Sei sulla pagina 1di 13

Food Hydrocolloids xxx (2017) 1e13

Contents lists available at ScienceDirect

Food Hydrocolloids
journal homepage: www.elsevier.com/locate/foodhyd

Hydrocolloids acting as emulsifying agents e How do they do it?


Eric Dickinson
School of Food Science and Nutrition, University of Leeds, Leeds, LS2 9JT, UK

a r t i c l e i n f o a b s t r a c t

Article history: We consider the essential compositional and molecular structural features controlling the formation and
Received 11 December 2016 stabilization of oil-in-water emulsions by hydrocolloid ingredients. The theoretical principles underlying
Received in revised form adsorption of polymers and steric stabilization by polymer layers are outlined with particular reference
12 January 2017
to copolymer morphology and chain branching. These basic concepts are used to interpret at the mo-
Accepted 19 January 2017
lecular level the experimental interfacial functionality of three classes of biopolymer emulsifying agents:
Available online xxx
gum arabic, pectin, and hydrophobically modified starch (OSA starch). Some inferences are made con-
cerning the mechanistic significance of certain generic features in relation to the surface activity and
Keywords:
Polysaccharide
emulsification characteristics of hydrocolloids d the presence of covalently bound protein, the diversity
Emulsifying agent of carbohydrate polymer structure, and the heterogeneity of natural ingredients.
Steric stabilization © 2017 Elsevier Ltd. All rights reserved.
OSA starch
Acacia gum
Sugar beet pectin

Foreword spirit, this researcher d like many others d has been enticed and
welcomed into the extended hydrocolloid research family. So it is a
This author's personal journey to try to understand hydrocolloid personal privilege and pleasure for me to dedicate this article to
functionality in food emulsions was to a great extent inspired by Professor Phillips on the occasion of his 90th birthday.
the influential research on gum arabic (Acacia senegal) carried out
by Professor Glyn Phillips and his collaborators. The intellectual 1. Introduction
challenges offered to me by this subject area were further sustained
through the stimulating discussions that took place every couple of Hydrocolloids are widely employed in the food industry as
years at the Gums and Stabilisers Conferences in Wrexham. Surely functional ingredients for preparing and stabilizing oil-in-water
one of Glyn's great strengths is his ability to bring together active emulsions. In fact, as judged by the technical literature that
scientists and industrialists in a congenial atmosphere of collabo- commonly accompanies the supply of commercial food ingredients,
ration and discovery. Especially memorable for me in the context of it would appear that almost any available hydrocolloid can provide
the present article are two noteworthy scientific events: a work- protection against emulsion breakdown; but clearly some are more
shop session on Emulsion Stabilization during the 4th Gums and effective than others. Over the 30-year lifetime of the journal Food
Stabilisers Conference (Phillips, Wedlock & Williams, 1988), and, Hydrocolloids there have been numerous published reports on the
some 20 years later, an International Forum on Natural Hydrocol- functionality of natural gums and individual biopolymers in
loid Emulsifiers, also in Wales, organized by Phillips Hydrocolloids emulsified systems. And various authors of these studies have
Research in conjunction with the Japanese ingredient company typically offered some kind of intuitive interpretation of this
San-Ei Gen F.F.I. Inc. (Phillips, 2008b). For observers of his bound- functionality in mechanistic terms. Over this same timescale, there
less enthusiasm and commitment to this field over the past 40 have been regular reports of studies aiming to compare the per-
years, both before and after his formal retirement from NEWI, it is formance of different hydrocolloid ingredients, and the ongoing
difficult to overstate Glyn's contribution to the promotion and value of these comparative studies has been steadily enhanced by
development of our global hydrocolloid research community. Un- the development of reliable experimental methodologies for the
der his compelling spell of Welsh charm and a warm generosity of quantitative characterization of emulsion stability (McClements,
2005, 2007). On the assumption that useful progress has been
made over this period, one may reasonably ask a couple of related
E-mail address: E.Dickinson@leeds.ac.uk. questions. Firstly, what is the current scientific consensus regarding

http://dx.doi.org/10.1016/j.foodhyd.2017.01.025
0268-005X/© 2017 Elsevier Ltd. All rights reserved.

Please cite this article in press as: Dickinson, E., Hydrocolloids acting as emulsifying agents e How do they do it?, Food Hydrocolloids (2017),
http://dx.doi.org/10.1016/j.foodhyd.2017.01.025
2 E. Dickinson / Food Hydrocolloids xxx (2017) 1e13

the underlying mechanistic basis of this type of hydrocolloid Lopez, 2001). For this uniquely hydrophilic food protein, the sur-
functionality? And, secondly, how can one explain the evident face activity is nevertheless attributable to the intrinsic hydro-
differences in emulsification performance between hydrocolloids phobicity of a small proportion of its amino-acid residues. Some
of different chemical structure and biological origin? food polysaccharides appear inherently surface-active due to the
A clue to the intrinsic functional identity of the hydrocolloid is, presence of short hydrophobic side-chains such as methyl and
of course, to be found in the name itself d a hydrophilic (poly- acetyl groups. Other carbohydrate polymers contain a small
meric) material with colloid-like character. That is, the hydrocolloid amount of proteinaceous material that is firmly linked to the
is a highly water-soluble (or water-dispersible) material that polysaccharide by covalent bonding or physical complexation. It is
readily dissolves (or disperses) to form highly hydrated entities of supposed that these specific molecular structural features enable
colloidal (nanoscale) dimensions. Each dissolved polymer molecule natural ingredients such as gum arabic (Acacia senegal) and certain
of a hydrocolloid ingredient is deemed to interact strongly via kinds of pectin to function as highly effective emulsifying agents. In
hydrogen bonding with its surrounding water molecules (and also addition, ingredient suppliers have introduced various kinds of
with any neighbouring hydrocolloid molecules). Due to the ten- hydrophobic chemical modifications into starch and cellulose
dency of these large hydrophilic macromolecules to overlap and polymers in order to prepare ‘green’ biopolymer-based surface-
join together into entangled networks and macroscopic gels, most active polymers having favourable emulsification properties. For
food hydrocolloids have the capability to function as viscosity the cases of both the natural and the chemically modified bio-
modifiers and thickeners in aqueous media at relatively low con- polymers, this article will assess how the work of some researchers
centrations. This universal behaviour allows these hydrocolloid has recently contributed to our improved understanding of their
ingredients to make a significant contribution to the provision of interfacial hydrocolloid structureefunction relationships.
long-term stability to aqueous suspensions of solid particles or oil- Once an oil-in-water emulsion has been prepared, the main
in-water emulsions. In particular, the generation of a hydrocolloid- function of the hydrocolloid emulsifier is to act as a steric stabi-
containing medium with a very high viscosity at a limiting low lizing agent to protect the dispersed droplets against aggregation
applied shear stress facilitates effective immobilization of dispersed and coalescence during subsequent processing and storage
oil droplets, thereby inhibiting gravity-induced creaming and any (Dickinson, 1992; McClements, 2005, 2009). There are three main
accompanying serum separation during extended emulsion stor- requirements for effective steric stabilization by interfacial poly-
age. Xanthan gum, for instance, is especially effective in providing mers: complete surface coverage, strong adsorption, and the exis-
stabilization by means of such rheological control (Dickinson, 2004, tence of a thick adsorbed layer. In the first place, there should be
2009b). Furthermore, this same non-adsorbing hydrocolloid poly- sufficient hydrocolloid present during emulsification to fully cover
mer is also a common participant in the closely related phenomena the newly formed oilewater interface. Secondly, the stabilizing
of phase separation and depletion flocculation, with further polymer molecules should remain effectively attached to the
important implications for emulsion serum separation and gravity interface over the life-time of the emulsion system. And, thirdly, the
induced creaming instability (Dickinson, Ma, & Povey, 1994; adsorbed layer should contain some bulky hydrated polymer
Moschakis, Murray, & Dickinson, 2006). chains that can extend well into the aqueous medium. In combi-
To be effective as an emulsifying agent, the hydrocolloid ingre- nation with steric stabilization, the coexistence of a fraction of non-
dient should possess some surface activity. That is, it should have adsorbed polymeric material can make an additional contribution
the capacity to lower the interfacial tension at the oilewater to colloid stability through the depletion stabilization mechanism
interface; and this tension lowering should occur over a timescale (Semenov & Shvets, 2015). Furthermore, the presence of charged
that is relevant to the process of emulsion preparation. In other groups on some hydrocolloid molecules provides the source of a
words, the rate of polymer adsorption at the freshly formed oile- supplementary electrostatic contribution to the stabilization
water interface, and the associated rate of development of the mechanism, especially under conditions of low ionic strength
transient adsorbed layer during emulsification, should exceed the (Dickinson, 1992; McClements, 2005; Semenova & Dickinson,
rate of dropletedroplet collisions caused by the hydrodynamic 2010).
forces of laminar and turbulent flow (Walstra & Smulders, 1998; To understand why one particular hydrocolloid ingredient
Walstra, 2003). In molecular terms, this means that the predomi- might possess superior emulsifying properties to another we must
nantly hydrophilic macromolecules of the hydrocolloid ingredient first have reliable knowledge of the chemical identity and distri-
should contain a sufficient proportion of accessible hydrophobic bution of the hydrophobic groups along the carbohydrate polymer
groups to enable them to stick to and spread out at the interface, chain. In addition, we need to have an appreciation of how the
thereby stabilizing the freshly formed emulsion droplets against molecular architecture of the hydrocolloid affects the interfacial
immediate recoalescence. A further practical consideration is that structure of the adsorbed polymer layer. Finally, we need to
the emulsifying agent should not be prone to strong aggregation or establish how the adsorbed layer structure influences the interac-
gelation; otherwise the resulting high viscosity of the aqueous tion forces between individual hydrocolloid-coated droplets and
phase will inhibit the disruption of the dispersed oil phase into hence the overall colloidal stability of the emulsion system. This
sufficiently small droplets. In practice, these criteria are fully article represents an attempt to assess how close we are to reaching
satisfied by only a small number of hydrocolloid ingredients. such a state of fundamental understanding for some of the main
Consequently, even though essentially all polysaccharide thick- classes of hydrocolloid emulsifying agents. Attention is directed
eners and gelling agents can contribute usefully to the long-term here towards the behaviour of soluble hydrocolloid ingredients that
stability of emulsions after preparation, only a limited number of exist as individual molecular species in aqueous media. The inter-
hydrocolloids are suitable as primary emulsifying agents. ested reader is invited to look elsewhere for information on the
The origin of the surface activity of certain hydrocolloids has emulsion-stabilizing properties of two important varieties of par-
previously been attributed to specific molecular features ticulate biopolymer ingredients d polysaccharide-based solid
(Dickinson, 1988, 2003, 2009a). The only food protein that can particles and hydrocolloid microgels (Dickinson, 2013, 2015a,
properly be described as a hydrocolloid d namely, gelatin d has 2016a,b; Berton-Carabin & Schroe €n, 2015; Lam, Velikov, & Velev,
been found to demonstrate significant surface activity and useful 2014; Rayner et al., 2014; Tavernier, Wijaya, van der Meeren,
emulsifying properties under favourable conditions (Dickinson & Dewettinck, & Patel, 2016).

Please cite this article in press as: Dickinson, E., Hydrocolloids acting as emulsifying agents e How do they do it?, Food Hydrocolloids (2017),
http://dx.doi.org/10.1016/j.foodhyd.2017.01.025
E. Dickinson / Food Hydrocolloids xxx (2017) 1e13 3

The rest of the article is organized as follows. We consider basic


generic principles describing the theory of interfacial stabilization
by model polymers with different sorts of molecular architecture.
Following on from this theoretical overview comes an analysis of
the relevance and applicability of these basic physico-chemical
principles to the known molecular structures and colloid stabiliz-
ing properties of three different hydrocolloid ingredients. The au-
thor's overall objective is to try to illuminate an underlying essence
of physical understanding whilst cautiously recognizing the di-
versity and molecular complexity of the individual systems
involved.

2. Adsorbed polymer layers and colloid stability

Our general understanding of polymer adsorption in relation to


colloid stability has been mainly derived from statistical thermo-
dynamic models of lattice chains adsorbing at solid impenetrable
surfaces. In particular, useful results for specific polymer models
have been generated numerically from the ScheutjenseFleer self-
consistent-field theory (Fleer, 2010; Fleer, Cohen Stuart,
Scheutjens, Cosgrove, & Vincent, 1993; Scheutjens & Fleer, 1985)
and by Monte Carlo simulation (Dickinson & Euston, 1992;
Dickinson & Lal, 1980). Where direct comparisons have been
made between these two alternative approaches, the reported level
of agreement is generally very satisfactory (Chen, Liu, & Hu, 2001;
Sun, Peng, Liu, Hu, & Jiang, 2007). In this article we shall assume the
validity of the commonly accepted view that “the ScheutjenseFleer
theory provides a useful framework for discussing the relative
interaction effects of food biopolymers” (Lips, Campbell, & Pelan,
1991). We also recognize, however, that such a simple statistical
approach is somewhat idealized in relation to the complexity of
food hydrocolloids. For instance, the theory typically assumes that
polymers are composed of flexible linear polymer chains, whereas
the polysaccharide backbone of a hydrocolloid is fairly stiff and is
commonly decorated with side-chains of variable length and
chemical complexity. In addition, polymer adsorption at an oile-
water interface is obviously not the same as polymer adsorption at
an impenetrable solid surface. Despite such differences, however,
there is a reasonably widespread consensus of opinion, supported
by extensive experimental evidence on well-defined synthetic
polymer systems, that the simplified statistical models do provide a
sound mechanistic basis for describing the essential features of
adsorption and colloidal stabilization by all sorts of macromole-
cules, including food biopolymers.
A key message from the statistical models is that steric stabili-
zation is most effectively achieved through the adsorption of co-
polymers. As illustrated schematically in Fig. 1(a), an adsorbing
copolymer molecule combines two key features: anchoring groups Fig. 1. Schematic representations of different kinds of polymer chains in the vicinity of
that stick directly to the surface, and chain regions that extend a planar solid surface: (a) random copolymer; (b) homopolymer (strong adsorption);
away from the surface. The hydrophobic anchoring groups have a (c) homopolymer (weak adsorption); (d) diblock copolymer; (e) triblock copolymer; (f)
branched copolymer. Non-adsorbing (hydrophilic) chain regions are denoted by thin
strong affinity for the ‘oil-like’ surface and a low affinity for the bulk
grey lines; adsorbing (hydrophobic) regions are denoted by thick black lines.
aqueous phase; and vice versa for the extended hydrophilic chains.
Repulsion between two closely approaching surfaces arises from
the entropically unfavourable overlap and compression of their hydrocolloid polymers due to their typically large molecular vol-
copolymer-coated adsorbed layers. Overlap of layers on different umes. Another important thermodynamic condition for effective
surfaces is unfavourable thermodynamically because it produces a stability is that the bulk aqueous phase should be a ‘good’ solvent
local osmotic pressure gradient from the mixing of chains in the for the hydrophilic dangling chains. Any shift in the value of an
overlap zone. Polymer layer compression is also entropically intensive variable like pH and temperature, which results in a
unfavourable because it restricts the volume available to each lowering of the effective solvent quality, is likely to be detrimental
chain, thereby causing a statistical bias away from the (most- to the ability of the adsorbed copolymer to inhibit flocculation.
probable) equilibrium configurations. Experience in the laboratory Homopolymers d that is, polymers which contain segments
has shown that, for an adsorbed layer to provide effective steric that are identical (or chemically very similar) d are not so effective
stabilization and thus prevent flocculation of copolymer-coated as steric stabilizers. This is because they cannot form adsorbed
droplets, it should be at least several nanometres thick. This sort layers of the requisite thickness and structural coherence. Suppose
of adsorbed monolayer thickness is readily achieved with most that all segments of a homopolymer have a strong tendency to stick

Please cite this article in press as: Dickinson, E., Hydrocolloids acting as emulsifying agents e How do they do it?, Food Hydrocolloids (2017),
http://dx.doi.org/10.1016/j.foodhyd.2017.01.025
4 E. Dickinson / Food Hydrocolloids xxx (2017) 1e13

to the surface: this would mean that the configuration produced on bridging interaction between the approaching surfaces is essen-
adsorption at low polymer concentration would be essentially flat, tially negligible.
as shown in Fig. 1(b). In a system with a high concentration of such Let us now consider in more detail how polymer chain archi-
adsorbing homopolymers, there would be a strong tendency to- tecture influences the steric stabilizing ability of a predominantly
wards bulk-phase aggregation and precipitation; and many of the hydrophilic copolymer. For this purpose we refer to numerical re-
polymers would inevitably become stuck to two (or more) droplet sults produced using the ScheutjenseFleer self-consistent-field
surfaces, causing bridging flocculation. Conversely, let us suppose theory for multi-block copolymers with and without side-chains
that the homopolymer consists of segments that are rather hy- (Ettelaie, Murray, & James, 2003). The authors’ basic model as-
drophilic: there would then be good solubility in the aqueous sumes a multi-block linear chain of 512 segments with the generic
medium, but the molecules would only have a weak tendency to chemical formula (AnBn)x (x ¼ 256/n). The type-A segments are
stick to the surface, producing an ill-defined polymer-decorated assumed to be weakly adsorbing, and their attractive interaction
interface with no more than partial surface coverage (see Fig. 1(c)). with the surface is characterized by an arbitrary low value of the
Such a transient and structurally incoherent adsorbed layer lacks FloryeHuggins adsorption energy, i.e. cAs ¼ e1 kT, where kT is the
the capability to confer effective stability to closely approaching thermal energy. This value of cAs is considered to be approximately
droplet surfaces. representative of the attractive energy associated with a single
There are many kinds of copolymers. A random copolymer like small hydrophobic residue. The hydrophilic type-B segments are
that illustrated schematically in Fig. 1(a) has several separate re- taken to be completely non-adsorbing (cBs ¼ 0). The calculated
gions available for anchoring to the surface. Computer simulation steric interaction energy Usteric between two flat plates immersed
has shown that the adsorbed layer structure formed from such a in a 0.1% solution of this multi-block copolymer is shown Fig. 2(a).
multi-block copolymer is influenced by the proportion of sticky The four separate curves of Usteric versus distance d correspond to
segments within the molecule, and also by the arrangement of the
‘stickers’ along the chain (Balazs, Gempe, & Lantman, 1991; Sun,
Peng, Liu, & Hu, 2007). Some real macromolecules can be
adequately described using a polymer chain model having only one
or two sticky anchoring regions. An adsorbed molecule of the
diblock copolymer AeB (with just one sticky type-A region) pref-
erentially adopts the simple trainetail configuration (see Fig. 1(d)).
Under good solvent conditions (for the type-B segments), the
resulting interaction for a pair of closely approaching AeB
copolymer-coated surfaces is purely repulsive. This is highly
favourable for effective steric stabilization. A food biopolymer that
is well approximated by the simple diblock copolymer model is the
milk protein b-casein; in contrast, another milk protein as1-casein
is more successfully modelled as a triblock copolymer (Dickinson,
1998, 2016a; Dickinson, Pinfield, Horne, & Leermakers, 1997). The
adsorption of a triblock copolymer AeBeA (containing two sticky
type-A regions) leads to the traineloopetrain configuration (see
Fig. 1(e)). While repulsion between AeBeA copolymer-coated
surfaces also confers steric stabilization, the resulting colloidal
interaction in this case may become net attractive under certain
conditions (Wijmans, Leermakers, & Fleer, 1994). For a triblock
copolymer with two type-A blocks that are both highly sticky, an
effective surfaceesurface attraction arises from the entropy gain
associated with the configurational change from a single loop on a
single surface to a polymer bridge between two surfaces. Macro-
molecules containing numerous anchoring regions often behave
like simple triblock copolymers. According to the self-consistent-
field calculations of Wijmans et al. (1994), the adsorption of a
simple triblock copolymer AeBeA is qualitatively similar to the
adsorption of a multi-block copolymer containing long blocks,
(AeB)neA (n > 1).
Our discussion so far has concerned linear polymers. But we
know that many carbohydrate polymers possess side-chains, and
that some are composed of long multi-branched chains. The typical
configuration for a branched block copolymer adsorbed on a solid
surface is illustrated schematically in Fig. 1(f). Even though the
branched copolymer may contain just a small proportion of sticky
hydrophobic groups, it can function as a highly effective steric
stabilizer so long as the potential anchoring region is readily Fig. 2. Calculated steric potentials for 0.1% solutions of multiblock copolymers be-
accessible to the surface and can bind sufficiently tightly to cause tween flat plates. The interaction energy Usteric (arbitrary units) is plotted against
permanent copolymer attachment. A high density of dangling hy- distance d between the surfaces (units of monomer lattice size). (a) Linear copolymer
drophilic chains in the gap between two approaching copolymer- (AnBn)x of 512 segments with variable block size: (A) n ¼ 4, (B) n ¼ 8, (C) n ¼ 16, (D)
n ¼ 32. (b) Symmetrical copolymer composed of main multiblock chain of 255 units
coated surfaces produces a strongly repulsive interaction poten- and side-chain of variable size: m ¼ 3 (short dash), m ¼ 15 (dot), m ¼ 31 (solid line),
tial. With just a small part of each copolymer involved in direct m ¼ 47 (dot-dash), m ¼ 63 (long dash). Reproduced with permission from Ettelaie et al.
attachment to the surface, the likelihood of an unfavourable (2003).

Please cite this article in press as: Dickinson, E., Hydrocolloids acting as emulsifying agents e How do they do it?, Food Hydrocolloids (2017),
http://dx.doi.org/10.1016/j.foodhyd.2017.01.025
E. Dickinson / Food Hydrocolloids xxx (2017) 1e13 5

differing block sizes (n ¼ 4, 8, 16 and 32). The shape of each plotted based on corn or maize starch, especially the waxy maize variety
curve Usteric(d) is primarily determined by the changing fractions of (Sweedman et al., 2013).
the weakly adsorbing type-A anchoring blocks as the two surfaces The impressive emulsifying behaviour of OSA starch derived
come close together (Ettelaie et al., 2003). We can observe that each from potato and barley sources was demonstrated by Nilsson and
curve in Fig. 2(a) has an energy minimum at some separation; as Bergenståhl (2006, 2007). Depending on the hydrodynamic con-
the block size increases, the depth of the minimum is reduced, and ditions during emulsification, and also on the individual ingredient
it moves to larger separations. In terms of the stability behaviour of properties (DS value, molar mass), it was found that the adsorbed
bulk colloidal systems with surfaces coated with these same co- amount at the emulsion droplet surface could be rather wide
polymers, as the block size increases one could infer a reduced ranging (from 1 to 2 mg m2 to >10 mg m2). These OSA starch
tendency towards flocculation and hence a greater effectiveness of samples were demonstrated to be surface-active at the macro-
the copolymer as a steric stabilizing agent. scopic triglycerideewater interface, although the reported values of
How does the addition of a hydrophilic side-chain affect the interfacial pressure (at a solution concentration of 100 mg l1) were
form of the steric interaction potential Usteric(d)? To address this rather modest (just 2e3 mN m1) (Nilsson & Bergenståhl, 2007).
question, Ettelaie et al. (2003) considered a symmetrical model Higher interfacial pressures at the tolueneewater interface (up to
copolymer of formula (A4B4)15(A4B3)B(B3A4)(B4A4)15 with a hy- ~10 mN m1) have been reported by Prochaska, Kedziora, Le Thanh,
drophilic side-chain Bm attached to the central type-B segment of and Lewandowicz (2007) for solutions of OSA potato starch
its 255-segment multi-block backbone. The hydrophobic type-A (DS ¼ 0.01).
segments are again taken as weakly adsorbing (cAs ¼ e1 kT), and Despite having the capability to lower the interfacial tension
the type-B segments as non-adsorbing (cBs ¼ 0). The five curves further than other kinds of chemically modified starches, the
plotted in Fig. 2(b) correspond to varying side-chain lengths as overall efficiency of adsorption of OSA starch is still relatively poor
defined by m ¼ 3, 15, 31, 47 and 63 (Ettelaie et al., 2003). For the (Prochaska et al., 2007). The adsorption efficiency is defined as the
case of copolymers with short single side-chains, there is a mini- minimum bulk phase concentration required to form a saturated
mum in Usteric(d) at separation distance d ~ 3 (measured in lattice interface. Although OSA starch can adsorb at the oilewater inter-
units). This would imply a tendency towards incipient flocculation face to produce a densely packed surface layer, the complete
of the corresponding colloidal system, as previously described for saturation of the interface requires the presence of a high hydro-
small-block copolymers with no side-chains (Fig. 2(a)). The depth colloid concentration in the bulk aqueous phase. That is to say, for
of this minimum in Usteric(d) diminishes as m increases. And for a the same total interfacial area created, the required emulsifier/oil
long single side-chain the pair potential becomes entirely positive, ratio has to be substantially higher than that required with, for
implying a purely repulsive steric force between the polymer- instance, a small-molecule surfactant or a food protein. However,
coated surfaces. The results of these calculations show how a once the state of full surface saturation has been reached, the
long hydrophilic side-chain is highly favourable towards steric droplets coated with OSA starch remain highly stable with respect
stabilization by an adsorbed multi-block copolymer even when the to flocculation and coalescence. In comparison with an emulsion
molecule's multi-block backbone sequence would otherwise imply stabilized electrostatically by a monolayer of globular protein (e.g.
flocculating conditions. On the basis of these and related compu- following emulsification with whey protein isolate), steric stabili-
tational results, one could reasonably believe that the presence of zation by a thick OSA starch layer produces an oil-in-water emul-
several extended side-chains on a single adsorbed polymer mole- sion that is more resilient to changes in key environmental
cule would be especially advantageous for copolymer steric stabi- conditions such as temperature, pH and ionic strength (Chanamai &
lizing functionality. This is the situation illustrated schematically in McClements, 2002; Charoen et al., 2011; Mao, Yang, Xu, Yuan, &
Fig. 1(f). Gao, 2010; Qian, Decker, Xiao, & McClements, 2011; Tesch,
Gerhards, & Schubert, 2002).
Native starch is a branched polydisperse homopolymer of
3. Emulsifying properties of three hydrocolloid ingredients glucose having limited water solubility. Two kinds of poly-
saccharides are involved: amylose, a polymer of moderately high
3.1. Hydrophobically modified starch molecular weight (~106 Da) containing just a few long branches,
and amylopectin, a more rigid macromolecule of very high mo-
For more than half a century, starch modified with octenyl lecular weight (107e108 Da) with numerous short branches (Mua &
succinic anhydride (OSA) has been used industrially as a food ad- Jackson, 1997). The number of glycoside residues in native starch
ditive for emulsification, stabilization and encapsulation polymer molecules can therefore range from several hundred
(Sweedman, Tizzotti, Scha €fer, & Gilbert, 2013). The formation of (amylose) to many tens of thousands (amylopectin). However, the
OSA starch involves the esterification reaction between a dicar- degree of polymerization in OSA starch is typically lower than that
boxylic acid anhydride and a small fraction of the hydroxyl groups in the original native starch: the high-molecular-weight fraction of
on carbons 2, 3 and 6 of the glucose units of the carbohydrate some OSA starch samples is degraded during processing by enzy-
polymer chain. A numerical parameter routinely used to define the matic or acid hydrolysis treatments (Hong, Li, Gu, Wang, & Pang,
material chemically is the degree of substitution (DS), i.e. the 2016; Li et al., 2014; Tizzotti, Sweedman, Schaefer, & Gilbert,
average number of octenyl succinate derivatives per glucose unit. 2013), and there is also mechanically-induced polymer chain
Processing conditions that affect the degree of substitution include disruption during emulsification (Nilsson, Leeman, Wahlund, &
factors such as pH, reaction time/temperature, and starch concen- Bergenståhl, 2006). As most of the available commercial samples
tration (Liu et al., 2008). The structural and functional character- of OSA starch are based on botanical sources containing a high
istics of OSA starch are further influenced by natural variations in proportion of amylopectin, it would appear from empirical expe-
polymeric chain length and amylose/amylopectin ratio, as deter- rience that amylopectin is the main component responsible for the
mined by the botanical source of the parent starch (He, Song, Ruan, ingredient's interfacial functionality. In the turbulent flow of the
& Chen, 2006; Song, Zhao, Li, Fu, & Dong, 2013; Sweedman, Hasjim, high-pressure homogenizer, there is rapid convective mass trans-
Sch€afer, & Gilbert, 2014) and the plant's growing conditions (Alvani, port of these big molecules to the newly created oilewater interface
Qi, Tester, & Snape, 2011). Commercially available samples of OSA (Nilsson, Leeman, Wahlund, & Bergenståhl, 2007). And once a
starch, with typical DS values in the range 0.01e0.03, are commonly saturated interfacial layer of OSA starch has been formed, three key

Please cite this article in press as: Dickinson, E., Hydrocolloids acting as emulsifying agents e How do they do it?, Food Hydrocolloids (2017),
http://dx.doi.org/10.1016/j.foodhyd.2017.01.025
6 E. Dickinson / Food Hydrocolloids xxx (2017) 1e13

structural attributes of the amylopectin molecule all contribute to it decays more slowly with separation. However, this notional
its steric stabilization effectiveness d its large molecular size, its advantage of the linear polymer is more than offset by its inferred
high backbone rigidity, and its extensive chain branching (Dokic, propensity to cause bridging flocculation, as indicated by the
Dokic, Dapcevic, & Krstonosic, 2008; Sweedman et al., 2013). presence of an energy minimum for the amylose-like polymer at a
The role of chain branching in relation to the stabilizing separation of ~6 nm. The bridging feature is absent for the branched
behaviour of hydrophobically modified starch has been recently amylopectin-like polymer, whose pair potential is increasingly
analysed using the ScheutjenseFleer self-consistent-field theory strongly repulsive at close separations (Ettelaie et al., 2016). Of
(Ettelaie, Holmes, Chen, & Farshchi, 2016). The authors have course, for any real OSA starch ingredient, the much greater average
compared the behaviour of two disordered polysaccharide models molecular weight of the amylopectin component means that the
each containing 1890 saccharide units: a linear amylose-like poly- bulkier branched polymers will be spatially more dominant at the
mer and branched amylopectin-like model. Whereas the amylose- interface than the linear polymers. Therefore, in reality, one expects
like polymer is assumed to be a flexible linear chain, the hypo- the repulsive range of the steric interaction of the amylopectin
thetical amylopectin-like polymer has 31 branch points with a component of OSA starch to extend out to larger relative separa-
degree of polymerization of 30 units between the branch points. tions than implied by the idealized predictions of Fig. 3. To sum-
Both these look-alike OSA starch polymers are assumed to have the marize this theoretical analysis of these look-alike OSA starch
same degree of substitution (DS ¼ 0.033), each having exactly 63 polymers, we can say that it confirms our previously inferred
hydrophobic groups per chain. The adsorption energy of an OSA expectation that the preferred molecular architecture for effective
side-chain group is set at cAs ¼ e12 kT, corresponding to 1.5 kT for emulsion stabilization is that of a bulky copolymer with extensive
every CH2 group in the side-chain. All other non-modified chain branching.
saccharide units, including branch points, are taken to be non- Like most hydrocolloids, OSA starch is heterogeneous in
adsorbing (cBs ¼ 0). composition. The ingredient is polydisperse, not only in terms of its
Calculated pair potentials for adsorbed layers of these amylose- molecular weight distribution, but also in its extent of chain
like or amylopectin-like polymers of identical molecular weight are branching. This means that, during and after emulsification, there
displayed in Fig. 3. The estimated energy of interaction between the must be some competition between the various adsorbing species
surfaces of a pair of micron-sized oil droplets coated with one or for occupancy of the oilewater interface, resulting presumably in a
other of the two model polymers is plotted against separation mixed biopolymer interface (Dickinson, 2011). An interesting
distance. This calculation of the colloidal interaction energy by observation from the analysis of Ettelaie et al. (2016) is the pre-
Ettelaie et al. (2016) takes account of the attractive van der Waals diction of an enhancement in effectiveness of steric stabilization as
interaction between the droplets as well as the self-consistent-field a consequence of the coexistence of linear and branched polymers
contribution from the overlap of the adsorbed polymer layers. in the adsorbed layer. That is to say, a mixed layer of amylose-like
Predicted pair potentials for both model polymers are consistent and amylopectin-like polymers appears capable of generating a
with a quasi-equilibrium state of colloidal stability: that is, there is steric interaction of longer repulsive range than is achievable with
strong repulsion between the polymer-coated surfaces for separa- either component alone at the same overall bulk concentration. A
tion distances less than ~10 nm. With oil droplets of this size, the somewhat analogous type of interfacial synergism was previously
predicted shallow minima (~2 kT) at larger separations predicted by Parkinson, Ettelaie, and Dickinson (2005) for a model
(15e25 nm) are considered to be small enough to be readily adsorbed layer containing both globular and disordered protein-
overcome by Brownian motion or weak applied shear forces. For like polymers as a way of explaining experimental stability data
these two idealized OSA starch polymers of identical molecular for emulsions made from mixed milk proteins (Parkinson &
weight and the same DS value, Fig. 3 indicates that the adsorbed Dickinson, 2004, 2007).
layer formed by the linear polymer (amylose) is of longer range and Historically speaking, a major industrial motivation behind the
development of OSA starch as a food emulsifier has been its com-
mercial promotion as a viable replacement for gum arabic (Schultz,
2010; Tan, 2004). Despite the lack of any obvious compositional
similarity between OSA starch and gum arabic, it seems pertinent
here to examine in detail what essential structural features they
might have in common.

3.2. Gum arabic

Gum arabic is a tree gum exudate. The natural raw material is


derived predominantly from the Acacia senegal species, although
commercial samples routinely contain gums from other tree spe-
cies, most notably Acacia seyal (Williams & Phillips, 2009). Gum
arabic has a unique status amongst food technologists as the
traditional hydrocolloid emulsifying agent of the food industry. In
particular, it is considered to be the ‘gold standard’ of beverage
emulsion manufacture (Piorkowski & McClements, 2014; Schultz,
2010).
In terms of overall composition, gum arabic is a heterogeneous
Fig. 3. Comparison of colloidal interactions induced by the adsorption of linear and complex polysaccharide containing a small amount of nitrogenous
branched OSA-starch polymers of the same molecular weight (~105 Da) and degree of material that cannot be recovered by purification. Most of the ni-
substitution (DS ~ 0.03). The calculated interaction potential between a pair of oil trogen content is associated with a high-molecular-weight fraction
droplets (diameter 1 mm) is plotted against surface-to-surface separation. Mediated
interactions are calculated for a dilute bulk solution of amylose-like polymer (dashed
(10e30% of the total gum) which is considered to be mainly
line) or amylopectin-like polymer (solid line). Reproduced with permission from responsible for its interfacial functionality in emulsions (Dickinson,
Ettelaie et al. (2016). Galazka, & Anderson, 1991b; Randall, Phillips, & Williams, 1988).

Please cite this article in press as: Dickinson, E., Hydrocolloids acting as emulsifying agents e How do they do it?, Food Hydrocolloids (2017),
http://dx.doi.org/10.1016/j.foodhyd.2017.01.025
E. Dickinson / Food Hydrocolloids xxx (2017) 1e13 7

The solid gum readily dissolves in water to give polyelectrolyte chain (~250 residues) through both serine and hydroxyproline
solutions of low viscosity and a pH around 4.5. There is some residues (Connolly, Fenyo, & Vandevelde, 1988; Mahendran,
rheological evidence for self-association of the molecules in Williams, Phillips, Al-Assaf, & Baldwin, 2008; Osman, Menzies,
aqueous solution (Sanchez, Renard, Robert, Schmitt, & Lefebvre, Williams, Phillips, & Baldwin, 1993). There also exists in the liter-
2002). However, due to the exceptionally compact arrangement ature an alternative structural description of the AGP complex
of its constituent macromolecules and their associated low hy- known as the ‘twisted hairy rope’ model (Alftre n, Pen~ arrieta,
drodynamic molecular volume, gum arabic solutions exhibit close Bergenståhl, & Nilsson, 2012; Qi, Fong, & Lamport, 1991). The ma-
to Newtonian flow behaviour at gum concentrations up to terial corresponding to peak 3, having a protein content of 20e50%,
25e30 wt%. This important functional characteristic enables the is known as the glycoprotein component (GP). Its proteinaceous
ingredient to be employed as an effective emulsifying agent at high region has a substantially different amino-acid sequence from that
gum/oil ratios. of the AG and AGP components; and, unlike the other fractions, it is
The detailed composition of the gum has been established not degraded by proteolytic enzymes.
chromatographically. Separation by gel permeation chromatog- There is a broad scientific consensus that the presence of a small
raphy with refractive index detection has revealed two distinct proportion of covalently bound protein (~2 wt% of Acacia senegal) is
fractions: a main component (peak 1) comprising about 90% of the responsible for conferring surface activity to gum arabic at the
total sample (~2.5  105 Da for Acacia senegal) and a minor oilewater interface, thereby enabling the ingredient to generate
component (peak 2, ~1e2  106 Da) comprising the remainder stabilizing macromolecular films of substantial viscoelasticity
(Williams & Phillips, 2009). A further minor component (peak 3) (Dickinson, 1988; Dickinson, Murray, & Stainsby, 1988; Elmanan,
composed of protein-rich material (~2  105 Da) was revealed by Al-Assaf, Phillips, & Williams, 2008; Erni et al., 2007). Once estab-
chromatographic UV detection to constitute around 1 wt% of the lished at the oilewater interface, the gum arabic adsorbed layer is
total gum. The carbohydrate polymers corresponding to all three characterized by surface shear rheological behaviour that is highly
absorbance peaks have a galactose-linked main chain. There is resilient to dilution of the bulk aqueous phase (Dickinson, Elverson,
extensive chain branching via galactose and arabinose, with & Murray, 1989). The experimental investigation of several
rhamnose and glucuronic acid residues terminating some of the different Acacia gum species with widely variable nitrogen contents
branches (Williams, Phillips, & Stephen, 1990; Williams, Phillips, (0.1e7.5%) has suggested a strong correlation between the amount
Stephen, & Churms, 2006). Based on electron microscopy and of proteinaceous material in the material and its functional prop-
small-angle neutron scattering, it has been established that the erties at the oilewater interface (Dickinson & Euston, 1991;
main arabinogalactan component (AG) (peak 1) has a thin disc-like Dickinson, Murray & Stainsby, 1988; Dickinson, Murray, Stainsby
structure (radius 6.5 nm) (Sanchez et al., 2008). The material cor- & Anderson, 1988). Nevertheless, the overall protein content is
responding to peak 2, with a protein content of ~10%, is known as not necessarily the overriding factor controlling the gum's surface
the arabinogalactaneprotein complex (AGP). The inferred activity or the mechanical properties of its interfacial film. Also of
morphology of this complex, as illustrated schematically in Fig. 4, is considerable importance to its interfacial functionality is the dis-
a compact ‘wattle-blossom’ structure (radius ~36 nm) with disc- tribution of the proteinaceous material between low- and high-
like carbohydrate blocks (~4  104 Da) linked to a polypeptide molecular-weight entities, as well the protein's structural accessi-
bility within the AGP complex for adsorption at the oilewater
interface.
Gum arabic is recognized to be a reliable emulsifier of essential
oils in the presence of added colouring agents under conditions of
low pH and moderately high ionic strength (Tan, 2004). As with
OSA starch, oil-in-water emulsions based on gum arabic can be
stabilized over a wide range of pH, temperature and ionic strength
(Chanamai & McClements, 2002; Charoen et al., 2011). A recent
comparison of emulsification properties determined at high and
low pH values suggests that the ratio of acidic to basic amino acids
plays an important role in relation to the gum's interfacial func-
tionality (Ma, Bell, & Davis, 2015). An especially attractive attribute
of gum arabic to the soft drink manufacturer is the ability to sta-
bilize the flavour oil emulsion as both a concentrate (~20 vol% oil)
and after extensive dilution (say 100 times). Aside from the occa-
sional inconvenience of intermittent supply, its main commercial
disadvantage as an emulsifying agent is that it is regarded as being
a relatively expensive ingredient. This is because of the high gum/
oil ratio (~1:1) that is typically needed to generate uniform fine
droplets (mean diameter ≪ 1 mm) (McNamee, O'Riordan, &
O'Sullivan, 1998; Nakauma et al., 2008). This required usage level
corresponds to a much higher emulsifier concentration than that
typically needed, for instance, with a milk protein ingredient
(typically ~1:10 protein/oil ratio). The relative inefficiency of gum
arabic as an emulsifying agent is a consequence of the fact that only
the protein-containing entities of the mixed ingredient are directly
involved during emulsification (Randall et al., 1988; Ray, Bird,
Fig. 4. Schematic representation of the ‘wattle blossom’ model of the arabinoga-
Iacobucci, & Clark, 1995). Indeed, when the gum's covalently
lactaneprotein complex (AGP) of gum arabic. The postulated structure consists of a
main polypeptide chain with intermittent attachments of short arabinose side-chains
bound protein is deliberately degraded with a proteolytic enzyme,
(AAA) and large blocks of highly branched carbohydrate. Reproduced with permission there is substantial loss of emulsification efficiency. Nevertheless,
from Mahendran et al. (2008).

Please cite this article in press as: Dickinson, E., Hydrocolloids acting as emulsifying agents e How do they do it?, Food Hydrocolloids (2017),
http://dx.doi.org/10.1016/j.foodhyd.2017.01.025
8 E. Dickinson / Food Hydrocolloids xxx (2017) 1e13

the protein content alone is not to be regarded as the single biological process initially involves the formation of sub-units of
definitive indicator of the effectiveness of the interfacial function- low molecular mass and reduced protein content. Later on, as the
ality of Acacia gums. Samples of gum arabic possessing the same acacia tree grows, these sub-units associate into larger AGP ag-
overall protein content (~2%) have been found to exhibit wide gregates. By designing and optimizing physical processes to accel-
variability in their emulsifying properties (Dickinson, Galazka, & erate and enhance this natural aggregation mechanism, Professor
Anderson, 1991a). Phillips and his collaborators have developed a series of ‘Supergum’
The large overall molecular size of the AGP entities and the products (Al-Assaf, Katayama, Phillips, Sasaki, & Williams, 2003;
branched character of their polysaccharide chains are two key Phillips, 2008a). In essence, the patented ‘Supergum’ maturation
structural factors that appear highly favourable towards the pro- technology involves a thermal treatment of the raw dry gum under
vision of effective steric stabilization through the formation of a strictly controlled conditions of temperature and humidity. By
thick adsorbed layer. Fig. 5 shows a sketch of the inferred structural inducing functionally significant changes in the distribution of
character of the gum arabic interfacial layer according to the widely proteinaceous components and carbohydrate sub-units, the emul-
accepted ‘wattle blossom’ model (Evans, Ratcliffe, & Williams, sifying performance of the gum material becomes substantially
2013; Islam, Phillips, Sljivo, Snowden, & Williams, 1997). The hy- enhanced following the thermal maturation process. With respect
drophobic polypeptide chains of the AGP molecules are the sticky to its overall chemical composition and inherent molecular archi-
regions that anchor the macromolecules to the oilewater interface, tecture, however, the industrially matured gum is considered to be
and the compact arrangement of blocks of branched carbohydrate ‘identical’ to that of native gum arabic (Phillips, 2008a).
chains provides a strong steric barrier against droplet flocculation
and coalescence. The relatively high adsorbed amount determined 3.3. Pectin
experimentally on limonene oil droplets (5e6 mg m2) has been
interpreted as indicative of some multilayer adsorption (Evans Pectin belongs to a family of heterogeneous polysaccharides
et al., 2013). Since gum arabic is a polyelectrolyte, we can plau- located in the primary cell wall of plants. It is traditionally used in
sibly envisage the possibility, under pH conditions where many of the food industry as a gelling agent under acidic conditions and at
the glucuronic acid groups are ionized (pH > 4), of significant high sugar contents (Endress, 2009; Voragen, Pilnik, Thibault,
intermolecular association within the adsorbed layer. In addition, Axelos, & Renard, 1995). Whereas interest in the emulsification
charged groups are expected to make a direct electrostatic contri- properties of pectin has a fairly long history (Lotzkar & Maclay,
bution to the repulsive interdroplet pair potential. However, the 1943), it is only in recent times that it has gained acceptance as
zeta potentials reported for acidic beverage emulsion formulations an effective emulsifying agent and emulsion-stabilizing agent (Alba
are generally of a rather modest magnitude (in the range from e 10 & Kontogiorgos, 2016; Ngouemazong, Christiaens, Shpigelman, van
to 20 mV) (Jayme, Dunstan, & Gee, 1999; Ray et al., 1995). So it Loey, & Hendrickx, 2015). In particular, due to its widespread
seems reasonable to assume that any contribution from electro- availability and regulatory acceptability, sugar beet pectin has often
static stabilization is of no more than secondary importance. been put forward as the obvious natural replacement for gum
By applying the simple principle of “giving Nature a helping arabic in beverage emulsions (Schultz, 2010). Certainly, in well-
hand” (Phillips, 2008a), a new generation of ingredients has been defined model systems at pH ¼ 3, sugar beet pectin can be suc-
developed based on mechanistic knowledge of the biological origin cessfully used to prepare fine uniform emulsion droplets (diameter
of the variable functionality of different batches of native gum ~0.5 mm) at a much lower bulk emulsifier concentration than is
arabic (Williams & Phillips, 2009). During normal tree growth in required with gum arabic (Nakauma et al., 2008). But unfortunately
the Sudan, the average molecular mass of the exuded gum arabic these pectin-stabilized emulsions exhibit poorer stability on long-
increases from ~2.5  105 Da (after 5 years) to ~4.5  105 Da (after term storage than the equivalent gum arabic formulations. To
15 years), and the proportions of protein and high-molecular- date, then, in practical situations, pectin is yet to find widespread
weight AGP fraction also increase over same time period (Idris, industrial application as an emulsifying agent due to “its limited
Williams, & Phillips, 1998). It appears that the underlying ability to stabilize emulsions with sufficient robustness in the hard
world of beverage emulsions” (Phillips, 2008a).
Current usage applies the term ‘pectin’ to a group of hetero-
polysaccharides having a backbone composed mainly of D-gal-
acturonic acid bonded with a(1 / 4) glycosidic linkages (Alba &
Kontogiorgos, 2016). Pectin's diversity of polysaccharide structure
arises from wide variations in molecular weight distribution and
extent of esterification, and in the size, nature, and degree of
branching of its neutral side-chains. Further structural modifica-
tions arise as a consequence of chemical and enzymatic trans-
formations during the extraction process (Christiaens et al., 2016).
The most abundant component, homogalacturonan (HG), corre-
sponding to ~65% of the total material, is a linear homopolymer of
~200 galacturonic acid units. Depending on the plant species, some
of the carboxyl groups of HG are esterified with methyl groups and/
or acetyl groups. By convention, samples with extents of methyl
esterification below or above 50% are described, respectively, as low
methyl-esterified (LM) or high methyl-esterified (HM). Pectin is
therefore a polyelectrolyte with a charge distribution that is
dependent on the degree of esterification and on the arrangement
of the methyl and acetyl residues along the chain. The HG polymer
molecules may experience some depolymerization or de-
Fig. 5. Sketch of the adsorbed layer structure of gum arabic at the oilewater interface esterification during acidic extraction of the native material and
according to the widely accepted ‘wattle blossom’ model. its subsequent processing (thermal, mechanical, enzymatic).

Please cite this article in press as: Dickinson, E., Hydrocolloids acting as emulsifying agents e How do they do it?, Food Hydrocolloids (2017),
http://dx.doi.org/10.1016/j.foodhyd.2017.01.025
E. Dickinson / Food Hydrocolloids xxx (2017) 1e13 9

Another type of modification involves replacement of methoxyl evidence for the contribution of acetyl groups to pectin interfacial
groups with amide groups. functionality is possibly a little more compelling (Dea & Madden,
Rhamnogalacturonan I and II (RG-I and RG-II) are two other 1986; Schmidt, Koch et al., 2015): for instance, an increase in
classes of pectic polymers corresponding, respectively, to 20e35% acetylation degree enhances the emulsification behaviour of citrus
and ~10% of the total pectin cell wall material. The polysaccharide pectin. On the other hand, the deacetylation of sugar beet pectin
backbone of RG-I is composed of repeating disaccharide units of does not lead to any significant reduction in its emulsifying ca-
rhamnoseegalacturonic acid (Voragen et al., 1995). The minor pacity, suggesting that the presence of acetyl groups is not essential
component RG-II has a structure like a piece of the HG backbone for its interfacial functionality (Leroux et al., 2003). More
(7e9 units) with four heteropolymeric side-chains attached (Caffall convincingly, perhaps, there is strong evidence for the role of
& Mohnen, 2009). Due to its diverse range of short and polymeric phenolic groups in the emulsifying properties of sugar beet pectin,
side-chains, RG-I is the most branched and structurally heteroge- since measurements of adsorbed fractions show considerable
neous of the pectic polysaccharides. In contrast to the ‘smooth’ enrichment in ferulic acid as well as in protein (Siew & Williams,
character of the HG homopolymer, the RG-I polymer is commonly 2008). Hence, one may reasonably infer that some kind of ferulic
referred to as ‘hairy’ because of its numerous lateral side-chains. acidearabinogalactaneprotein complex plays a key mechanistic
The chemical composition and the heterogeneous branched role in the emulsifying properties of sugar beet pectin (Chen, Fu
structure of the RG-I side-chains are found to vary considerably et al., 2016). However, the lack of any significant correlation be-
from source to source. In the case of sugar beet pectin, some of the tween the overall ferulic acid content and the amount of adsorbed
side-chains are esterified with ferulic acid (Rombouts & Thibault, pectin has been interpreted to imply that the protein content is a
1986). more important indicator of emulsifying capability than the
Owing to its inherent heterogeneity, and the intricate variations phenolic ester content (Funami et al., 2011).
in composition and molecular structure possessed by samples from As with other hydrocolloid emulsifiers, once the constituent
different natural sources, the mechanistic origin of pectin's func- polymeric entities are strongly anchored to the oilewater interface,
tionality is less clearly established than for OSA starch or even gum the long-term stability is provided by repulsive steric interactions
arabic. Chemical analysis of a pectin sample invariably reveals the between the adsorbed pectin layers on different emulsion droplets
presence of some protein, either as a contaminant or an intrinsic (Gromer, Penfold, Gunning, Kirby, & Morris, 2010). An overview
part of the material. The amount of protein detected depends on analysis of the available evidence has led to the suggestion that, for
the extraction conditions and the botanical source d it is as low as the robust maintenance of stability, the constituent pectin poly-
just 1% for apple pectin, but higher for citrus pectin (~3%) and sugar mers should have molar masses in the range 1e2  105 g mol1, and
beet pectin (>5%). The presence of bound proteinaceous moieties there should be a substantial proportion of RG-I polymer chains
acting as hydrophobic groups adsorbing at the oilewater interface (Alba & Kontogiorgos, 2016). The repulsive steric interactions are
is considered to be the main origin of the surface activity and assumed to be enhanced by the presence of neutral sugar side-
emulsifying properties of sugar beet pectin (Funami et al., 2007; chains along the hairy RG-I polysaccharide chain backbone. The
Siew & Williams, 2008; Siew, Williams, Cui, & Wang, 2008; steric stabilizing ability of sugar beet pectin can be further
Williams et al., 2005) and depolymerized citrus pectin (Akhtar, enhanced by inserting ferulic acid cross-links using the oxidative
Dickinson, Mazoyer, & Langendorff, 2002; Leroux, Langendorff, enzyme laccase (Jung & Wicker, 2012).
Schick, Vaishnav, & Mazoyer, 2003). Consistent with this interpre- As a consequence of the ionization of carboxyl groups on
tation is the detection of a greater concentration of protein at the smooth HG polymers, a contribution from electrostatic stabilization
oilewater interface than in the continuous phase of these pectin- in pectin-containing emulsions is to be expected. For any particular
stabilized emulsions. The existence of discrete entities of pro- pectin sample, the charge distribution along the chain is deter-
teinepectin complex has been established directly by atomic force mined by the fraction of methyl groups and their degree of
microscopy (Fishman et al., 2015; Kirby, MacDougall, & Morris, ‘blockiness’. Depending on the detailed molecular structure, there
2008). Furthermore, it has been found that enzymatic degrada- is also the possibility of multilayer adsorption as a consequence of
tion of the protein component of sugar beet pectin leads to a sub- attractive electrostatic interactions between negatively charged
stantial reduction in its interfacial activity and its emulsifying galacturonic acid groups and positively charged protein residues
capacity (Chen, Qiu et al., 2016; Funami et al., 2007). As well as the (Siew et al., 2008). The relative contribution of these electrostatic
overall proportion of protein, two other molecular factors are effects is obviously dependent on the pH and the ionic strength of
presumed to affect pectin interfacial functionality: the distribution the aqueous medium. At pH values well below 3.5 (the pKa of
of the hydrophobic residues along the polypeptide chain, and the galacturonic acid), the uncharged polymer chains can readily self-
accessibility of the protein component for adsorption at the oile- associate, thereby allowing more of the hydrophobic groups to
water interface during emulsification. locate themselves in close proximity to the oilewater interface.
Various alternative explanations for the emulsifying character of This supposedly leads to a thicker and denser adsorbed layer with a
pectin derived from a range of different sources have recently been greater capacity for steric stabilization (Alba, Sagis, & Kontogiorgos,
critically assessed (Alba & Kontogiorgos, 2016; Ngouemazong et al., 2016; Castellani, Al-Assaf, Axelos, Phillips, & Anton, 2010). The
2015). In particular, there is a general recognition of the role of molecular configuration of the adsorbed pectin polymer in this
potentially adsorbing hydrophobic residues on the carbohydrate low-pH environment is illustrated schematically in Fig. 6(a). Under
polymer in the form of methyl and acetyl groups, and also, in the less acidic solution conditions (pH > 3.5), electrostatic repulsion
case of sugar beet pectin, the phenolic groups of ferulic acid. between the anionic carboxylic acid groups causes the polymer
Reducing the molecular weight of citrus pectin by limited acid chains to extend further from the emulsion droplet surface, as
hydrolysis produces an enhancement in interfacial activity and a illustrated in Fig. 6(b). According to Alba and Kontogiorgos (2016),
consequent reduction in average emulsion droplet size (Akhtar the macromolecular configuration shown in Fig. 6(b) is not so
et al., 2002; Leroux et al., 2003). Nevertheless, it appears that the effective in preventing droplet coalescence owing to the reduced
degree of methylation may have only a minor effect on pectin's number of hydrophobic groups attached to the surface and hence
emulsifying properties (Akhtar et al., 2002; Chen, Fu & Luo, 2016; the correspondingly greater likelihood of spontaneous polymer
Schmidt, Schmidt, Kurz, Endress, & Schuchmann, 2015b), whereas desorption.

Please cite this article in press as: Dickinson, E., Hydrocolloids acting as emulsifying agents e How do they do it?, Food Hydrocolloids (2017),
http://dx.doi.org/10.1016/j.foodhyd.2017.01.025
10 E. Dickinson / Food Hydrocolloids xxx (2017) 1e13

Fig. 6. Idealized illustration of the effect of pH on pectin adsorbed layer structure at the surface of an oil droplet. (a) At lower pH (<3.5) the protonation of the carboxylate anions
generates a compact configuration of adsorbed chains with many hydrophobic groups in close proximity to the oilewater interface. (b) At higher pH (>3.5) electrostatic repulsion
between negatively charged carboxylate ions leads to an extended layer of adsorbed chains with fewer hydrophobic groups in contact with the interface. Label P denotes a
proteinaceous region. Reproduced with permission from Alba and Kontogiorgos (2016).

4. Concluding remarks this stabilizing role. Furthermore, the wide diversity of contribu-
tions to biological function in the original plant sources is consis-
Reviewing the emulsifying properties of food hydrocolloids, an tent with the presence on these biopolymers of a range of different
attempt has been made here to address the underlying mechanistic hydrophobic groups potentially available for attachment to oile-
question d how do they do it? We have outlined some basic water interfaces. Two main categories of anchoring species have
principles of polymer adsorption and have assessed the accumu- been clearly identified: (i) hydrophobic groups attached to certain
lated evidence relating to three practically relevant case studies. glycosidic residues along the carbohydrate chains (i.e. octenyl,
Two alternative and apparently conflicting interpretations have methyl, acetyl, phenolic, etc.), and (ii) short regions of covalently
emerged. bound protein or polypeptide. Beyond the case studies highlighted
in this article, one can readily find other instances in the scientific
Answer 1: Interfacial functionality arises from the hydrocolloid's
literature of the involvement of these same two categories of
capacity to form a dense protective adsorbed layer of large
surface-anchoring species. For instance, in analogy with OSA starch,
branched copolymer molecules. In order to confer effective
the incorporation of hydrophobic alkyl side-chains has been re-
long-term steric stabilization on the dispersed emulsion drop-
ported to confer much enhanced emulsifying properties on the
lets, the adsorbed hydrocolloid has to possess a sufficient pro-
predominantly hydrophilic cellulose-based polymer hydrox-
portion of accessible hydrophobic groups to enable its polymer
yethylcellulose (Sun, Sun, Wei, Liu, & Zhang, 2007). Furthermore, in
chains to remain permanently attached to the oilewater
analogy with gum arabic (and pectin), the mechanistic role of
interface.
bound protein has been regularly implicated in the interfacial
Answer 2: Each hydrocolloid ingredient is unique in terms of its functionality of a wide range of naturally occurring hydrocolloids,
biological origin, its compositional variability, and its complex e.g., fenugreek gum (Brummer, Cui, & Wang, 2003), soy soluble
polymeric structure. Accordingly each ingredient confers inter- polysaccharide (Nakamura, Yoshida, Maeda, Furuta, & Corredig,
facial functionality in its own special way. 2004, 2006), mesquite gum (Roman-Guerrero et al., 2009), gum
ghatti (Al-Assaf, Amar, & Phillips, 2008; Kang, Guo, Phillips, & Cui,
2014), corn fibre gum (Yadav, Cooke, Johnston, & Hicks, 2010),
We believe that both these answers are true. That is to say, a
durian seed gum (Amid & Mirhosseini, 2012) and basil seed gum
common generic mechanism applies, but with rather intricate
(Hosseini-Parvar, Osano, & Matia-Merino, 2016; Naji-Tabasi &
macromolecular variations. Reconciliation of the two alternative
Razavi, 2016; Osano, Hosseini-Parvar, Matia-Merino, & Golding,
answers may seem more straightforward and logical when we
2014). The reader should also be made aware of the less well
recognize that, while the functionality of an ingredient like chem-
established hypothesis of Yadav, Igartuburu, Yan, and Nothnagel
ically modified starch is in part a product of human design, the
(2007) that bound lipid species make a predominant contribution
emulsifying properties of gum arabic and pectin are merely the
to gum emulsification activity.
accidental consequences of biological variation. Even for these
There has been an ongoing trend in the food sector towards the
‘accidental’ hydrocolloid emulsifiers, however, there is the oppor-
use of ‘natural’ emulsifiers (Dickinson, 1993; Ozturk & McClements,
tunity for the food scientist “to give nature a helping hand” through
2016). In accordance with perceived consumer preference, it is
the use of controlled thermal processing (Phillips, 2008a) or some
increasingly the manufacturer's ultimate strategy that the pro-
other physical treatment such as high pressure processing (Ma
cessing of ingredients intended for eventual human consumption
et al., 2015).
should not involve any chemically induced modification. In
The long-term steric stabilization of an emulsion droplet re-
contrast, traditional types of physical processing like heating and
quires an appropriate balance between attractive and repulsive
drying are generally considered to be acceptable. An indirect
forces d strong attraction between the droplet surface and some
consequence of the discovery of the mechanistic role of bound
small part of the adsorbed polymer, and strong repulsion between
protein in the functionality of gum arabic has been the steadily
the rest of the polymer and equivalent chains on other droplets.
growing interest in proteinepolysaccharide complexes as novel
Their large molecular size makes hydrocolloids especially suited to

Please cite this article in press as: Dickinson, E., Hydrocolloids acting as emulsifying agents e How do they do it?, Food Hydrocolloids (2017),
http://dx.doi.org/10.1016/j.foodhyd.2017.01.025
E. Dickinson / Food Hydrocolloids xxx (2017) 1e13 11

ingredients for emulsion stabilization. Most notably, the attention appear to provide some modest support and justification for this
of food researchers has been directed towards emulsifying in- intuitive interpretation.
gredients based on Maillard-type proteinepolysaccharide conju- Without any doubt, there is much research still to be done to
gates (Dickinson, 2009a, 2015b; Evans et al., 2013; Lee et al., 2016). resolve some of the outstanding issues raised in this article. We
These covalent conjugates are typically prepared by the controlled trust that those who will take on the challenge of involving
heat treatment of binary mixtures of protein þ polysaccharide themselves in this future work can be inspired by the steady
under conditions of low water activity (Kato, Sasaki, Furuta, & worthwhile progress made so far. Especially noteworthy in this
Kobayashi, 1990). In analogy with the AGP complex of gum latter regard are the many contributions from Professor Glyn
arabic, the essential concept lying behind the preparation of the Phillips and his illustrious band of international collaborators.
Maillard-type conjugate is the formation of an amphiphilic pro-
teinepolysaccharide copolymer (or, more realistically, a heteroge-
References
neous mixture of such copolymers) with the capability to anchor
strongly at the oilewater interface through its proteinaceous Adjonu, R., Doran, G., Torley, P., & Agboola, S. (2014). Whey protein peptides as
element, and furthermore, once having become so attached, to components of nanoemulsions: A review of emulsifying and biological func-
provide effective steric stabilization through its bulky hydrophilic tionalities. Journal of Food Engineering, 122, 15e27.
Akhtar, M., Dickinson, E., Mazoyer, J., & Langendorff, V. (2002). Emulsion stabilizing
polysaccharide component. Where covalent conjugates have a clear properties of deploymerized pectin. Food Hydrocolloids, 16, 249e256.
advantage over electrostatic proteinepolysaccharide complexes is Al-Assaf, S., Amar, V., & Phillips, G. O. (2008). Characterization of gum ghatti and
their ability to maintain permanent macromolecular structural comparison with gum arabic. In P. A. Williams, & G. O. Phillips (Eds.), Gums and
stabilisers for the food industry d 14 (pp. 280e290). Cambridge, UK: Royal So-
integrity over a wide range of pH and electrolyte conditions ciety of Chemistry.
(Dickinson, 2008; Evans et al., 2013; Liu, Ma, Gao, & McClements, Al-Assaf, S., Katayama, T., Phillips, G. O., Sasaki, Y., & Williams, P. A. (2003). Quality
2017; Semenova & Dickinson, 2010). control of gum arabic. Foods & Food Ingredients Journal of Japan, 208, 771e780.
Alba, K., & Kontogiorgos, V. (2016). Pectin at the oilewater interface: Relationship of
In contrast to our esteemed colleagues working in laboratories molecular composition and structure to functionality. Food Hydrocolloids.
of synthetic chemistry or polymer physics, the food scientist/ http://dx.doi.org/10.1016/j.foodhyd.2016.07.026.
technologist is limited to the living natural world for the supply of Alba, K., Sagis, L. M. C., & Kontogiorgos, V. (2016). Engineering of acidic O/W
emulsions with pectin. Colloids and Surfaces B, 145, 301e308.
polymer-based ingredients. Having evolved to perform a specific n, J., Pen
Alftre ~ arrieta, J. M., Bergenståhl, B., & Nilsson, L. (2012). Comparison of
functional role in some individual biological niche, the natural molecular and emulsifying properties of gum arabic and mesquite gum using
macromolecular material has a tendency to be intrinsically com- asymmetrical flow field-flow fractionation. Food Hydrocolloids, 26, 54e62.
Alvani, K., Qi, X., Tester, R. F., & Snape, C. E. (2011). Physico-chemical properties of
plex in its composition and molecular architecture. In the words of
potato starches. Food Chemistry, 125, 958e965.
Professor Glyn Phillips, “natural polymers are never uniform or Amid, B. T., & Mirhosseini, H. (2012). Emulsifying activity, particle uniformity and
simple; their functionality depends on more than one structural rheological properties of a natural polysaccharideeprotein biopolymer from
feature” (Phillips, 2008a). In the context of interfacial functionality, durian seed. Food Biophysics, 7, 317e328.
Balazs, A. C., Gempe, M., & Lantman, C. W. (1991). Effect of molecular architecture
one apparent consequence of this compositional ‘messiness’ is that on the adsorption of copolymers. Macromolecules, 24, 168e176.
the adsorption efficiency of the hydrocolloid is substantially lower Berton-Carabin, C. C., & Schroe €n, K. (2015). Pickering emulsions for food applica-
than that normally found with a conventional synthetic surfactant tions background, trends, and challenges. Annual Review of Food Science and
Technology, 6, 263e297.
or a peptide-based emulsifier (Adjonu, Doran, Torley, & Agboola, Brummer, Y., Cui, W., & Wang, Q. (2003). Extraction, purification and physico-
2014). In practice, what this means is that a higher concentration chemical characterization of fenugreek gum. Food Hydrocolloids, 17, 229e236.
of the emulsifying ingredient is required to make an oil-in-water Caffall, K. H., & Mohnen, D. (2009). The structure, function and biosynthesis of plant
cell wall pectic polysaccharides. Carbohydrate Research, 344, 1879e1900.
emulsion of similar average droplet size. Once prepared, however, Castellani, O., Al-Assaf, S., Axelos, M., Phillips, G. O., & Anton, M. (2010). Hydro-
the hydrocolloid-stabilized emulsion is typically more resilient to colloids with emulsifying capacity. 2. Adsorption properties at the n-hex-
subsequent environmental changes. This enhanced resilience is adecaneewater interface. Food Hydrocolloids, 24, 121e130.
Chanamai, R., & McClements, D. J. (2002). Comparison of gum arabic, modified
especially evident under the severe demands of commercial
starch, and whey protein isolate as emulsifiers: Influence of pH, CaCl2 and
emulsion product formulation d namely, the potentially disturbing temperature. Journal of Food Science, 67, 120e125.
effects of thermal processing, pH adjustment, and elevated salt Charoen, R., Jangchud, A., Jangchud, K., Harnsilawat, T., Naivikul, O., &
McClements, D. J. (2011). Influence of biopolymer emulsifier type on formation
content, and the requirement for physico-chemical compatibility
and stability of rice bran oil-in-water emulsions: Whey protein, gum arabic, and
with other ingredients such as flavours and colour additives. modified starch. Journal of Food Science, 76, 165e172.
Let us end on a more speculative note. Regarding compositional Chen, H.-M., Fu, X., & Luo, Z.-G. (2016). Effect of molecular structure on emulsifying
messiness and structural complexity, we might reasonably be properties of sugar beet pulp pectin. Food Hydrocolloids, 54, 99e106.
Chen, T., Liu, H., & Hu, Y. (2001). Monte Carlo simulation for the adsorption of
tempted to assert that the existence of a complicated mixture of diblock copolymers. Journal of Chemical Physics, 114, 5937e5948.
heterogeneous and polydisperse macromolecules within a single Chen, H., Qiu, S., Gan, J., Liu, Y., Zhu, Q., & Yin, L. (2016). New insights into the
hydrocolloid ingredient ought not to be entirely detrimental to its functionality of protein to the emulsifying properties of sugar beet pectin. Food
Hydrocolloids, 57, 262e270.
interfacial functionality. Indeed, on the contrary, there is much Christiaens, S., van Buggenhout, S., Houben, K., Kermani, Z. J., Moelants, K. R. N.,
empirical and anecdotal evidence to support the view that the Ngouemazong, E. D., et al. (2016). Processestructureefunction relations of
compositional and structural messiness of a macromolecular pectin in food. Critical Reviews in Food Science and Nutrition, 56, 1021e1042.
Connolly, S., Fenyo, J.-C., & Vandevelde, M.-C. (1988). Effect of a proteinase on the
ingredient should be regarded as a positive flexible attribute. Such a macromolecular distribution of Acacia Senegal gum. Carbohydrate Polymers, 8,
tentative hypothesis would be broadly consistent with intuitive 23e32.
theoretical insight which suggests that, so long as the constituents Dea, I. C. M., & Madden, J. K. (1986). Acetylated pectic polysaccharides of sugar beet.
Food Hydrocolloids, 1, 71e88.
remain soluble, and are not prone to extensive aggregation or Dickinson, E. (1988). The role of hydrocolloids in stabilizing particulate dispersions
gelation, the presence of a mixture of macromolecular entities and emulsions. In G. O. Phillips, D. J. Wedlock, & P. A. Williams (Eds.), Gums and
which differ in shape, size and composition should be beneficial for stabilisers for the food industry d 4 (pp. 249e263). Oxford: IRL Press.
Dickinson, E. (1992). An introduction to food colloids. Oxford: University Press.
rapid development of an adsorbed stabilizing layer during emul-
Dickinson, E. (1993). Towards more natural emulsifiers. Trends in Food Science &
sification. Moreover, a heterogeneous system of polymers having Technology, 4, 330e334.
different extents of chain branching could well be imagined to lead Dickinson, E. (1998). Proteins at interfaces and in emulsions: Stability, rheology and
to a structured mixed layer which is optimized for long-term interactions. Journal of the Chemical Society, Faraday Transactions, 94,
1657e1669.
emulsion stability. The predictions of the self-consistent-field cal- Dickinson, E. (2003). Hydrocolloids at interfaces and the influence on the stability of
culations of Ettelaie et al. (2016) highlighted in this article would dispersed systems. Food Hydrocolloids, 17, 25e39.

Please cite this article in press as: Dickinson, E., Hydrocolloids acting as emulsifying agents e How do they do it?, Food Hydrocolloids (2017),
http://dx.doi.org/10.1016/j.foodhyd.2017.01.025
12 E. Dickinson / Food Hydrocolloids xxx (2017) 1e13

Dickinson, E. (2004). Effect of hydrocolloids on emulsion stability. In P. A. Williams, of the proteinaceous moiety on the emulsifying properties of sugar beet pectin.
& G. O. Phillips (Eds.), Gums and stabilisers for the food industry d 12 (pp. Food Hydrocolloids, 21, 1319e1329.
394e404). Cambridge, UK: Royal Society of Chemistry. Gromer, A., Penfold, R., Gunning, A. P., Kirby, A. R., & Morris, V. J. (2010). Molecular
Dickinson, E. (2008). Interfacial structure and stability of food emulsions as affected basis for the emulsifying properties of sugar beet pectin studied by atomic force
by proteinepolysaccharide interactions. Soft Matter, 4, 932e942. microscopy and force spectroscopy. Soft Matter, 6, 3957e3969.
Dickinson, E. (2009a). Hydrocolloids as emulsifiers and emulsion stabilizers. Food He, G.-Q., Song, X.-Y., Ruan, H., & Chen, F. (2006). Octenyl succinic anhydride
Hydrocolloids, 23, 1473e1482. modified early indica rice starches differing in amylose content. Journal of
Dickinson, E. (2009b). Hydrocolloids and emulsion stability. In G. O. Phillips, & Agricultural and Food Chemistry, 54, 2775e2779.
P. A. Williams (Eds.), Handbook of hydrocolloids (2nd ed., pp. 23e49). Cam- Hong, Y., Li, Z., Gu, Z., Wang, Y., & Pang, Y. (2016). Structure and emulsification
bridge, UK: Woodhead. properties of octenyl succinic anhydride starch using acid-hydrolysed method.
Dickinson, E. (2011). Mixed biopolymers at interfaces: Competitive adsorption and Starch/Sta €rke, 68, 1e9.
multilayer structures. Food Hydrocolloids, 25, 1966e1983. Hosseini-Parvar, S. H., Osano, J. P., & Matia-Merino, L. (2016). Emulsifying properties
Dickinson, E. (2013). Stabilizing emulsion-based colloidal structures with mixed of basil seed gum: Effect of pH and ionic strength. Food Hydrocolloids, 52,
food ingredients. Journal of the Science of Food and Agriculture, 93, 710e721. 838e847.
Dickinson, E. (2015a). Microgels d an alternative colloidal ingredient for stabili- Idris, O. H. M., Williams, P. A., & Phillips, G. O. (1998). Characterization of the gums
zation of food emulsions. Trends in Food Science & Technology, 43, 178e188. from Acacia Senegal trees of different age and location using multi-detection gel
Dickinson, E. (2015b). Colloids in food: Ingredients, structure, and stability. Annual permeation chromatography. Food Hydrocolloids, 12, 379e389.
Review of Food Science and Technology, 6, 211e233. Islam, A. M., Phillips, G. O., Sljivo, A., Snowden, M. J., & Williams, P. A. (1997).
Dickinson, E. (2016a). Exploring the frontiers of colloidal behaviour where polymers A review of recent developments on the regulatory, structural and functional
and particles meet. Food Hydrocolloids, 52, 497e509. aspects of gum arabic. Food Hydrocolloids, 11, 493e505.
Dickinson, E. (2016b). Biopolymer-based particles as stabilizing agents for emul- Jayme, M. L., Dunstan, D. E., & Gee, M. L. (1999). Zeta potentials of gum arabic
sions and foams. Food Hydrocolloids. http://dx.doi.org/10.1016/ stabilized oil-in-water emulsions. Food Hydrocolloids, 13, 459e465.
j.foodhyd.2016.06.024. Jung, J., & Wicker, L. (2012). Laccase mediated conjugation of sugar beet pectin and
Dickinson, E., Elverson, D. J., & Murray, B. S. (1989). On the film-forming and the effect on emulsion stability. Food Hydrocolloids, 28, 168e173.
emulsion-stabilizing properties of gum arabic: Dilution and flocculation as- Kang, J., Guo, Q., Phillips, G. O., & Cui, S. W. (2014). Understanding the structur-
pects. Food Hydrocolloids, 3, 101e114. eeemulsification relationship of gum ghatti e a review of recent advances. Food
Dickinson, E., & Euston, S. R. (1991). Stability of food emulsions containing both Hydrocolloids, 42, 187e195.
protein and polysaccharide. In E. Dickinson (Ed.), Food polymers gels and colloids Kato, A., Sasaki, Y., Furuta, R., & Kobayashi, K. (1990). Functional protein/poly-
(pp. 132e146). Cambridge, UK: Royal Society of Chemistry. saccharide conjugate prepared by controlled dry heating of ovalbumin/dextran
Dickinson, E., & Euston, S. R. (1992). Monte Carlo simulation of colloidal systems. mixtures. Agricultural and Biological Chemistry, 54, 107e112.
Advances in Colloid and Interface Science, 42, 89e148. Kirby, A., MacDougall, A., & Morris, V. (2008). Atomic force microscopy of tomato
Dickinson, E., Galazka, V. B., & Anderson, D. M. W. (1991a). Emulsifying behaviour of and sugar beet pectin molecules. Carbohydrate Polymers, 71, 640e647.
gum arabic. 1. Effect of the nature of the oil phase on the emulsion droplet-size Lam, S., Velikov, K. P., & Velev, O. D. (2014). Pickering stabilization of foams and
distribution. Carbohydrate Polymers, 14, 373e383. emulsions with particles of biological origin. Current Opinion in Colloid &
Dickinson, E., Galazka, V. B., & Anderson, D. M. W. (1991b). Emulsifying behaviour of Interface Science, 19, 490e500.
gum arabic. 2. Effect of the gum molecular weight on the emulsion droplet-size Lee, Y.-Y., Tang, T.-K., Phuah, E.-T., Alitheen, N. B. M., Tan, C.-P., & Lai, O.-M. (2016).
distribution. Carbohydrate Polymers, 14, 385e392. New functionalities of maillard reaction products as emulsifiers and encapsu-
Dickinson, E., & Lal, M. (1980). Statistical mechanics of physical adsorption in lating agents, and the processing parameters: A brief review. Journal of the
condensed phases. Advances in Molecular and Related Interaction Processes, 17, Science of Food and Agriculture. http://dx.doi.org/10.1002/jsfa.8124.
1e87. Leroux, J., Langendorff, V., Schick, G., Vaishnav, V., & Mazoyer, J. (2003). Emulsion
Dickinson, E., & Lopez, G. (2001). Comparison of the emulsifying properties of fish stabilizing properties of pectin. Food Hydrocolloids, 17, 455e462.
gelatin and commercial milk proteins. Journal of Food Science, 66, 118e123. Li, Z., Hong, Y., Gu, Z., Tian, Y., Li, Z., & Li, C. (2014). Emulsification properties of
Dickinson, E., Ma, J., & Povey, M. J. W. (1994). Creaming of concentrated oil-in-water enzymatically treated octenyl succinic anhydride starch. Starch/Sta €rke, 66,
emulsions containing xanthan. Food Hydrocolloids, 8, 481e497. 1089e1095.
Dickinson, E., Murray, B. S., & Stainsby, G. (1988). Protein adsorption at airewater Lips, A., Campbell, I. J., & Pelan, E. G. (1991). Aggregation mechanisms in food col-
and oilewater interfaces. In E. Dickinson, & G. Stainsby (Eds.), Advances in food loids and the role of biopolymers. In E. Dickinson (Ed.), Food polymers, gels and
emulsions and foams (pp. 123e162). London: Elsevier Applied Science. colloids (pp. 1e21). Cambridge, UK: Royal Society of Chemistry.
Dickinson, E., Murray, B. S., Stainsby, G., & Anderson, M. W. (1988). Surface activity Liu, Z., Li, Y., Cui, F., Ping, L., Song, J., Ravee, Y., et al. (2008). Production of octenyl
and emulsifying behaviour of some Acacia gums. Food Hydrocolloids, 2, succinic anhydride-modified waxy corn starch and its characterization. Journal
477e490. of Agricultural and Food Chemistry, 56, 11499e11506.
Dickinson, E., Pinfield, V. J., Horne, D. S., & Leermakers, F. A. M. (1997). Self- Liu, F., Ma, C., Gao, Y., & McClements, D. J. (2017). Food-grade covalent complexes
consistent- field modelling of adsorbed casein: Interaction between two and their application as nutraceutical delivery systems: A review. Comprehen-
protein-coated surfaces. Journal of the Chemical Society, Faraday Transactions, 93, sive Reviews in Food Science and Food Safety, 16, 76e95.
1785e1790. Lotzkar, H., & Maclay, W. D. (1943). Pectin as an emulsifying agent. Industrial and
Dokic, P., Dokic, L., Dapcevic, T., & Krstonosic, V. (2008). Colloid characteristics and Engineering Chemistry, 35, 1294e1297.
emulsifying properties of OSA starches. Progress in Colloid and Polymer Science, Ma, F., Bell, A. E., & Davis, F. J. (2015). Effects of high hydrostatic pressure and pH
135, 48e56. treatments on the emulsification properties of gum arabic. Food Chemistry, 184,
Elmanan, M., Al-Assaf, S., Phillips, G. O., & Williams, P. A. (2008). Studies on gum 114e121.
exudates. Part VI: Interfacial rheology of Acacia senegal and Acacia seyal. Food Mahendran, T., Williams, P. A., Phillips, G. O., Al-Assaf, S., & Baldwin, T. C. (2008).
Hydrocolloids, 22, 682e689. New insights into the structural characteristics of the arabinogalactaneprotein
Endress, H.-U. (2009). Pectins. In G. O. Phillips, & P. A. Williams (Eds.), Handbook of (AGP) fraction of gum arabic. Journal of Agricultural and Food Chemistry, 56,
hydrocolloids (2nd ed., pp. 274e297). Cambridge, UK: Woodhead. 9269e9276.
Erni, P., Windhab, E. J., Gunde, R., Graber, M., Pfister, B., Parker, A., et al. (2007). Mao, L., Yang, J., Xu, D., Yuan, F., & Gao, Y. (2010). Effects of homogenization models
Interfacial rheology of surface-active biopolymers: Acacia Senegal gum versus and emulsifiers on the physico-chemical properties of b-carotene nano-
hydrophobically modifed starch. Biomacromolecules, 8, 3458e3466. emulsions. Journal of Dispersion Science and Technology, 31, 986e993.
Ettelaie, R., Holmes, M., Chen, J., & Farshchi, A. (2016). Steric stabilizing properties of McClements, D. J. (2005). Food emulsions (2nd ed.). Boca Raton, FL: CRC Press.
hydrophobically modified starch: Amylose versus amylopectin. Food Hydrocol- McClements, D. J. (2007). Critical review of techniques and methodologies for
loids, 58, 364e377. characterization of emulsion stability. Critical Reviews in Food Science and
Ettelaie, R., Murray, B. S., & James, E. L. (2003). Steric interactions mediated by Nutrition, 47, 611e649.
multiblock copolymers and biopolymers: Role of block size and addition of McClements, D. J. (2009). Biopolymers in food emulsions. In S. Kasapis, I. T. Norton,
hydrophilic side chains. Colloids and Surfaces B, 31, 195e206. & J. B. Ubbink (Eds.), Modern biopolymer science (pp. 129e166). San Diego:
Evans, M., Ratcliffe, I., & Williams, P. A. (2013). Emulsion stabilization using pro- Academic Press.
teinepolysaccharide complexes. Current Opinion in Colloid & Interface Science, McNamee, B. F., O'Riordan, E. A., & O'Sullivan, M. (1998). Emulsification and
18, 272e282. encapsulation properties of gum arabic. Journal of Agricultural and Food
Fishman, M. L., Chau, H. K., Qi, P. X., Hotchkiss, A. T., Garcia, R. A., & Cooke, P. H. Chemistry, 46, 4551e4555.
(2015). Characterization of the global structure of low methoxyl pectin in so- Moschakis, T., Murray, B. S., & Dickinson, E. (2006). Particle tracking using confocal
lution. Food Hydrocolloids, 46, 153e159. microscopy to probe the microrheology in a phase-separating emulsion con-
Fleer, G. J. (2010). Polymers at interfaces and in colloidal dispersions. Advances in taining non-adsorbing polysaccharide. Langmuir, 22, 4710e4719.
Colloid and Interface Science, 159, 99e116. Mua, J. P., & Jackson, D. S. (1997). Fine structure of corn amylose and amylopectin
Fleer, G. J., Cohen Stuart, M. A., Scheutjens, J. M. H. M., Cosgrove, T., & Vincent, B. fractions with various molecular weights. Journal of Agricultural and Food
(1993). Polymers at interfaces. London: Chapman & Hall. Chemistry, 45, 3840e3847.
Funami, T., Nakauma, M., Ishihara, S., Tanaka, R., Inoue, T., & Phillips, G. O. (2011). Naji-Tabasi, S., & Razavi, S. M. A. (2016). New studies on basil (Ocimum bacilicum L.)
Structural modifications of sugar beet pectin and the relationship of structure seed gum. 2. Emulsifying and foaming characterization. Carbohydrate Polymers,
to functionality. Food Hydrocolloids, 25, 221e229. 149, 140e150.
Funami, T., Zhang, G., Hiroe, M., Noda, S., Nakauma, M., Asai, I., et al. (2007). Effects Nakamura, A., Yoshida, R., Maeda, H., & Corredig, M. (2006). Soy soluble

Please cite this article in press as: Dickinson, E., Hydrocolloids acting as emulsifying agents e How do they do it?, Food Hydrocolloids (2017),
http://dx.doi.org/10.1016/j.foodhyd.2017.01.025
E. Dickinson / Food Hydrocolloids xxx (2017) 1e13 13

polysaccharide stabilization at oilewater interfaces. Food Hydrocolloids, 20, Sanchez, C., Schmitt, C., Kolodziejczyk, E., Lapp, A., Gaillard, C., & Renard, D. (2008).
277e283. The acacia gum arabinogalactan fraction is a thin oblate ellipsoid. Biophysical
Nakamura, A., Yoshida, R., Maeda, H., Furuta, H., & Corredig, M. (2004). Study of the Journal, 94, 629e639.
role of the carbohydrate and protein moieties of soy soluble polysaccharides in Scheutjens, J. M. H. M., & Fleer, G. J. (1985). Interaction between two adsorbed
their emulsifying properties. Journal of Agricultural and Food Chemistry, 52, polymer layers. Macromolecules, 18, 1882e1990.
5506e5512. Schmidt, U. S., Koch, L., Rentschler, C., Kurz, T., Endress, H. U., & Schuchmann, H.
Nakauma, M., Funami, T., Noda, S., Ishihara, S., Al-Assaf, S., Nishinari, K., et al. (2008). (2015). Effect of molecular weight reduction, acetylation and esterification on
Comparison of sugar beet pectin, soybean soluble polysaccharide, and gum the emulsification properties of citrus pectin. Food Biophysics, 10, 217e227.
arabic as emulsifiers. 1. Effect of concentration, pH and salts on the emulsifying Schmidt, U. S., Schmidt, K., Kurz, T., Endress, H. U., & Schuchmann, H. P. (2015).
properties. Food Hydrocolloids, 22, 1254e1267. Pectins of different origin and their performance in forming and stabilizing oil-
Ngouemazong, E. D., Christiaens, S., Shpigelman, A., van Loey, A., & Hendrickx, M. in-water-emulsions. Food Hydrocolloids, 46, 59e66.
(2015). The emulsifying and emulsion-stabilizing properties of pectin: A review. Schultz, M. (2010). Industry requirements for hydrocolloids in beverage emulsions.
Comprehensive Reviews in Food Science and Food Safety, 14, 705e718. In P. A. Williams, & G. O. Phillips (Eds.), Gums and stabilisers for the food industry
Nilsson, L., & Bergenståhl, B. (2006). Adsorption of hydrophobically modified starch d 15 (pp. 257e266). Cambridge, UK: Royal Society of Chemistry.
at oil/water interfaces during emulsification. Langmuir, 22, 8770e8776. Semenova, M. G., & Dickinson, E. (2010). Biopolymers in food colloids: Thermody-
Nilsson, L., & Bergenståhl, B. (2007). Emulsification and adsorption properties of namics and molecular interactions. Leiden: Brill.
hydrophobically modified potato and barley starch. Journal of Agricultural and Semenov, A. N., & Shvets, A. A. (2015). Theory of colloid depletion stabilization by
Food Chemistry, 55, 1469e1474. unattached and adsorbed polymers. Soft Matter, 11, 8863e8878.
Nilsson, L., Leeman, M., Wahlund, K. G., & Bergenståhl, B. (2006). Mechanical Siew, C. K., & Williams, P. A. (2008). Role of protein and ferulic acid in the emul-
degradation and changes in conformation of hydrophobically modified starch. sification properties of sugar beet pectin. Journal of Agricultural and Food
Biomacromolecules, 7, 2671e2679. Chemistry, 56, 4164e4171.
Nilsson, L., Leeman, M., Wahlund, K. G., & Bergenståhl, B. (2007). Competitive Siew, C. K., Williams, P. A., Cui, S. W., & Wang, Q. (2008). Characterization of the
adsorption of a polydisperse polymer during emulsification: Experiments and surface-active components of sugar beet pectin and the hydrodynamic thick-
modelling. Langmuir, 23, 2346e2351. ness of the adsorbed pectin layer. Journal of Agricultural and Food Chemistry, 56,
Osano, J. P., Hosseini-Parvar, S. H., Matia-Merino, L., & Golding, M. (2014). Emulsi- 8111e8120.
fying properties of a novel polysaccharide extracted from basil seed (Ocimum Song, X., Zhao, Q., Li, Z., Fu, D., & Dong, Z. (2013). Effects of amylose content on the
bacilicum L.): Effect of polysaccharide and protein content. Food Hydrocolloids, paste properties and emulsification of octenyl succinic starch esters. Starch/
37, 40e48. €rke, 65, 112e122.
Sta
Osman, M. E., Menzies, A. R., Williams, P. A., Phillips, G. O., & Baldwin, T. C. (1993). Sun, L., Peng, C., Liu, H., & Hu, Y. (2007). Influence of polymer structure on
The molecular characterisation of the polysaccharide gum from Acacia Senegal. adsorption behaviour at solideliquid interface by Monte Carlo simulation.
Carbohydrate Research, 246, 303e318. Molecular Simulation, 33, 989e997.
Ozturk, B., & McClements, D. J. (2016). Progress in natural emulsifiers for utilization Sun, L., Peng, C., Liu, H., Hu, Y., & Jiang, J. (2007). Analogy in the adsorption of
in food emulsions. Current Opinion in Food Science, 7, 1e6. random copolymers and homopolymers at solideliquid interface: A Monte
Parkinson, E. L., & Dickinson, E. (2004). Inhibition of heat-induced aggregation of a Carlo simulation study. Journal of Chemical Physics, 126, 094905.
b-lactoglobulin-stabilized emulsion by very small additions of casein. Colloids Sun, W., Sun, D., Wei, Y., Liu, S., & Zhang, S. (2007). Oil-in-water emulsions stabilized
and Surfaces B, 39, 23e30. by hydrophobically modified hydroxyethyl cellulose: Adsorption and thick-
Parkinson, E. L., & Dickinson, E. (2007). Synergistic stabilization of heat-treated ening effect. Journal of Colloid and Interface Science, 311, 228e236.
emulsions containing mixtures of milk proteins. International Dairy Journal, Sweedman, M. C., Hasjim, J., Scha €fer, C., & Gilbert, R. G. (2014). Structures of octenyl
17, 95e103. succinylated starches: Effects on emulsions containing b-carotene. Carbohydrate
Parkinson, E. L., Ettelaie, R., & Dickinson, E. (2005). Using self-consistent-field Polymers, 112, 85e93.
theory to understand enhanced steric stabilization by casein-like copolymers Sweedman, M. C., Tizzotti, M. J., Sch€ afer, C., & Gilbert, R. G. (2013). Structure and
at low surface coverage in mixed protein layers. Biomacromolecules, 6, physico-chemical properties of octenyl succinic anhydride modified starches: A
3018e3029. review. Carbohydrate Polymers, 92, 905e920.
Phillips, G. O. (2008a). Giving nature a helping hand. In P. A. Williams, & Tan, C.-T. (2004). Beverage emulsions. In S. E. Friberg, K. Larsson, & J. Sjo € blom (Eds.),
G. O. Phillips (Eds.), Gums and stabilisers for the food industry d 14 (pp. 3e26). Food emulsions (4th ed., pp. 485e524). New York: Marcel Dekker.
Cambridge, UK: Royal Society of Chemistry. Tavernier, I., Wijaya, W., van der Meeren, P., Dewettinck, K., & Patel, A. R. (2016).
Phillips, G. O. (2008b). International hydrocolloids forum d natural hydrocolloid Food-grade particles for emulsion stabilization. Trends in Food Science & Tech-
emulsifiers. Foods & Food Ingredients Journal of Japan, 213(3 & 4). nology, 50, 159e174.
Phillips, G. O., Wedlock, D. J., & Williams, P. A. (Eds.). (1988). Gums and stabilisers for Tesch, S., Gerhards, C., & Schubert, H. (2002). Stabilization of emulsions by OSA
the food industry d 4 (pp. 507e522). Cambridge, UK: Royal Society of starches. Journal of Food Engineering, 54, 167e174.
Chemistry. Tizzotti, M. J., Sweedman, M. C., Schaefer, C., & Gilbert, R. G. (2013). The influence of
Piorkowski, D. T., & McClements, D. J. (2014). Beverage emulsions: Recent de- macromolecular architecture on the critical aggregation concentration of large
velopments in formulation, production and applications. Food Hydrocolloids, 42, amphiphilic starch derivatives. Food Hydrocolloids, 31, 365e374.
5e41. Voragen, A. G. J., Pilnik, W., Thibault, J.-F., Axelos, M. A. V., & Renard, C. M. G. C.
Prochaska, K., Kedziora, P., Le Thanh, J., & Lewandowicz, G. (2007). Surface activity (1995). Pectins. In A. M. Stephen (Ed.), Food polysaccharides and their applica-
of commercial food grade modified starches. Colloids and Surfaces B, 60, tions (pp. 287e339). New York: Marcel Dekker.
187e194. Walstra, P. (2003). Physical chemistry of foods. New York: Marcel Dekker.
Qian, C., Decker, E. A., Xiao, H., & McClements, D. J. (2011). Comparison of Walstra, P., & Smulders, P. E. A. (1998). Emulsion formation. In B. P. Binks (Ed.),
biopolymer emulsifier performance in formation and stabilization of orange oil- Modern aspects of emulsion science (pp. 56e99). Cambridge, UK: Royal Society of
in-water emulsions. Journal of the American Oil Chemists’ Society, 88, 47e55. Chemistry.
Qi, W., Fong, C., & Lamport, D. T. A. (1991). Gum arabic glycoprotein is a twisted Wijmans, C. M., Leermakers, F. A. M., & Fleer, G. J. (1994). Multiblock copolymers
hairy rope d a new model based on o-galactosylhydroxyproline as the poly- and colloidal stability. Journal of Colloid and Interface Science, 167, 124e134.
saccharide attachment site. Plant Physiology, 96, 848e855. Williams, P. A., & Phillips, G. O. (2009). Gum arabic. In G. O. Phillips, & P. A. Williams
Randall, R. C., Phillips, G. O., & Williams, P. A. (1988). The role of the proteinaceous (Eds.), Handbook of hydrocolloids (2nd ed., pp. 252e273). Cambridge, UK:
component on the emulsifying properties of gum arabic. Food Hydrocolloids, 2, Woodhead.
131e140. Williams, P. A., Phillips, G. O., & Stephen, A. M. (1990). Spectroscopic and molecular
Ray, A. K., Bird, P. B., Iacobucci, G. A., & Clark, B. C., Jr. (1995). Functionality of gum comparisons of three fractions from Acacia Senegal gum. Food Hydrocolloids, 4,
arabic: Fractionation, characterization and evaluation of gum fractions in citrus 305e311.
oil emulsions and model beverages. Food Hydrocolloids, 9, 123e131. Williams, P. A., Phillips, G. O., Stephen, A. M., & Churms, S. C. (2006). Gums and
Rayner, M., Marku, D., Eriksson, M., Sjo €o€, M., Dejmek, P., & Wahlgren, M. (2014). mucilages. In A. M. Stephen, G. O. Phillips, & P. A. Williams (Eds.), Food poly-
Biomass-based particles for the formulation of Pickering type emulsions in food saccharides and their applications (2nd ed., pp. 455e496). Boca Raton, FL: CRC/
and topical applications. Colloids and Surfaces A, 458, 48e62. Taylor & Francis.
Roman-Guerrero, A., Orozco-Villafuerte, J., Perez-Orozco, J. P., Cruz-Sosa, F., Jime- Williams, P. A., Sayers, C., Viebke, C., Senan, C., Mazoyer, J., & Boulenguer, P. (2005).
nez-Alvarado, R., & Vernon-Carter, E. J. (2009). Application and evaluation of Elucidation of the emulsification properties of sugar beet pectin. Journal of
mesquite gum and its fractions as interfacial film formers and emulsifiers of Agricultural and Food Chemistry, 53, 3592e3597.
orange peel oil. Food Hydrocolloids, 23, 708e713. Yadav, M. P., Cooke, P., Johnston, D. B., & Hicks, K. B. (2010). Effect of protein-rich
Rombouts, F. M., & Thibault, J.-F. (1986). Feruloylated pectin substances from sugar components on the emulsifying properties of corn fibre gum. Cereal
beet pulp. Carbohydrate Research, 154, 177e187. Chemistry, 87, 89e94.
Sanchez, C., Renard, D., Robert, P., Schmitt, C., & Lefebvre, J. (2002). Structural and Yadav, M. P., Igartuburu, J. M., Yan, Y., & Nothnagel, E. A. (2007). Chemical investi-
rheological properties of acacia gum dispersions. Food Hydrocolloids, 16, gation of the structural basis of the emulsifying activity of gum arabic. Food
257e267. Hydrocolloids, 21, 297e308.

Please cite this article in press as: Dickinson, E., Hydrocolloids acting as emulsifying agents e How do they do it?, Food Hydrocolloids (2017),
http://dx.doi.org/10.1016/j.foodhyd.2017.01.025

Potrebbero piacerti anche