glicksman, M.; Ed.; CRC Press Inc.: Boca
Raton, 1983; 205–229.
86. May, C.D. Pectins. In Handbook of Hydrocolloids; Phillips, G.O.; Williams, P.A.; Eds.; CRC
Press: Boca Raton, 2000; 169–188.
87. Hoefler, A.C. Hydrocolloids; Eagen Press: Minnesota, 2004; 111 pp.
88. Nussinovitch, A. Hydrocolloid Applications; Chapman and Hall: London, 1997; 354 pp.
89. Willats, W.G.T.; Knox, P.; Mikkelsen, J.D. Pectin: new insights into an old polymer are start-
Downloaded by [University of Central Florida] at 09:01 17 October 2014

ing to gel. Trends Food Sci. Tech. 2006, 17 (3), 97–104.


90. Hino, T.; Ishimoto, H.; Shimabayashi, S. Thermal gelation of aqueous curdlan suspension:
preparation of curdlan jelly. J. Pharm. Pharmacol. 2003, 55 (4), 435–441(7).
91. Gibson, W.; Sanderson, G.R. Gellan Gum. In Thickening and gelling agents for food; Imeson,
A.; Ed.; Blackie Academic & Professional: London, New York 1997; 119–143.
92. Laroche, C.; Michaud, P. New Developments and Prospective Applications for β (1,3) Glucans.
Recent Pat. Biotechnol. 2007, 1, 59–73.
93. Williams, C.T. Chocolate and Confectionery; Leonard Hill: London, 1964; 221 pp.
94. Byars, J.A.; Fanta, G.F.; Felker, F.C. The effect of cooling conditions on jet-cooked normal
corn starch dispersions. Carbohyd. Polym. 2003, 54 (3), 321–326.
95. Kelder, J.D.H.; Ptasinski, K.J.; Kerkhof, P.J.A.M. Starch gelatinization in coiled heaters.
Biotechnol. Progr. 2004, 20 (3), 921–929.
96. Cayot, N.; Bounie, D.; Baussart, H. Dynamic Modelling for a Twin Screw Food Extruder:
Analysis of the Dynamic Behaviour Through Process Variables. J. Food Eng. 1995, 25 (2),
245–260.
97. Rohn, C. Analytical Polymer Rheology – Structure-Processing-Property Relationships;
Hanser: Munich., 1995; 314 pp.
98. Akdogan, H. High moisture food extrusion. Int. J. Food Sci. Tech. 1999, 34 (3), 195–207.
99. Elsey, J.; Riepenhausen, J.; McKay, B.; Barton, G.W.; Willis, M. Modeling and Control of a
Food Extrusion Process. Comput. Chem. Eng. 1997, 21 (Suppl.), s361–s366.
100. Harper, J. Extrusion of Foods, Vol. 1; CRC Press Inc.: Boca Raton., 1981; 212 pp.
101. Berins, M.L., ed. SPI Plastics Engineering handbook of the Society of the Plastics Industry Inc.
5th ed., Chapman & Hall: New York. 1991; 78–132 pp.
102. van Zuilichem, D.J.; Kuiper, E.; Stolp, W.; Jager, T. Mixing effects of constituting elements of
mixing screws in single and twin screw extruder. Powder Technol. 1999, 106 (3), 147–159.
103. Parker, R.; Ollett, A.-L.; Lai-Fook, R.A.; Smith, A.C. The rheology of food ‘melts’ and its applica-
tion to Extrusion Processing. In Rheology of food, pharmaceutical and biological materials with
general rheology; Carter, R.E.; Ed.; Elsevier Applied Science: London, New York, 1990; 57–73.
104. van de Velde, F.; van Riel, J.; Tromp, R.H. Visualisation of starch granule morphologies using
confocal scanning laser microscopy (CLSM) J. Sci. Food Agr. 2002, 82 (1–4), 1528–1536.
105. Barron, C.; Buleon, A.; Colonna, P.; Della Valle, G. Structural modifications of low hydrated
pea starch subjected to high thermomechanical processing. Carbohyd. Polym. 2000, 43 (2),
171–181.
106. Lai, L.S.; Kokini, J.L. Physicochemical changes and rheological properties of starch during
extrusion. Biotechnol. Progr. 1991, 7 (3), 251–266.
107. Wulansari, R.; Mitchell, J.R.; Blanshard, J.M.V. Starch conversion during extrusion as affected
by added gelatin. J. Food Sci. 1999, 64 (6), 1055–1058.
108. McHugh, T.H.; Huxsoll, C.C. Extrusion Processing of Restructured Peach and Peach/Starch
Gels. Food Sci. Technol.-Leb. 1999, 32 (8), 513–520.
208 BUREY ET AL.

109. Collyer, A.A.; Bennett, M.J.; Clegg, D.W.; Elson, C.R.; Jones, S.A. A Feasibility Study on the
Injection Moulding of Fruit Gums at 85% Solids. In Rheology of Food, Pharmaceuticals and
Biological Materials with General Rheology; Carter, R.E.; Ed.; Elsevier Applied Science:
London, New York, 1990; 13–26.
110. Barres, C.; Vergnes, B.; Tayeb, J.; Della valle, G. Transformation of wheat flour by extrusion
cooking – influence of screw configuration and operating conditions Cereal Chem. 1990, 67
(5), 427–433.
111. Aguilera, J.M. Microstructure and Food Product Engineering. Food Technol. 2000, 54 (11),
56–65.
112. Langton, M.; Aström, A.; Hermansson, A.-M. Texture as a reflection of microstructure. Food
Qual. Prefer. 1996, 7 (3–4), 185–191.
113. Lofgren, C.; Walkenström, P.; Hermansson, A.-M. Microstructure and rheological behaviour
of pure and mixed pectin gels. Biomacromol. 2002, 3 (5), 1144–1153.
Downloaded by [University of Central Florida] at 09:01 17 October 2014

114. Olsson, C.; Langton, M.; Hermansson, A.-M. Microstructures of beta-lactoglobulin/amylopectin


gels on different length scales and their significance for rheological properties. Food Hydrocolloid.
2002, 16 (2), 111–126.
115. Kaláb, M.; Allan-Wojtas, P.; Miller, S.S. Microscopy and other imaging techniques in food
structure analysis. Trends Food Sci. Tech. 1995, 6 (6), 177–186.
116. Sun, M.; Griess, G.A.; Serwer, P. Light Microscopy of the Structure of a Gel. J. Struct. Biol.
1994, 113 (1), 56–63.
117. Autio, K.; Laurikainen, T. Relationships between flour/dough microstructure and dough han-
dling and baking properties. Trends Food Sci. Technol. 1997, 8 (6), 181–185.
118. Stokes, D.J.; Donald, A.M. In-situ mechanical testing of dry and hydrated breadcrumb in the
environmental scanning electron microscope (ESEM). J. Mater. Sci. 2000, 35 (3), 599–607.
119. Bhatnagar, S.; Hanna, M.A. Modification of Microstructure of Starch Extruded with Selected
Lipids. Starch. 1997, 49 (1), 12–20.
120. Barron, C.; Bouchet, B.; Della Valle, G.; Gallant, D.G.; Planchot, V. Microscopical Study of the
Destructuring of Waxy Maize and Smooth Pea Starches by Shear and Heat at Low Hydration. J.
Cereal Sci. 2001, 33 (3), 289–300.
121. Hayat, M.A. Stains and Cytochemical Methods; Plenum Press: New York, 1993; 455 pp.
122. Burey (nee Ramesh Sukha), P., Characterisation and Processing of Starch / Sugar / Gelatin
Confectionery Gels, in School of Engineering. The University of Queensland: Brisbane, 2005.
123. Lorén, N.; Altskär, A.; Hermansson, A.-M. Structure Evolution during Gelation at Later Stages
of Spinodal Decomposition in Gelatin/Maltodextrin Mixtures. Macromolecules. 2001, 34 (23),
8117–8128.
124. Alevisopoulos, S.; Kasapis, S. Molecular weight effects on the gelatin/maltodextrin gel. Carbohyd.
Polym. 1999, 40 (1), 83–87.
125. Djabourov, M.; Bonnet, N.; Kaplan, H.; Favard, N.; Favard, P.; Lechaire, J.P.; Maillard, M. 3D
Analysis of gelatin gel networks from transmission electron microscopy imaging. J. Phys. II.
1993, 3 (5), 611–624.
126. Bonell, D., Ed. Scanning tunneling microscopy and spectroscopy : theory, techniques, and
applications. Wiley-VCH: New York. 1993pp.
127. Starostina, N.; West, P., Part II: Sample Preparation for AFM Particle Characterization.
Pacific Nanotechnology: Santa Clara, 2006.
128. Fishman, M.L.; Cooke, P.H.; Chau, H.K.; Coffin, D.R.; Hotchkiss, A.T. Global structures of
high methoxyl pectin from solution and in gels. Biomacomol. 2007, 8 (2), 573–578.
129. Morris, V.J.; Gunning, A.P.; Faulds, C.B.; Williamson, G.; Svensson, B. AFM images of com-
plexes between amylose and Aspergillus niger glucoamylase mutants, native, and mutant
starch binding domains: A model for the action of glucoamylase. Starch. 2005, 57 (1), 1–7.
130. Uricanu, V.I.; Duits, M.H.G.; Nelissen, R.M.F.; Bennink, M.L.; Mellema, J. Local structure
and elasticity of soft gelatin gels studied with atomic force microscopy. Langmuir. 2003, 19
(20), 8182–8194.
CONFECTIONERY GELS FORMULATION 209

131. Benmouna, F.; Johannsmann, D. Viscoelasticity of gelatin surfaces probed by AFM noise anal-
ysis. Langmuir. 2004, 20 (1), 188–193.
132. Stanley, D.; Stone, A.P.; Tung, M.A. Mechanical Properties of Food. In Handbook of Food
Analysis: Physical Characterisation and Nutrient Analysis; Mollet, L.M.L., Ed. Marcel Dekker:
New York, 1996; 93–113.
133. Voisey, P. Modernisation of texture instrumentation. In Theory, Determination and Control of
Physical Properties of Food Materials; Rha, C., Ed. Reidel: Dordrecht, 1975; 63–130.
134. Anand, A.; Scanlon, M.G. Dimensional effects on the prediction of texture-related mechanical
properties of foods by indentation. T. ASAE. 2002, 45 (4), 1045–1050.
135. Teratsubo, M.; Tanaka, Y.; Saeki, S. Measurement of stress and strain during tensile testing of
gellan gum gels: effect of deformation speed. Carbohyd. Polym. 2002, 47 (1), 1–5.
136. Ross-Murphy, S.B. Rheological characterisation of gels. Rheological characterisation of gels.
1995, 26 (4), 391–400.
Downloaded by [University of Central Florida] at 09:01 17 October 2014

137. Olkku, J.E.; Sherman, P. Compression Testing of Cylindrical Samples with an INSTRON
Universal Testing Machine. In Food texture and viscosity : concept and measurement; Bourne,
M.C.; Ed.; Academic Press: New York, 1982; 157–175.
138. Borwankar, R.P. Food Texture and Rheology: A Tutorial Review. J. Food Eng. 1992, 16 (1–2),
1–16.
139. Jackman, R.L.; Stanley, D.W. Perspectives in the textural evaluation of plant foods. Trends
Food Sci. Tech. 1995, 6 (6), 187–194.
140. Jowitt, R. An engineering approach to some aspects of food texture. In Food Texture and
Viscosity: Concept and Measurement; Bourne, M., Ed. Academic Press Inc: New York, 1982;
142–155.
141. Wilkinson, C.; Dijksterhuis, G.B.; Minekus, M. From food structure to texture. Trends Food
Sci. Tech. 2000, 11 (12), 442–450.
142. Bourne, M.C. Food Texture and Viscosity: Concept and Measurement; Academic Press Inc.:
New York, 2002; 427 pp.
143. Bourne, M.C. Calibration of Rheological Techniques Used for Foods. J. Food Eng. 1992, 16
(1–2), 151–163.
144. Bagley, E.B.; Christianson, D.D.; Wolf, W.J. Friction Effects in Compressional Deformation
of Gelatin and Starch Gels and Comparison of Material Response in Simple Shear, Torsion and
Lubricated Uniaxial Compression. J. Rheol. 1985, 29 (1), 103–108.
145. Friedman, H.H.; Whitney, J.E.; Sczczesniak, A.S. The Texturometer — A new instrument for
objective texture measurement. J. Food Sci. 1963, 28 (4), 390–396.
146. Antoniou, K.D.; Petridis, D.; Raphaelides, S.; Ben Omar, Z.; Kesteloot, R. Texture Assessment
of French Cheeses. J. Food Sci. 2000, 65 (1), 168–173.
147. Mao, R.; Tang, J.; Swanson, B.G. Texture properties of high and low acyl mixed gellan gels.
Carboyhyd. Polym. 2000, 41 (4), 331–338.
148. Moritaka, H.; Naito, S.; Nishinari, K.; Ishihara, M.; Fukuba, H. Effects of gellan gum, citric
acid and sweetener on the texture of lemon jelly. J. Texture Stud. 1999, 30 (1), 29–41.
149. Bourne, M.C. Effect of Water Activity on Texture Profile Parameters of Apple Flesh. J. Textur.
Stud. 1986, 17 331–340.
150. Olkku, J.; Rha, C.K. Textural Parameters of Candy Licorice. J. Food Sci. 1975, 40 (5), 1050–1054.
151. Hough, G.; Califano, A.N.; Bertola, N.C.; Bevilacqua, A.E.; Martinez, E.; Jose Vega, M.;
Zaritzky, N.E. Partial least squares correlations between sensory and instrumental measure-
ments of flavor and texture for reggianito grating cheese. Food Qual. Pref. 1996, 7 (1), 47–53.
152. Ronn, B.B.; Hyldig, G.; Wienberg, L.; Qvist, K.B.; Laustsen, A.M. Predicting sensory properties
from rheological measurements of low-fat spreads. Food Qual. Pref. 1998, 9 (4), 187–196.
153. Rao, M.A., ed. Rheology of Fluid and Semisolid Foods — Principles and applications. Aspen
Publishers. Inc.: Maryland. 1999; 433 pp.
154. Larson, R.G. The Structure and Rheology of Complex Fluids; Oxford University Press, Inc.:
New York, 1999; 663 pp.
210 BUREY ET AL.

155. Menard, K.P. Dynamic Mechanical Analysis: A Practical Introduction; CRC Press: Boca
Raton, 1999; 208 pp.
156. Rezler, R.; Baranowska, H.M.; Surma, S.; Poliszko, S. Investigation of starch gels by means of
the relaxation method. Zywnosc. 2000, 2 (23 Supp), 194–203.
157. Gregson, C.M.; Hill, S.E.; Mitchell, J.R.; Smewing, J. Measurement of the Rheology of
Polysaccharide Gels by Penetration. Carbohyd. Polym. 1999, 38 (3), 255–259.
158. Mousia, Z.F., I.A.; Blachot, J.F.; Mitchell, J.R. Effect of water partitioning on the glass-transition
behaviour of phase separated amylopectin-gelatin mixtures. Polym. 2000, 41 (5), 1841–1848.
159. Steffe, J.F.; Morgan, R.G. On-line measurement of dynamic rheological properties during food
extrusion. J. Food. Proc. Eng. 1987, 10 (1), 21–26.
160. Langley, K.R.; Green, M.L.; Brooker, B.E. Mechanical properties and structure of model com-
posite foods. In Food Polymers, Gels and Colloids; Dickinson, E., Ed. The Royal Society of
Chemistry: Cambridge, 1991; 383–391.
Downloaded by [University of Central Florida] at 09:01 17 October 2014

161. Maurer, K.; Hardy, J. Rheological and sensory characterization of viscous and pasty food products:
application of fractal concepts and fourier analysis. J. Text. Stud. 1996, 27 (1), 41–59.
162. Morikawa, K.; Nishinari, K. Rheological and DSC studies of gelatinization of chemically mod-
ified starch heated at various temperatures. Carbohyd. Polym. 2000, 43 (3), 241–247.
163. Noël, Y.; Zannoni, M.; Hunter, E.A. Texture of Parmigiano Reggiano cheese: Statistical rela-
tionships between rheological and sensory variates. Lait. 1996, 76 (3), 243–254.
164. Taneya, S.; Izutsu, T.; Kimura, T.; Shioya, T.; Buchheim, W.; Kindstedt, P.S.; Pechak, D.G.
Structure and Rheology of String Cheese. Food Struct. 1992, 11 (1), 61–71.
165. Vega, C.; Goff, H. Phase separation in soft-serve ice cream mixes: rheology and microstruc-
ture. Int. Dairy J. 2005, 15 (33), 249–254.
166. Tolstoguzov, F.B. The importance of glassy biopolymer components in food. Food. 2000, 44
(2), 76–84.
167. de Bont, P.W.; Hendriks, C.L.L.; van Kempen, G.M.P.; Vreeker, R. Time evolution of phase
separating milk protein and amylopectin mixtures. Food Hydrocolloid. 2004, 8 (6), 1023–1031.
168. Mackley, M.R.; Marshall, R.T.J.; Smeulders, J.B.A.F. The multipass rheometer. J.Rheol. 1995,
39 (6), 1293–1309.
169. Páques, M.; Tromp, R.H.; van Aken, G.A. Microrheology, a New Approach in the Study of
Structure-/Function-relationships in Food Systems. Scanning. 2000, 22 (2), 114–115.
170. Nicolas, Y.; Paques, M.; van den Ende, D.; Dhont, J.K.G.; van Polanen, R.; Knaebel, A.;
Steyer, A.; Munch, J.-P. Microrheology: new methods to approach the functional properties of
food. Food Hydrocolloid. 2003, 17 (6), 907–913.
171. Stokes, J.R.; Wolf, B.; Frith, W.J. Phase-separated biopolymer mixture rheology: Prediction
using a viscoelastic emulsion model. J. Rheol. 2001, 45 (5), 1173–1191.
172. Marcotte, M.; Hoshahili, A.R.T.; Ramaswamy, H.R. Rheological properties of selected hydro-
colloids as a function of concentration and temperature. Food Res. Int. 2001, 34 (8), 695–703.
173. Williams, P.A.; Ahmad, F.B. Effect of dextran on the thermal and rheological properties of
starch. In Hydrocollids Part. 2. Fundamentals and applications in food, biology, and medicine;
Nishinari, K.; Ed.; Elsevier Science: Amsterdam, 2000; 165–176.
174. Walkenström, P.; Hermansson, A.-M. Effects of shear on pure and mixed gels of gelatin and
particulate whey protein. Food Hydrocolloid. 1998, 12 (1), 77–87.
175. Marrs, W.M. Gelatin carbohydrate interactions and their effect on the structure and texture of
confectionery gels. Progr. Food Nutr. Sci. 1982, 6 (1–6), 259–268.
176. te Nijenhuis, K.T. Thermoreversible Networks — Viscoelastic properties and structure of gels.
Adv. Polym. Sci. 1997, 130, 1–12.

You might also like