Sei sulla pagina 1di 15

Food Hydrocolloids 17 (2003) 2539

www.elsevier.com/locate/foodhyd

Review

Hydrocolloids at interfaces and the inuence on the properties of


dispersed systems q
Eric Dickinson*
Procter Department of Food Science, University of Leeds, Leeds LS2 9JT, UK
Received 20 August 2001; revised 29 October 2001; accepted 16 November 2001

Abstract
Although traditionally associated with thickening and gelation behaviour, food hydrocolloids also inuence the properties of dispersed
systems through their interfacial properties. Hence, surface-active hydrocolloids may act as emulsiers and emulsion stabilisers through
adsorption of protective layers at oilwater interfaces, and interactions of hydrocolloids with emulsion droplets may affect rheology and
stability with respect to aggregation and serum separation. A review of literature evidence suggests that much of the reported emulsifying
capability of polysaccharides is explicable in terms of complexation or contamination with a small fraction of surface-active protein. To
support this point of view, the specic cases of gum arabic, galactomannans and pectin are considered in some detail. In mixed protein 1
polysaccharide systems, associative electrostatic interactions can lead to coacervation or soluble complex formation depending on the nature
of the biopolymers and the solution conditions (pH and ionic strength). Proteinhydrocolloid complexation at interfaces can be associated
with bridging occulation or steric stabilisation. As well as controlling rheology, the presence of a non-adsorbing hydrocolloid can affect
creaming stability by inducing depletion occulation.
q 2002 Elsevier Science Ltd. All rights reserved.
Keywords: Hydrocolloids; Biopolymer emulsiers; Surface activity; Emulsion stability; Proteinhydrocolloid interactions; Flocculation

1. Introduction
Food hydrocolloids are high-molecular-weight hydrophilic biopolymers used as functional ingredients in the food
industry for the control of microstructure, texture, avour
and shelf-life. The term `hydrocolloid' embraces all the
many polysaccharides that are extracted from plants,
seaweeds and microbial sources, as well as gums derived
from plant exudates, and modied biopolymers made by the
chemical or enzymatic treatment of starch or cellulose. In
addition, due to its polydisperse and highly hydrophilic
character, the unique protein gelatin has become accepted
as an exceptional member of this polysaccharide club. But
other food proteins, like casein and gluten, are traditionally
not classied as hydrocolloidseven though they certainly
do exhibit functional properties that overlap considerably
with those of the food polysaccharides.
The general molecular and functional properties of
q
Based on the author's presentation at the Masterclass on `Measuring
Performance' at the 11th Gums and Stabilisers for the Food Industry
Conference (Wrexham, UK, 26 July 2001).
* Tel.: 144-1132-332956; fax: 144-1132-332982.
E-mail address: e.dickinson@leeds.ac.uk (E. Dickinson).

proteins and polysaccharides are compared in Table 1.


Both classes of biopolymer contribute to the structural and
textural properties of foods through their aggregation and
gelation behaviour. Furthermore, proteins are known for
their emulsication and foaming properties, and polysaccharides for their water-holding and thickening properties.
In reality, of course, Table 1 is a gross simplication. Large
differences in functional properties exist between the
various food biopolymers depending on the detailed chemical structure and sensitivity to solution conditions (pH, ionic
strength, specic ions). The functionality of an individual
biopolymer in foods is also affected by its interactions with
other food componentsproteins, polysaccharides, lipids,
sugars, salts, and so forth.
Food colloids are multi-phase food systems containing
particles or other structures with characteristic spatial
dimensions in the colloidal size range (Dickinson, 1992).
The term `colloid' can be applied to particulate dispersions,
foams, gels and emulsions (oil-in-water, water-in-oil, waterin-water). Therefore, most manufactured foodstuffs can be
classied as food colloids, and many contain hydrocolloid
ingredients added by the manufacturer for control of stability and rheological properties. An important class of food
colloids are oil-in-water emulsions such as salad dressing,

0268-005X/03/$ - see front matter q 2002 Elsevier Science Ltd. All rights reserved.
PII: S 0268-005 X(01)00 120-5

26

E. Dickinson / Food Hydrocolloids 17 (2003) 2539

Table 1
The common general characteristics and the differences between proteins
and polysaccharides as functional biopolymers in food systems
Similarities
Natural polymers
Widespread in food colloids
Used in pharmaceuticals, cosmetics, personal products
Environmentally friendly polymers
Complicated structure
Complex aggregation behaviour
Gelling/stabilising agents
Differences
Proteins
Wide-ranging structures
Reactive
Monodisperse
Many segment types
Linear chain
Flexible chain
Medium molecular weight
Small molecular volume
Amphiphilic
Surface-active
Polyelectrolyte
Emulsifying/foaming
Temperature sensitive a
Strong surfactant binding

Polysaccharides
Similar structures
Unreactive
Polydisperse
Few segment types
Linear or branched
Stiff chain
High molecular weight
Large molecular volume
Hydrophilic
Not surface-active
Non-ionic or charged
Thickening/waterholding
Temperature insensitive
Weak surfactant binding

That is, the structure and properties of most proteins can change drastically when heated above a characteristic `denaturation temperature'.

Table 2
Principal factors affecting oil-in-water emulsion stability
Droplet-size distribution
Initially determined by
Emulsication equipment
Concentration of emulsier
Type of emulsier
Oil/water ratio
Other factors (temperature, pH, viscosity)
Nature of interfacial adsorbed layer
Determined by
Concentration and type of emulsier
Interactions of adsorbed species
Competition between adsorbed species
Nature of continuous aqueous phase
Rheology, solvent quality, ionic environment, unadsorbed polymers
and amphiphiles
Nature of dispersed oil phase
Solid/liquid content
Solubility in continuous phase

ice-cream, cream liqueurs or soft drinks. The primary stabilising mechanism may occur in the bulk aqueous phase or at
the surface of the droplets, depending on the chemical
nature of the particular ingredient(s) involved (Lips,
Campbell, & Pelan, 1991). In this article, we consider the
extent to which food hydrocolloids have a role in conferring
stability on dispersions and emulsions through interfacial
action.
In the formulation of emulsion systems, one normally
distinguishes between two types of ingredient: the `emulsifying agent' (or `emulsier') and the `stabiliser' (Dickinson,
1992). The emulsifying agent is the single chemical species
(or mixture of species) that promotes emulsion formation
and short-term stabilisation by interfacial action. The term
`bioemulsier' (or `biosurfactant') is also commonly used in
the eld of biotechnology and applied microbiology
(Navon-Venezia et al., 1995; Ron & Rosenberg, 2001;
Rosenberg & Ron, 1999), although it is rarely used amongst
food technologists (Shepherd, Rockey, Sutherland, &
Roller, 1995).
There are two broad classes of emulsifying agents used in
food processingsmall-molecule surfactants (monoglycerides, polysorbates, sucrose esters, lecithin, etc.) and macromolecular emulsiers (usually proteins, especially from
milk and eggs). A small-molecule surfactant is a distinctly
amphiphilic molecule having both polar and non-polar
parts. For reasons of tradition and marketing, proteins are
not commonly classied as emulsiers, even though they do
often operate as such during food manufacture. So, in practice, the term `emulsier' in the food industry is conned
almost exclusively to low-molecular-weight amphiphiles.
Confusingly, the term also includes those small molecules
that affect texture or shelf-life in ways other than through
emulsication, e.g. by modifying fat crystallisation or by
interacting with starch.
To confer substantial shelf-life we require the presence of
a stabiliser. This can be dened (Dickinson, 1992) as a
single chemical component (or mixture) conferring long-term
emulsion stability, possibly by an adsorption mechanism,
but not necessarily so. Stabilisers are normally biopolymersproteins or polysaccharides; small-molecule
surfactants are not so effective in conferring long-term
stability. The main stabilising action of food polysaccharides is via viscosity modication or gelation in the aqueous
continuous phase. Proteins, on the other hand, have a strong
tendency to adsorb at oilwater interfaces to form stabilising layers around oil droplets, and so they are able to full
both the emulsifying and the stabilising roles. Emulsions
can also be stabilised by particles (e.g. casein micelles and
fat crystals).
A stable emulsion is one with no discernible change in the
size distribution of the droplets, or their state of aggregation,
or their spatial arrangement within the sample vessel, over
the time-scale of observation. This time-scale may vary
from hours to months depending on the situation. The
dominant mechanisms of instability are gravity creaming,

E. Dickinson / Food Hydrocolloids 17 (2003) 2539

Ostwald ripening, occulation and droplet coalescence


(Dickinson, 1992). Table 2 gives the main factors affecting
stability. The state of occulation of the droplets is dependent on the interactions between stabilising layers, which in
turn depends on factors such as the biopolymer surface
coverage, the layer thickness, the surface charge density,
and the aqueous solution conditions (especially pH, ionic
strength, and divalent ion content). For a freshly prepared
ne triglyceride oil-in-water emulsion, the most obvious
initial manifestation of instability is creaming, which typically leads on to macroscopic phase separation into separate
discernible regions of cream and serum. This may then be
followed by droplet coalescence within the cream and
`oiling off' at the top of the sample. All of these processes
can be affected by interactions of the hydrocolloid stabiliser,
both in the bulk aqueous phase and at the surface of the
emulsion droplets.
2. What makes a good emulsifying agent or a good
emulsion stabilising agent?
For a polymer to be effective as an emulsifying agent, it
must be surface-active. That is, it must have the capacity to
lower the tension at the oilwater interface, both substantially and rapidly when present at the concentrations typically used during emulsication. Generally speaking, the
lower the interfacial tension, the greater the extent to
which droplets can be broken up during intense shearing
or turbulent ow (Walstra, 1983; Walstra & Smulders,
1998). During high-pressure homogenisation, for instance,
most of the processes occur on time-scales of milliseconds
or less: droplet deformation, emulsier adsorption, emulsier spreading, and droplet collision. To retain small droplets
during emulsication, the time between droplet collisions
should be long compared with the time for emulsier to
adsorb at the new oilwater interface and to create a transient stabilising layer (not necessarily with full saturation
coverage). Stabilisation of ne oil droplets during emulsication can be ascribed predominantly to the GibbsMarangoni
mechanism (Dickinson, 1994), which to be effective requires
adsorption of a water-soluble emulsier at a sufcient surface
concentration to generate substantial interfacial tension
gradients during close approach of colliding droplets.
For a biopolymer to be surface-active, it clearly must
have amphiphilic character. So, if it is a hydrocolloid, it
must contain hydrophobic groups that are numerous enough
and sufciently accessible on a short timescale to enable the
adsorbing molecules to adhere to and spread out at the interface, thereby protecting the newly formed droplets. The
required time may be too long for very large macromolecules, or species that are strongly aggregated in the bulk
aqueous phase. Hence, an ideal emulsifying agent, capable
of making small droplets, is typically composed of species
of relatively low molecular mass with good solubility in the
aqueous continuous phase (e.g. a water-soluble surfactant of

27

high HLB number). The limited emulsifying capacity of


some biopolymers can be attributed to poor solubility and/
or insufcient amphiphilic character to produce rapid and
substantial lowering of the interfacial tension during droplet
break-up. The well-established improvement in emulsifying
properties of a protein like gluten by degradative enzyme
treatment (Larre, Huchet, Berot, & Popineau, 2001) can
therefore be attributed to the improved solubility and greater
surfactant-type character of the derived peptide fragments.
Irrespective of the degree of amphiphilic character, though,
the large molecular mass of a typical hydrocolloid makes it
an unlikely candidate for an ingredient of the very highest
emulsifying ability.
One distinctly identiable group of surface-active polysaccharides are the hydrophobically modied cellulose
derivativesnotably methylcellulose and hydroxypropyl
methylcellulose. Such amphiphilic biopolymers can certainly be used as emulsiers in their own right to prepare
stable soybean oil-in-water emulsions (Gaonkar, 1991). But
the droplets produced are much coarser than those made
with low-molecular-mass emulsiers or proteins under
similar conditions (Darling & Birkett, 1987). This poorer
emulsifying performance is attributable to the relatively
high molecular weight of these modied cellulose polymers.
Similar arguments can be applied in relation to the emulsifying performance of hydrophobically modied starches
and polypropylene glycol alginate.
The reliability of the published literature on the apparent
surface activity of certain hydrocolloids is undermined
somewhat by the fact that many commercial gum samples
contain a small amount of protein, either as contaminant or
as an intrinsic part of the molecular structure. As this proteinaceous material is typically strongly hydrophobic, it can
adsorb strongly at liquid interfaces, thereby giving an erroneous impression of the intrinsic surface activity of the
polysaccharide hydrocolloid itself. So, for instance, the
stabilising function of xanthan gum in mayonnaise has
been attributed (Hennock, Rahalkar, & Richmond, 1984)
to its acting as an emulsier by co-adsorbing with egg
yolk at the oilwater interface. This interpretation was
argued on the basis that a 1 wt% solution of the gum had
been found (Prud'homme & Long, 1983) to lower the
tension at the airwater surface by 30 mM m 21 or more!
However, it is likely that the gum sample used in the aforementioned research contained as much as 5% of contaminating protein (Anderson, 1986). And it is not only in the `old
literature' that one may nd disconcertingly confusing
information. Reports of highly surface-active xanthan
samples still persist in some fairly recent publications
(Garti, Slavin, & Aserin, 1999).
Once an emulsion of small droplets has been successfully
prepared, considerations of surface activity or interfacial
tension gradients are no longer relevant. What matters for
subsequent long-term stability is how well the molecular
characteristics of the adsorbed biopolymer conform to the
requirements of producing a robust macromolecular barrier

28

E. Dickinson / Food Hydrocolloids 17 (2003) 2539

at the interface. The physico-chemical processes involved in


preventing droplets from aggregating or coalescing are the
classical colloidal stability mechanisms of steric stabilisation and electrostatic stabilisation (Dickinson, 1988a,b,
1992). For a biopolymer to be most effective in stabilising
dispersed particles or emulsion droplets, it should exhibit
the following four characteristics:
(i) Strong adsorption. This implies that the amphiphilic
polymer has a substantial degree of hydrophobic character (e.g. non-polar side chains or a peptide/protein
moiety) to keep it permanently anchored to the interface.
(ii) Complete surface coverage. This implies that there is
sufcient polymer present to fully saturate the surface.
(iii) Formation of a thick steric stabilising layer. This
implies that the polymer is predominantly hydrophilic
and of high molecular weight (10 4 10 6 Da) within an
aqueous medium with good solvent properties.
(iv) Formation of a charged stabilising layer. This
implies the presence of charged groups on the polymer
that contribute to the net repulsive electrostatic interaction between particle surfaces, especially at low ionic
strength. (This characteristic is only essential if the polymer layer is not sufciently thick.).
Conditions (i)(iii) cannot be met simultaneously for a
homopolymer in which all the chain segments are chemically similar, and so such uniform polymers do not make
good steric stabilisers (Dickinson, 1988a). What is needed
for good steric stabilisation is a block copolymer, composed
of a small fraction of strongly adsorbing hydrophobic
segments (to keep the macromolecule permanently attached
to the surface) and a large fraction of non-adsorbing hydrophilic segments (to stick it away from the surface and confer
upon the protective layer some appreciable thickness). Such
steric stabilisation, combined with electrostatic stabilisation
(i.e. condition (iv)) if the macromolecule contains ionisable
groups, provides a repulsive energy barrier which prevents
pairs of particles from becoming strongly stuck together at
close separations under the strongly attractive inuence of
ubiquitous van der Waals forces (Dickinson, 1992).
So, let us turn to four basic questions that lie at the heart
of the subject of this article.
Question 1: Does the hydrophilic character of polysaccharides mean that they have little surface activity at oil
water (or airwater) interfaces, and so are not useful as
emulsifying agents? Answer: Yes.
Question 2: Can one say then that hydrocolloids do not
adsorb at surfaces in foods? Answer: Yes, it is true that
hydrocolloids do not normally adsorb at hydrophobic surfacesbut with some exceptions.
Question 3: What are the `exceptions'? Answer: For
instance
gelatin (a protein!).
gum arabic (contains protein!).

chemically modied starch or cellulose derivatives (e.g.


highly substituted methyl celluloses).
some naturally occurring galactomannan hydrocolloids
(e.g. guar gum, fenugreek gum).
acetylated pectin from sugar beet.
depolymerised citrus pectin.
possibly any hydrocolloid adsorbing at a hydrophilic (or
amphiphilic) surface (e.g. at an existing adsorbed protein
layer).
Question 4: Can the reported surface activity and emulsifying capability of some carbohydrate polymers be attributed mainly to the presence of protein, present either as a
contaminant, or as a covalently linked (or physically associated) proteinpolysaccharide complex? Answer: In most
cases, probably yes.
In order to justify the bare answers given above, it seems
appropriate to discuss a few of the exceptions in more detail.
We shall begin with the most well known of all the food
hydrocolloid emulsiersgum arabic.
3. Some hydrocolloids with surface activity and
emulsication properties
3.1. Gum arabic
The most commonly recognised hydrocolloid emulsier
is gum arabic. It is widely used in the soft drinks industry for
emulsifying avour oils (e.g. orange oil) under acidic conditions. About three-quarters of the gum production comes
from the species Acacia senegal, with Acacia seyal providing most of the rest (Thevenet, 1988). Gum arabic is the
established benchmark hydrocolloid emulsier, and, because
it is a fairly expensive ingredient, there have been many
suggestions in the literature for its replacement by other
polymeric emulsierslike hydrophobically modied
starch (Trubiano, 1995) or other gums, e.g. mesquite gum
(Vernon-Carter, 1998).
The well-known lm-forming ability of gum arabic has
long since unambiguously established that this is indeed a
genuine emulsier that confers functionality not by modifying the rheology of the aqueous phase but by leading to
formation of a macromolecular stabilising layer around oil
droplets. Hence, the gum can stabilise the avour oil emulsion both as a concentrate and in the form of a diluted
(carbonated) beverage (Tan, 1990). Nevertheless, the level
of surface activity is actually rather low in comparison with
typical food protein emulsiers. To compensate for this in
generating stable sub-micron-sized droplets, it is necessary
in practice (McNamee, O'Riordan, & O'Sullivan, 1998) to
use a rather high gum-to-oil weight ratio, i.e. approximately
1:1, as compared with 1:10 for equivalent protein-stabilised
emulsions. Once formed by adsorption at a macroscopic
oilwater interface, the high surface shear viscosity of the
gum arabic lm is little affected by extensive dilution of the

E. Dickinson / Food Hydrocolloids 17 (2003) 2539

Fig. 1. Schematic representation of the wattle blossom model representing


the functionally active component of A. senegal gum (a) in solution and (b)
adsorbed at the oilwater interface. Separate hydrophilic carbohydrate
blocks (C) of molecular mass ca. 2 105 Da are attached to the backbone
chain of hydrophobic protein (P).

aqueous sub-phase (Dickinson, Elverson, & Murray, 1989).


The lm viscoelasticity is therefore maintained even when a
major proportion of the hydrocolloid has been removed
from the aqueous phase in contact with the adsorbed
layer. This is consistent with the observation that only a
small proportion of gum arabic used in emulsion preparation
is actually involved in the stabilisation process.
Gum arabic (A. senegal) is a complex branched heteropolyelectrolyte with a backbone of 1,3-linked b-galactopyranose units and side-chains of 1,6-linked galactopyranose
units terminating in glucuronic acid or 4-O-methylglucuronic acid residues. It also contains ca. 2% protein, covalently
linked to carbohydrate through serine and hydroxyproline
residues, resulting in a mixture of arabinogalactanprotein
complexes, each containing several polysaccharide units
linked to a common protein core. This so-called `wattle
blossom' model (Connolly, Fenyo, & Vandevelde, 1988;
Fincher, Stone, & Clark, 1983) is shown schematically in
Fig. 1(a). It has been demonstrated that the gum is a
complex mixture of at least three distinct fractions with
different chemical structures (Williams, Phillips, & Randall,
1990) with a major component containing little or no proteinaceous material. The protein element appears to be
mainly associated with a high-molecular-mass fraction
representing less than 30% of the total gum (Vandevelde
& Fenyo, 1985). It is this fraction that appears to be predominantly responsible for the special emulsifying and
stabilising properties of the hydrocolloid (Randall, Phillips,
& Williams, 1988; Ray, Bird, Iacobucci, & Clark, 1995).
Such heterogeneity of macromolecular structure and functionality is by no means unique to gum arabic. Molecular
weight fractions with differing emulsication properties

29

have also been identied for other gums, e.g. gum talha
(Underwood & Cheetham, 1994).
Fig. 1(b) shows the high-molecular-mass protein-rich fraction of A. senegal gum adsorbing at the oilwater interface
according to the wattle blossom representation (Islam,
Phillips, Sljivo, Snowden, & Williams, 1997; Randall,
Phillips, & Williams, 1989). The more hydrophobic protein
chain rmly anchors the proteinpolysaccharide hybrid at the
interface, and the protruding hydrophilic carbohydrate blocks
attached to this chain provide a strong steric barrier towards
occulation and coalescence. Whilst charged groups provide
the basis for some electrostatic contribution to the colloidal
stabilisation, the relatively low value of the (negative) zeta
potential, 1020 mV under beverage emulsion conditions
(Jayme, Dunstan, & Gee, 1999; Ray et al., 1995), indicates
that the primary stabilisation mechanism is steric in character.
Reports of the surface activity of various Acacia gums with
nitrogen contents in the range 0.17.5% (Dickinson, Murray,
Stainsby, & Anderson, 1988) suggest a good correlation
between the Acacia gum protein content and the limiting
long-time interfacial tension, and also between emulsifying
capacity and the initial rate of change of tension with time.
Consistent with this generalisation, A. senegal lms have been
found to be more elastic and of higher stability than those
composed of A. seyal (Fauconnier et al., 2000). Notwithstanding that, nitrogen content alone is really no reliable indicator
of the effectiveness of Acacia gums for emulsication. Gum
arabic (A. senegal) samples of similar nitrogen content
(,0.3%, i.e. corresponding to ,2% protein) have been
found (Dickinson, Galazka, & Anderson, 1991a) to exhibit
substantial differences in emulsifying capacity and emulsion
stability. Furthermore, some samples of A. seyal, having a
considerably lower protein content (,0.8%), have actually
been found to give better emulsion stability than A. senegal
(Buffo, Reineccius, & Oehlert, 2001).
On the other hand, there does appear to be a good correlation between emulsion stability and gum arabic average
molecular weight. It was observed (Dickinson, Galazka, &
Anderson, 1991b) that the 10% fraction of a gum arabic
sample corresponding to the highest molecular weight
(0.38% N), as separated by gel permeation chromatography,
could produce a more stable emulsion than the remaining
90% fraction (0.35% N). Additionally, in separate experiments, when the average molecular mass of a gum arabic
sample (,0.35% N) was gradually reduced from 3:1 105
to 2:2 105 Da by controlled irradiative degradation, the
resulting emulsion stability was correspondingly reduced
(Dickinson, Galazka, & Anderson, 1991c). This trend is
consistent with the formation of thicker and more effective
steric stabilising layers by adsorption of polymers of greater
molecular mass.
3.2. Galactomannans
A galactomannan is a rather rigid hydrophilic biopolymer
with a polymannose backbone and grafted galactose units.

30

E. Dickinson / Food Hydrocolloids 17 (2003) 2539

Fig. 2. Comparison of the emulsion stabilising ability of depolymerised


citrus pectin fractions (70% DE) with that of gum arabic (GA) (Akhtar et
al., 2002). Fine emulsions were prepared by high-pressure homogenisation
of 10% rapeseed oil and 4 wt% hydrocolloid at pH 4.7 and ionic strength
0.2 M. Percentage serum separation is shown for four pectin samples with
different degrees of depolymerisation (different values of average molecular mass, in Da) after 1 month (A), 2 months and 3 months (B).

Galactomannans such as guar and locust bean gum are


widely used in the food industry as thickening, waterholding and stabilising agents. As the carbohydrate structure
gives no suggestion of the presence of any signicant
proportion of hydrophobic groups, it is generally assumed
that this type of hydrocolloid functions by modifying the
rheological properties of the aqueous phase between the
dispersed particles or droplets.
Puried guar or locust bean gum is reported not to exhibit
any surface activity (Gaonkar, 1991). Nevertheless, there
are some statements in the literature (Garti, 1999, 2001;
Garti & Reichman, 1993, 1994) that these polymers have
the capacity to emulsify oils and stabilise fairly coarse emulsions (,10 mm mean droplet size, 34 mg m 22 surface
coverage) at a moderately low gum/oil ratio (say, 1:5). It
has been suggested by Garti and Reichman (1994), based on
light microscopy observations of strong birefringency at the
oilwater interface, that these gums stabilise emulsions by
forming liquid crystalline layers around the droplets. We are
reminded here of the suggestion of Friberg (1971) concerning the stabilising role of small-molecule emulsiers by
multi-layer lamellar liquid crystals. Nevertheless, it has
been commented upon before (Bergenstahl, 1997; Dickinson,
1986) that the applicability of this stabilisation mechanism
to food emulsions is likely to be limited in practice.
A lowering of the interfacial tension by guar and locust
bean gums by up to ca. 20 mM m 21 has been reported
(Garti, 1999; Garti & Reichman, 1994). While these authors
are genuinely convinced that this surface activity is an
intrinsic property of the polysaccharide itself, and not due
to the presence of contaminating protein, it is fair to say that
all such samples probably do contain a residual strongly
bound protein/peptide fraction, which is likely to be the
origin of much of the apparent emulsifying functionality
of galactomannan gums. Based on a study of 10 commercial
samples, Anderson (1986) reported an average nitrogen
content of 0.57% for guar gum. This corresponds to
,2.5% protein, which is actually higher than typically

found for gum arabic! Garti and Reichman (1994) puried


their guar gum down to 0.8% protein, but still found a similar degree of surface activity and emulsication ability.
Hence, based on this evidence, it is conceivable that this
polysaccharide may confer some emulsion stabilising
property in its own right bound to the interface through
the slight hydrophobicity of the polymannose backbone.
Even if this is the case, however, it should be noted that
this putative adsorption of the gum to the oilwater interface is reportedly rather weak and reversible. That is, the
birefringency around the droplets and the associated emulsion stability is lost on diluting the aqueous phase of the
emulsion with water (Garti & Reichman, 1994).
Another galactomannan that has received particular
attention recently is fenugreek gum. Garti, Madar, Aserin,
and Sternheim (1997) have reported that puried fenugreek
gum has substantial surface activity and that stable emulsions of moderately low average droplet size (,3 mm) can
be produced; moreover, physical separation of the contaminating protein from the crude gum material did not
apparently reduce the measured surface activity. Conversely, however, in a separate recent study (Brummer,
Cui, & Wang, 2001) where a fenugreek gum sample
containing 2.4% protein was enzymatically treated with
pronase, the ability of the resulting hydrolyzate (containing
0.6% protein, but of the same overall molecular weight) was
found to be substantially reduced. It can therefore be reasonably inferred that the surface activity of fenegreek gum, like
that of gum arabic, is due to the presence of a small fraction
of protein closely associated (chemically or physically) with
the polysaccharide structure.
3.3. Pectin
Pectins extracted from plant cell walls are commonly
used as a gelling and thickening agent in foods. High methoxyl pectin ($50% DE) forms gels under acidic conditions
in aqueous media of high sugar content, whereas low methoxyl pectin forms gels in the presence of calcium ions.
Typically, commercial pectins extracted from citrus peel
or apple pomace are not effective as emulsifying agents
irrespective of the degree of esterication. There are, however, some exceptions to this statement.
Highly acetylated pectin from sugar beet has been
reported (Dea & Madden, 1986) to be much more surfaceactive than commercial high-methoxyl or low-methoxyl
pectins, and hence to be readily capable of producing and
stabilising ne vegetable oil-in-water emulsions. According
to Dea and Madden (1986), the molecular origin of the
considerable emulsifying capacity of sugar beet pectin is
the substantially hydrophobic character of the acetyl groups
(29%). Nevertheless, it should be noted that the experiments upon which this statement was based were carried out
on samples containing .2 wt% protein.
Even with a low acetyl content (,0.8%), pectins from
citrus fruits and apples can also exhibit good surface activity

E. Dickinson / Food Hydrocolloids 17 (2003) 2539

and emulsion stabilising characteristics if the average


molecular weight is reduced to ,80 kDa by severe acid
hydrolysis (Mazoyer, Leroux, & Bruneau, 1999). Based
on time-dependent changes in emulsion droplet-size distribution and serum separation, depolymerised pectin of molecular weight 70 kDa and 70% degree of esterication has
been used to make highly stable ne oil-in-water emulsions
at a relatively low gum/oil ratio (1:5) (Akhtar, Dickinson,
Mazoyer, & Langendorff, 2002). Emulsion stability was
found to be reduced on increasing the pH from 4.7 to 7,
or by substantially increasing (or decreasing) the pectin
molecular weight (Fig. 2).
Independent of the amount of depolymerised pectin used,
the proportion of the hydrocolloid adsorbed at the oilwater
interface during emulsication was found to remain roughly
constant at ca. 25% (Akhtar et al., 2002). The aqueous
phases of the original pectin solutions, and the serum layers
of the emulsions separated by centrifugation, were both
analysed with respect to protein content. This showed that
a negligible proportion of the original protein content of the
sample (0.6 wt%) was detectable in the aqueous phase after
emulsication. Thus, it appears that the protein fraction of
the hydrocolloid emulsier was almost entirely associated
with the adsorbed layer around the droplets. Furthermore,
the fraction remaining in the serum phase was found to be
quite ineffective for subsequent emulsication. We therefore can infer that the presence of the protein moiety in
the depolymerised pectin plays a major functional role in
enhancing the emulsication properties of pectin following
depolymerisation. The analogy with gum arabic is again
quite noteworthy. That is, the hydrophobic protein component is responsible for the surface activity of the hydrocolloid emulsier, acting as the strong anchor point at the
oilwater interface, with the hydrophilic polysaccharide
chains providing the protective layer that confers effective
steric stabilisation during extended storage.

4. Interactions of hydrocolloids with proteins


4.1. Interactions in aqueous solution
The nature and strength of proteinhydrocolloid interactions in bulk solution and at interfaces have an important
inuence on the stability properties of food dispersions and
emulsions (Dickinson, 1993; Dickinson & Euston, 1991). In
aqueous solution, a binary mixture of protein 1 hydrocolloid can exhibit one of three different equilibrium situations (Albertsson, 1971): (a) miscibility, (b) thermodynamic
incompatibility, and (c) complex coacervation (or complexation). Miscibility occurs commonly at low biopolymer
concentrations, and either incompatibility or coacervation at
high concentrations, depending on whether the protein
hydrocolloid interaction is net repulsive or net attractive,
respectively. Thermodynamic incompatibility implies the
separation into two distinct aqueous phases, one rich in

31

protein and the other rich in hydrocolloid. Two distinct


phases also arise in complex coacervation, except that
now one phase is rich in the two biopolymers and the
other phase is depleted in both.
A useful parameter to quantify the relative strength of the
binary proteinhydrocolloid interaction in dilute solution is
the cross second virial coefcient (Aph) as may be determined experimentally by static light scattering from a polymer mixture (Kaddur & Strazielle, 1986). Positive and
negative values of Aph are indicative of net repulsive and
net attractive interactions, respectively. The overall thermodynamic behaviour of the mixed bipolymer solution
depends on the relative values of Aph, App and Ahh,
where App and Ahh are the pure protein and pure hydrocolloid virial coefcients, representing the contributions
from proteinprotein and hydrocolloidhydrocolloid interactions, respectively. The low polymeric entropy of mixing
means that, if the proteinhydrocolloid interaction is only
slightly net repulsive (i.e. Aph . App Ahh 1=2 ), the
system prefers to exist as two separate phases when the
overall biopolymer concentration reaches just a few percent
(Grinberg & Tolstoguzov, 1997). Values of Aph for many
combinations of food proteins and hydrocolloids are such
that the system exhibits thermodynamic incompatibility.
There are only a very few casese.g. serum albumin 1
pectin (Semenova, Bolotina, & Dmitrochenko, 1991)for
which genuinely complete miscibility has been established
in the non-dilute mixed biopolymer solution. The rest exhibit some kind of soluble complexation, complex coacervation or amorphous co-precipitation, depending on the
biopolymer structures involved and the solution conditions
(pH, ionic strength, etc.). At one extreme, with weak or
specic unlike attractive interactions Aph , 0; we may
have a one-phase non-ideal dilute solution of soluble
mixed complexes; and at the other extreme, with strong
unlike attractions Aph p 0 and stoichiometric release of
microions, we may get a mixed precipitate formed with very
little water present. Strictly speaking, complex coacervation
represents the middle ground of a separated liquid-like
phase of mixed biopolymers, which are strongly hydrated
and osmotically swelled, for the retention of microions
(Dubin, 2001).
Unless protein and hydrocolloid carry opposite net
charges, complex coacervation generally does not take
place. The phenomenon tends to be suppressed also at low
biopolymer charge densities. Precipitation and/or gelation
may occur at high charge densities. The maximum coacervation yield occurs for the case of an equal ratio mixture (by
weight) of biopolymers at the pH where they carry equal and
opposite charges (Schmitt, Sanchez, Desobry-Banon, &
Hardy, 1998). The extent to which a particular biopolymer
is involved in Coulombic interactions depends on how far
its isoelectric point pI differs from the solution pH. Most
food proteins pIp , 5 form complex coacervates with
anionic hydrocolloids pIh , 3 in the intermediate region
of pH where the two macromolecules carry opposite net

32

E. Dickinson / Food Hydrocolloids 17 (2003) 2539

Fig. 3. Distribution of charge on the surface of bovine serum albumin


(BSA) at pH 7 as calculated by Hattori et al. (2001). Pictures (a)(c)
show the protein molecule viewed from three different directions. Surface
regions coloured red and blue carry local net negative and net positive
charges, respectively, and white regions are locally electrically neutral.
(Reproduced with permission from Hattori et al., (2001)).

charges pIp , pH , pIh : According to Ledward (1994),


the strength of complexation between protein and polysaccharide depends on the distribution of ionisable groups on
the surface of the protein, the ease of unfolding of the
protein's native structure, and the backbone exibility and
charge distribution on the polysaccharide. The extent of
reversibility of complexation depends on the aqueous
environment and on the mixing conditions (Tolstoguzov,
1986). The tendency towards the non-equilibrium formation
of an insoluble coacervate is enhanced at low salt concentrations and at pH values signicantly below pIp.
When both biopolymers carry a net negative charge,
soluble complexes may be produced (Dickinson, 1995b).
In this case, any electrostatic interaction involves the anionic polyelectrolyte interacting with positively charged local
patches on the protein (Park, Muhoberac, Dubin, & Xia,
1992). The importance of such charged patches has long
been recognised in protein chromatography (Regnier,
1987). Increasing the net negative charge on the protein
has two effects: (a) it enhances proteinprotein electrostatic
repulsion (i.e. it increases App) and (b) it reduces protein
hydrocolloid attraction by screening the interactions of positively charged groups (i.e. it increases Aph). The positively
charged groups on a protein (NH31) interact more strongly
with OSO32 groups than with 2CO22 groups. This means

that, in terms of the relative contributions of effects (a) and


(b), a useful distinction can be made between a sulfated
hydrocolloid like carrageenan and a carboxylated hydrocolloid like pectin. At low ionic strength, sulfated hydrocolloids of relatively high charged density form fairly strong
reversible complexes with proteins, even at neutral or alkaline pH (i.e. well above pIp). In contrast, at pH . pIp ; any
interaction of a protein with a carboxylated hydrocolloid
interaction is quite weak or non-existentor the system
may even exhibit thermodynamic incompatibility.
Computer modelling of the electrostatic binding of
human serum albumin (HSA) to the strongly anionic polysaccharide hyaluronic acid (HA) has recently been reported
(Grymonpre, Staggemeier, Dubin, & Mattison, 2001).
Under incipient binding conditions pH . pIp ; these calculations reveal a region of positive electric potential 0.5 nm
from the protein's van der Waals surface. This charged
domain diminishes in intensity with increasing pH or
ionic strength, and its size and curvature are such that it
can readily accommodate a 12 nm long section of the polyelectrolyte. The electrostatic interaction energy of the
HSAHA complex under such incipient binding conditions
has been estimated as ca. 1 kT (Grymonpre et al., 2001).
Fig. 3 shows the calculated charge density distribution on
the surface of bovine serum albumin (BSA) as viewed from
three different directions (Hattori, Kimura, Seyrek, &
Dubin, 2001). Although the protein carries a net negative
charge, there is a clearly dened patch (size 23 nm) of
positive charge on the BSA molecule surface even at pH
7. Involvement of this patch in electrostatic binding at pH .
pIp allows BSA to form soluble complexes at low ionic
strength with dextran sulfate, i-carrageenan and k-carrageenan.
The electrostatic character of the complexes has been
demonstrated (Galazka, Smith, Ledward, & Dickinson,
1999) by their reversible dissociation in the presence of
added electrolyte (0.020.1 M NaCl). At low ionic strength,
the `critical' upper pH for effective complexation with BSA
is ca. 6.5 for k-carrageenan, as compared with ca. 7 for icarrageenan, and .7 for dextran sulfate. The value of the
protein net charge at the critical pH for complexation has
been shown (Mattison, Dubin, & Brittain, 1998) to be
proportional to the square root of ionic strength, and to
depend on polyelectrolyte charge density (Dickinson,
1998). With respect to adjustment of pH or ionic strength,
the relative strengths of the BSApolysaccharide interactions correlate well with the relative densities of charged
groups along the different polysaccharide chains: dextran
sulfate . i-carrageenan . k-carrageenan.
An extreme type of proteinpolysaccharide interaction
occurs when a covalent linkage is formed between the two
kinds of biopolymers (Dickinson, 1993). The motivation to
construct such a mixed biopolymer conjugate might be the
attempt to generate a new type of `natural' amphiphilic
hydrocolloid having emulsication properties as good as
or better than gum arabic. Without using any chemicals,
covalent proteinpolysaccharide conjugates can be prepared

E. Dickinson / Food Hydrocolloids 17 (2003) 2539

33

by linking protein 1-amino-groups to the reducing-end


carbonyl groups of polysaccharides through controlled
Maillard reaction (Kato, Sasaki, Furuta, & Kobayashi,
1990). For instance, the slow dry heating of mixtures of
b-lactoglobulin 1 dextran produces soluble conjugates
with substantially improved solubility and emulsifying
properties compared to the protein alone (Dickinson &
Galazka, 1991). From the hydrocolloid functionality point
of view, a major advantage of covalent bonding over
electrostatic proteinpolysaccharide complexation is the
maintenance of a strong association over a wide range of
pH and ionic strength without any incipient coacervation or
precipitation. So, in contrast to casein, which starts to precipitate below pH < 5:5; a caseinmaltodextrin conjugate
can be prepared (Shepherd, Robertson, & Ofman, 2000)
which has good solubility and is an effective emulsier
under acidic conditions.
4.2. Interactions of hydrocolloids with adsorbed protein
layers

Fig. 4. Schematic representation of three alternative effects of the adsorption of stiff hydrocolloid polymers on the surface of spherical emulsion
droplets depending on the hydrocolloid concentration and the nature of
the hydrocolloidprotein interaction: (a) a sterically stabilised system, (b)
an emulsion gel, and (c) a system occulated by macromolecular bridging.

Fig. 5. Inuence of i-carrageenan on the state of occulation of BSA-stabilised emulsions (20 vol% oil, 1.7 wt% protein, pH 6, ionic strength
0.005 M) stored at 25 8C for 40 h. The average apparent droplet size d32
is plotted against the hydrocolloid concentration ch added to the freshly
made emulsion. (Reproduced with permission from Dickinson and
Pawlowsky (1997)).

Some food hydrocolloids achieve interfacial functionality


in emulsions, not so much by adsorbing directly at the oil
water interface during emulsication, but rather by forming
an associative interaction after emulsion formation with the
stabilising protein layer. Partial unfolding of globular
proteins may make them more susceptible to complexation
with hydrocolloids in the adsorbed state than in aqueous
solution at the same pH and ionic strength (Dickinson,
1995a). Rather direct evidence for proteinpolysaccharide
complexation at the oilwater interface is provided by
surface shear rheology measurements, e.g. for BSA 1
dextran sulfate at neutral pH (Dickinson & Galazka,
1992), and by particle electrophoretic mobility measurements, e.g. for BSA 1 sodium alginate below neutral pH
(Ward-Smith, Hey, & Mitchell, 1994). More indirect
evidence for interfacial complexation comes from bulk
rheological and occulation studies of emulsions.
Strong attractive interaction Aph p 0 between
adsorbed protein and hydrocolloid present at sufciently
high total concentration typically produces a secondary
layer of steric stabilising polymer around protein-coated
emulsion droplets (see Fig. 4(a)). At an excess hydrocolloid
concentration high enough to cause gelation in the aqueous
phase, the hydrocolloid-coated droplets may become immobilised in the polymer gel network (Fig. 4(b)) leading to
increased stabilisation (an emulsion gel). Conversely, the
same protein-coated emulsion droplets may become
destabilised by bridging occulation (Fig. 4(c)) if the hydrocolloidprotein association is weak Aph , 0; or if the
hydrocolloid is not present at high enough total concentration to saturate all the surfaces of the droplets. Bridging
occulation of emulsion droplets may also be induced by
adhesion of oppositely charged colloidal particles or even
bacterial cells (Li, McClements, & McLandsborough,
2001).

34

E. Dickinson / Food Hydrocolloids 17 (2003) 2539

Fig. 6. Inuence of high-methoxyl pectin on the rheology of pure caseinstabilised emulsions (40 vol% oil, 2 wt% protein, pH 5.5, ionic strength
0.01 M). The limiting zero-shear-rate viscosity at 22 8C is plotted against
hydrocolloid concentration for emulsions made separately with as1-casein
(W) and b-casein (X). (After Dickinson et al. (1998)).

Bridging occulation of a BSA-stabilised emulsion has


been found to be induced by addition of i-carrageenan under
conditions of soluble complex formation in aqueous solution (Dickinson & Pawlowsky, 1997). Fig. 5 shows the
effect of added hydrocolloid concentration ch on the average
apparent particle size d32 of a BSA-stabilised emulsion with
an aqueous phase of pH 6 and low ionic strength. In the
emulsion sample without added hydrocolloid, or in samples
with ch # 1023 wt%, the low value of d32 < 0:5 mm corresponds to the actual sizes of individual dispersed droplets.
However, for ch $ 5 1023 wt%, there is a jump in the
apparent particle size by more than an order of magnitude
due to bridging occulation of the BSA-coated droplets by
i-carrageenan. This high value of d32 < 10 mm is a rough
measure of the average oc size in emulsions with hydrocolloid added in the concentration range 0.010.1 wt%.
Partial restabilisation at higher i-carrageenan contents is
indicated by the considerably reduced values of d32 for
ch . 0:1 wt%. At this ionic strength (0.005 M), changing
the aqueous phase conditions to pH 7 leads to a protein
hydrocolloid interaction that is too weak to induce
signicant occulation (Dickinson & Pawlowsky, 1996).
However, upon addition of the more highly charged (and
therefore more strongly protein-associating) dextran sulfate
instead of i-carrageenan, substantial bridging occulation
was found to occur at pH 7 (and higher).
Proteinpolyelectrolyte interactions are sensitive to
details of protein structure as well as to the charge density
on the polyelectrolyte. For instance, in distinct contrast to
the behaviour of BSA, the globular protein b-lactoglobulin
(with a similar pIp) shows little evidence of signicant
electrostatic complexation with dextran sulfate at neutral
pH and low ionic strength, based on measurements of the
electrophoretic mobility of the protein-coated droplets
(Dickinson, 1995a,b). Furthermore, again in contrast to
BSA, the same polyelectrolyte appears to cause no discernible bridging occulation of b-lactoglobulin-stabilised

emulsions or enhancement of protein layer surface shear


viscosity at the oilwater interface.
High-methoxyl pectin adsorbs on the surface of caseincoated oil droplets and stabilises them under pH conditions
where the precipitation of the emulsions would otherwise
take place (Dalgleish & Hollocou, 1997). Even though the
attractive interaction of carboxylated hydrocolloid with
adsorbed protein may be quite weak at pH . pIp ; it can
nonetheless have a substantial inuence on the stability
and rheology of a concentrated emulsion. This has been
illustrated (Dickinson, Semenova, Antipova, & Pelan,
1998) for the case of as1-casein-stabilised emulsions
containing high-methoxyl pectin at pH 5.5 and low ionic
strength. The very high bulk viscosity of these occulated
emulsions for ch $ 1 wt% (see Fig. 6) is attributable to
signicant attractive proteinpolysaccharide interaction at
the surface of the droplets. Associative interfacial interaction has been conrmed at the macroscopic oilwater
interface by a ve-fold increase in the surface shear viscosity of an adsorbed as1-casein layer on introducing pectin into
the aqueous sub-phase. This behaviour is thermodynamically consistent with the negative cross second virial coefcient Aph derived from light scattering for an aqueous
solution of as1-casein 1 pectin at pH 5.5 and ionic
strength 0.01 M (Semenova et al., 1999). In contrast, the
value of Aph is positive for the equivalent aqueous solution
of b-casein 1 pectin, and the viscosity of the (non-occulated) pectin-containing b-casein-stabilised emulsion is
very much lower than that of the equivalent as1-casein
system, as shown in Fig. 6.
With certain hydrocolloids, interfacial complexation is
more likely to have its origin in hydrogen bonding rather
than in electrostatic interactions. This seems especially the
case with the uniquely hydrophilic protein gelatin which has
the ability to form gel-like layers at a range of different
surfaces (Izmailova et al., 2001). In particular, a thin layer
of gelatin network may accumulate at a surface already
coated with another more surface-active biopolymer.
Consider, for example, the mixed system of b-casein
(pI < 5) 1 acid processed type-A gelatin (pI < 9) at neutral
pH. Whereas surface tension experiments indicate that the
primary interfacial layer is totally dominated by the much
more surface-active b-casein, measurements of surface
shear viscosity are at least an order of magnitude higher
for b-casein 1 gelatin than for b-casein alone, indicating
the presence of a considerable amount of gelatin at the
interfacial region (Galazka & Dickinson, 1995). Due to
the opposite net charges of the two biopolymers, the
association between molecules in primary and secondary
layers could reasonably be interpreted solely in terms of
electrostatic interaction. However, there is separate
evidence available (Castle, Dickinson, Murray, & Stainsby,
1987) of accumulation of secondary layers of negatively
charged alkali-processed type-B gelatin pI < 5 below
a primary layer of negatively charged casein at neutral
pH, again based on surface shear viscometry, but in

E. Dickinson / Food Hydrocolloids 17 (2003) 2539

35

5. Effect of non-adsorbing hydrocolloids on colloidal


stability

Fig. 7. Inuence of xanthan gum on the creaming of sodium caseinatestabilised emulsions (10 wt% oil, 0.5 wt% protein, pH 7, ionic strength
0.005 M) stored at 25 8C in 10 cm tubes. The serum layer height H is
plotted against the storage time t for various concentrations of hydrocolloid
added to the freshly made emulsion: P, 0 wt% and $0.125 wt%; V,
0.025 wt%; W, 0.05 wt%; L, 0.0625 wt%; O, 0.1 wt%. (Reproduced with
permission from Cao et al., (1990)).

this case with sodium caseinate instead of b-casein. Gelatin is well known for its thermoreversible gelation via
cooperative hydrogen-bonded interactions within the
collagen-type helices that are essential in providing
cross-links (Izmailova et al., 2001). Hence, it would
appear reasonable to speculate that such cooperative
intermolecular hydrogen-bonding forces, probably supplemented by attractive ionic forces, are mainly responsible
for holding together the mixed casein/gelatin interfacial
lm.
Thermal or hydrostatic pressure processing of a proteinstabilised emulsion, which induces changes in globular
protein structure, can also inuence proteinhydrocolloid
interactions at the surface of protein-coated droplets,
with signicant implications for stability and rheology.
Substantial effects of high-pressure treatment on emulsion
properties have recently been found for systems containing
b-lactoglobulin 1 low methoxy pectin (Dickinson & James,
2000) and ovalbumin 1 sulfated polysaccharides (Galazka,
Dickinson, & Ledward, 2000). Following partial denaturation, the number of available interacting sites on the protein
increases as buried groups are liberated (Ledward, 1994).
As well as inducing unfolding of the globular proteins, the
application of high hydrostatic pressure (a few kilobars)
leads to disruption of intermolecular ionic bonds and
hence to dissociation of electrostatic complexes. When pressure is released, the ionic attractive interactions are restored,
and rapid recomplexation of it occurs, but now between the
polysaccharide and the partly denatured protein (Galazka,
Ledward, Sumner, & Dickinson, 1997). The greater molecular exibility of the denatured state allows congurational
adjustments to occur which maximise favourable ionic
interactions resulting in more stable complexes than with
the native protein.

For completeness in this review, it is appropriate to


consider the inuence of non-adsorbing hydrocolloids on
the stability of food colloids. Like surfactant micelles or
even unadsorbed milk proteins, hydrocolloids located in
the aqueous phase can induce the destabilisation of dispersions and emulsions by the mechanism of depletion occulation (Dickinson, 1995b; Dickinson & Euston, 1991). This
arises when pairs of particle surfaces approach within a
distance less than the mean diameter of the free polymer
molecule in aqueous solution. The exclusion of the polymer
from the intervening gap is associated with an attractive
interparticle force, resulting from the tendency of solvent
to ow out from the gap under the inuence of the local
osmotic pressure gradient (Napper, 1983). At very low
concentrations of free polymer, the entropy loss linked to
particle aggregation outweighs the depletion effect and the
system remains stable. However, beyond a certain critical
polymer concentration, which decreases with increasing
particle volume fraction and increasing molecular mass of
the polymer, reversible depletion occulation occurs, leading to enhanced sedimentation and phase separation. Hence,
a concentrated protein-stabilised emulsion is highly susceptible to destabilisation by depletion occulation following
addition of a small amount of non-adsorbing hydrocolloid
Aph . 0: The emulsion destabilisation is manifest in rapid
creaming and serum separation (Dickinson, 1993). Hydrocolloids inducing strong depletion occulation, however,
may produce apparent restabilisation with respect to
serum separation due to the formation of a long-lasting
gel-like network of aggregated droplets (Parker, Gunning,
Ng, & Robins, 1995).
For non-adsorbing macromolecular species present at
xed volume fraction, it has been shown theoretically
(Piech & Walz, 2000) that the magnitude of the depletion
attraction energy between particles increases with the
degree of non-sphericity of the macromolecular species.
Hence, non-adsorbing hydrocolloids with extended and
stiff polysaccharide backbones are especially effective at
inducing depletion occulation. This is shown in Fig. 7
for the case of xanthan gum added at various low concentrations to a model protein-stabilised emulsion system
(10 vol% oil) of sample height 10 cm. The kinetics of
destabilisation is indicated by plotting serum layer thickness
against storage time. The reference emulsion (no added
hydrocolloid) shows no serum separation discernible by
eye during storage at 25 8C for three days (Cao, Dickinson,
& Wedlock, 1990). The presence of a very low concentration of xanthan (0.025 or 0.05 wt%) is enough to induce
depletion occulation, leading to rapid separation (complete
within 12 h) of the low-viscosity liquid-like emulsion to
form a 6570% serum layer. The same degree of phase
separation takes about two days to complete at 0.0625 wt%
xanthan. Furthermore, there is only ,10% serum separation

36

E. Dickinson / Food Hydrocolloids 17 (2003) 2539

Fig. 8. Rheology of small-molecule surfactant-stabilised emulsions


(30 wt% oil, 1 wt% surfactant, pH 7, ionic strength 0.05 M) containing
0.1 wt% rhamsan gum incorporated after emulsication. The apparent
shear viscosity h at 25 8C is plotted against the shear stress S on a log
log scale for two different surfactants as emulsifying agent: W, Tween 20;
X, SDS. (Reproduced with permission from Dickinson et al. (1993)).

after three days in the presence of 0.1 wt% xanthan, and


none at all after three days at 0.125 wt% xanthan. Stability
towards cream or serum separation over a period of several
weeks could be attained by increasing the stabiliser content
to just 0.25 wt% (Cao, Dickinson, & Wedlock, 1991).
The inhibition of creaming at stabiliser concentrations
$0.125 wt% in Fig. 7 can be attributed to immobilisation
of the protein-coated oil droplets in a weak gel-like network
with a very high low-stress shear viscosity. Partial stabilisation at 0.0625 and 0.1 wt% xanthan can be mainly attributed
to slow movement of the occulated droplets under gravity
due to the high viscosity of the continuous phase (Dickinson,
1993).
In an emulsion containing a mixture of hydrocolloid and
small-molecule surfactant, the nature of the hydrocolloid
surfactant interactions can inuence both rheology and

Fig. 9. Inuence of rhamsan gum on the creaming of small-molecule surfactant-stabilised emulsions (30 wt% oil, 1 wt% surfactant, pH 7, ionic
strength 0.05 M) stored at 25 8C in 6 cm tubes. The time ts for the serum
layer to become visible to the unaided eye is plotted against the hydrocolloid concentration c: W, Tween 20; X, SDS. (Reproduced with permission
from Dickinson et al. (1993)).

stability. This is shown in Fig. 8 for two different emulsions


(30 wt% oil, 1 wt% surfactant) of similar average droplet
size (d43 < 0:6 mm), containing the same concentration
(0.1 wt%) of another anionic microbial polysaccharide,
rhamsan, but made with two different emulsiers: (i) nonionic Tween 20 and (ii) anionic sodium dodecyl sulfate
(SDS). Without emulsied oil present, there was observed
no signicant difference in shear-thinning rheology of the
two mixed hydrocolloid 1 surfactant aqueous phases
(rhamsan 1 Tween or rhamsan 1 SDS); and neither emulsion showed any measurable change in mean droplet size on
storage for several days (Dickinson, Goller, & Wedlock,
1993). Fig. 8 shows that, for moderate applied stresses
(.1 Pa), the apparent viscosities of the two emulsions are
also similar, but at low stresses (,10 2 Pa) the Tween 20
emulsion viscosity is two orders of magnitude higher than
that of the SDS emulsion. This difference can be attributed
to the different effect of the charged polyelectrolyte on the
electrostatically stabilised SDS-coated emulsion droplets as
compared with its effect on the sterically stabilised Tweencoated droplets (Dickinson et al., 1993).
There is often found to be a correlation between emulsion
rheology and emulsion creaming stability. For example, for
the same sets of emulsion systems containing rhamsan and
Tween 20 or SDS, Fig. 9 shows the creaming stability as a
function of rhamsan concentration (Dickinson et al., 1993).
The serum separation time ts is dened as the time when a
distinct serum layer becomes rst visible to the naked eye.
According to this criterion, whereas the 1% SDS system
becomes unstable to creaming in just one day, the equivalent Tween system is stable for several days. Comparison of
Figs. 8 and 9 shows that the higher low-stress viscosity of
the Tween 20 emulsion correlates well with its better creaming stability. The stronger gel-like network of the concentrated Tween-stabilised emulsion (30% oil), as measured by
the much higher h value at low S in Fig. 8, can more effectively retard the thermodynamic drive towards depletionbased phase separation. It is interesting to contrast this
case with the one shown in Fig. 6, where the enhanced
low-stress viscosity from bridging occulation arises as a
result of an associative interaction between hydrocolloid
(pectin) and adsorbed protein (as1-casein). In the bridging
occulation case, the higher low-stress viscosity in the
concentrated emulsion (40% oil) correlates with poorer
creaming stability (i.e. a faster rate of serum separation) in
the diluted emulsion (Dickinson et al., 1998). This opposite
behaviour to that given in Fig. 9 for depletion occulation
shows the complexity of interpreting stability/rheology relationships in hydrocolloid-containing emulsion systems.
Finally, it seems worthwhile to stress that occulation
is a ubiquitous phenomenon in hydrocolloid-containing
systems. As illustrated by Lips et al. (1991) for a dilute
model colloid of polystyrene latex particles, a biopolymer
will tend either to cause occulation by bridging (at about
half saturation coverage) or by depletion (at rather higher
concentrations). Which mechanism actually occurs in any

E. Dickinson / Food Hydrocolloids 17 (2003) 2539

particular system depends sensitively on the nature of the


hydrocolloid-particle interaction, which in turn is dependent
on the particle surface properties and the solution conditions. In emulsions, the most likely situations favouring
destabilisation by occulation (bridging or depletion) are
those where the oil volume fraction is low (say, 510%)
and where the aqueous phase contains a non-gelling hydrocolloid (or a gelling hydrocolloid at a concentration below
the gelation threshold). In practice, of course, biopolymer
adsorption and occulation effects in concentrated emulsions are commonly masked by the very high low-stress
viscosity of the hydrocolloid-containing aqueous phase or
by hydrocolloid gelation between the droplets.
6. Concluding remarks
Some food hydrocolloids are sufciently surface-active in
their own right to act as sole emulsifying agents during the
formation of oil-in-water emulsions. In most cases, it
appears that the origin of the emulsifying ability is the
presence of proteinaceous material covalently bound or
physically associated with the carbohydrate polymer. Typically, the emulsions produced are coarser than those that can
be made with low-molecular-weight water-soluble surfactants or milk proteins at the same emulsier/oil ratio. Once
strongly adsorbed at the oilwater interface, however,
hydrocolloids can be even more effective than surfactants
or proteins in conferring long-term stability on emulsions.
This is due to the very favourable attributes of hydrophilicity and large molecular size in connection with the efciency of steric stabilisation. Hydrocolloids are especially
useful in practice for stabilisation of emulsions under the
acidic conditions where milk proteins are much less effective.
Hydrocolloids have an important role, both positive and
negative, in relation to the storage stability of food emulsions. Of course, this may simply be by modication of the
small-deformation rheology of the continuous aqueous
phase. But, commonly also, the hydrocolloid inuences
the state of occulation of the droplets via either a bridging
or depletion mechanism, depending on whether the hydrocolloid has a net attractive or repulsive interaction with the
surface. As the latter typically consists of adsorbed protein,
an understanding of the factors affecting proteinpolysaccharide interactions in solution seems extremely valuable
for understanding and controlling the physico-chemical
phenomena involved. Even small changes in weak attractive
or repulsive interactions between different biopolymers in a
colloidal system can have a profound effect on physical
stability with respect to creaming, sedimentation or serum
separationwith drastic consequences for product quality
and appearance.
References
Akhtar, M., Dickinson, E., Mazoyer, J., & Langendorff, V. (2002). Emul-

37

sion stabilizing properties of depolymerized pectin. Food Hydrocolloids, 16.


Albertsson, P. -A. (1971). Partition of cell particles and macromolecules,
New York: Wiley.
Anderson, D. M. W. (1986). The amino acid components of some commercial gums. In G. O. Phillips, D. J. Wedlock & P. A. Williams (Eds.),
Gums and stabilisers for the food industry (pp. 7986). Vol. 3. London:
Elsevier.
Bergenstahl, B. (1997). Physicochemical aspects of emulsier functionality. In G. L. Hasenhuettl & R. W. Hartel, Food emulsiers and their
applications (pp. 147172). New York: Chapman & Hall.
Brummer, Y., Cui, S, Wang, Q. (2001). Extraction, purication and
physicochemical characterization of fenugreek gum. Poster presentation at Gums and Stabilisers for the Food Industry Conference.
Wrexham, UK, July 26.
Buffo, R. A., Reineccius, G. A., & Oehlert, G. W. (2001). Factors affecting
the emulsifying and rheological properties of gum Acacia in beverage
emulsions. Food Hydrocolloids, 15, 5366.
Cao, Y., Dickinson, E., & Wedlock, D. J. (1990). Creaming and occulation in emulsions containing polysaccharide. Food Hydrocolloids, 4,
185195.
Cao, Y., Dickinson, E., & Wedlock, D. J. (1991). Inuence of polysaccharide on the creaming of casein-stabilized emulsions. Food Hydrocolloids, 5, 443454.
Castle, J., Dickinson, E., Murray, B. S., & Stainsby, G. (1987). Mixed
protein lms adsorbed at the oil/water interface. American Chemical
Society Symposium Series, 343, 118134.
Connolly, S., Fenyo, J. C., & Vandevelde, M. C. (1988). Effect of a proteinase on the macromolecular distribution of Acacia senegal gum. Carbohydrate Polymers, 8, 2332.
Dalgleish, D. G., & Hollocou, A. -L. (1997). Stabilization of protein-based
emulsions by means of interacting polysaccharides. In E. Dickinson &
B. Bergenstahl, Food colloids: proteins, lipids and polysaccharides
(pp 236244). Cambridge, UK: Royal Society of Chemistry.
Darling, D. F., & Birkett, R. J. (1987). Food colloids in practice. In
E. Dickinson, Food emulsions and foams (pp. 129). London: Royal
Society of Chemistry.
Dea, I. C. M., & Madden, J. K. (1986). Acetylated pectic polysaccharides of
sugar beet. Food Hydrocolloids, 1, 7188.
Dickinson, E. (1986). Mixed proteinaceous emulsiers: review of competitive protein adsorption and the relationship to food colloid stabilization. Food Hydrocolloids, 1, 323.
Dickinson, E. (1988a). The role of hydrocolloids in stabilizing particulate
dispersions and emulsions. In G. O. Phillips, D. J. Wedlock & P. A.
Williams, (Eds.), Gums and stabilisers for the food industry (pp. 249
263). , Vol. 4. Oxford, UK: IRL Press.
Dickinson, E. (1988b). The structure and stability of emulsions. In J. M. V.
Blanshard & J. R. Mitchell, Food structureits creation and evaluation (pp. 4157). London: Butterworths.
Dickinson, E. (1992). An introduction to food colloids, Oxford, UK:
University Press chapter 1.
Dickinson, E. (1993). Proteinpolysaccharide interactions in food colloids.
In E. Dickinson & P. Walstra, Food colloids and polymers: stability and
mechanical properties (pp. 7793). Cambridge, UK: Royal Society of
Chemistry.
Dickinson, E. (1994). Emulsions and droplet size control. In D. J. Wedlock,
Controlled particle, droplet and bubble formation (pp. 191216).
Oxford, UK: Butterworth-Heinemann.
Dickinson, E. (1995a). Mixed biopolymers at interfaces. In S. E. Harding,
S. E. Hill & J. R. Mitchell, Biopolymer mixtures (pp. 349372). Leicestershire, UK: Nottingham University Press.
Dickinson, E. (1995b). Emulsion stabilization by polysaccharides and
proteinpolysaccharide complexes. In A. M. Stephen, Food polysaccharides and their applications (pp. 501515). New York: Dekker.
Dickinson, E. (1998). Stability and rheological implications of electrostatic
milk proteinpolysaccharide interactions. Trends in Food Science and
Technology, 9, 347354.

38

E. Dickinson / Food Hydrocolloids 17 (2003) 2539

Dickinson, E., & Euston, S. R. (1991). Stability of food emulsions containing both protein and polysaccharide. In E. Dickinson, Food polymers,
gels and colloids (pp. 132146). Cambridge, UK: Royal Society of
Chemistry.
Dickinson, E., & Galazka, V. B. (1991). Emulsion stabilization by ionic and
covalent complexes of b-lactoglobulin with polysaccharides. Food
Hydrocolloids, 5, 281296.
Dickinson, E., & Galazka, V. B. (1992). Emulsion stabilization by protein/
polysaccharide complexes. In G. O. Phillips, D. J. Wedlock & P. A.
Williams, (Eds.), Gums and stabilisers for the food industry (pp. 351
362), Vol. 6. Oxford, UK: IRL Press.
Dickinson, E., & James, J. D. (2000). Inuence of high-pressure treatment
on b-lactoglobulinpectin associations in emulsions and gels. Food
Hydrocolloids, 14, 365376.
Dickinson, E., & Pawlowsky, K. (1996). Rheology as a probe of protein
polysaccharide interactions in oil-in-water emulsions. In G. O. Phillips,
P. A. Williams & D. J. Wedlock, (Eds.), Gums and stabilisers for the
food industry (pp. 181191), Vol. 8. Oxford, UK: IRL Press.
Dickinson, E., & Pawlowsky, K. (1997). Effect of -carrageenan on occulation, creaming and rheology of a protein-stabilized emulsion. Journal
of Agricultural and Food Chemistry, 45, 37993806.
Dickinson, E., Murray, B. S., Stainsby, G., & Anderson, D. J. (1988).
Surface activity and emulsifying behaviour of some Acacia gums.
Food Hydrocolloids, 2, 477490.
Dickinson, E., Elverson, D. J., & Murray, B. S. (1989). On the lm-forming
and emulsion-stabilizing properties of gum arabic: dilution and occulation aspects. Food Hydrocolloids, 3, 101114.
Dickinson, E., Galazka, V. B., & Anderson, D. M. W. (1991a). Emulsifying
behaviour of gum arabic. 1. Effect of the nature of the oil phase on the
emulsion droplet-size distribution. Carbohydrate Polymers, 14, 373
383.
Dickinson, E., Galazka, V. B., & Anderson, D. M. W. (1991b). Emulsifying
behaviour of gum arabic. 2. Effect of the gum molecular weight on the
emulsion droplet-size distribution. Carbohydrate Polymers, 14, 385
392.
Dickinson, E., Galazka, V. B., & Anderson, D. M. W. (1991c). Effect of
molecular weight on the emulsifying behaviour of gum arabic. In
E. Dickinson, Food polymers, gels and colloids (pp. 490493).
Cambridge, UK: Royal Society of Chemistry.
Dickinson, E., Goller, M. I., & Wedlock, D. J. (1993). Creaming and
rheology of emulsions containing polysaccharide and non-ionic or
anionic surfactants. Colloids and Surfaces A, 75, 195201.
Dickinson, E., Semenova, M. G., Antipova, A. S., & Pelan, E. G. (1998).
Effect of high-methoxy pectin on properties of casein-stabilized emulsions. Food Hydrocolloids, 12, 425432.
Dubin, P. L. (2001). Microscopic and macroscopic phase transitions in
polyelectrolytemicelle systems. In Y. Iwasawa, N. Oyama & H.
Kunieda, (Eds.), Studies in surface science and catalysis (pp. 979
984), Vol. 132. Amsterdam: Elsevier.
Fauconnier, M. -L., Blecker, C., Groyne, J., Razandralambo, H., Vanzeveren, E., Marlier, M., & Paquot, M. (2000). Characterization of two
Acacia gums and their fractions using a Langmuir lm balance. Journal
of Agricultural and Food Chemistry, 48, 27092712.
Fincher, G. B., Stone, B. A., & Clarke, A. E. (1983). Arabinogalactanproteins: structure, biosynthesis and function. Annual Review of Plant
Physiology, 34, 4770.
Friberg, S. (1971). Liquid crystalline phases in emulsions. Journal of
Colloid and Interface Science, 37, 291295.
Galazka, V. B., & Dickinson, E. (1995). Surface properties of protein layers
adsorbed from mixtures of gelatin with various caseins. Journal of
Texture Studies, 26, 401409.
Galazka, V. B., Ledward, D. A., Sumner, I. G., & Dickinson, E. (1997).
Inuence of high pressure on bovine serum albumin and its complex
with dextran sulfate. Journal of Agricultural and Food Chemistry, 45,
34653471.
Galazka, V. B., Smith, D., Ledward, D. A., & Dickinson, E. (1999).
Complexes of bovine serum albumin with sulfated polysaccharides:

effects of pH, ionic strength and high pressure treatment. Food Chemistry, 64, 303310.
Galazka, V. B., Dickinson, E., & Ledward, D. A. (2000). Emulsifying
properties of ovalbumin in mixtures with sulfated polysaccharides:
effects of pH, ionic strength, heat and high-pressure treatment. Journal
of the Science of Food and Agriculture, 80, 111.
Gaonkar, A. G. (1991). Surface and interfacial activities and emulsion
characteristics of some food hydrocolloids. Food Hydrocolloids, 5,
329337.
Garti, N. (1999). What can nature offer from an emulsier point of view:
trends and progress? Colloids and Surfaces A, 152, 125146.
Garti, N. (2001). Food emulsiers and stabilizers. In N. A. M. Eskin & D. S.
Robinson, Food shelf life stability (pp. 211263). Boca Raton, FL: CRC
Press.
Garti, N., & Reichman, D. (1993). Hydrocolloids as food emulsiers and
stabilizers. Food Microstructure, 12, 411426.
Garti, N., & Reichman, D. (1994). Surface properties and emulsication
activity of galactomannans. Food Hydrocolloids, 8, 155173.
Garti, N., Madar, Z., Aserin, A., & Sternheim, B. (1997). Fenugreek
galactomannans as food emulsiers. Food Science and Technology,
30, 305311.
Garti, N., Slavin, Y., & Aserin, A. (1999). Surface and emulsication
properties of a new gum extracted from Portulaca oleracea L. Food
Hydrocolloids, 13, 145155.
Grinberg, V. Y., & Tolstoguzov, V. B. (1997). Thermodynamic incompatibility of proteins and polysaccharides in solutions. Food Hydrocolloids, 11, 145148.
Grymonpre, K. R., Staggemeier, B. A., Dubin, P. L., & Mattison, K. W.
(2001). Identication by integrated computer modelling and light
scattering studies of an electrostatic serum albuminhyaluronic acid
binding site. Biomacromolecules, 2, 422429.
Hattori, T., Kimura, K., Seyrek, E., & Dubin, P. L. (2001). Binding of
bovine serum albumin to heparin determined by turbidometric titration
and frontal analysis continuous capillary electrophoresis. Analytical
Biochemistry, 295, 158167.
Hennock, M., Rahalkar, R. R., & Richmond, P. (1984). Effect of xanthan
gum upon the rheology and stability of oil-in-water emulsions. Journal
of Food Science, 49, 12711274.
Islam, A. M., Phillips, G. O., Sljivo, A., Snowden, M. J., & Williams, P. A.
(1997). A review of recent developments on the regulatory, structural
and functional aspects of gum arabic. Food Hydrocolloids, 11, 493
505.
Izmailova, V. N., Yampolskaya, G. P., Levachev, S. M., Derkatch, S. R.,
Tulovskaya, Z. D., & Voronko, N. G. (2001). Bulk and interfacial sol
gel transitions in systems containing gelatin. In E. Dickinson & R.
Miller, Food colloids: fundamentals of formulation (pp. 376383).
Cambridge, UK: Royal Society of Chemistry.
Jayme, M. L., Dunstan, D. E., & Gee, M. L. (1999). Zeta potentials of gum
arabic stabilized oil-in-water emulsions. Food Hydrocolloids, 13, 459
465.
Kaddur, L. O., & Strazielle, C. (1986). Experimental investigation of light
scattering by a solution of two polymers. Polymer, 28, 459468.
Kato, A., Sasaki, Y., Furuta, R., & Kobayashi, K. (1990). Functional
protein/polysaccharide conjugate prepared by controlled dry heating
of ovalbumin/dextran mixtures. Agricultural and Biological Chemistry,
54, 107112.
Larre, C., Huchet, B., Berot, S., & Popineau, Y. (2001). Functional properties of peptides derived from wheat storage proteins by limited enzymatic hydrolysis and ultraltration. In E. Dickinson & R. Miller, Food
colloids: fundamentals of formulation (pp. 262271). Cambridge, UK:
Royal Society of Chemistry.
Ledward, D. A. (1994). Proteinpolysaccharide interactions. In N. S.
Hettiarachchy & G. R. Ziegler, Protein functionality in food systems
(pp. 225259). New York: Dekker.
Li, J., McClements, D. J., & McLandsborough, L. A. (2001). Interaction
between emulsion droplets and Escherichia coli cells. Journal of Food
Science, 66, 570574.

E. Dickinson / Food Hydrocolloids 17 (2003) 2539


Lips, A., Campbell, I. J., & Pelan, E. G. (1991). Aggregation mechanisms in
food colloids and the role of biopolymers. In E. Dickinson, Food polymers, gels and colloids (pp. 121). Cambridge, UK: Royal Society of
Chemistry.
Mattison, K. W., Dubin, P. L., & Brittain, I. J. (1998). Complex formation
between bovine serum albumin and strong polyelectrolytes: effect of polymer charge density. Journal of Physical Chemistry B, 102, 38303836.
Mazoyer, J., Leroux, J., Bruneau, G., (1999). Use of depolymerized citrus
fruit and apple pectins as emulsiers and emulsion stabilizers. US
Patent No. 5,900,268.
McNamee, B. F., O'Riordan, E. D., & O'Sullivan, M. (1998). Emulsication and encapsulation properties of gum arabic. Journal of Agricultural
and Food Chemistry, 46, 45514555.
Napper, D. H. (1983). Polymeric stabilization of colloidal dispersions,
London: Academic Press.
Navon-Venezia, S., Zosim, Z., Gottlieb, A., Legmann, R., Carmeli, S., Ron,
E. Z., & Rosenberg, E. (1995). Alsan, a new bioemulsier from Acinetobacter radioresistens. Applied and Environmental Microbiology, 61,
32403244.
Park, J. M., Muhoberac, B. B., Dubin, P. L., & Xia, J. (1992). Effects of
protein charge heterogeneity in proteinpolyelectrolyte complexation.
Macromolecules, 25, 290295.
Parker, A., Gunning, P. A., Ng, K., & Robins, M. M. (1995). How does
xanthan stabilize salad dressing? Food Hydrocolloids, 9, 333342.
Piech, M., & Walz, J. Y. (2000). Analytical expressions for calculating the
depletion interaction produced by charged spheres and spheroids. Langmuir, 16, 78957899.
Prud'homme, R. K., & Long, R. E. (1983). Surface tensions of concentrated
xanthan and polyacrylamide solutions with added surfactants. Journal
of Colloid and Interface Science, 93, 274276.
Randall, R. C., Phillips, G. O., & Williams, P. A. (1988). The role of the
proteinaceous component on the emulsifying properties of gum arabic.
Food Hydrocolloids, 2, 131140.
Randall, R. C., Phillips, G. O., & Williams, P. A. (1989). Fractionation and
characterization of gum from Acacia senegal. Food Hydrocolloids, 3,
6575.
Ray, A. K., Bird, P. B., Iacobucci, G. A., & Clark Jr., B. C. (1995). Functionality of gum arabic: fractionation, characterization and evaluation
of gum fractions in citrus oil emulsions and model beverages. Food
Hydrocolloids, 9, 123131.
Regnier, F. E. (1987). The role of protein structure in chromatographic
behaviour. Science, 238, 319323.
Ron, E. Z., & Rosenberg, E. (2001). Natural roles of biosurfactants.
Environmental Microbiology, 3, 229236.
Rosenberg, E., & Ron, E. Z. (1999). High-and low-molecular-mass microbial surfactants. Applied Microbiology and Biotechnology, 52, 154
162.

39

Schmitt, C., Sanchez, C., Desobry-Banon, S., & Hardy, J. (1998). Structure
and technofunctional properties of proteinpolysaccharide complexes.
Critical Reviews in Food Science and Nutrition, 38, 689753.
Semenova, M. G., Bolotina, V. S., & Dmitrochenko, A. P. (1991). The
factors affecting the compatibility of serum albumin and pectinate in
aqueous medium. Carbohydrate Polymers, 15, 367385.
Semenova, M. G., Antipova, A. S., Belyakova, L. E., Dickinson, E., Brown,
R., Pelan, E. G., & Norton, I. T. (1999). Effect of pectinate on properties
of oil-in-water emulsions stabilised by as1-casein and b-casein. In
E. Dickinson & J. M. Rodriguez Patino, Food emulsions and foams:
interfaces, interactions and stability (pp. 163175). Cambridge, UK:
Royal Society of Chemistry.
Shepherd, R., Rockey, J., Sutherland, I. W., & Roller, S. (1995). Novel
bioemulsiers from microorganisms for use in foods. Journal of
Biotechnology, 40, 207217.
Shepherd, R., Robertson, A., & Ofman, D. (2000). Dairy glycoconjugate
emulsiers: caseinmaltodextrins. Food Hydrocolloids, 14, 281286.
Tan, C. -T. (1990). Beverage emulsions. In K. Larsson & S. E. Friberg,
Food emulsions (2nd ed.) (pp. 445478). New York: Dekker.
Thevenet, F. (1988). Acacia gums. American Chemical Society Symposium
Series, 370, 3744.
Tolstoguzov, V. B. (1986). Functional properties of proteinpolysaccharide mixtures. In D. A. Ledward & J. R. Mitchell, Functional properties
of food macromolecules (pp. 385415). London: Elsevier.
Trubiano, P. C. (1995). The role of specialty food starches in avour emulsions. American Chemical Society Symposium Series, 610, 199209.
Underwood, D. R., & Cheetham, P. S. J. (1994). The isolation of active
emulsier components by the fractionation of gum talha. Journal of the
Science of Food and Agriculture, 66, 217224.
Vandevelde, M. -C., & Fenyo, J. -C. (1985). Macromolecular distribution
of Acacia senegal gum (gum arabic) by size-exclusion chromatography.
Carbohydrate Polymers, 5, 251273.
Vernon-Carter, E. J. (1998). Stability of Capsicum annuum oleoresin-inwater emulsions containing Prosopis and Acacia gums. Journal of
Texture Studies, 29, 553567.
Walstra, P. (1983). Formation of emulsions. In P. Becher, (Ed.), Encyclopedia of emulsion technology (pp. 57127), Vol. 1. New York: Dekker.
Walstra, P., & Smulders, P. A. E. (1998). Emulsion formation. In B. P.
Binks, Modern aspects of emulsion science (pp. 5699). Cambridge,
UK: Royal Society of Chemistry.
Ward-Smith, R. S., Hey, M. J., & Mitchell, J. R. (1994). Protein/polysaccharide interactions at the oilwater interface. Food Hydrocolloids, 8,
309315.
Williams, P. A., Phillips, G. O., & Randall, R. C. (1990). Structure
function relationships of gum arabic. In G. O. Phillips, D. J. Wedlock
& P. A. Williams, (Eds.), Gums and stabilisers for the food industry
(pp. 2536), Vol. 5. Oxford, UK: IRL Press.

Potrebbero piacerti anche