Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Journal of Food Engineering 67 (2005) 225–234

www.elsevier.com/locate/jfoodeng

Biopolymers and emulsifiers at the air–water interface.


Implications in food colloid formulations
Cecilio Carrera Sánchez, Mª. Rosario Rodrı́guez Niño, Ana Lucero Caro,
Juan M. Rodrı́guez Patino *
Departamento de Ingeniera Quı́mica, Facultad de Quı́mica, Universidad de Sevilla, C/. Prof. Garcı́a González, 1, 41012 Sevilla, Spain

Received 10 October 2003; accepted 1 May 2004

Abstract

In this paper we are concerned with adsorption, structure, topography, and dynamic properties (relaxation phenomena and
surface dilatational rheology) of food dairy proteins (b-casein, caseinate, and whey protein isolate, WPI), water-insoluble lipids
(monopalmitin, monoolein, and monolaurin) and phospholipids (dipalmitoyl-phosphatidyl-choline, DPPC, and dioleoyl-phosphat-
idyl-choline, DOPC) at the air–water interface. Combined surface chemistry (surface film balance and static and dynamic tensiom-
etry) and microscopy (Brewster angle microscopy, BAM) techniques have been used to determine the static and dynamic
characteristics of these emulsifiers and their mixtures at the air–water interface. The derived information shows that biopolymer
(proteins) and low-molecular-weight-emulsifier (LMWE, monoglycerides and phospholipids) type and their mixtures affect the inter-
facial characteristics of adsorbed and spread films. Important functional differences have been established between proteins, lipids
and phospholipids. The static and dynamic characteristics of mixed films depend on the interfacial composition and the surface pres-
sure (p). At higher surface pressures, collapsed protein residues may be displaced from the interface by LMWE molecules with
important repercussions on the interfacial characteristics of the mixed films.
Ó 2004 Elsevier Ltd. All rights reserved.

Keywords: Fluid interfaces; Adsorption; Interfacial rheology; Monolayer; Food emulsifier; Biopolymer; Low-molecular weight emulsifier; Milk
protein; Monoglyceride; Phospholipid

1. Introduction stability, shelf-life, or product texture. For up-to-date


reviews, readers are directed to recent references
Food dispersions (emulsions and foams) are complex (Damodaran & Paraf, 1997; Friberg & Larsson, 1997;
multicomponent systems containing many biopolymers Hartel & Hasenhuette, 1997). Manufacturers employ
and low-molecular weight emulsifiers (LMWE) which two types of emulsifiers or foaming agents in food
may show surface activity by themselves or by associa- (Dickinson, 1992), namely: LMWE (mainly mono-
tion with other components (polysaccharides). In addi- and diglycerides, phospholipids, etc.) and macro-
tion food dispersions contain many other organic molecules (proteins and certain hydrocolloids). The
(ethanol, sugars, etc.) and inorganic (salts) components emulsifier film adsorbed at the oil–water or air–water
which may interact with biopolymers and LMWE in interface is the source of many of the unique properties
different complex fashion depending on pH, tempera- of food dispersions, particularly their stability and
ture, processing history, etc., all of which intensify the interactions, which translate into the shelf-life and tex-
problems of the manufacturer attempting to control tural properties so desired by manufacturers and appre-
ciated by consumers.
*
Corresponding author. Tel.: +34 954 556446; fax: +34 954 557134. Proteins and LMWE have an important physical prop-
E-mail address: jmrodri@us.es (J.M. Rodrı́guez Patino). erty in common, their amphiphilic nature (Horne &

0260-8774/$ - see front matter Ó 2004 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jfoodeng.2004.05.065
226 C.C. Sánchez et al. / Journal of Food Engineering 67 (2005) 225–234

Rodrı́guez Patino, 2003; Rodriguez Niño, Rodrı́guez DANR ML 90) were supplied by Danisco Ingredients
Patino, Carrera, Cejudo, & Navarro, 2003). This prop- (Braban, Denmark) with over 95–98% of purity. DL -a-
erty provides the potential for association, adsorption, dipalmitoyl-phosphatidyl-choline (DPPC, Sigma (St.
and reorientation at fluid interfaces, depending on the Louis, MO, USA), 99%) and L -a-dioleoyl-phosphat-
properties of the components and the protein–LMWE ra- idyl-choline (DOPC, Sigma, 99%) were used as supplied.
tio (Rodrı́guez Niño & Rodrı́guez Patino, 1998a, 1998b; Ninety-nine percent pure b-casein was supplied and puri-
Rodrı́guez Niño, Carrera, Cejudo, & Rodrı́guez Patino, fied from bulk milk from the Hannah Research Institute
2001; Rodrı́guez Patino, Rodrı́guez Niño, & Carrera, (Ayr, Scotland). Caseinate (a mixture of 38% b-casein,
2003a). LMWE stabilize the dispersed droplets or bub- 39% as1-casein, 12% j-casein, and 11% as2-casein)
bles by formation of a densely packed but much less rigid was supplied and purified from bulk milk from Unilever
monomolecular layer, which is stabilized by dynamic Research Laboratories (Colworth, UK). Whey protein
processes (i.e. Gibbs–Marangoni effect). LMWE adsorb isolate (WPI), a native protein with high content of b-
strongly to fluid interfaces giving close molecular packing lactoglobulin (protein 92 ± 2%, b-lactoglobulin >95%,
at the interface to produce low surface and interfacial ten- a-lactalbumin <5%) obtained by fractionation, was sup-
sions (Rodrı́guez Niño & Rodrı́guez Patino, 1998a, plied by Danisco Ingredients (Brabran, Denmark). To
1998b). In contrast, proteins act as polymeric emulsifiers form the surface film, monoglyceride and phospholipid
with multiple anchoring sites at the interface that, to- were spread in the form of a solution, using hexane:eth-
gether with the unfolding process of the adsorbing pro- anol (9:1, v:v) and chloroform:ethanol (4:1, v:v), respec-
tein molecule, stabilize the interfacial layer kinetically. tively, as a spreading solvent. Analytical grade hexane
This behavior contributes significantly to the interfacial (Merck, 99%), ethanol (Merck, >99.8%), and chloroform
rheological properties and immobilizes proteins in the ad- (Sigma, 99%), were used without further purification.
sorbed layer (Bos, Nylander, Arnebrant, & Clark, 1997). Samples for interfacial characteristics of protein films
However more important in some products is the effect of were prepared using Milli-Q ultrapure water at pH 7.
the LMWE in destabilizing the emulsion (Goff & Jordan, The water used as subphase was purified by means of a
1989). In the formulation of ice cream the LMWE (typi- Millipore filtration device, Milli-Q (Milford, MA,
cally, mono- and diglycerides) are added to break the ad- USA). To adjust subphase pH, buffer solutions were
sorbed layer of protein and allow the adsorption of fat to used. Acetic acid/sodium acetate aqueous solution
the surface of the air bubble. Thus, an important action (CH3COOH/CH3COONa) was used to achieve pH 5,
of LMWE is to promote the displacement of proteins and a commercial buffer solution called trizma
(mainly caseins) from the interface. The competitive ((CH2OH)3CNH2/(CH2OH)3CNH3Cl) for pH 7. All
adsorption and/or displacement between LMWE and these products were supplied by Sigma (>99.5%). Ionic
proteins at fluid–fluid interfaces have been studied strength was 0.05 M in all the experiments.
in detail in several investigations (Bos et al., 1997;
Nylander, 1998; Wilde, 2000; Dickinson, 2001; Rodrı́-
2.2. Methods
guez Patino et al., 2003a). However, so far, little is known
about the structure that biopolymers and LMWE adopt
2.2.1. Equilibrium surface pressure and adsorption
at fluid interfaces.
isotherm
This paper will concentrate on the interfacial behav-
The equilibrium surface pressure (pe) is a key parameter
ior of milk proteins and LMWE (monoglycerides and
for the analysis of the mechanisms that trigger the relaxa-
phospholipids). Emphasis will be on the air–water inter-
tion phenomena in spread monolayers at the air–water
face as a three-dimensional dynamic entity. We will con-
interface (Gaines, 1966). The equilibrium spreading pres-
sider emulsifier (protein, LMWE and their mixtures)
sure is the maximum surface pressure to which a spread
adsorption, structure and topography at the interface,
monolayer may be compressed before monolayer collapse.
relaxation phenomena, and interfacial rheology, as re-
Equilibrium surface pressure of protein and LMWE at the
lated to the formation and stability of food dispersions
air–water interface was measured by the Wilhelmy plate
(emulsions and foams).
method as described elsewhere (Carrera, Rodrı́guez Pati-
no, & Rodrı́guez Niño, 1999). The adsorption isotherm
of proteins and LMWE–protein films was studied by tensi-
2. Experimental
ometry as described elsewhere (Rodrı́guez Niño & Rodrı́-
guez Patino, 1998a; Rodrı́guez Niño et al., 2001).
2.1. Chemicals

Synthetic 1-monohexadecanoyl-rac-glycerol (mono- 2.2.2. Surface film balance


palmitin, DIMODANR PA 90), 1-mono(cis-9-octa- Measurements of the surface pressure (p) versus aver-
decanoyl)glycerol (monoolein, RYLOTMMG 19), and age area per molecule (A), the so-called p–A isotherm,
1-monododecanoyl-rac-glycerol (monolaurin, DIMO- were performed on a fully automated Langmuir- or
C.C. Sánchez et al. / Journal of Food Engineering 67 (2005) 225–234 227

Wilhelmy-type film balance as described elsewhere


(Rodrı́guez Niño, Carrera, & Rodrı́guez Patino, 1999).

2.2.3. Brewster angle microscopy


For microscopic observation of the monolayer struc-
ture, the Brewster angle microscope, BAM 2 plus (NFT,
Germany) was used as described elsewhere (Rodrı́guez
Patino, Carrera, & Rodrı́guez Niño, 1999a, 1999b). To
measure the relative thickness of the film a previous
camera calibration is necessary in order to determine
the relationship between the gray level (GL) and the rel-
ative reflectivity (I), according to a procedure described
previously (Rodrı́guez Patino et al., 1999a, Rodrı́guez
Patino, Carrera, & Rodrı́guez Niño, 1999b).

2.2.4. Surface relaxation measurements


Measurements of surface relaxation in protein or
LMWE films at the air–water interface were performed
on a fully automated Langmuir-type film balance. The
method has been described previously (Carrera et al.,
1999).

2.2.5. Surface dilatational rheology


To obtain surface rheological parameters at the air–
water interface––such as surface dilatational modulus
(E), elastic (Ed) and viscous (Ev) components, and loss
angle tangent (tan h)––a modified Wilhelmy-type film
balance (KSV 3000) was used as described elsewhere
(Rodrı́guez Patino, Carrera, Rodrı́guez Niño, & Cejudo, Fig. 1. (A) Equilibrium surface pressure of monopalmitin (MP),
monoolein (MO), dipalmitoyl-phosphatidyl-choline (DPPC), dioleoyl-
2001c, 2001d). phosphatidyl-choline (DOPC), b-casein (BC) and WPI at pH 7 and at
20 °C. (B) The effect of spreading of (s) MP, (n) MO and (j)
monolaurin (ML) on a film of WPI previously adsorbed on the air–
3. Results and discussion water interface. The arrows indicate the equilibrium surface pressures
(pe) for MP, MO and ML. C is the concentration of WPI in the bulk
phase. The amount of monoglyceride spread on the WPI film is enough
3.1. Protein and LMWE films at equilibrium to saturate the monolayer by itself. Temperature 20 °C. pH 5.

Equilibrium spreading pressure of monoglycerides


(monopalmitin, and monoolein), phospholipids (DPPC ments it has been observed that the interfacial
and DOPC) and proteins (b-casein and WPI) at the characteristics of mixed protein and LMWE films at
air–water interface, at pH 7 and at 20 °C is shown in air–water interfaces depend at least on the interfacial
Fig. 1A. The magnitude of pe was dependent on the composition and on the protein/LMWE ratio (Fig.
LMWE and the protein. pe was higher for monoglycer- 1B). At higher concentrations of WPI in the bulk phase
ides (Rodriguez Niño et al., 2003) and phospholipids the surface activity of the mixed film is similar to that for
(unpublished data) and lower for proteins (Rodrı́guez pure WPI while at lower concentrations the surface
Niño et al., 2001). The effect of temperature (data not activity of the mixed film is similar to pe of the mono-
shown) was also different for LMWE and proteins, espe- glyceride (MP or MO). The solubility of monolaurin
cially for an ordered protein (WPI). pe for LMWE and proves that the mixed film is dominated by the protein
b-casein was not affected by temperature within the within the overall range, because for the monolaurin–
range of 5 and 40 °C. However, pe for WPI increased protein mixed film the p–log C plot is practically the
with temperature, especially at temperatures higher than same as that for pure protein (Fig. 1B). In general, the
25 °C. Higher pe values correlate with higher interac- surface activity of the protein + LMWE mixed films
tions within the film forming components thus, with a is determined by the LMWE as p of the mixed film is
more condensed film structure at equilibrium. the same as the pe of LMWE and the monolayer is
Protein–LMWE interactions at the air–water inter- not saturated by the protein (Rodrı́guez Niño & Rodrı́-
face can be studied by tensiometry. From these experi- guez Patino, 1998a, 2001). However, the protein deter-
228 C.C. Sánchez et al. / Journal of Food Engineering 67 (2005) 225–234

mines the surface activity of mixed films as the protein different structures and the collapse phase. As for
saturates the monolayer. LMWE, d increases with p and is maximum at the col-
lapse (at the highest p). At p lower than pe, the relative
3.2. Structure and topography of protein and LMWE film thickness is independent of the protein but, at the
films collapse point, d for b-casein is higher than for WPI
(Rodrı́guez Patino et al., 1999b). The differences ob-
Structural and topographical characteristics of pro- served between lipids (Rodrı́guez Patino et al., 1999a)
teins and LMWE spread films at the air–water interface and proteins (Rodrı́guez Patino et al., 1999b) in p–A iso-
can be deduced from p–A isotherms (Rodrı́guez Patino therms and BAM images is of great utility for the appli-
& Rodrı́guez Niño, 1999) coupled to information from cation of BAM to the analysis of more complicated
BAM (Rodrı́guez Patino et al., 1999a, 1999b). From systems in which proteins and lipids are spread at the
the p–A isotherm, different structures can be deduced interface (Rodrı́guez Patino, Carrera, & Rodrı́guez
for LMWE monolayers as a function of LMWE, temper- Niño, 1999c, 1999d; Rodrı́guez Patino, Rodrı́guez Niño,
ature, and surface density or surface pressure. For Carrera, & Cejudo, 2001a, Rodrı́guez Patino, Rodrı́guez
instance, monopalmitin (and most saturated phospholi- Niño, Carrera, & Cejudo, 2001b).
pids) monolayers (Fig. 2) show that liquid expanded From a systematic study centered on the p–A iso-
(LE), liquid-condensed (LC) and solid (S) structures, therm of protein–monoglyceride mixed monolayers––
and, finally, the collapse at a p higher than pe, take place including the application of the additivity rule on misci-
as a function of surface pressure. In contrast with mono- bility and the quantification of interactions between
palmitin, monoolein and DOPC monolayers (data not monolayer components by excess free energy––it has
shown) presents only the liquid expanded structure and been concluded that, at a macroscopic level, these com-
the collapse at pe. BAM allows direct visualization of pounds form a practically immiscible monolayer at the
changes in morphology and collapse of monopalmitin air–water interface, at p lower than that for the protein
monolayer (as an example) at the air–water interface collapse (Rodrı́guez Patino et al., 2001a, 2001b) (Fig.
(Fig. 2). Monopalmitin monolayer at 10 mN/m shows cir- 2). At higher p the collapsed protein is displaced from
cular LC domains from the homogeneous ambient phase the interface by LMWE. The existence of low protein
with a LE structure. The LC domains grow in size and the interactions in disordered proteins (b-casein and casei-
monolayer is covered with LC domains as p is increased. nate) facilitates the protein displacement by LMWE
At the highest p, the LC domains are so closely packed from the air–water interface. On the other hand, the low-
that they occupy the entire field of view, the contrast van- er surface activity of unsaturated LMWE explains the
ishes suddenly, and the presence of monolayer fractures fact that this LMWE has a lower capacity than saturated
can be observed in different zones (Rodrı́guez Patino et LMWE for protein displacement. Different proteins and
al., 1999a). BAM images corroborate that only the homo- LMWE show different interfacial interactions, miscibil-
geneous LE phase is present during the compression of ity and topography, confirming the importance of pro-
a monoolein and DOPC monolayers (data not shown). tein and LMWE structure in determining the
The evolution with the monolayer compression of the mechanism of interfacial interactions (Rodrı́guez Patino
film thickness (d) gives complementary information et al., 2001a, 2001b). Thus, displacement of proteins
about the structural characteristics of LMWE during by LMWE from the air–water interface depend on
monolayer compression (Rodrı́guez Patino et al., the particular protein–LMWE system (Fig. 3). The dis-
1999a). The film thickness increases as the monolayer is placement surface pressure (pd) value for monopalm-
compressed, passes through a maximum and then de- itin–b-casein mixed films is lower than those for
creases at the monolayer collapse point. The evolution monopalmitin–caseinate and monopalmitin–WPI mixed
of d for saturated LMWE monolayer with monolayer films. Thus, protein displacement by monopalmitin is
compression also shows important differences with easier for b-casein than it is for caseinate and WPI, in
unsaturated LMWE monolayers and their d is lower this order. In the same order increases the surface elastic-
than for saturated LMWE (Rodrı́guez Patino et al., ity of the protein (see last section). Thus, the more elastic
1999a). WPI film is more resistant than the less elastic b-casein
Results of BAM (specially the relative reflectivity) as films. Monoolein has a lower capacity than mono-
a function of p obtained with protein monolayers clearly palmitin for protein displacement due to the fact that
show the same structural characteristics as those de- monoolein requires higher pd for protein displacement.
duced from the p–A isotherm (Fig. 2). The domains that Phospholipids and proteins mixed films behave in a
residues of protein molecules adopt at the air–water more complicated fashion (unpublished data). Phosp-
interface appeared to be of uniform reflectivity (Fig. holipids and b-casein form a practically immiscible
2), suggesting homogeneity in thickness and film isot- monolayer at the air–water interface on neutral or acidic
ropy. The results of p–A isotherms (Fig. 2) confirm that aqueous subphases. However, some degree of interac-
protein monolayers at the air–water interface adopt two tion exists between phospholipids and b-casein in the
C.C. Sánchez et al. / Journal of Food Engineering 67 (2005) 225–234 229

Fig. 2. p–A isotherm and visualization by BAM of (n) b-casein, () monopalmitin, and (––) monopalmitin–b-casein mixed film at XMP = 0.5. BAM
images of monopalmitin––(a) LE phase at p < 5 mN/m, (b) coexistence of LE and LC domains at 5 mN/m < p < 30 mN/m, (c) LC domains at
p > 35 mN/m, and (d) fracture of a collapse monolayer at p ffi pe (monopalmitin)––and saturated LMWE, b-casein––homogeneous topography at (a)
p < pe and (b) at p ffi pe––and unsaturated LMWE, and monopalmitin–b-casein mixed monolayer at XMP = 0.5 and at p < pe––(a) segregated LE–LC
monopalmitin and b-casein domains, (b) homogeneous LE monopalmitin–b-casein domains, and (c) segregated LC monopalmitin and b-casein
domains––at p > pe––(d) region of monopalmitin LC domains, (e) coexistence of monopalmitin and collapse b-casein, and (f) squeezing out of b-
casein by monopalmitin––and at the collapse point––(g) a region of collapsed monopalmitin dominates the topography of the interface, (h) fracture
of collapsed monopalmitin, and (i) coexistence of collapse monopalmitin and islands of collapsed b-casein.

mixed film and these interactions, of an electrostatic Competitive adsorption of proteins and LMWE at
character, are more pronounced at pH 9 as the phospho- fluid interfaces can affect the stability of food disper-
lipid molecules become completely ionized. sions (Bos et al., 1997; Wilde, 2000). Thus, knowledge
230 C.C. Sánchez et al. / Journal of Food Engineering 67 (2005) 225–234

4B), which occurs at p near to and above pe of the pro-


tein, involves a buckling of the monolayer and reorder-
ing of the molecules as the protein film gets thicker in
response to the decreasing surface coverage. Finally, at
sufficiently high p the protein network begins to fail
(Fig. 4C), freeing proteins, which then desorb from the
interface (Rodrı́guez Patino et al., 2001a, 2001b). But,
for spread water-insoluble emulsifier monolayers, the
protein displacement is not total even at the highest sur-
face pressure, at the collapse point of the mixed film.
The orogenic displacement mechanism is a conse-
quence of the low level of interaction between pro-
teins and LMWE at the air–water interface (Rodrı́guez
Patino et al., 2001a, 2001b). This model has been shown
to work for a range of proteins with different secondary
and tertiary structures and different types of LMWE
(non-ionic, cationic, anionic and zwitterionic). This
mechanism also explains the destabilization of beer
foams by lipids with a range of chain lengths (Wilde
Fig. 3. The displacement surface pressure (pd) of (h) b-casein, (s) et al., 2003).
caseinate, and (n) WPI by (A) monopalmitin and (B) monoolein from
spread mixed monolayers at the air–water interface as a function of the
3.3. Relaxation phenomena in protein and LMWE films
mass fraction of monoglyceride in the mixture. The horizontal lines
represent the equilibrium surface pressures of (––, pe MP) monopalm-
itin, (  , pe MO) monoolein, (- - -, pe BC) b-casein, (- Æ -, pe CS) Non-equilibrium processes occurring in systems con-
caseinate, and (-    -, pe WPI) WPI pure monolayers at 20 °C and at taining fluid interfaces with a surfactant present are of
pH 7.

of the proteins, LMWE, and their mixtures at fluid–fluid


interfaces is a key factor for the formation and stability
of food dispersions (emulsions and foams). Recent
works have allowed the indirect study of competitive (A)
adsorption at a molecular level (Gunning, Mackie,
Wilde, & Morris, 1999; Mackie, Gunning, Wilde, &
Morris, 1999). The Norwich group (Gunning et al.,
1999; Mackie et al., 1999) has demonstrated how surfac-
tants disrupt and displace proteins from an interface by (B)
a three-stage ‘‘orogenic’’ mechanism. Briefly, emulsifier
penetrates into the protein film, forming separate ad-
sorbed domains that exert a lateral p which mechani-
cally compresses the protein network until it fails and
is displaced from the interface. The mechanism also re-
veals that protein adsorbed layers with a high surface
(C)
elasticity are more resistant to displacement. The
BAM images (Fig. 2) appear to support the idea that
displacement takes place via an orogenic mechanism,
but there exist differences between adsorbable water-
soluble emulsifiers (Gunning et al., 1999; Mackie et al., LMWE PROTEIN
1999) and spread water-insoluble emulsifiers (Mackie,
Gunning, Ridout, Wilde, & Rodrı́guez Patino, 2001;
Rodrı́guez Patino et al., 2001a, 2001b). The results sug- Fig. 4. Displacement of proteins by LMWE spread at the air–water
gest that for spread water-insoluble emulsifiers (Fig. 4) interface according to the orogenic mechanism. (A) At p < pprotein
e a
the first stage of the orogenic mechanism, which occurs displacement front of emulsifier domains is produced. (B) At
p ffi pprotein a buckling of the protein monolayer and reordering of
at p lower than pe of the protein, involves a displace- e
the molecules is produced and the protein film gets thicker in response
ment front of emulsifier domains (Fig. 4A) instead of to the decreasing surface coverage. (C) At p > pprotein the protein
e
the adsorption of water-soluble surfactant molecules at network begins to fail and desorbs from the interface, forming collapse
defects in the protein network. The second stage (Fig. protein multilayers in the aqueous bulk phase near to the interface.
C.C. Sánchez et al. / Journal of Food Engineering 67 (2005) 225–234 231

great practical significance and include important tech- From a practical point of view it must be emphasized
nological operations such as emulsification and foam- that under these conditions the mixed film is more stable
ing. Two experimental approaches can be used for the in relation to monolayer molecular loss than that of the
analysis of long-term relaxation phenomena in emulsi- pure components. At the collapse point of the mixed
fier monolayers (Gaines, 1966). In a first approach, the film, the relaxation phenomena may be due either to
surface pressure (p) is kept constant, and the area A is nucleation and growth of critical nuclei of monoglycer-
measured as a function of time. In the second approach, ide or to a complex mechanism including competition
area is kept constant (at the monolayer collapse) and the between desorption and monolayer collapse. The rea-
decrease in p is monitored as a function of time. Infor- sons for these behaviors may be associated again with
mation on various relaxation paths (Marangoni effect, the immiscibility between protein and monoglyceride
chemical reaction, polar group hydration, conforma- at the air–water interface and to the protein displace-
tion/organization changes, film dissolution by desorp- ment by the monoglyceride at surface pressures higher
tion and/or diffusion, collapse, etc.) can be derived than that for protein collapse.
from these data (Rodriguez Niño et al., 2003; Rodrı́guez
Patino et al., 2003a). 3.4. Interfacial rheological characteristics of protein and
Desorption of spread LMWE monolayers at any con- LMWE films
stant surface pressure, at p < pe, involves two stages
(Rodriguez Niño et al., 2003; Rodrı́guez Patino et al., The breaking of drops and bubbles during emulsifica-
2003a). The first is dissolution into the bulk aqueous tion and foaming requires rapid and substantial stretch-
phase to form a saturated aqueous layer. The second ing of the drops or bubbles, and consequently, the
stage occurs when, after a time, the concentration gradi- surface tension may be far from equilibrium. Thus, dil-
ent within the diffusion layer becomes constant and
desorption reaches a steady state. The monolayer molec-
ular loss was lower for saturated than for unsaturated 1.0
LMWE. At p > pe the relaxation phenomena in LMWE
films are due to the transformation of a homogeneous
0.9
monolayer phase into a heterogeneous monolayer-col-
lapse phase system. However, some differences exist be-
tween saturated (monopalmitin or DPPC) and 0.8
A/A o

β-casein
unsaturated (monoolein or DOPC) LMWE monolayers.
monopalmitin
Relaxation phenomena in saturated LMWE monolayer 0.7
0.2
are controlled predominantly by the collapse mechanism
because of the p values relaxed to pe value. For unsatu- 0.4.
0.6
rated LMWE monolayer p relaxed from the collapse 0.6.
(A)
value, which is close to pe, towards lower p values at 0.8.

longer times. This phenomenon should be ascribed to 0.5


0 50 100 15 0 200 25 0
the concurrence of different phenomena, such as desorp-
time (min)
tion and collapse. Protein monolayer behaves in a differ-
ent way to LMWE under the same experimental 60
conditions. At p < pe the relaxation is a reversible proc-
55
ess, which include monolayer organization/reorienta- πe
MP

tion. The relaxation in relative molecular area and in 50


surface pressure at p > pe should be attributed to proc- 45
π(m N/m )

esses related to monolayer organization/reorientation


40
and collapse, respectively.
The strength of interactions between protein and 35
LMWE can also be studied by relaxation experiments. 30
Long-term relaxation phenomena in protein–monoglyc- 25
eride mixed films (Fig. 5) at the air–water interface have
20 (B)
been analyzed according to models for desorption, col-
lapse and/or organization/reorganization changes 15
0 20 40 60 80 100 120 140
(Rodrı́guez Patino, Rodrı́guez Niño, & Carrera, 2002a,
2002b). At lower surface pressures (i.e., p lower than time (min)
protein collapse pressure), the organization/reorganiza- Fig. 5. (A) Relaxation at constant surface pressure (p = 20 mN/m) and
tion change of protein molecules in the mixed film is (B) relaxation at constant area (at the collapse point) for of b-casein–
the mechanism that controls the relaxation process. monopalmitin mixed monolayers on water at pH 7 and at 20 °C.
232 C.C. Sánchez et al. / Journal of Food Engineering 67 (2005) 225–234

atational properties of adsorbed emulsifier layers are crease is higher than for the more expanded monoolein
also important. The viscoelastic properties of the surface monolayer. This indicates that E is not only determined
have often been correlated with functionality (Bos & van by the interactions between spread monoglyceride (or
Vliet, 2001; Dickinson, 1999, 2001; Murray, 2002). The phospholipid) molecules (which depend on p), but that
ability of the protein to resist displacement by emulsifi- the structure of the spread molecule also plays an impor-
ers is closely linked to the surface dilatational rheology, tant role. In fact, for the more aggregated monopalmitin
whereas the precise form of the displacement is consid- molecules in LC domains (see Fig. 2) E is higher than
ered to be more closely related to the surface shear that of monoolein molecules with LE structure, at the
behavior (Mackie et al., 2003; Murray, 2002; Roth, same p.
Murray, & Dickinson, 2000). Different and complemen- As for LMWE, for WPI monolayers (Fig. 6) E in-
tary interfacial techniques (surface film balance, BAM, creased with increasing p up to the collapse point.
and interfacial dilatational rheology) are useful in the This increase is the result of an increase in the interac-
analysis of the structural and dynamic characteristics tions between the monolayer molecules, as deduced
of protein, LMWE, and their mixtures at the air–water from p–A isotherms and BAM image. However, for
interface (Rodrı́guez Patino, Rodrı́guez Niño, & the more disordered proteins (b-casein and caseinate)
Carrera, 2002b; Rodrı́guez Patino, Rodrı́guez Niño, & the E–p dependence is more complex. E increases to
Carrera, 2003b). a maximum with p, but decreases with p and passes
A common trend of the p dependence of dilatational to a minimum. Finally, E increases up to the collapse
modulus (E) for monopalmitin and monoolein mono- point (Fig. 6). This inflection in the E–p curve may
layers (Fig. 6) is that E increased with increasing p up be attributed to the transition from an ‘‘all-train’’ con-
to the collapse point. This increase is a result of an in- figuration to a ‘‘train-and-loop’’ conformation of the
crease in the interactions between the monolayer mole- b-casein molecule (Lucassen-Reynders & Benjamins,
cules (that is, of its structure), as deduced from p–A 1999). The results with protein monolayers indicate
isotherms and BAM images (Fig. 2). However, for the that E is not only determined by the structure of pro-
more condensed monolayer (monopalmitin) this in- tein molecules, but the internal nature of the protein

β-casein + LMWE WPI + LMWE

250 (A) 250 (C)

200 200
E (mN/m)

E (mN/m)

150 150

100 100

50 50

0 0
100 100
(B) (D)

80 80
E (mN/m)

E (mN/m)

60 60

40
40
20
20

0 10 20 30 40 50 0 10 20 30 40 50

Fig. 6. Surface pressure dependence of surface dilatational modulus for protein + monoglyceride mixed films at the air–water interface at pH 7. (A)
b-casein + monopalmitin mixed films. (B) b-casein + monoolein mixed films. (C) WPI + monopalmitin mixed films. (D) WPI + monoolein mixed
films. Temperature: 20 °C; frequency: 50 mHz; amplitude: 5%. Monolayer composition (mass fraction of monoglyceride): (s) 0, (n) 0.2, (,) 0.4, (})
0.6, (+) 0.8, and ( ) 1.0.
C.C. Sánchez et al. / Journal of Food Engineering 67 (2005) 225–234 233

spread molecule also plays an important role. In fact, (Eds.), Food emulsifiers and their applications (pp. 95–146). New
for the more ordered b-lactoglobulin molecules in WPI York: Chapman and Hall.
E is higher than that for b-casein or caseinate mole- Bos, M. A., & van Vliet, T. (2001). Interfacial rheological properties of
adsorbed protein layers and surfactants: A review. Advances in
cules (Fig. 6) with disordered structure, at the same Colloid and Interface Science, 91, 437–471.
surface pressures. Carrera, C., Rodrı́guez Patino, J. M., & Rodrı́guez Niño, Mª. R.
Surface dilatational rheology is a very sensitive tech- (1999). Relaxation phenomena in monoglyceride films at the air–
nique to analyze the competitive adsorption/displace- water interface. Colloids Surfaces B, 42, 175–192.
ment of protein and LMWE emulsifier at the air–water Damodaran, S., & Paraf, A. (1997). Food proteins and their applica-
tions. New York: Marcel Dekker.
interface. At higher p, the collapsed protein residues dis- Dickinson, E. (1992). An introduction to food colloids. Oxford: Oxford
placed from the interface by LMWE molecules have University Press.
important repercussions on the dilatational characteris- Dickinson, E. (1999). Adsorbed protein layers at fluid interfaces:
tics of the mixed films (Rodrı́guez Patino et al., 2002b, Interactions structure and surface rheology. Colloids and Surfaces
B: Biointerfaces, 15, 161–176.
2003b). However, the mechanical properties of the mixed
Dickinson, E. (2001). Milk protein interfacial layers and the relation-
films also demonstrate that, even at the highest p, the ship to emulsion stability and rheology. Colloids and Surfaces B:
monoglyceride is unable to displace completely protein Biointerfaces, 20, 197–210.
molecules from the air–water interface (Fig. 6). The sur- Friberg, S. E., & Larsson, K. (1997). Food emulsions (3rd ed.). New
face dilatational properties of mixed protein–emulsifier York: Marcel Dekker.
films also depend on the presence of some food compo- Gaines, G. L. (1966). Insoluble monolayers at liquid–gas interface. New
York: Wiley.
nents (ethanol and sucrose) in the aqueous phase. In gen- Goff, H. D., & Jordan, W. K. (1989). Action of emulsifiers in
eral, a decrease in the dilatational rheological properties promoting fat destabilization during the manufacture of ice cream.
on the addition of ethanol was found for protein–water- Journal of Dairy Science, 72, 18–29.
insoluble emulsifiers, whereas the opposite was observed Gunning, A. P., Mackie, A. R., Wilde, P. J., & Morris, V. J. (1999). In-
for protein–water-soluble emulsifiers (Rodrı́guez Niño, situ observation of surfactant-induced displacement of protein by
atomic force microscopy. Langmuir, 15, 4636–4640.
Wilde, Clark, & Rodrı́guez Patino, 1998c). Hartel, R., & Hasenhuette, G. R. (1997). Food emulsifiers and their
applications. New York: Chapman and Hall.
Horne, D. S., & Rodrı́guez Patino, J. M. (2003). Adsorbed biopoly-
3.5. Final considerations mers: Behavior in food applications. In M. Malmsten (Ed.),
Biopolymers at interfaces (pp. 857–900). New York: Marcel
Dekker.
In this paper, we have analyzed the structure, topog- Lucassen-Reynders, E. H., & Benjamins, J. (1999). Dilatational
raphy, adsorption, interactions, miscibility, and dynamic rheology of proteins adsorbed at fluid interfaces. In E. Dickinson
properties of food dairy proteins (b-casein, caseinate, & J. M. Rodrı́guez Patino (Eds.), Food emulsions and foams:
and WPI) and LMWE (monoglycerides and phospholi- Interfaces, interactions and stability (pp. 195–206). Cambridge:
pids) at the air–water interface. The summary includes Royal Society of Chemistry.
Mackie, A. R., Gunning, P. A., Pugnaloni, L. A., Dickinson, E.,
an assessment of information derived from a variety of Wilde, P. J., & Morris, V. J. (2003). The growth of surfactant
chemical and physical techniques. The results demon- domains in protein films. Langmuir, 19, 6032–6038.
strate that protein and LMWE type affect the interfacial Mackie, A. R., Gunning, A. P., Ridout, M. J., Wilde, P. J., &
characteristics. The nature of biopolymer and LMWE Rodrı́guez Patino, J. M. (2001). In situ measurement of the
interactions at the interface has an important role on displacement of protein films from the air/water by surfactant.
Biomacromolecules, 2, 1001–1006.
their physicochemical characteristics, including their role Mackie, A. R., Gunning, A. P., Wilde, P. J., & Morris, V. J. (1999).
in conferring stability on emulsions and foams. Impor- The orogenic displacement of proteins from the air/water interface
tant functional differences have been demonstrated by surfactant. Journal of Colloid and Interface Science, 210,
between globular (WPI) and disordered (b-casein and 157–166.
caseinate) proteins, between proteins and LMWE, and Murray, B. S. (2002). Interfacial rheology of food emulsifiers and
proteins. Current Opinion in Colloid and Interface Science, 7,
between saturated and unsaturated LMWE. 426–431.
Nylander, T. (1998). Protein–lipid interactions. In D. Möbius & R.
Miller (Eds.), Proteins at liquid interfaces (pp. 365–431). Amster-
Acknowledgment dam: Elsevier.
Rodrı́guez Niño, Mª. R., Carrera, C., Cejudo, M., & Rodrı́guez Patino,
J. M. (2001). Protein and lipid films at equilibrium at air–water
This research was supported by CICYT through interface. Journal of the American Oil Chemists Society, 78, 873–879.
Grant AGL2001-3843-C02-01. Rodrı́guez Niño, Mª. R., Carrera, C., & Rodrı́guez Patino, J. M.
(1999). Interfacial characteristics of b-casein spread films at the air–
water interface. Colloids Surfaces B, 12, 161–173.
References Rodriguez Niño, Mª. R., Rodrı́guez Patino, J. M., Carrera, C.,
Cejudo, M., & Navarro, J. M. (2003). Physicochemical character-
Bos, M., Nylander, T., Arnebrant, T., & Clark, D. C. (1997). Protein/ istics of food lipids and proteins at fluid-fluid interfaces. Chemical
emulsifier interactions. In G. L. Hasenhuette & R. W. Hartel Engineering Communications, 190, 15–47.
234 C.C. Sánchez et al. / Journal of Food Engineering 67 (2005) 225–234

Rodrı́guez Patino, J. M., Carrera, C., & Rodrı́guez Niño, Mª. R. monoglyceride mixed films at the air–water interface. Industrial and
(1999a). Morphological and structural characteristics of mono- Engineering Chemistry Research, 41, 3169–3178.
glyceride monolayers at the air–water interface by Brewster angle Rodrı́guez Patino, J. M., Rodrı́guez Niño, M. R., & Carrera, C.
microscopy. Langmuir, 15, 2484–2492. (2002b). Static and dynamic properties of a whey protein
Rodrı́guez Patino, J. M., Carrera, C., & Rodrı́guez Niño, J. M. isolate and monoglyceride mixed films at the air–water
(1999b). Structural and morphological characteristics of b-casein interface. Industrial and Engineering Chemistry Research, 41,
monolayers at the air–water interface by Brewster angle micros- 2652–2661.
copy. Food Hydrocolloids, 13, 401–408. Rodrı́guez Patino, J. M., Rodrı́guez Niño, Mª. R., & Carrera, C.
Rodrı́guez Patino, J. M., Carrera, C., & Rodrı́guez Niño, Mª. R. (2003a). Protein–emulsifier interactions at the air–water interface.
(1999c). Analysis of b-casein–monopalmitin mixed films at the air– Current Opinion in Colloid and Interface Science, 8, 387–395.
water interface. Journal of Agriculture and Food Chemistry, 47, Rodrı́guez Patino, J. M., Rodrı́guez Niño, Mª. R., & Carrera, C.
4998–5008. (2003b). Structure miscibility and rheological characteristics of b-
Rodrı́guez Patino, J. M., Carrera, C., & Rodrı́guez Niño, Mª. R. casein–monoglyceride mixed films at the air–water interface.
(1999d). Is Brewster angle microscopy a useful technique to Journal of Agricultural and Food Chemistry, 51, 112–119.
distinguish between isotropic domains in b-casein–monoolein Rodrı́guez Patino, J. M., Rodrı́guez Niño, Mª. R., Carrera, C., &
mixed monolayers at the air–water interface?. Langmuir, 15, Cejudo, M. (2001a). The effect of pH on monoglyceride–caseinate
4777–4788. mixed monolayers at the air–water interface. Journal of Colloid and
Rodrı́guez Patino, J. M., Carrera, C., Rodrı́guez Niño, Mª. R., & Interface Science, 240, 113–126.
Cejudo, C. (2001c). Structural-dilatational characteristics relation- Rodrı́guez Patino, J. M., Rodrı́guez Niño, Mª. R., Carrera, C., &
ships of monoglyceride monolayers at the air–water interface. Cejudo, M. (2001b). Whey protein isolate-monoglyceride mixed
Langmuir, 17, 4003–4013. monolayers at the air–water interface. Structure morphology and
Rodrı́guez Patino, J. M., Carrera, C., Rodrı́guez Niño, Mª. R., & interactions. Langmuir, 17, 7545–7553.
Cejudo, M. (2001d). Structural and dynamic properties of milk Rodrı́guez Niño, Mª. R., Wilde, P. J., Clark, D. C., & Rodrı́guez
proteins at the air–water interface. Journal of Colloid and Interface Patino, J. M. (1998c). Surface dilational properties of protein and
Science, 242, 141–151. lipid films at the air–water interface. Langmuir, 14, 2160–2166.
Rodrı́guez Niño, Mª. R., & Rodrı́guez Patino, J. M. (1998a). Surface Roth, S., Murray, B. S., & Dickinson, E. (2000). Interfacial shear
tension of protein and insoluble lipids at the air–aqueous phase rheology of aged and heat-treated b-lactoglobulin films: Displace-
interface. Journal of the American Oil Chemists Society, 75, ment by non-ionic surfactant. Journal of Agricultural and Food
1233–1239. Chemistry, 48, 1491–1497.
Rodrı́guez Niño, Mª. R., & Rodrı́guez Patino, J. M. (1998b). Surface Wilde, P. J. (2000). Interfaces: Their role in foam and emulsion
tension of bovine serum albumin and Tween 20 at the air–aqueous behaviour. Current Opinion in Colloid and Interface Science, 5,
phase interface. Journal of the American Oil Chemists Society, 75, 176–181.
1241–1248. Wilde, P. J., Husband, F. A., Cooper, D., Ridout, M. J., Mackie, A.
Rodrı́guez Patino, J. M., & Rodrı́guez Niño, Mª. R. (1999). Interfacial R., Gunning, A. P., Morris, V. J., Woodward, N., & Clare Mills, E.
characteristics of food emulsifiers (proteins and lipids) at the air– N. (2003). Interfacial mechanisms underlying lipid damage of beer
water interface. Colloids Surfaces A, 15, 235–252. foams. In E. Dickinson & T. Van Vliet (Eds.), Food colloids
Rodrı́guez Patino, J. M., Rodrı́guez Niño, Mª. R., & Carrera, C. biopolymers and materials (pp. 200–206). Cambridge: Royal Society
(2002a). Relaxation phenomena in whey protein isolated and of Chemistry.

You might also like