You are on page 1of 15

Food Hydrocolloids 17 (2003) 25±39

www.elsevier.com/locate/foodhyd
Review

Hydrocolloids at interfaces and the in¯uence 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 in¯uence the properties of dispersed
systems through their interfacial properties. Hence, surface-active hydrocolloids may act as emulsi®ers and emulsion stabilisers through
adsorption of protective layers at oil±water 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 speci®c 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). Protein±hydrocolloid 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 emulsi®ers; Surface activity; Emulsion stability; Protein±hydrocolloid interactions; Flocculation

1. Introduction proteins and polysaccharides are compared in Table 1.


Both classes of biopolymer contribute to the structural and
Food hydrocolloids are high-molecular-weight hydrophi- textural properties of foods through their aggregation and
lic biopolymers used as functional ingredients in the food gelation behaviour. Furthermore, proteins are known for
industry for the control of microstructure, texture, ¯avour their emulsi®cation and foaming properties, and polysac-
and shelf-life. The term `hydrocolloid' embraces all the charides for their water-holding and thickening properties.
many polysaccharides that are extracted from plants, In reality, of course, Table 1 is a gross simpli®cation. Large
seaweeds and microbial sources, as well as gums derived differences in functional properties exist between the
from plant exudates, and modi®ed biopolymers made by the various food biopolymers depending on the detailed chemi-
chemical or enzymatic treatment of starch or cellulose. In cal structure and sensitivity to solution conditions (pH, ionic
addition, due to its polydisperse and highly hydrophilic strength, speci®c ions). The functionality of an individual
character, the unique protein gelatin has become accepted biopolymer in foods is also affected by its interactions with
as an exceptional member of this polysaccharide club. But other food componentsÐproteins, polysaccharides, lipids,
other food proteins, like casein and gluten, are traditionally sugars, salts, and so forth.
not classi®ed as hydrocolloidsÐeven though they certainly Food colloids are multi-phase food systems containing
do exhibit functional properties that overlap considerably particles or other structures with characteristic spatial
with those of the food polysaccharides. dimensions in the colloidal size range (Dickinson, 1992).
The general molecular and functional properties of The term `colloid' can be applied to particulate dispersions,
foams, gels and emulsions (oil-in-water, water-in-oil, water-
in-water). Therefore, most manufactured foodstuffs can be
q
Based on the author's presentation at the Masterclass on `Measuring classi®ed as food colloids, and many contain hydrocolloid
Performance' at the 11th Gums and Stabilisers for the Food Industry ingredients added by the manufacturer for control of stabil-
Conference (Wrexham, UK, 2±6 July 2001).
* Tel.: 144-1132-332956; fax: 144-1132-332982.
ity and rheological properties. An important class of food
E-mail address: e.dickinson@leeds.ac.uk (E. Dickinson). 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) 25±39

Table 1 ice-cream, cream liqueurs or soft drinks. The primary stabil-


The common general characteristics and the differences between proteins ising mechanism may occur in the bulk aqueous phase or at
and polysaccharides as functional biopolymers in food systems
the surface of the droplets, depending on the chemical
Similarities nature of the particular ingredient(s) involved (Lips,
Campbell, & Pelan, 1991). In this article, we consider the
Natural polymers
extent to which food hydrocolloids have a role in conferring
Widespread in food colloids
Used in pharmaceuticals, cosmetics, personal products stability on dispersions and emulsions through interfacial
Environmentally friendly polymers action.
Complicated structure In the formulation of emulsion systems, one normally
Complex aggregation behaviour distinguishes between two types of ingredient: the `emulsi-
Gelling/stabilising agents
fying agent' (or `emulsi®er') and the `stabiliser' (Dickinson,
Differences 1992). The emulsifying agent is the single chemical species
(or mixture of species) that promotes emulsion formation
Proteins Polysaccharides and short-term stabilisation by interfacial action. The term
Wide-ranging structures Similar structures
Reactive Unreactive
`bioemulsi®er' (or `biosurfactant') is also commonly used in
Monodisperse Polydisperse the ®eld of biotechnology and applied microbiology
Many segment types Few segment types (Navon-Venezia et al., 1995; Ron & Rosenberg, 2001;
Linear chain Linear or branched Rosenberg & Ron, 1999), although it is rarely used amongst
Flexible chain Stiff chain food technologists (Shepherd, Rockey, Sutherland, &
Medium molecular weight High molecular weight
Small molecular volume Large molecular volume
Roller, 1995).
Amphiphilic Hydrophilic There are two broad classes of emulsifying agents used in
Surface-active Not surface-active food processingÐsmall-molecule surfactants (monoglycer-
Polyelectrolyte Non-ionic or charged ides, polysorbates, sucrose esters, lecithin, etc.) and macro-
Emulsifying/foaming Thickening/waterholding molecular emulsi®ers (usually proteins, especially from
Temperature sensitive a Temperature insensitive
Strong surfactant binding Weak surfactant binding
milk and eggs). A small-molecule surfactant is a distinctly
amphiphilic molecule having both polar and non-polar
a
That is, the structure and properties of most proteins can change dras- parts. For reasons of tradition and marketing, proteins are
tically when heated above a characteristic `denaturation temperature'. not commonly classi®ed as emulsi®ers, even though they do
often operate as such during food manufacture. So, in prac-
tice, the term `emulsi®er' in the food industry is con®ned
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
Table 2 emulsi®cation, e.g. by modifying fat crystallisation or by
Principal factors affecting oil-in-water emulsion stability interacting with starch.
To confer substantial shelf-life we require the presence of
Droplet-size distribution
a stabiliser. This can be de®ned (Dickinson, 1992) as a
Initially determined by single chemical component (or mixture) conferring long-term
Emulsi®cation equipment emulsion stability, possibly by an adsorption mechanism,
Concentration of emulsi®er
but not necessarily so. Stabilisers are normally bio-
Type of emulsi®er
Oil/water ratio polymersÐproteins or polysaccharides; small-molecule
Other factors (temperature, pH, viscosity) surfactants are not so effective in conferring long-term
stability. The main stabilising action of food polysacchar-
Nature of interfacial adsorbed layer
ides is via viscosity modi®cation or gelation in the aqueous
Determined by continuous phase. Proteins, on the other hand, have a strong
Concentration and type of emulsi®er tendency to adsorb at oil±water interfaces to form stabilis-
Interactions of adsorbed species ing layers around oil droplets, and so they are able to ful®l
Competition between adsorbed species both the emulsifying and the stabilising roles. Emulsions
Nature of continuous aqueous phase can also be stabilised by particles (e.g. casein micelles and
fat crystals).
Rheology, solvent quality, ionic environment, unadsorbed polymers A stable emulsion is one with no discernible change in the
and amphiphiles
size distribution of the droplets, or their state of aggregation,
Nature of dispersed oil phase or their spatial arrangement within the sample vessel, over
the time-scale of observation. This time-scale may vary
Solid/liquid content from hours to months depending on the situation. The
Solubility in continuous phase
dominant mechanisms of instability are gravity creaming,
E. Dickinson / Food Hydrocolloids 17 (2003) 25±39 27

Ostwald ripening, ¯occulation and droplet coalescence high HLB number). The limited emulsifying capacity of
(Dickinson, 1992). Table 2 gives the main factors affecting some biopolymers can be attributed to poor solubility and/
stability. The state of ¯occulation of the droplets is depen- or insuf®cient amphiphilic character to produce rapid and
dent on the interactions between stabilising layers, which in substantial lowering of the interfacial tension during droplet
turn depends on factors such as the biopolymer surface break-up. The well-established improvement in emulsifying
coverage, the layer thickness, the surface charge density, properties of a protein like gluten by degradative enzyme
and the aqueous solution conditions (especially pH, ionic treatment (LarreÂ, Huchet, BeÂrot, & Popineau, 2001) can
strength, and divalent ion content). For a freshly prepared therefore be attributed to the improved solubility and greater
®ne triglyceride oil-in-water emulsion, the most obvious surfactant-type character of the derived peptide fragments.
initial manifestation of instability is creaming, which typi- Irrespective of the degree of amphiphilic character, though,
cally leads on to macroscopic phase separation into separate the large molecular mass of a typical hydrocolloid makes it
discernible regions of cream and serum. This may then be an unlikely candidate for an ingredient of the very highest
followed by droplet coalescence within the cream and emulsifying ability.
`oiling off' at the top of the sample. All of these processes One distinctly identi®able group of surface-active poly-
can be affected by interactions of the hydrocolloid stabiliser, saccharides are the hydrophobically modi®ed cellulose
both in the bulk aqueous phase and at the surface of the derivativesÐnotably methylcellulose and hydroxypropyl
emulsion droplets. methylcellulose. Such amphiphilic biopolymers can cer-
tainly be used as emulsi®ers in their own right to prepare
stable soybean oil-in-water emulsions (Gaonkar, 1991). But
2. What makes a good emulsifying agent or a good the droplets produced are much coarser than those made
emulsion stabilising agent? with low-molecular-mass emulsi®ers or proteins under
similar conditions (Darling & Birkett, 1987). This poorer
For a polymer to be effective as an emulsifying agent, it emulsifying performance is attributable to the relatively
must be surface-active. That is, it must have the capacity to high molecular weight of these modi®ed cellulose polymers.
lower the tension at the oil±water interface, both substan- Similar arguments can be applied in relation to the emulsi-
tially and rapidly when present at the concentrations typi- fying performance of hydrophobically modi®ed starches
cally used during emulsi®cation. Generally speaking, the and polypropylene glycol alginate.
lower the interfacial tension, the greater the extent to The reliability of the published literature on the apparent
which droplets can be broken up during intense shearing surface activity of certain hydrocolloids is undermined
or turbulent ¯ow (Walstra, 1983; Walstra & Smulders, somewhat by the fact that many commercial gum samples
1998). During high-pressure homogenisation, for instance, contain a small amount of protein, either as contaminant or
most of the processes occur on time-scales of milliseconds as an intrinsic part of the molecular structure. As this pro-
or less: droplet deformation, emulsi®er adsorption, emulsi- teinaceous material is typically strongly hydrophobic, it can
®er spreading, and droplet collision. To retain small droplets adsorb strongly at liquid interfaces, thereby giving an erro-
during emulsi®cation, the time between droplet collisions neous impression of the intrinsic surface activity of the
should be long compared with the time for emulsi®er to polysaccharide hydrocolloid itself. So, for instance, the
adsorb at the new oil±water interface and to create a tran- stabilising function of xanthan gum in mayonnaise has
sient stabilising layer (not necessarily with full saturation been attributed (Hennock, Rahalkar, & Richmond, 1984)
coverage). Stabilisation of ®ne oil droplets during emulsi®- to its acting as an emulsi®er by co-adsorbing with egg
cation can be ascribed predominantly to the Gibbs±Marangoni yolk at the oil±water interface. This interpretation was
mechanism (Dickinson, 1994), which to be effective requires argued on the basis that a 1 wt% solution of the gum had
adsorption of a water-soluble emulsi®er at a suf®cient surface been found (Prud'homme & Long, 1983) to lower the
concentration to generate substantial interfacial tension tension at the air±water surface by 30 mM m 21 or more!
gradients during close approach of colliding droplets. However, it is likely that the gum sample used in the afore-
For a biopolymer to be surface-active, it clearly must mentioned research contained as much as 5% of contamin-
have amphiphilic character. So, if it is a hydrocolloid, it ating protein (Anderson, 1986). And it is not only in the `old
must contain hydrophobic groups that are numerous enough literature' that one may ®nd disconcertingly confusing
and suf®ciently accessible on a short timescale to enable the information. Reports of highly surface-active xanthan
adsorbing molecules to adhere to and spread out at the inter- samples still persist in some fairly recent publications
face, thereby protecting the newly formed droplets. The (Garti, Slavin, & Aserin, 1999).
required time may be too long for very large macromole- Once an emulsion of small droplets has been successfully
cules, or species that are strongly aggregated in the bulk prepared, considerations of surface activity or interfacial
aqueous phase. Hence, an ideal emulsifying agent, capable tension gradients are no longer relevant. What matters for
of making small droplets, is typically composed of species subsequent long-term stability is how well the molecular
of relatively low molecular mass with good solubility in the characteristics of the adsorbed biopolymer conform to the
aqueous continuous phase (e.g. a water-soluble surfactant of requirements of producing a robust macromolecular barrier
28 E. Dickinson / Food Hydrocolloids 17 (2003) 25±39

at the interface. The physico-chemical processes involved in ² chemically modi®ed starch or cellulose derivatives (e.g.
preventing droplets from aggregating or coalescing are the highly substituted methyl celluloses).
classical colloidal stability mechanisms of steric stabilisa- ² some naturally occurring galactomannan hydrocolloids
tion and electrostatic stabilisation (Dickinson, 1988a,b, (e.g. guar gum, fenugreek gum).
1992). For a biopolymer to be most effective in stabilising ² acetylated pectin from sugar beet.
dispersed particles or emulsion droplets, it should exhibit ² depolymerised citrus pectin.
the following four characteristics: ² possibly any hydrocolloid adsorbing at a hydrophilic (or
amphiphilic) surface (e.g. at an existing adsorbed protein
(i) Strong adsorption. This implies that the amphiphilic layer).
polymer has a substantial degree of hydrophobic charac-
ter (e.g. non-polar side chains or a peptide/protein Question 4: Can the reported surface activity and emul-
moiety) to keep it permanently anchored to the interface. sifying capability of some carbohydrate polymers be attrib-
(ii) Complete surface coverage. This implies that there is uted mainly to the presence of protein, present either as a
suf®cient polymer present to fully saturate the surface. contaminant, or as a covalently linked (or physically asso-
(iii) Formation of a thick steric stabilising layer. This ciated) protein±polysaccharide complex? Answer: In most
implies that the polymer is predominantly hydrophilic cases, probably yes.
and of high molecular weight (10 4 ±10 6 Da) within an In order to justify the bare answers given above, it seems
aqueous medium with good solvent properties. appropriate to discuss a few of the exceptions in more detail.
(iv) Formation of a charged stabilising layer. This We shall begin with the most well known of all the food
implies the presence of charged groups on the polymer hydrocolloid emulsi®ersÐgum arabic.
that contribute to the net repulsive electrostatic inter-
action between particle surfaces, especially at low ionic
strength. (This characteristic is only essential if the poly- 3. Some hydrocolloids with surface activity and
mer layer is not suf®ciently thick.). emulsi®cation properties

Conditions (i)±(iii) cannot be met simultaneously for a 3.1. Gum arabic


homopolymer in which all the chain segments are chemi-
cally similar, and so such uniform polymers do not make The most commonly recognised hydrocolloid emulsi®er
good steric stabilisers (Dickinson, 1988a). What is needed is gum arabic. It is widely used in the soft drinks industry for
for good steric stabilisation is a block copolymer, composed emulsifying ¯avour oils (e.g. orange oil) under acidic condi-
of a small fraction of strongly adsorbing hydrophobic tions. About three-quarters of the gum production comes
segments (to keep the macromolecule permanently attached from the species Acacia senegal, with Acacia seyal provid-
to the surface) and a large fraction of non-adsorbing hydro- ing most of the rest (Thevenet, 1988). Gum arabic is the
philic segments (to stick it away from the surface and confer established benchmark hydrocolloid emulsi®er, and, because
upon the protective layer some appreciable thickness). Such it is a fairly expensive ingredient, there have been many
steric stabilisation, combined with electrostatic stabilisation suggestions in the literature for its replacement by other
(i.e. condition (iv)) if the macromolecule contains ionisable polymeric emulsi®ersÐlike hydrophobically modi®ed
groups, provides a repulsive energy barrier which prevents starch (Trubiano, 1995) or other gums, e.g. mesquite gum
pairs of particles from becoming strongly stuck together at (Vernon-Carter, 1998).
close separations under the strongly attractive in¯uence of The well-known ®lm-forming ability of gum arabic has
ubiquitous van der Waals forces (Dickinson, 1992). long since unambiguously established that this is indeed a
So, let us turn to four basic questions that lie at the heart genuine emulsi®er that confers functionality not by modify-
of the subject of this article. ing the rheology of the aqueous phase but by leading to
Question 1: Does the hydrophilic character of polysac- formation of a macromolecular stabilising layer around oil
charides mean that they have little surface activity at oil± droplets. Hence, the gum can stabilise the ¯avour oil emul-
water (or air±water) interfaces, and so are not useful as sion both as a concentrate and in the form of a diluted
emulsifying agents? Answer: Yes. (carbonated) beverage (Tan, 1990). Nevertheless, the level
Question 2: Can one say then that hydrocolloids do not of surface activity is actually rather low in comparison with
adsorb at surfaces in foods? Answer: Yes, it is true that typical food protein emulsi®ers. To compensate for this in
hydrocolloids do not normally adsorb at hydrophobic sur- generating stable sub-micron-sized droplets, it is necessary
faces¼but with some exceptions. in practice (McNamee, O'Riordan, & O'Sullivan, 1998) to
Question 3: What are the `exceptions'? Answer: For use a rather high gum-to-oil weight ratio, i.e. approximately
instance¼ 1:1, as compared with 1:10 for equivalent protein-stabilised
emulsions. Once formed by adsorption at a macroscopic
² gelatin (a protein!). oil±water interface, the high surface shear viscosity of the
² gum arabic (contains protein!). gum arabic ®lm is little affected by extensive dilution of the
E. Dickinson / Food Hydrocolloids 17 (2003) 25±39 29

have also been identi®ed for other gums, e.g. gum talha
(Underwood & Cheetham, 1994).
Fig. 1(b) shows the high-molecular-mass protein-rich frac-
tion of A. senegal gum adsorbing at the oil±water 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 protein±polysaccharide 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, 10±20 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 0.1±7.5% (Dickinson, Murray,
Stainsby, & Anderson, 1988) suggest a good correlation
Fig. 1. Schematic representation of the wattle blossom model representing
the functionally active component of A. senegal gum (a) in solution and (b)
between the Acacia gum protein content and the limiting
adsorbed at the oil±water interface. Separate hydrophilic carbohydrate long-time interfacial tension, and also between emulsifying
blocks (C) of molecular mass ca. 2 £ 105 Da are attached to the backbone capacity and the initial rate of change of tension with time.
chain of hydrophobic protein (P). Consistent with this generalisation, A. senegal ®lms have been
found to be more elastic and of higher stability than those
aqueous sub-phase (Dickinson, Elverson, & Murray, 1989). composed of A. seyal (Fauconnier et al., 2000). Notwithstand-
The ®lm viscoelasticity is therefore maintained even when a ing that, nitrogen content alone is really no reliable indicator
major proportion of the hydrocolloid has been removed of the effectiveness of Acacia gums for emulsi®cation. Gum
from the aqueous phase in contact with the adsorbed arabic (A. senegal) samples of similar nitrogen content
layer. This is consistent with the observation that only a (,0.3%, i.e. corresponding to ,2% protein) have been
small proportion of gum arabic used in emulsion preparation found (Dickinson, Galazka, & Anderson, 1991a) to exhibit
is actually involved in the stabilisation process. substantial differences in emulsifying capacity and emulsion
Gum arabic (A. senegal) is a complex branched hetero- stability. Furthermore, some samples of A. seyal, having a
polyelectrolyte with a backbone of 1,3-linked b-galactopyr- considerably lower protein content (,0.8%), have actually
anose units and side-chains of 1,6-linked galactopyranose been found to give better emulsion stability than A. senegal
units terminating in glucuronic acid or 4-O-methylglucuro- (Buffo, Reineccius, & Oehlert, 2001).
nic acid residues. It also contains ca. 2% protein, covalently On the other hand, there does appear to be a good correla-
linked to carbohydrate through serine and hydroxyproline tion between emulsion stability and gum arabic average
residues, resulting in a mixture of arabinogalactan±protein molecular weight. It was observed (Dickinson, Galazka, &
complexes, each containing several polysaccharide units Anderson, 1991b) that the 10% fraction of a gum arabic
linked to a common protein core. This so-called `wattle sample corresponding to the highest molecular weight
blossom' model (Connolly, Fenyo, & Vandevelde, 1988; (0.38% N), as separated by gel permeation chromatography,
Fincher, Stone, & Clark, 1983) is shown schematically in could produce a more stable emulsion than the remaining
Fig. 1(a). It has been demonstrated that the gum is a 90% fraction (0.35% N). Additionally, in separate experi-
complex mixture of at least three distinct fractions with ments, when the average molecular mass of a gum arabic
different chemical structures (Williams, Phillips, & Randall, sample (,0.35% N) was gradually reduced from 3:1 £ 105
1990) with a major component containing little or no pro- to 2:2 £ 105 Da by controlled irradiative degradation, the
teinaceous material. The protein element appears to be resulting emulsion stability was correspondingly reduced
mainly associated with a high-molecular-mass fraction (Dickinson, Galazka, & Anderson, 1991c). This trend is
representing less than 30% of the total gum (Vandevelde consistent with the formation of thicker and more effective
& Fenyo, 1985). It is this fraction that appears to be pre- steric stabilising layers by adsorption of polymers of greater
dominantly responsible for the special emulsifying and molecular mass.
stabilising properties of the hydrocolloid (Randall, Phillips,
& Williams, 1988; Ray, Bird, Iacobucci, & Clark, 1995). 3.2. Galactomannans
Such heterogeneity of macromolecular structure and func-
tionality is by no means unique to gum arabic. Molecular A galactomannan is a rather rigid hydrophilic biopolymer
weight fractions with differing emulsi®cation properties with a polymannose backbone and grafted galactose units.
30 E. Dickinson / Food Hydrocolloids 17 (2003) 25±39

found for gum arabic! Garti and Reichman (1994) puri®ed


their guar gum down to 0.8% protein, but still found a simi-
lar degree of surface activity and emulsi®cation 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 oil±water inter-
face is reportedly rather weak and reversible. That is, the
birefringency around the droplets and the associated emul-
Fig. 2. Comparison of the emulsion stabilising ability of depolymerised sion stability is lost on diluting the aqueous phase of the
citrus pectin fractions (70% DE) with that of gum arabic (GA) (Akhtar et
emulsion with water (Garti & Reichman, 1994).
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 Another galactomannan that has received particular
0.2 M. Percentage serum separation is shown for four pectin samples with attention recently is fenugreek gum. Garti, Madar, Aserin,
different degrees of depolymerisation (different values of average molecu- and Sternheim (1997) have reported that puri®ed fenugreek
lar mass, in Da) after 1 month (A), 2 months and 3 months (B). gum has substantial surface activity and that stable emul-
sions of moderately low average droplet size (,3 mm) can
Galactomannans such as guar and locust bean gum are be produced; moreover, physical separation of the con-
widely used in the food industry as thickening, water- taminating protein from the crude gum material did not
holding and stabilising agents. As the carbohydrate structure apparently reduce the measured surface activity. Con-
gives no suggestion of the presence of any signi®cant versely, however, in a separate recent study (Brummer,
proportion of hydrophobic groups, it is generally assumed Cui, & Wang, 2001) where a fenugreek gum sample
that this type of hydrocolloid functions by modifying the containing 2.4% protein was enzymatically treated with
rheological properties of the aqueous phase between the pronase, the ability of the resulting hydrolyzate (containing
dispersed particles or droplets. 0.6% protein, but of the same overall molecular weight) was
Puri®ed guar or locust bean gum is reported not to exhibit found to be substantially reduced. It can therefore be reason-
any surface activity (Gaonkar, 1991). Nevertheless, there ably inferred that the surface activity of fenegreek gum, like
are some statements in the literature (Garti, 1999, 2001; that of gum arabic, is due to the presence of a small fraction
Garti & Reichman, 1993, 1994) that these polymers have of protein closely associated (chemically or physically) with
the capacity to emulsify oils and stabilise fairly coarse emul- the polysaccharide structure.
sions (,10 mm mean droplet size, 3±4 mg m 22 surface
coverage) at a moderately low gum/oil ratio (say, 1:5). It 3.3. Pectin
has been suggested by Garti and Reichman (1994), based on
light microscopy observations of strong birefringency at the Pectins extracted from plant cell walls are commonly
oil±water interface, that these gums stabilise emulsions by used as a gelling and thickening agent in foods. High meth-
forming liquid crystalline layers around the droplets. We are oxyl pectin ($50% DE) forms gels under acidic conditions
reminded here of the suggestion of Friberg (1971) concern- in aqueous media of high sugar content, whereas low meth-
ing the stabilising role of small-molecule emulsi®ers by oxyl pectin forms gels in the presence of calcium ions.
multi-layer lamellar liquid crystals. Nevertheless, it has Typically, commercial pectins extracted from citrus peel
been commented upon before (BergenstaÊhl, 1997; Dickinson, or apple pomace are not effective as emulsifying agents
1986) that the applicability of this stabilisation mechanism irrespective of the degree of esteri®cation. There are, how-
to food emulsions is likely to be limited in practice. ever, some exceptions to this statement.
A lowering of the interfacial tension by guar and locust Highly acetylated pectin from sugar beet has been
bean gums by up to ca. 20 mM m 21 has been reported reported (Dea & Madden, 1986) to be much more surface-
(Garti, 1999; Garti & Reichman, 1994). While these authors active than commercial high-methoxyl or low-methoxyl
are genuinely convinced that this surface activity is an pectins, and hence to be readily capable of producing and
intrinsic property of the polysaccharide itself, and not due stabilising ®ne vegetable oil-in-water emulsions. According
to the presence of contaminating protein, it is fair to say that to Dea and Madden (1986), the molecular origin of the
all such samples probably do contain a residual strongly considerable emulsifying capacity of sugar beet pectin is
bound protein/peptide fraction, which is likely to be the the substantially hydrophobic character of the acetyl groups
origin of much of the apparent emulsifying functionality (2±9%). Nevertheless, it should be noted that the experi-
of galactomannan gums. Based on a study of 10 commercial ments upon which this statement was based were carried out
samples, Anderson (1986) reported an average nitrogen on samples containing .2 wt% protein.
content of 0.57% for guar gum. This corresponds to Even with a low acetyl content (,0.8%), pectins from
,2.5% protein, which is actually higher than typically citrus fruits and apples can also exhibit good surface activity
E. Dickinson / Food Hydrocolloids 17 (2003) 25±39 31

and emulsion stabilising characteristics if the average protein and the other rich in hydrocolloid. Two distinct
molecular weight is reduced to ,80 kDa by severe acid phases also arise in complex coacervation, except that
hydrolysis (Mazoyer, Leroux, & Bruneau, 1999). Based now one phase is rich in the two biopolymers and the
on time-dependent changes in emulsion droplet-size distri- other phase is depleted in both.
bution and serum separation, depolymerised pectin of mole- A useful parameter to quantify the relative strength of the
cular weight 70 kDa and 70% degree of esteri®cation has binary protein±hydrocolloid interaction in dilute solution is
been used to make highly stable ®ne oil-in-water emulsions the cross second virial coef®cient (Ap±h) as may be deter-
at a relatively low gum/oil ratio (1:5) (Akhtar, Dickinson, mined experimentally by static light scattering from a poly-
Mazoyer, & Langendorff, 2002). Emulsion stability was mer mixture (Kaddur & Strazielle, 1986). Positive and
found to be reduced on increasing the pH from 4.7 to 7, negative values of Ap±h are indicative of net repulsive and
or by substantially increasing (or decreasing) the pectin net attractive interactions, respectively. The overall thermo-
molecular weight (Fig. 2). dynamic behaviour of the mixed bipolymer solution
Independent of the amount of depolymerised pectin used, depends on the relative values of Ap±h, Ap±p and Ah±h,
the proportion of the hydrocolloid adsorbed at the oil±water where Ap±p and Ah±h are the pure protein and pure hydro-
interface during emulsi®cation was found to remain roughly colloid virial coef®cients, representing the contributions
constant at ca. 25% (Akhtar et al., 2002). The aqueous from protein±protein and hydrocolloid±hydrocolloid inter-
phases of the original pectin solutions, and the serum layers actions, respectively. The low polymeric entropy of mixing
of the emulsions separated by centrifugation, were both means that, if the protein±hydrocolloid interaction is only
analysed with respect to protein content. This showed that slightly net repulsive (i.e. Ap±h . …Ap±p £ Ah±h †1=2 ), the
a negligible proportion of the original protein content of the system prefers to exist as two separate phases when the
sample (0.6 wt%) was detectable in the aqueous phase after overall biopolymer concentration reaches just a few percent
emulsi®cation. Thus, it appears that the protein fraction of (Grinberg & Tolstoguzov, 1997). Values of Ap±h for many
the hydrocolloid emulsi®er was almost entirely associated combinations of food proteins and hydrocolloids are such
with the adsorbed layer around the droplets. Furthermore, that the system exhibits thermodynamic incompatibility.
the fraction remaining in the serum phase was found to be There are only a very few casesÐe.g. serum albumin 1
quite ineffective for subsequent emulsi®cation. We there- pectin (Semenova, Bolotina, & Dmitrochenko, 1991)Ðfor
fore can infer that the presence of the protein moiety in which genuinely complete miscibility has been established
the depolymerised pectin plays a major functional role in in the non-dilute mixed biopolymer solution. The rest exhi-
enhancing the emulsi®cation properties of pectin following bit some kind of soluble complexation, complex coacerva-
depolymerisation. The analogy with gum arabic is again tion or amorphous co-precipitation, depending on the
quite noteworthy. That is, the hydrophobic protein com- biopolymer structures involved and the solution conditions
ponent is responsible for the surface activity of the hydro- (pH, ionic strength, etc.). At one extreme, with weak or
colloid emulsi®er, acting as the strong anchor point at the speci®c unlike attractive interactions …Ap±h , 0†; we may
oil±water interface, with the hydrophilic polysaccharide have a one-phase non-ideal dilute solution of soluble
chains providing the protective layer that confers effective mixed complexes; and at the other extreme, with strong
steric stabilisation during extended storage. unlike attractions …Ap±h p 0† and stoichiometric release of
microions, we may get a mixed precipitate formed with very
little water present. Strictly speaking, complex coacervation
4. Interactions of hydrocolloids with proteins represents the middle ground of a separated liquid-like
phase of mixed biopolymers, which are strongly hydrated
4.1. Interactions in aqueous solution and osmotically swelled, for the retention of microions
(Dubin, 2001).
The nature and strength of protein±hydrocolloid inter- Unless protein and hydrocolloid carry opposite net
actions in bulk solution and at interfaces have an important charges, complex coacervation generally does not take
in¯uence on the stability properties of food dispersions and place. The phenomenon tends to be suppressed also at low
emulsions (Dickinson, 1993; Dickinson & Euston, 1991). In biopolymer charge densities. Precipitation and/or gelation
aqueous solution, a binary mixture of protein 1 hydro- may occur at high charge densities. The maximum coacer-
colloid can exhibit one of three different equilibrium situa- vation yield occurs for the case of an equal ratio mixture (by
tions (Albertsson, 1971): (a) miscibility, (b) thermodynamic weight) of biopolymers at the pH where they carry equal and
incompatibility, and (c) complex coacervation (or com- opposite charges (Schmitt, Sanchez, Desobry-Banon, &
plexation). Miscibility occurs commonly at low biopolymer Hardy, 1998). The extent to which a particular biopolymer
concentrations, and either incompatibility or coacervation at is involved in Coulombic interactions depends on how far
high concentrations, depending on whether the protein± its isoelectric point pI differs from the solution pH. Most
hydrocolloid interaction is net repulsive or net attractive, food proteins …pIp , 5† form complex coacervates with
respectively. Thermodynamic incompatibility implies the anionic hydrocolloids …pIh , 3† in the intermediate region
separation into two distinct aqueous phases, one rich in of pH where the two macromolecules carry opposite net
32 E. Dickinson / Food Hydrocolloids 17 (2003) 25±39

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 hydrocol-
loid like pectin. At low ionic strength, sulfated hydrocol-
loids of relatively high charged density form fairly strong
reversible complexes with proteins, even at neutral or alka-
line 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-existentÐor the system
may even exhibit thermodynamic incompatibility.
Computer modelling of the electrostatic binding of
human serum albumin (HSA) to the strongly anionic poly-
saccharide hyaluronic acid (HA) has recently been reported
(GrymonpreÂ, Staggemeier, Dubin, & Mattison, 2001).
Under incipient binding conditions …pH . pIp †; these calcu-
lations 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 poly-
electrolyte. The electrostatic interaction energy of the
HSA±HA 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
Fig. 3. Distribution of charge on the surface of bovine serum albumin
the surface of bovine serum albumin (BSA) as viewed from
(BSA) at pH 7 as calculated by Hattori et al. (2001). Pictures (a)±(c) three different directions (Hattori, Kimura, Seyrek, &
show the protein molecule viewed from three different directions. Surface Dubin, 2001). Although the protein carries a net negative
regions coloured red and blue carry local net negative and net positive charge, there is a clearly de®ned patch (size 2±3 nm) of
charges, respectively, and white regions are locally electrically neutral. positive charge on the BSA molecule surface even at pH
(Reproduced with permission from Hattori et al., (2001)).
7. Involvement of this patch in electrostatic binding at pH .
pIp allows BSA to form soluble complexes at low ionic
charges …pIp , pH , pIh †: According to Ledward (1994), strength with dextran sulfate, i-carrageenan and k-carrageenan.
the strength of complexation between protein and polysac- The electrostatic character of the complexes has been
charide depends on the distribution of ionisable groups on demonstrated (Galazka, Smith, Ledward, & Dickinson,
the surface of the protein, the ease of unfolding of the 1999) by their reversible dissociation in the presence of
protein's native structure, and the backbone ¯exibility and added electrolyte (0.02±0.1 M NaCl). At low ionic strength,
charge distribution on the polysaccharide. The extent of the `critical' upper pH for effective complexation with BSA
reversibility of complexation depends on the aqueous is ca. 6.5 for k-carrageenan, as compared with ca. 7 for i-
environment and on the mixing conditions (Tolstoguzov, carrageenan, and .7 for dextran sulfate. The value of the
1986). The tendency towards the non-equilibrium formation protein net charge at the critical pH for complexation has
of an insoluble coacervate is enhanced at low salt concen- been shown (Mattison, Dubin, & Brittain, 1998) to be
trations and at pH values signi®cantly below pIp. proportional to the square root of ionic strength, and to
When both biopolymers carry a net negative charge, depend on polyelectrolyte charge density (Dickinson,
soluble complexes may be produced (Dickinson, 1995b). 1998). With respect to adjustment of pH or ionic strength,
In this case, any electrostatic interaction involves the anio- the relative strengths of the BSA±polysaccharide inter-
nic polyelectrolyte interacting with positively charged local actions correlate well with the relative densities of charged
patches on the protein (Park, Muhoberac, Dubin, & Xia, groups along the different polysaccharide chains: dextran
1992). The importance of such charged patches has long sulfate . i-carrageenan . k-carrageenan.
been recognised in protein chromatography (Regnier, An extreme type of protein±polysaccharide interaction
1987). Increasing the net negative charge on the protein occurs when a covalent linkage is formed between the two
has two effects: (a) it enhances protein±protein electrostatic kinds of biopolymers (Dickinson, 1993). The motivation to
repulsion (i.e. it increases Ap±p) and (b) it reduces protein± construct such a mixed biopolymer conjugate might be the
hydrocolloid attraction by screening the interactions of posi- attempt to generate a new type of `natural' amphiphilic
tively charged groups (i.e. it increases Ap±h). The positively hydrocolloid having emulsi®cation properties as good as
charged groups on a protein (±NH31) interact more strongly or better than gum arabic. Without using any chemicals,
with ±OSO32 groups than with 2CO22 groups. This means covalent protein±polysaccharide conjugates can be prepared
E. Dickinson / Food Hydrocolloids 17 (2003) 25±39 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 protein±polysaccharide 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 preci-
pitate below pH < 5:5; a casein±maltodextrin conjugate
can be prepared (Shepherd, Robertson, & Ofman, 2000)
which has good solubility and is an effective emulsi®er
under acidic conditions.

4.2. Interactions of hydrocolloids with adsorbed protein


layers

Some food hydrocolloids achieve interfacial functionality


in emulsions, not so much by adsorbing directly at the oil±
water interface during emulsi®cation, 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
Fig. 4. Schematic representation of three alternative effects of the adsorp- with hydrocolloids in the adsorbed state than in aqueous
tion of stiff hydrocolloid polymers on the surface of spherical emulsion solution at the same pH and ionic strength (Dickinson,
droplets depending on the hydrocolloid concentration and the nature of
1995a). Rather direct evidence for protein±polysaccharide
the hydrocolloid±protein interaction: (a) a sterically stabilised system, (b)
an emulsion gel, and (c) a system ¯occulated by macromolecular bridging. complexation at the oil±water 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 measure-
ments, 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 …Ap±h p 0† between
adsorbed protein and hydrocolloid present at suf®ciently
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 immo-
bilised 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 hydro-
colloid±protein association is weak …Ap±h , 0†; or if the
hydrocolloid is not present at high enough total concentra-
Fig. 5. In¯uence of i-carrageenan on the state of ¯occulation of BSA-stabi- tion to saturate all the surfaces of the droplets. Bridging
lised emulsions (20 vol% oil, 1.7 wt% protein, pH 6, ionic strength
¯occulation of emulsion droplets may also be induced by
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 adhesion of oppositely charged colloidal particles or even
made emulsion. (Reproduced with permission from Dickinson and bacterial cells (Li, McClements, & McLandsborough,
Pawlowsky (1997)). 2001).
34 E. Dickinson / Food Hydrocolloids 17 (2003) 25±39

emulsions or enhancement of protein layer surface shear


viscosity at the oil±water interface.
High-methoxyl pectin adsorbs on the surface of casein-
coated 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 in¯uence 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
Fig. 6. In¯uence of high-methoxyl pectin on the rheology of pure casein- emulsions for ch $ 1 wt% (see Fig. 6) is attributable to
stabilised 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
signi®cant attractive protein±polysaccharide interaction at
hydrocolloid concentration for emulsions made separately with as1-casein the surface of the droplets. Associative interfacial inter-
(W) and b-casein (X). (After Dickinson et al. (1998)). action has been con®rmed at the macroscopic oil±water
interface by a ®ve-fold increase in the surface shear viscos-
Bridging ¯occulation of a BSA-stabilised emulsion has ity of an adsorbed as1-casein layer on introducing pectin into
been found to be induced by addition of i-carrageenan under the aqueous sub-phase. This behaviour is thermodynami-
conditions of soluble complex formation in aqueous solu- cally consistent with the negative cross second virial coef®-
tion (Dickinson & Pawlowsky, 1997). Fig. 5 shows the cient Ap±h derived from light scattering for an aqueous
effect of added hydrocolloid concentration ch on the average solution of as1-casein 1 pectin at pH 5.5 and ionic
apparent particle size d32 of a BSA-stabilised emulsion with strength 0.01 M (Semenova et al., 1999). In contrast, the
an aqueous phase of pH 6 and low ionic strength. In the value of Ap±h is positive for the equivalent aqueous solution
emulsion sample without added hydrocolloid, or in samples of b-casein 1 pectin, and the viscosity of the (non-¯occu-
with ch # 1023 wt%, the low value of d32 < 0:5 mm corre- lated) pectin-containing b-casein-stabilised emulsion is
sponds to the actual sizes of individual dispersed droplets. very much lower than that of the equivalent as1-casein
However, for ch $ 5 £ 1023 wt%, there is a jump in the system, as shown in Fig. 6.
apparent particle size by more than an order of magnitude With certain hydrocolloids, interfacial complexation is
due to bridging ¯occulation of the BSA-coated droplets by more likely to have its origin in hydrogen bonding rather
i-carrageenan. This high value of d32 < 10 mm is a rough than in electrostatic interactions. This seems especially the
measure of the average ¯oc size in emulsions with hydro- case with the uniquely hydrophilic protein gelatin which has
colloid added in the concentration range 0.01±0.1 wt%. the ability to form gel-like layers at a range of different
Partial restabilisation at higher i-carrageenan contents is surfaces (Izmailova et al., 2001). In particular, a thin layer
indicated by the considerably reduced values of d32 for of gelatin network may accumulate at a surface already
ch . 0:1 wt%. At this ionic strength (0.005 M), changing coated with another more surface-active biopolymer.
the aqueous phase conditions to pH 7 leads to a protein± Consider, for example, the mixed system of b-casein
hydrocolloid interaction that is too weak to induce (pI < 5) 1 acid processed type-A gelatin (pI < 9) at neutral
signi®cant ¯occulation (Dickinson & Pawlowsky, 1996). pH. Whereas surface tension experiments indicate that the
However, upon addition of the more highly charged (and primary interfacial layer is totally dominated by the much
therefore more strongly protein-associating) dextran sulfate more surface-active b-casein, measurements of surface
instead of i-carrageenan, substantial bridging ¯occulation shear viscosity are at least an order of magnitude higher
was found to occur at pH 7 (and higher). for b-casein 1 gelatin than for b-casein alone, indicating
Protein±polyelectrolyte interactions are sensitive to the presence of a considerable amount of gelatin at the
details of protein structure as well as to the charge density interfacial region (Galazka & Dickinson, 1995). Due to
on the polyelectrolyte. For instance, in distinct contrast to the opposite net charges of the two biopolymers, the
the behaviour of BSA, the globular protein b-lactoglobulin association between molecules in primary and secondary
(with a similar pIp) shows little evidence of signi®cant layers could reasonably be interpreted solely in terms of
electrostatic complexation with dextran sulfate at neutral electrostatic interaction. However, there is separate
pH and low ionic strength, based on measurements of the evidence available (Castle, Dickinson, Murray, & Stainsby,
electrophoretic mobility of the protein-coated droplets 1987) of accumulation of secondary layers of negatively
(Dickinson, 1995a,b). Furthermore, again in contrast to charged alkali-processed type-B gelatin …pI < 5† below
BSA, the same polyelectrolyte appears to cause no discern- a primary layer of negatively charged casein at neutral
ible bridging ¯occulation of b-lactoglobulin-stabilised pH, again based on surface shear viscometry, but in
E. Dickinson / Food Hydrocolloids 17 (2003) 25±39 35

5. Effect of non-adsorbing hydrocolloids on colloidal


stability

For completeness in this review, it is appropriate to


consider the in¯uence 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 disper-
sions and emulsions by the mechanism of depletion ¯occu-
lation (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
Fig. 7. In¯uence of xanthan gum on the creaming of sodium caseinate-
stabilised emulsions (10 wt% oil, 0.5 wt% protein, pH 7, ionic strength
to ¯ow out from the gap under the in¯uence of the local
0.005 M) stored at 25 8C in 10 cm tubes. The serum layer height H is osmotic pressure gradient (Napper, 1983). At very low
plotted against the storage time t for various concentrations of hydrocolloid concentrations of free polymer, the entropy loss linked to
added to the freshly made emulsion: P, 0 wt% and $0.125 wt%; V, particle aggregation outweighs the depletion effect and the
0.025 wt%; W, 0.05 wt%; L, 0.0625 wt%; O, 0.1 wt%. (Reproduced with system remains stable. However, beyond a certain critical
permission from Cao et al., (1990)).
polymer concentration, which decreases with increasing
particle volume fraction and increasing molecular mass of
this case with sodium caseinate instead of b-casein. Gela- the polymer, reversible depletion ¯occulation occurs, lead-
tin is well known for its thermoreversible gelation via ing to enhanced sedimentation and phase separation. Hence,
cooperative hydrogen-bonded interactions within the a concentrated protein-stabilised emulsion is highly suscep-
collagen-type helices that are essential in providing tible to destabilisation by depletion ¯occulation following
cross-links (Izmailova et al., 2001). Hence, it would addition of a small amount of non-adsorbing hydrocolloid
appear reasonable to speculate that such cooperative …Ap±h . 0†: The emulsion destabilisation is manifest in rapid
intermolecular hydrogen-bonding forces, probably supple- creaming and serum separation (Dickinson, 1993). Hydro-
mented by attractive ionic forces, are mainly responsible colloids inducing strong depletion ¯occulation, however,
for holding together the mixed casein/gelatin interfacial may produce apparent restabilisation with respect to
®lm. serum separation due to the formation of a long-lasting
Thermal or hydrostatic pressure processing of a protein- gel-like network of aggregated droplets (Parker, Gunning,
stabilised emulsion, which induces changes in globular Ng, & Robins, 1995).
protein structure, can also in¯uence protein±hydrocolloid For non-adsorbing macromolecular species present at
interactions at the surface of protein-coated droplets, ®xed volume fraction, it has been shown theoretically
with signi®cant implications for stability and rheology. (Piech & Walz, 2000) that the magnitude of the depletion
Substantial effects of high-pressure treatment on emulsion attraction energy between particles increases with the
properties have recently been found for systems containing degree of non-sphericity of the macromolecular species.
b-lactoglobulin 1 low methoxy pectin (Dickinson & James, Hence, non-adsorbing hydrocolloids with extended and
2000) and ovalbumin 1 sulfated polysaccharides (Galazka, stiff polysaccharide backbones are especially effective at
Dickinson, & Ledward, 2000). Following partial denatura- inducing depletion ¯occulation. This is shown in Fig. 7
tion, the number of available interacting sites on the protein for the case of xanthan gum added at various low concen-
increases as buried groups are liberated (Ledward, 1994). trations to a model protein-stabilised emulsion system
As well as inducing unfolding of the globular proteins, the (10 vol% oil) of sample height 10 cm. The kinetics of
application of high hydrostatic pressure (a few kilobars) destabilisation is indicated by plotting serum layer thickness
leads to disruption of intermolecular ionic bonds and against storage time. The reference emulsion (no added
hence to dissociation of electrostatic complexes. When pres- hydrocolloid) shows no serum separation discernible by
sure is released, the ionic attractive interactions are restored, eye during storage at 25 8C for three days (Cao, Dickinson,
and rapid recomplexation of it occurs, but now between the & Wedlock, 1990). The presence of a very low concentra-
polysaccharide and the partly denatured protein (Galazka, tion of xanthan (0.025 or 0.05 wt%) is enough to induce
Ledward, Sumner, & Dickinson, 1997). The greater mole- depletion ¯occulation, leading to rapid separation (complete
cular ¯exibility of the denatured state allows con®gurational within 1±2 h) of the low-viscosity liquid-like emulsion to
adjustments to occur which maximise favourable ionic form a 65±70% serum layer. The same degree of phase
interactions resulting in more stable complexes than with separation takes about two days to complete at 0.0625 wt%
the native protein. xanthan. Furthermore, there is only ,10% serum separation
36 E. Dickinson / Food Hydrocolloids 17 (2003) 25±39

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 emulsi®ers: (i) non-
ionic Tween 20 and (ii) anionic sodium dodecyl sulfate
(SDS). Without emulsi®ed oil present, there was observed
no signi®cant difference in shear-thinning rheology of the
two mixed hydrocolloid 1 surfactant aqueous phases
(rhamsan 1 Tween or rhamsan 1 SDS); and neither emul-
sion 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
Fig. 8. Rheology of small-molecule surfactant-stabilised emulsions emulsion viscosity is two orders of magnitude higher than
(30 wt% oil, 1 wt% surfactant, pH 7, ionic strength 0.05 M) containing
that of the SDS emulsion. This difference can be attributed
0.1 wt% rhamsan gum incorporated after emulsi®cation. The apparent
shear viscosity h at 25 8C is plotted against the shear stress S on a log± to the different effect of the charged polyelectrolyte on the
log scale for two different surfactants as emulsifying agent: W, Tween 20; electrostatically stabilised SDS-coated emulsion droplets as
X, SDS. (Reproduced with permission from Dickinson et al. (1993)). compared with its effect on the sterically stabilised Tween-
coated droplets (Dickinson et al., 1993).
after three days in the presence of 0.1 wt% xanthan, and There is often found to be a correlation between emulsion
none at all after three days at 0.125 wt% xanthan. Stability rheology and emulsion creaming stability. For example, for
towards cream or serum separation over a period of several the same sets of emulsion systems containing rhamsan and
weeks could be attained by increasing the stabiliser content Tween 20 or SDS, Fig. 9 shows the creaming stability as a
to just 0.25 wt% (Cao, Dickinson, & Wedlock, 1991). function of rhamsan concentration (Dickinson et al., 1993).
The inhibition of creaming at stabiliser concentrations The serum separation time ts is de®ned as the time when a
$0.125 wt% in Fig. 7 can be attributed to immobilisation distinct serum layer becomes ®rst visible to the naked eye.
of the protein-coated oil droplets in a weak gel-like network According to this criterion, whereas the 1% SDS system
with a very high low-stress shear viscosity. Partial stabilisa- becomes unstable to creaming in just one day, the equiva-
tion at 0.0625 and 0.1 wt% xanthan can be mainly attributed lent Tween system is stable for several days. Comparison of
to slow movement of the ¯occulated droplets under gravity Figs. 8 and 9 shows that the higher low-stress viscosity of
due to the high viscosity of the continuous phase (Dickinson, the Tween 20 emulsion correlates well with its better cream-
1993). ing stability. The stronger gel-like network of the concen-
In an emulsion containing a mixture of hydrocolloid and trated Tween-stabilised emulsion (30% oil), as measured by
small-molecule surfactant, the nature of the hydrocolloid± the much higher h value at low S in Fig. 8, can more effec-
surfactant interactions can in¯uence both rheology and tively retard the thermodynamic drive towards depletion-
based 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 rela-
tionships in hydrocolloid-containing emulsion systems.
Finally, it seems worthwhile to stress that ¯occulation
is a ubiquitous phenomenon in hydrocolloid-containing
Fig. 9. In¯uence of rhamsan gum on the creaming of small-molecule surfac- systems. As illustrated by Lips et al. (1991) for a dilute
tant-stabilised emulsions (30 wt% oil, 1 wt% surfactant, pH 7, ionic
model colloid of polystyrene latex particles, a biopolymer
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 hydrocol- will tend either to cause ¯occulation by bridging (at about
loid concentration c: W, Tween 20; X, SDS. (Reproduced with permission half saturation coverage) or by depletion (at rather higher
from Dickinson et al. (1993)). concentrations). Which mechanism actually occurs in any
E. Dickinson / Food Hydrocolloids 17 (2003) 25±39 37

particular system depends sensitively on the nature of the sion stabilizing properties of depolymerized pectin. Food Hydrocol-
hydrocolloid-particle interaction, which in turn is dependent loids, 16.
Albertsson, P. -A. (1971). Partition of cell particles and macromolecules,
on the particle surface properties and the solution condi- New York: Wiley.
tions. In emulsions, the most likely situations favouring Anderson, D. M. W. (1986). The amino acid components of some commer-
destabilisation by ¯occulation (bridging or depletion) are cial gums. In G. O. Phillips, D. J. Wedlock & P. A. Williams (Eds.),
those where the oil volume fraction is low (say, 5±10%) Gums and stabilisers for the food industry (pp. 79±86). Vol. 3. London:
and where the aqueous phase contains a non-gelling hydro- Elsevier.
BergenstaÊhl, B. (1997). Physicochemical aspects of emulsi®er functional-
colloid (or a gelling hydrocolloid at a concentration below ity. In G. L. Hasenhuettl & R. W. Hartel, Food emulsi®ers and their
the gelation threshold). In practice, of course, biopolymer applications (pp. 147±172). New York: Chapman & Hall.
adsorption and ¯occulation effects in concentrated emul- Brummer, Y., Cui, S, Wang, Q. (2001). Extraction, puri®cation and
sions are commonly masked by the very high low-stress physicochemical characterization of fenugreek gum. Poster presenta-
viscosity of the hydrocolloid-containing aqueous phase or tion at Gums and Stabilisers for the Food Industry Conference.
Wrexham, UK, July 2±6.
by hydrocolloid gelation between the droplets.
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, 53±66.
6. Concluding remarks Cao, Y., Dickinson, E., & Wedlock, D. J. (1990). Creaming and ¯occula-
tion in emulsions containing polysaccharide. Food Hydrocolloids, 4,
Some food hydrocolloids are suf®ciently surface-active in 185±195.
their own right to act as sole emulsifying agents during the Cao, Y., Dickinson, E., & Wedlock, D. J. (1991). In¯uence of polysacchar-
formation of oil-in-water emulsions. In most cases, it ide on the creaming of casein-stabilized emulsions. Food Hydrocol-
loids, 5, 443±454.
appears that the origin of the emulsifying ability is the Castle, J., Dickinson, E., Murray, B. S., & Stainsby, G. (1987). Mixed
presence of proteinaceous material covalently bound or protein ®lms adsorbed at the oil/water interface. American Chemical
physically associated with the carbohydrate polymer. Typi- Society Symposium Series, 343, 118±134.
cally, the emulsions produced are coarser than those that can Connolly, S., Fenyo, J. C., & Vandevelde, M. C. (1988). Effect of a protein-
be made with low-molecular-weight water-soluble surfac- ase on the macromolecular distribution of Acacia senegal gum. Carbo-
hydrate Polymers, 8, 23±32.
tants or milk proteins at the same emulsi®er/oil ratio. Once Dalgleish, D. G., & Hollocou, A. -L. (1997). Stabilization of protein-based
strongly adsorbed at the oil±water interface, however, emulsions by means of interacting polysaccharides. In E. Dickinson &
hydrocolloids can be even more effective than surfactants B. BergenstaÊhl, Food colloids: proteins, lipids and polysaccharides
or proteins in conferring long-term stability on emulsions. (pp 236±244). Cambridge, UK: Royal Society of Chemistry.
This is due to the very favourable attributes of hydrophili- Darling, D. F., & Birkett, R. J. (1987). Food colloids in practice. In
E. Dickinson, Food emulsions and foams (pp. 1±29). London: Royal
city and large molecular size in connection with the ef®-
Society of Chemistry.
ciency of steric stabilisation. Hydrocolloids are especially Dea, I. C. M., & Madden, J. K. (1986). Acetylated pectic polysaccharides of
useful in practice for stabilisation of emulsions under the sugar beet. Food Hydrocolloids, 1, 71±88.
acidic conditions where milk proteins are much less effective. Dickinson, E. (1986). Mixed proteinaceous emulsi®ers: review of compe-
Hydrocolloids have an important role, both positive and titive protein adsorption and the relationship to food colloid stabiliza-
tion. Food Hydrocolloids, 1, 3±23.
negative, in relation to the storage stability of food emul-
Dickinson, E. (1988a). The role of hydrocolloids in stabilizing particulate
sions. Of course, this may simply be by modi®cation of the dispersions and emulsions. In G. O. Phillips, D. J. Wedlock & P. A.
small-deformation rheology of the continuous aqueous Williams, (Eds.), Gums and stabilisers for the food industry (pp. 249±
phase. But, commonly also, the hydrocolloid in¯uences 263). , Vol. 4. Oxford, UK: IRL Press.
the state of ¯occulation of the droplets via either a bridging Dickinson, E. (1988b). The structure and stability of emulsions. In J. M. V.
Blanshard & J. R. Mitchell, Food structureÐits creation and evalua-
or depletion mechanism, depending on whether the hydro-
tion (pp. 41±57). London: Butterworths.
colloid has a net attractive or repulsive interaction with the Dickinson, E. (1992). An introduction to food colloids, Oxford, UK:
surface. As the latter typically consists of adsorbed protein, University Press chapter 1.
an understanding of the factors affecting protein±polysac- Dickinson, E. (1993). Protein±polysaccharide interactions in food colloids.
charide interactions in solution seems extremely valuable In E. Dickinson & P. Walstra, Food colloids and polymers: stability and
mechanical properties (pp. 77±93). Cambridge, UK: Royal Society of
for understanding and controlling the physico-chemical
Chemistry.
phenomena involved. Even small changes in weak attractive Dickinson, E. (1994). Emulsions and droplet size control. In D. J. Wedlock,
or repulsive interactions between different biopolymers in a Controlled particle, droplet and bubble formation (pp. 191±216).
colloidal system can have a profound effect on physical Oxford, UK: Butterworth-Heinemann.
stability with respect to creaming, sedimentation or serum Dickinson, E. (1995a). Mixed biopolymers at interfaces. In S. E. Harding,
S. E. Hill & J. R. Mitchell, Biopolymer mixtures (pp. 349±372). Leices-
separationÐwith drastic consequences for product quality
tershire, UK: Nottingham University Press.
and appearance. Dickinson, E. (1995b). Emulsion stabilization by polysaccharides and
protein±polysaccharide complexes. In A. M. Stephen, Food polysac-
charides and their applications (pp. 501±515). New York: Dekker.
References Dickinson, E. (1998). Stability and rheological implications of electrostatic
milk protein±polysaccharide interactions. Trends in Food Science and
Akhtar, M., Dickinson, E., Mazoyer, J., & Langendorff, V. (2002). Emul- Technology, 9, 347±354.
38 E. Dickinson / Food Hydrocolloids 17 (2003) 25±39

Dickinson, E., & Euston, S. R. (1991). Stability of food emulsions contain- effects of pH, ionic strength and high pressure treatment. Food Chem-
ing both protein and polysaccharide. In E. Dickinson, Food polymers, istry, 64, 303±310.
gels and colloids (pp. 132±146). Cambridge, UK: Royal Society of Galazka, V. B., Dickinson, E., & Ledward, D. A. (2000). Emulsifying
Chemistry. properties of ovalbumin in mixtures with sulfated polysaccharides:
Dickinson, E., & Galazka, V. B. (1991). Emulsion stabilization by ionic and effects of pH, ionic strength, heat and high-pressure treatment. Journal
covalent complexes of b-lactoglobulin with polysaccharides. Food of the Science of Food and Agriculture, 80, 1±11.
Hydrocolloids, 5, 281±296. Gaonkar, A. G. (1991). Surface and interfacial activities and emulsion
Dickinson, E., & Galazka, V. B. (1992). Emulsion stabilization by protein/ characteristics of some food hydrocolloids. Food Hydrocolloids, 5,
polysaccharide complexes. In G. O. Phillips, D. J. Wedlock & P. A. 329±337.
Williams, (Eds.), Gums and stabilisers for the food industry (pp. 351± Garti, N. (1999). What can nature offer from an emulsi®er point of view:
362), Vol. 6. Oxford, UK: IRL Press. trends and progress? Colloids and Surfaces A, 152, 125±146.
Dickinson, E., & James, J. D. (2000). In¯uence of high-pressure treatment Garti, N. (2001). Food emulsi®ers and stabilizers. In N. A. M. Eskin & D. S.
on b-lactoglobulin±pectin associations in emulsions and gels. Food Robinson, Food shelf life stability (pp. 211±263). Boca Raton, FL: CRC
Hydrocolloids, 14, 365±376. Press.
Dickinson, E., & Pawlowsky, K. (1996). Rheology as a probe of protein± Garti, N., & Reichman, D. (1993). Hydrocolloids as food emulsi®ers and
polysaccharide interactions in oil-in-water emulsions. In G. O. Phillips, stabilizers. Food Microstructure, 12, 411±426.
P. A. Williams & D. J. Wedlock, (Eds.), Gums and stabilisers for the Garti, N., & Reichman, D. (1994). Surface properties and emulsi®cation
food industry (pp. 181±191), Vol. 8. Oxford, UK: IRL Press. activity of galactomannans. Food Hydrocolloids, 8, 155±173.
Dickinson, E., & Pawlowsky, K. (1997). Effect of õ-carrageenan on ¯occu- Garti, N., Madar, Z., Aserin, A., & Sternheim, B. (1997). Fenugreek
lation, creaming and rheology of a protein-stabilized emulsion. Journal galactomannans as food emulsi®ers. Food Science and Technology,
of Agricultural and Food Chemistry, 45, 3799±3806. 30, 305±311.
Dickinson, E., Murray, B. S., Stainsby, G., & Anderson, D. J. (1988). Garti, N., Slavin, Y., & Aserin, A. (1999). Surface and emulsi®cation
Surface activity and emulsifying behaviour of some Acacia gums. properties of a new gum extracted from Portulaca oleracea L. Food
Food Hydrocolloids, 2, 477±490. Hydrocolloids, 13, 145±155.
Dickinson, E., Elverson, D. J., & Murray, B. S. (1989). On the ®lm-forming Grinberg, V. Y., & Tolstoguzov, V. B. (1997). Thermodynamic incompat-
and emulsion-stabilizing properties of gum arabic: dilution and ¯occu- ibility of proteins and polysaccharides in solutions. Food Hydrocol-
lation aspects. Food Hydrocolloids, 3, 101±114. loids, 11, 145±148.
Dickinson, E., Galazka, V. B., & Anderson, D. M. W. (1991a). Emulsifying GrymonpreÂ, K. R., Staggemeier, B. A., Dubin, P. L., & Mattison, K. W.
behaviour of gum arabic. 1. Effect of the nature of the oil phase on the (2001). Identi®cation by integrated computer modelling and light
emulsion droplet-size distribution. Carbohydrate Polymers, 14, 373± scattering studies of an electrostatic serum albumin±hyaluronic acid
383. binding site. Biomacromolecules, 2, 422±429.
Dickinson, E., Galazka, V. B., & Anderson, D. M. W. (1991b). Emulsifying Hattori, T., Kimura, K., Seyrek, E., & Dubin, P. L. (2001). Binding of
behaviour of gum arabic. 2. Effect of the gum molecular weight on the bovine serum albumin to heparin determined by turbidometric titration
emulsion droplet-size distribution. Carbohydrate Polymers, 14, 385± and frontal analysis continuous capillary electrophoresis. Analytical
392. Biochemistry, 295, 158±167.
Dickinson, E., Galazka, V. B., & Anderson, D. M. W. (1991c). Effect of Hennock, M., Rahalkar, R. R., & Richmond, P. (1984). Effect of xanthan
molecular weight on the emulsifying behaviour of gum arabic. In gum upon the rheology and stability of oil-in-water emulsions. Journal
E. Dickinson, Food polymers, gels and colloids (pp. 490±493). of Food Science, 49, 1271±1274.
Cambridge, UK: Royal Society of Chemistry. Islam, A. M., Phillips, G. O., Sljivo, A., Snowden, M. J., & Williams, P. A.
Dickinson, E., Goller, M. I., & Wedlock, D. J. (1993). Creaming and (1997). A review of recent developments on the regulatory, structural
rheology of emulsions containing polysaccharide and non-ionic or and functional aspects of gum arabic. Food Hydrocolloids, 11, 493±
anionic surfactants. Colloids and Surfaces A, 75, 195±201. 505.
Dickinson, E., Semenova, M. G., Antipova, A. S., & Pelan, E. G. (1998). Izmailova, V. N., Yampolskaya, G. P., Levachev, S. M., Derkatch, S. R.,
Effect of high-methoxy pectin on properties of casein-stabilized emul- Tulovskaya, Z. D., & Voronko, N. G. (2001). Bulk and interfacial sol±
sions. Food Hydrocolloids, 12, 425±432. gel transitions in systems containing gelatin. In E. Dickinson & R.
Dubin, P. L. (2001). Microscopic and macroscopic phase transitions in Miller, Food colloids: fundamentals of formulation (pp. 376±383).
polyelectrolyte±micelle systems. In Y. Iwasawa, N. Oyama & H. Cambridge, UK: Royal Society of Chemistry.
Kunieda, (Eds.), Studies in surface science and catalysis (pp. 979± Jayme, M. L., Dunstan, D. E., & Gee, M. L. (1999). Zeta potentials of gum
984), Vol. 132. Amsterdam: Elsevier. arabic stabilized oil-in-water emulsions. Food Hydrocolloids, 13, 459±
Fauconnier, M. -L., Blecker, C., Groyne, J., Raza®ndralambo, H., Vanze- 465.
veren, E., Marlier, M., & Paquot, M. (2000). Characterization of two Kaddur, L. O., & Strazielle, C. (1986). Experimental investigation of light
Acacia gums and their fractions using a Langmuir ®lm balance. Journal scattering by a solution of two polymers. Polymer, 28, 459±468.
of Agricultural and Food Chemistry, 48, 2709±2712. Kato, A., Sasaki, Y., Furuta, R., & Kobayashi, K. (1990). Functional
Fincher, G. B., Stone, B. A., & Clarke, A. E. (1983). Arabinogalactan- protein/polysaccharide conjugate prepared by controlled dry heating
proteins: structure, biosynthesis and function. Annual Review of Plant of ovalbumin/dextran mixtures. Agricultural and Biological Chemistry,
Physiology, 34, 47±70. 54, 107±112.
Friberg, S. (1971). Liquid crystalline phases in emulsions. Journal of LarreÂ, C., Huchet, B., BeÂrot, S., & Popineau, Y. (2001). Functional proper-
Colloid and Interface Science, 37, 291±295. ties of peptides derived from wheat storage proteins by limited enzym-
Galazka, V. B., & Dickinson, E. (1995). Surface properties of protein layers atic hydrolysis and ultra®ltration. In E. Dickinson & R. Miller, Food
adsorbed from mixtures of gelatin with various caseins. Journal of colloids: fundamentals of formulation (pp. 262±271). Cambridge, UK:
Texture Studies, 26, 401±409. Royal Society of Chemistry.
Galazka, V. B., Ledward, D. A., Sumner, I. G., & Dickinson, E. (1997). Ledward, D. A. (1994). Protein±polysaccharide interactions. In N. S.
In¯uence of high pressure on bovine serum albumin and its complex Hettiarachchy & G. R. Ziegler, Protein functionality in food systems
with dextran sulfate. Journal of Agricultural and Food Chemistry, 45, (pp. 225±259). New York: Dekker.
3465±3471. Li, J., McClements, D. J., & McLandsborough, L. A. (2001). Interaction
Galazka, V. B., Smith, D., Ledward, D. A., & Dickinson, E. (1999). between emulsion droplets and Escherichia coli cells. Journal of Food
Complexes of bovine serum albumin with sulfated polysaccharides: Science, 66, 570±574.
E. Dickinson / Food Hydrocolloids 17 (2003) 25±39 39

Lips, A., Campbell, I. J., & Pelan, E. G. (1991). Aggregation mechanisms in Schmitt, C., Sanchez, C., Desobry-Banon, S., & Hardy, J. (1998). Structure
food colloids and the role of biopolymers. In E. Dickinson, Food poly- and technofunctional properties of protein±polysaccharide complexes.
mers, gels and colloids (pp. 1±21). Cambridge, UK: Royal Society of Critical Reviews in Food Science and Nutrition, 38, 689±753.
Chemistry. Semenova, M. G., Bolotina, V. S., & Dmitrochenko, A. P. (1991). The
Mattison, K. W., Dubin, P. L., & Brittain, I. J. (1998). Complex formation factors affecting the compatibility of serum albumin and pectinate in
between bovine serum albumin and strong polyelectrolytes: effect of poly- aqueous medium. Carbohydrate Polymers, 15, 367±385.
mer charge density. Journal of Physical Chemistry B, 102, 3830±3836. Semenova, M. G., Antipova, A. S., Belyakova, L. E., Dickinson, E., Brown,
Mazoyer, J., Leroux, J., Bruneau, G., (1999). Use of depolymerized citrus R., Pelan, E. G., & Norton, I. T. (1999). Effect of pectinate on properties
fruit and apple pectins as emulsi®ers and emulsion stabilizers. US of oil-in-water emulsions stabilised by as1-casein and b-casein. In
Patent No. 5,900,268. E. Dickinson & J. M. Rodriguez Patino, Food emulsions and foams:
McNamee, B. F., O'Riordan, E. D., & O'Sullivan, M. (1998). Emulsi®ca- interfaces, interactions and stability (pp. 163±175). Cambridge, UK:
tion and encapsulation properties of gum arabic. Journal of Agricultural Royal Society of Chemistry.
and Food Chemistry, 46, 4551±4555. Shepherd, R., Rockey, J., Sutherland, I. W., & Roller, S. (1995). Novel
Napper, D. H. (1983). Polymeric stabilization of colloidal dispersions, bioemulsi®ers from microorganisms for use in foods. Journal of
London: Academic Press. Biotechnology, 40, 207±217.
Navon-Venezia, S., Zosim, Z., Gottlieb, A., Legmann, R., Carmeli, S., Ron, Shepherd, R., Robertson, A., & Ofman, D. (2000). Dairy glycoconjugate
E. Z., & Rosenberg, E. (1995). Alsan, a new bioemulsi®er from Acine- emulsi®ers: casein±maltodextrins. Food Hydrocolloids, 14, 281±286.
tobacter radioresistens. Applied and Environmental Microbiology, 61, Tan, C. -T. (1990). Beverage emulsions. In K. Larsson & S. E. Friberg,
3240±3244. Food emulsions (2nd ed.) (pp. 445±478). New York: Dekker.
Park, J. M., Muhoberac, B. B., Dubin, P. L., & Xia, J. (1992). Effects of Thevenet, F. (1988). Acacia gums. American Chemical Society Symposium
protein charge heterogeneity in protein±polyelectrolyte complexation. Series, 370, 37±44.
Macromolecules, 25, 290±295. Tolstoguzov, V. B. (1986). Functional properties of protein±polysacchar-
Parker, A., Gunning, P. A., Ng, K., & Robins, M. M. (1995). How does ide mixtures. In D. A. Ledward & J. R. Mitchell, Functional properties
xanthan stabilize salad dressing? Food Hydrocolloids, 9, 333±342. of food macromolecules (pp. 385±415). London: Elsevier.
Piech, M., & Walz, J. Y. (2000). Analytical expressions for calculating the Trubiano, P. C. (1995). The role of specialty food starches in ¯avour emul-
depletion interaction produced by charged spheres and spheroids. Lang- sions. American Chemical Society Symposium Series, 610, 199±209.
muir, 16, 7895±7899. Underwood, D. R., & Cheetham, P. S. J. (1994). The isolation of active
Prud'homme, R. K., & Long, R. E. (1983). Surface tensions of concentrated emulsi®er components by the fractionation of gum talha. Journal of the
xanthan and polyacrylamide solutions with added surfactants. Journal Science of Food and Agriculture, 66, 217±224.
of Colloid and Interface Science, 93, 274±276. Vandevelde, M. -C., & Fenyo, J. -C. (1985). Macromolecular distribution
Randall, R. C., Phillips, G. O., & Williams, P. A. (1988). The role of the of Acacia senegal gum (gum arabic) by size-exclusion chromatography.
proteinaceous component on the emulsifying properties of gum arabic. Carbohydrate Polymers, 5, 251±273.
Food Hydrocolloids, 2, 131±140. Vernon-Carter, E. J. (1998). Stability of Capsicum annuum oleoresin-in-
Randall, R. C., Phillips, G. O., & Williams, P. A. (1989). Fractionation and water emulsions containing Prosopis and Acacia gums. Journal of
characterization of gum from Acacia senegal. Food Hydrocolloids, 3, Texture Studies, 29, 553±567.
65±75. Walstra, P. (1983). Formation of emulsions. In P. Becher, (Ed.), Encyclo-
Ray, A. K., Bird, P. B., Iacobucci, G. A., & Clark Jr., B. C. (1995). Func- pedia of emulsion technology (pp. 57±127), Vol. 1. New York: Dekker.
tionality of gum arabic: fractionation, characterization and evaluation Walstra, P., & Smulders, P. A. E. (1998). Emulsion formation. In B. P.
of gum fractions in citrus oil emulsions and model beverages. Food Binks, Modern aspects of emulsion science (pp. 56±99). Cambridge,
Hydrocolloids, 9, 123±131. UK: Royal Society of Chemistry.
Regnier, F. E. (1987). The role of protein structure in chromatographic Ward-Smith, R. S., Hey, M. J., & Mitchell, J. R. (1994). Protein/polysac-
behaviour. Science, 238, 319±323. charide interactions at the oil±water interface. Food Hydrocolloids, 8,
Ron, E. Z., & Rosenberg, E. (2001). Natural roles of biosurfactants. 309±315.
Environmental Microbiology, 3, 229±236. Williams, P. A., Phillips, G. O., & Randall, R. C. (1990). Structure±
Rosenberg, E., & Ron, E. Z. (1999). High-and low-molecular-mass micro- function relationships of gum arabic. In G. O. Phillips, D. J. Wedlock
bial surfactants. Applied Microbiology and Biotechnology, 52, 154± & P. A. Williams, (Eds.), Gums and stabilisers for the food industry
162. (pp. 25±36), Vol. 5. Oxford, UK: IRL Press.

You might also like