Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Soft Matter

View Article Online


PAPER View Journal | View Issue

Rheological and structural study of electrostatic


cross-linked xanthan gum hydrogels induced by
Cite this: Soft Matter, 2013, 9, 3063
b-lactoglobulin†
Xuan T. Le and Sylvie L. Turgeon*

This study examines for the first time the role of xanthan gum (XG) and b-lactoglobulin (blg) in network
formation induced by electrostatic attractive interaction. The gelation processes of blg–XG mixtures
were monitored by viscoelastic measurements as a function of the blg–XG ratio and blg and XG
concentrations. The structural characterization of the gels was addressed by means of confocal laser
scanning microscopy. It was found that the initial tenuous network of XG provided a frame for gel
organization, the blg aggregated along the XG chains and could be regarded as a crosslinking agent,
and more elastic gels were obtained at high XG concentrations. The lowest XG concentration at which
gelation is possible is estimated to be 4.8  103 wt%. The blg–XG ratio strongly affects gelation
kinetics and is a main factor controlling the gelation process and the gel structure. The optimal ratio for
electrostatic interaction at final pH of gels was estimated to be approximately 3.5. The decreasing
repulsive interaction between XG chains and the increasing attractive interaction between blg and XG
Received 2nd November 2012
Accepted 9th January 2013
with decreasing pH resulted in the formation of soluble complexes followed by the formation of
interpolymer complexes. The electrostatic cross-linking of XG chains by blg results in a sol–gel transition
DOI: 10.1039/c3sm27528k
at the point of gelation. This mechanism may be applicable in a wide variety of protein–polysaccharide
www.rsc.org/softmatter systems in which the structure of these systems is mainly stabilized by electrostatic attractive interaction.

Introduction several protein–polysaccharide mixtures has been associated


with interpenetrating5 and phase-separated networks.4,11 An
The gelation of naturally occurring macromolecules is of great interpenetrating network is formed when both biopolymers
interest with respect to the production of popular biomaterials associate independently to form separate networks. This type of
due to their high biocompatibility and large variety of applica- network is observed in BSA–LM pectin mixtures in the presence
tions in the cosmetics, pharmaceutical, biomedical, food, and of calcium ions. Proteins aggregate by heating to form a very
coating industries.1,2 The simultaneous use of proteins and weak gel interpenetrated by an LM pectin network, which is
polysaccharides is commonly encountered in food products to formed by linkages between LM pectin chains via calcium ions.5
control the structure, texture, and stability of the products.3–8 If some degree of demixing occurs prior to gelation, then both
Several types of gel structures can be formed depending on the biopolymer networks will be separated spatially, resulting in a
characteristics of the proteins and polysaccharides used and on phase-separated network. Generally, the gelation of mixtures of
the environmental conditions. Mixed gels containing more than proteins and polysaccharides in aqueous dispersion occurs
one gelling agent can be classied into four types: swollen aer phase separation and arrests phase separation.12–14 Phase
networks, interpenetrating networks, phase-separated separation is either segregative (thermodynamic incompati-
networks, and coupled gels.9,10 The gelation mechanism of bility) or associative (thermodynamic compatibility), depending
mainly on the electrical charges of the associated biopolymers
and thus on the factors affecting them, such as the ionic
STELA Dairy Research Center and Institute of Nutraceuticals and Functional Foods,
strength and pH.7,15 The microstructure of gels is the result of a
Faculty of Agriculture and Food Science, Université Laval, Pavillon Paul-Comtois,
2425 rue de l0 agriculture, Québec, Canada G1V-0A6. E-mail: Sylvie.Turgeon@fsaa. balance between demixing and gelation processes.11,16 A
ulaval.ca; Fax: +1-418-656-3353; Tel: +1-418-656-2131 ext. 4970 coupled network is formed under associative conditions when
† Electronic supplementary information (ESI) available: Details of a hydrogel both biopolymers are linked together to form junction zones.
formation process for the mixture at a ratio of 5, and 0.30 wt% total solid Some studies have explored coupled networks of two poly-
concentrations by the confocal laser scanning microscope and for the mixture
saccharides. A typical example of this network has been repor-
at a ratio of 2, and 0.10 wt% total solid concentrations by a phase contrast
optical microscope (Olympus, Tokyo, Japan) and the time dependence of
ted for galactomannan–xanthan binary gels.17,18 Fewer studies
storage modulus and loss modulus during gelation for some blg–XG mixtures have reported the gelation of protein–polysaccharide mixtures
are provided. See DOI: 10.1039/c3sm27528k under associative conditions and complex, coacervate

This journal is ª The Royal Society of Chemistry 2013 Soft Matter, 2013, 9, 3063–3073 | 3063
View Article Online

Soft Matter Paper

structures are frequently obtained for such mixtures. There is A previous study on b-lactoglobulin–xanthan gum (blg–XG)
therefore a need to understand the gelation mechanisms of as a model system proposed that the gelation process is initi-
protein–polysaccharide mixtures under associative conditions ated by the formation of soluble complexes followed by inter-
to produce new materials for the improvement of conventional polymeric electrostatic complexes that are able to form junction
foods and for the development of novel so biomaterials. zones, which result in a sol–gel transition at the point of gela-
Recently, it has been found that electrostatic gels can be tion.12 The system studied had a total concentration of 0.1 wt%.
obtained from the associative interaction between native Light scattering is largely used to monitor the kinetics of gela-
proteins and polysaccharides at very low concentrations.12 The tion but is limited to low concentrations because of the very
novelty of this study resides in the fact that no thermal, enzy- strong multiple scattering events that are observed at high
matic or any other denaturing treatment was applied to the concentrations and the large uctuations in intensity and
protein or the mixture at any time. Gelation was induced by relaxation time due to the onset of the gel network.27 Some
in situ acidication at room temperature to a pH at which the complementary techniques are needed to verify the gelation
molecules carry net opposite charges. The authors also noted mechanism of mixed blg–XG systems. Techniques that are
potential uses of these gels in the food industry to enhance the frequently used for structural characterization, e.g., gas/vapor
stability of foods and to protect micronutrients, and in the adsorption, mercury porosimetry, and conventional scanning
pharmaceutical industry to deliver and to protect drugs or active electron microscopy (SEM), can only be applied to a dry mate-
molecules. However, microstructural details of this gel remain rial. However, drying hydrogels of such low concentration can
unknown until now and the molecular mechanisms by which result in the shrinkage, deformation and collapse of the entire
proteins interact with XG chains are not fully understood. pore network.28
Therefore, understanding the sol–gel transition of these Confocal laser scanning microscopy (CLSM) and small-
protein–polysaccharide systems is of great interest so that the deformation rheology have oen been used to monitor the
desired structure and texture can be established in the design of structural development of protein–polysaccharide mixed
new products. Specically, reversible hydrogels are sought for systems without altering the gel structures.4,5,11,29 The objective
diverse applications, including cell delivery, tissue engineering, of this study was to investigate the role of blg and XG in gel
coating, and in the elds of biomedicine, cosmetics and phar- formation and to describe in greater detail the gelation mech-
maceuticals.19–21 Here, we report on the gelation of mixed anism of blg–XG mixtures under associative conditions using
aqueous solutions of the globular protein b-lactoglobulin (blg) both techniques to monitor the structural development of
and polysaccharide xanthan gum (XG). Both biopolymers are coupled gels as they are affected by the biopolymer concentra-
extensively used because they are naturally occurring. blg is the tion and ratios. We show that the role of XG is to provide a
main protein component of the whey fraction and has a molar framework for gel organization, whereas the role of blg is to
mass of 18.6 kg mol1. It has an isoelectric point (Ip) of 5.1.22 participate in network development as a crosslinking agent.
Above a critical protein concentration, irreversible aggregation
leads to the formation of a gel when a blg dispersion is heated
above approximately 60  C. The gelation of a blg dispersion has
Experimental section
not been observed at ambient temperatures, even when the pH Materials
has been brought to the isoelectric point of the protein.12 blg is Whey protein isolate (High-Beta, lot # JE 002-8-922, 98.2 wt%
able to form electrostatic complexes with anionic poly- protein, 85 wt% of which is blg, 1.8 wt% minerals, and 4 wt%
electrolytes at pH levels above its Ip through its positively moisture) was obtained from Davisco Foods International Inc.
charged patches.23 Due to the high content of blg in this powder, it was assumed
XG is an anionic polysaccharide produced commercially by that its behavior was governed by that of blg;12,30 therefore, this
bacterial fermentation. It is widely used in the food industry as powder will be referred to as blg hereaer. Xanthan gum (XG)
a stabilizer and thickener of food products due to its specic (Keltrol RD, lot # 4G3367K, 96.36 wt% total sugar, 3.02 wt%
physical (viscosity, pseudo-plasticity) and chemical (water protein) was provided by CP. Kelco UK Ltd. Glucono-d-lactone
solubility, pH stability) properties. This polysaccharide (GDL, lot # 70H0163, USA) and rhodamine B isothiocyanate (lot
consists of a linear (1–4)-b-D-glucose backbone with a charged # G39595, USA) were purchased from Sigma and J.T. Baker,
trisaccharide side chain on every second glucose residue. XG respectively.
exhibits pronounced pseudoplastic behavior due to its strong
self-associative behavior.24 The chain–chain association is
modulated by changes in the conformation of xanthan gum Sample preparation
with salt content and temperature. In distilled water, XG XG and blg powders were dissolved in ltered deionized water
appears elongated due to the electrostatic repulsion from the (Modulab Analytical, Fisher Scientic) with continuous stirring
charged groups in the lateral chains; the coiling or bending of at 25  C for at least 1 h for protein and 2 h for polysaccharide
XG molecules appeared to occur when some salt was present.25 and le overnight at 4  C. Before preparing the mixtures for
Heating a XG solution above the transition temperature, Tm, analysis, the initial pH of the protein and polysaccharide
results in the melting of the ordered structure. This structure dispersions was adjusted to 6.60  0.08 using 0.1 N HCl and
returns to its original state upon cooling, depending on the 0.1 N NaOH. Four blg–XG ratios (r) were studied: 2 : 1, 5 : 1,
salt environment.26 10 : 1 and 20 : 1 (r ¼ 2, 5, 10, and 20, respectively). The mixed

3064 | Soft Matter, 2013, 9, 3063–3073 This journal is ª The Royal Society of Chemistry 2013
View Article Online

Paper Soft Matter

dispersions were prepared at (1) a xed total solid concentration stirring for 1 h before mixing with XG dispersions. The FA-XG
of 0.3 wt% and different blg–XG ratios, (2) a xed XG concen- prepared according to the method of Donato et al.5 was
tration of 0.03 wt% with different concentrations of blg (0.06, provided by Laneuville and was used in the mixture with RITC-
0.15, 0.30 and 0.60 wt%), (3) a xed protein concentration of 0.3 blg in order to determine the location of blg and XG in the gel.
wt% and different XG concentrations, and (4) different total Preliminary trials revealed that labeling did not change the
solid concentrations to study the effect of protein–poly- rheological behavior of the systems (results not shown).
saccharide ratio and protein, polysaccharide and total solid Mixtures of RITC-blg–XG and RITC-blg–FA-XG were prepared as
concentration on gel formation. Protein and polysaccharide described above. Aer GDL addition, the sample was poured
dispersions were mixed at the desired r and stirred gently for 30 between a 0.5 mm deep well concavity slide and a cover slip,
min. The gelation of the mixed system was induced by the which was hermetically sealed with nail enamel. Micrographs
addition of GDL. The amount of GDL used in each formulation were taken during gelation and approximately 24 h aer GDL
was related to the total biopolymer concentrations in the addition. The previous study reported that the structure of the
mixture to obtain a nal pH of 4.4  0.08 at room temperature. gels was fully developed aer z18 h of acidication.12 Digital
The pH of the dispersions was continuously measured using a images were acquired at a pixel resolution of 512  512 or
DL53 titrator (Mettle Toledo, Switzerland). 1024  1024.

Dynamic oscillatory measurements


z-Potential measurement
The small-deformation properties of the mixed gels during
To estimate the optimal ratio of blg–XG for electrostatic inter-
acidication were determined in dynamic oscillation mode with
action, the net charge of complexes formed under shear
a controlled-strain rheometer (ARES-100, TA-Instrument, Pis-
conditions was measured. The electrical charges (z-potential) of
cataway, NJ, USA) equipped with a circulating water bath
individual biopolymers and biopolymer complexes as a func-
temperature controller. A Couette device with a cup (33.93 mm
tion of pH and at the nal pH of the systems in an aqueous
diameter) and bob system (32.05 mm diameter, 33.29 mm
solution were determined using a Zetasizer 2000 (Malvern
length) was used. Thirty seconds aer GDL addition, the
Instruments Ltd, Worcs, UK). Before analysis, samples were
samples were transferred to the rheometer. The development of
diluted in ltered deionized water to obtain solutions of 0.05
the storage modulus (G0 ) and the loss modulus (G00 ) during
wt% solid total concentration. Measurements were carried out
gelation was recorded for 15 h at 25  C at a frequency of 0.1 Hz
in duplicate, and the results are reported as means and stan-
and a strain of 0.5%. Following oscillation, a strain sweep test
dard deviations.
was recorded at increasing strains from 0.1 to 100% to verify the
linear region. Samples were covered with a thin lm of castor oil
to prevent evaporation during measurement. The interconnec- Results and discussions
tion point, pH4, at which interpolymeric complexes formed, Gelation process
was determined as the last point before the exponential
increase in G0 , and the gelation point, pHg, was arbitrarily A typical evolution of G0 , G00 versus pH is shown in Fig. 1a. Three
chosen as the pH at which G0 was equal to 1 Pa (ref. 31) (see stages (denoted as I, II, and III in Fig. 1a) can be clearly iden-
more details in the Gelation process section). tied. In stage I, when the pH is higher than 5.5 and the protein
net charge is negative, the mixture remains in the sol state.
Turbidity measurements Although both G0 and G00 showed large deviations due to low
torque signals, G0 appeared to be higher than G00 and neither
The development of turbidity in the mixed dispersions aer changed during stage I. However, the absorbance was observed
GDL addition was monitored by measuring absorbance varia- to increase slightly (Fig. 1b) in stage I.
tions as a function of time at 800 nm using an HP 8453 UV- As pH is further reduced (stage II), G0 , G00 and the absorbance
visible spectrophotometer (Agilent Technologies, Inc. Santa increase more rapidly and then approach a plateau (stage III).
Clara, CA, USA) at 25  C. A blank correction was made under the No crossover between G0 and G00 was observed during the gela-
same conditions in the absence of both biopolymers. The tion process; in fact, G0 was already higher than G00 at the
inection point in the turbidity curves was taken as the inter- beginning of the measurement. It may be possible that the
connection point, pH4.30 mixture was already highly structured well before the gelation
point. This behavior may be derived from the pseudo-plastic
Confocal laser scanning microscopy properties of XG over a wide range of pH, through which XG
The structural features of the gels were investigated using a solutions exhibit dominant elastic responses to frequency.24
Nikon Eclipse TE2000-E Confocal Laser Scanning Microscope Previous studies12,32 reported that regardless of the experi-
(Tokyo, Japan) equipped with an inverted microscope. The mental conditions (under shear or quiescent conditions), elec-
objective lens used was 60X WI, NA 1.20. The excitation wave- trostatic attraction between blg and XG can occur at pHc [ Ip
length was 543 nm, with an emission maximum at 605/675 nm. of blg due to positively charged patches on the protein molecule
Because proteins do not exhibit intrinsic uorescence at this surface, which form soluble complexes followed by the forma-
wavelength, proteins were stained by adding rhodamine B iso- tion of interbiopolymer complexes with a further decrease
thiocyanate (RITC) to protein dispersions under magnetic in pH.23,33 The pHc for this system under shear was equal to

This journal is ª The Royal Society of Chemistry 2013 Soft Matter, 2013, 9, 3063–3073 | 3065
View Article Online

Soft Matter Paper

paralleled with the increase in gel strength observed by


rheology, indicating interpolymeric formation by the intercon-
nection of soluble complexes. In the rheological prole (Fig. 1a
and S1 – see ESI†), this point is characterized by an initial
increase in G0 (point 1); the associated pH and time values were
denoted as pH4 and t4, respectively. Beyond this pH value, open
and stringy structures are observed at pH 5.46 (Fig. 2) and G0
exceeds the value of 1 Pa at which gelation occurs. Comparing
inset (a) and inset (d) in Fig. 2, we can nd that a highly inter-
linked, permanent network instead of the sol state is formed at
pH 4.4, which strongly agrees with the high G0 observed in stage
III (Fig. 1a and S1†). The data of frequency sweep tests showed
that storage modulus and loss modulus were slightly dependent
on frequency which indicated an almost uniform solid response
over the frequency window studied.41 This is consistent with
network bond relaxation moving to much longer timescales and
G00 /G0 ratio varies from 0.01 to 0.1 suggesting that the gel
properties are intermediate between entanglement network and
true gel, and so are characteristic of a weak gel.42
The microstructure during the unraveling of the gelation
processes (see ESI†) revealed that the nal microstructure was
obtained at the end of stage II. In addition, no change in G0 was
observed in stage III (Fig. 1a and S1†); thus, stage III is called the
quasi-equilibrium stage. The pH at which G0 did not change is
called the electrical charge equivalence pH (EEP). De Jong et al.
Fig. 1 pH dependence of (a) storage modulus and loss modulus, and (b) also reported no further change in the microstructure of mixed
absorbance for the blg–XG mixture at r ¼ 5 and a total solid concentration of 0.30
whey protein–polysaccharide cold-set gels aer gelation.11
wt%. The white squares (Fig. 1a) indicate the point at which the gel is stable and
the deviation of maximum G0 does not exceed 5%.

Effects of blg–XG ratio


5.76  0.03. Similar effects have also been observed in other
32
The effects of the biopolymer ratio on gelation processes are
protein–polyelectrolyte systems.30,34–36 Thus, the slight increase presented in Fig. 3. The evolution of storage modulus as a
in turbidity observed in stage I may correspond to the formation function of time with different ratios is also presented in Fig. S2
of soluble complexes and the rapid increase in G0 , G00 , and the (see ESI†)
turbidity observed early in stage II may reect the formation of As the ratio increases, interpolymer complexes are formed
interbiopolymer complexes. In addition, Fig. 1a and b show that and gelation occurs at higher pH and sooner (see also Table 1
turbidity measurements are more sensitive to structural and Fig. S2†). This behavior can be explained by the faster
changes than rheological measurements in stage I, stage II was diffusion of proteins to polysaccharide chains in systems
initiated at pH4 ¼ 5.66 (Fig. 1b) instead of pH4 ¼ 5.50 (Fig. 1a). featuring smaller densities of polysaccharides with lower
To determine the gelation point, Winter and Chambon's viscosities. In addition, more proteins were available for inter-
criteria37 seem to be the most rigorous among those prescribed action, and the systems were neutralized sooner.12,30,32 More-
by several methods. However, as sol–gel transition takes place over, stronger gels were obtained as the XG fraction
from the solution state, this method is oen very difficult to incorporated into the gel formulation increased. This result
apply due to the low viscoelastic moduli and low torque signals suggests that the gel network may be based on XG chains. In
near the gelation point.38 Therefore, another method, i.e., one addition, Fig. 3 shows that a maximum and constant G0 is
that tracks the time or the pH at which the G0 rose above a obtained for r ¼ 5 and r ¼ 10 at the electrical charge equivalence
threshold value of 1 Pa, was used. This method is now widely pHs, indicating the existence of an electrostatic equilibrium in
used in the literature.31,39,40 The gelation pH, pHg, represents a the gel, where different biopolymers carry similar but opposite
gelation point determined as a function of pH (point 2 in charges and their interaction is maximal, and the gel structure
Fig. 1a) and tg is the corresponding time of gelation (point 2 in may be stabilized at this pH and does not change upon the
Fig. S1†). decrease of pH; The electrostatic equilibrium is attained more
The corresponding microstructures obtained for a blg–XG quickly (higher EEP is obtained at higher ratios – Table 1) as the
mixture at r ¼ 5 and a total solid concentration of 0.30 wt% ratio increases due to the availability of a greater number of
during acidication are shown in Fig. 2. proteins for interaction. However, the gradual evolution of G0
Fig. 2 clearly shows that macroscopic phase separation continued to take place for r ¼ 2 (Fig. 3), and G0 did not expe-
occurs at pH 5.50 and that there is a coarsening of the network rience a plateau aer reaching the Ip of the proteins. This is
structure. This coarsening of the network with decreasing pH is possibly due to the interaction between protein-unsaturated XG

3066 | Soft Matter, 2013, 9, 3063–3073 This journal is ª The Royal Society of Chemistry 2013
View Article Online

Paper Soft Matter

Fig. 2 Microstructure formation for RITC-blg–XG mixture at r ¼ 5 and total solid concentration 0.30 wt% observed by confocal laser scanning microscopy during
acidification. The scale bar is 20 mm in all images. The light areas are protein rich.

concentration. The concentration of blg strongly affects the


gelation kinetics (see Fig. S3 in the ESI† for time evolution of
storage modulus).
An increase in protein concentration from 0.06 to 0.6 wt%
results in an increase in pHg from pH 5.0 to pH 5.5 and a
decrease in tg from 210 minutes to 35 minutes (Fig. S3†). In
addition, in the blg concentration range between 0.06 and 0.30
wt%, which corresponds to a range of blg–XG ratios from 2 to
10, the gel obtained at r ¼ 5 presented the highest solid-like
character, indicating that this ratio is close to the optimal ratio
that fosters the blg–XG interaction. The weakness of the gel at a
ratio of 10 may be due to the effect of excess protein,12 which
Fig. 3 Evolution of the storage modulus during gelation for blg–XG mixtures at
promoted the association of proteins with individual XG
a total solid concentration of 0.30 wt% and different ratios. molecules to form individual complexes, with a consequent
reduction in the junction zones of the gel network. However, at
very high protein concentrations (0.60 wt%), the gel obtained
Table 1 Gelation kinetic parameters for blg–XG mixtures at a total solid was stronger than at r ¼ 5, which may be explained by the
concentration of 0.30 wt% and different ratios cooperative binding of proteins at a pH close to the Ip of the
protein. At this pH and as the local concentration of protein
blg–XG
increases, the repulsive forces between protein molecules are
ratios t4 (min) tg (min) pH4 pHg EEPa
negligible, which may favor the hydrophobic/hydrogen inter-
2:1 100 110 5.37  0.03 5.22  0.05 N/Ab actions between protein particles resulting in the formation of a
5:1 98  4 108  4 5.42  0.05 5.37  0.05 4.55  0.03 second protein network in a network of coupled gels. The
10 : 1 60  7 72  17 5.60  0.14 5.55  0.17 4.62
a b
EEP (electrical charge equivalence pH). Not attained under the
current condition.

chains by hydrogen bonds or to the higher viscosity of this


system,12 which slowed down the interactions between inter-
polymeric complexes. To verify the role of viscosity and the role
of protein availability on the kinetics of gelation, some blg–XG
mixtures at a xed polysaccharide concentration and different
blg concentrations were tested by dynamic oscillatory
measurements.

Effects of protein concentrations


Fig. 4 shows the protein concentration dependence of gelation Fig. 4 Evolution of storage modulus as a function of pH for blg–XG mixtures at a
processes for blg–XG mixtures at a xed polysaccharide fixed polysaccharide (0.03 wt%) and different protein concentrations.

This journal is ª The Royal Society of Chemistry 2013 Soft Matter, 2013, 9, 3063–3073 | 3067
View Article Online

Soft Matter Paper

attractive interaction between blg molecules, especially as they negative net charge on the XG molecules should drop to zero
complex with pectin at pH levels ranging from 4 to 6, was and the electrostatic interaction is maximal under shear
reported by Girard, Turgeon et al. 2003.23 Faster gelation at high conditions may occur between 2 and 5. At a ratio of 2, the
protein concentrations (see Fig. S3†) indicates that as more insoluble complexes exhibited a net negative charge, indicating
protein is available, more of the gel structure is rapidly devel- that the negative charge of the XG dominated the charge of the
oped. This also suggests that the crosslinking function of blg. At higher ratios (r $ 5), the net charge of all systems is
proteins in network development and the availability of positive and similar to that of the protein, indicating an excess
proteins are determining factors for gelation kinetics. of blg. As the gel is mainly stabilized by the electrostatic inter-
The z-potentials of individual components and three mixed action, the maximum of electrostatic interaction should give a
blg–XG solutions under shear conditions as a function of pH gel with the highest solid-like character. As expected, the
and at pH 4.40 are shown in Fig. 5. highest solid-like character was obtained at r ¼ 3.5 for nal gels
As XG is an acidic polysaccharide having the carboxyl groups (Fig. 6), so the optimal ratio for electrostatic interactions was
on the side chains, the negative zeta potential decreased with estimated to be approximately 3.5. The highest G0 obtained at
decrease in pH. At pH 4.8, the net charge on blg is close to zero r ¼ 2 in Fig. 3 is explained by the highest proportion of XG in the
because this pH is near to the electrostatic point of the gel formulation where the total solid concentration was xed at
protein.22 When the pH of solution was decreased from 6.0 to 0.30 wt%.
4.4 the z-potentials of mixed solutions were more positive than
that of the pure XG solution, indicating that the electrostatic
complexes had been formed. Moreover, at higher ratios blg–XG
Effects of polysaccharide concentrations at a xed protein
(>2), the systems were neutralized sooner at the electrical charge
concentration
equivalence pHs (EEPs), and aer these pHs the net charges of
the mixed solutions were similar to that of the protein, indi- The gelation processes as a function of XG concentrations are
cating that the electrostatic complexes had been neutralized shown in Fig. 7. At 0.03 wt% XG and a ratio of 10, the gelation
and the net charge of systems came from that of the free was reached at pHg 5.5, which is higher than that (pHg 5.3)
proteins in the mixed solutions. Fig. 5b suggests that at pH 4.4 reached by 0.06 wt% XG at a ratio of 5. This behavior is
the stoichiometric equivalence ratio blg–XG at which an initially

Fig. 6 Storage modulus at pH 4.4 for b-lg–xanthan gels at a fixed concentration


of xanthan (0.03 wt%) and different blg–XG ratios. Different letters indicate
significant difference between ratios evaluated by the Tukey procedure (p < 0.05).

Fig. 5 z-Potentials for individual biopolymers, blg–XG complexes at different


ratios and a constant polysaccharide concentration of 0.03 wt% (a) as a function Fig. 7 Evolution of storage modulus as a function of pH for blg–XG mixtures at
of pH and (b) at pH 4.4. fixed protein (0.3 wt%) and different XG concentrations.

3068 | Soft Matter, 2013, 9, 3063–3073 This journal is ª The Royal Society of Chemistry 2013
View Article Online

Paper Soft Matter

explained by the availability of blg at a ratio of 10 as described concentration of blg is 0.024 wt%. This critical XG concentra-
above. tion is in the vicinity of the overlap concentration for an XG
The doubling of the XG concentration from 0.03 wt% to 0.06 solution at which the intermolecular interactions between XG
wt% increased the G0 6.8-fold, from 53 to 362 Pa. The stronger polymer chains through hydrogen bonding and polymer
gel obtained for an XG concentration of 0.06 wt% again suggests entanglement result in the formation of a complex network of
that XG plays a key role in the formation of gel networks. entangled rod-like molecules.43–45 The overlap concentration
has been reported to be as low as 103, 5.0  103 and 12.6 
103 wt%.44,46,47 This indicates that the XG chain overlap is a pre-
Effects of total solid concentrations at ratio 5
requisite for mixtures to gel.
The effects of the solid total concentration on the gelation
processes and the effects of the XG concentration on the storage
modulus at the nal pH are presented in Fig. 8a and b, Role of the protein and polysaccharide in network formation
respectively. Fig. 9 shows different networks formed from different quanti-
The gelation point, pHg, seemed to be independent of the ties of XG and blg in the gel formulations. The bright zones
solid total concentration; gelation occurred at pH z 5.3. The correspond to zones enriched in proteins and dark zones
ratio or the availability of proteins was the same for all formu- correspond to zones devoid of blg.
lations. The effect of solid total concentration is a combination Fig. 9a and b show a gel obtained with the same quantity of
of the effects of the viscosity of a mixture and the volume blg but a different quantity of XG. The network became denser
occupied by XG. At higher concentrations, the gelation may slow with an increase in the XG concentration, which explains the
down due to high viscosity, but it may also be enhanced by the higher G0 observed at pH 4.4 for the gel at an XG concentration
volume occupied by XG. However, stronger gels were obtained of 0.06 wt% and a ratio of 5 (Fig. 7). A network with larger dark
0
at higher biopolymer concentrations. The dependence of GpH4.4 zones is obtained in the gel formulation with high protein
on the XG concentration was marked, particularly at low content at a constant polysaccharide concentration (compare
0
concentrations (Fig. 8b). The value of GpH4.4 was extrapolated to Fig. 9b with Fig. 9c). The network density is proportional to the
1 Pascal for an XG concentration of 4.8  103 wt%. This value XG concentration, which suggests that the blg aggregated on XG
(Co) can be taken as an estimate of the lowest XG concentration chains, so dark zones may correspond to zones devoid of
at a ratio of 5 at which gelation is possible; the corresponding biopolymers, and pores of the gel network. An open network
shown in Fig. 9b may have resulted from the effect of excess
protein in the system. At a ratio of 10, the repulsive forces
caused by the excess proteins in the porous network are more
predominant, which may result in a larger pore size. This result
is in agreement with those presented in Fig. 4, where the gel at a
ratio of 5 is observed to be stronger than the gel at a ratio of 10.
At a very low XG concentration (CXG ¼ 4.50  103 wt% < Co)
and a high blg concentration (4.50  102 wt%), a complex was
formed instead of a gel (Fig. 10a). However, at a higher XG
concentration (CXG ¼ 8.30  103 wt% > Co), the gelation of the
blg–XG mixture occurred (Fig. 10b).
To complete these observations and to ensure that the blg
aggregated on XG chains, observations on mixtures containing
both RITC-blg and FA-XG were performed. Fig. 11 shows
the nal microstructures obtained for the mixture of 0.06 wt%
FA-XG and 0.30 wt% RITC-blg.
The uorescence of each polymer was coded: green for XG
uorescence, red for blg uorescence. The red and green
networks obtained in Fig. 11a and b respectively correspond to
protein and polysaccharide networks. Fig. 11c shows the nal
microstructure obtained for 0.30 wt% blg–0.06 wt% XG mixture
when the two networks were superimposed. The three networks
are identical. So, dark zones correspond to zones devoid of
biopolymers and proteins and XG being located in the same
location. These observations conrmed that XG provided a
frame for gel organization, and blg aggregated along the XG
chains and could be regarded as crosslinking agents. Iijima
Fig. 8 (a) Evolution of storage modulus as a function of pH for blg–XG mixtures
et al.48 reported that aer annealing of XG solution at 1.0 wt%
at ratio 5 and different solid total concentrations. (b) G0 as a function of XG for 24 h at 40  C and subsequently cooling it, a homogeneous
concentrations for gels at pH 4.4 and a ratio of 5. network structure was obtained that is comparable to the

This journal is ª The Royal Society of Chemistry 2013 Soft Matter, 2013, 9, 3063–3073 | 3069
View Article Online

Soft Matter Paper

Fig. 9 Confocal micrograph of RITC-blg–XG gels obtained under different conditions. The images represent an area of 91.87  91.87 mm. The light areas are protein
rich.

103, 5.0  103 and 12.6  103 wt%.44,46,47 In addition, the


highly ordered network of entangled stiff molecules results in a
XG solution with the viscoelastic characteristics of a weak gel
and a dominant elastic response to frequency.24 Moreover,
Iijima et al.48 reported that large-scale molecular assemblies are
formed and act as colloidal dispersions when XG is dissolved in
water. Kobori et al.50 also reported the aggregation of sodium
caseinate along XG chains at a pH level below the Ip of this
protein.
These results conrm the hypothesis that the gel networks
are based on XG chains. Therefore, the density of the gels
Fig. 10 Confocal micrograph of RITC-blg–XG: (a) complex formation at a total increased with increasing XG content (Fig. 9a and b). This is not
solid concentration of 0.05 wt% and a ratio of 10; (b) gel formation at 0.05 wt% in agreement with the previous result that indicated that the
and a ratio of 5. The images represent an area of 212.17  212.17 mm. The light density of the gels, based on fractal dimensions of systems,
areas are protein rich. increased with increasing protein content.12 In recent report, we
found that, by image analysis, the fractal dimension increases
with increasing XG concentration.41 These ndings suggest also
network obtained in our study by CLSM (Fig. 9). In distilled that there exists a limiting molecular weight of XG required for
water and at a very dilute solution (0.02  103 wt%) XG chains the formation of a gel below which particulate electrostatic
exist as single strands and isolated molecules are observed by complexes instead of a stable network will form, as reported by
AFM.49 The molecules observed are approximately 0.8 mm in Laneuville et al.51
length, 2 nm in width and 8 nm in thickness. In their quiescent
state and at the critical overlap concentration (c*), the inter-
molecular interactions between XG polymer chains through Identication of gelation mechanism
hydrogen bonding and polymer entanglement result in the It is commonly accepted that XG does not gel on its own
formation of a complex network of entangled rod-like mole- through the typical gelation mechanism, although there is
cules.43–45 This concentration has been reported to be as low as considerable evidence of strong intermolecular interactions in

Fig. 11 Localization of blg and xanthan in the mixture of 0.30 wt% blg and 0.06 wt% XG observed by confocal laser scanning microscopy. Scale bar corresponds to
30 mm and is appropriate to all images.

3070 | Soft Matter, 2013, 9, 3063–3073 This journal is ª The Royal Society of Chemistry 2013
View Article Online

Paper Soft Matter

solution.24,25,46,48,50,52–57 In the presence of divalent or trivalent formation of soluble and interpolymeric electrostatic
metal ions, the sol–gel transition occurred.58–60 These authors complexes. However, the interconnection between the inter-
consider the metal ion crosslinking agents that link the nega- polymeric complexes seems to spontaneously occur during the
tively charged xanthan chains to the gel network. However, the second stage due to the entanglement of XG chains and the
limiting concentration of XG58 needed for gelation with 5 mM decrease in the repulsive interaction between them as the pH is
Cr(III) at 20  C is higher than that in this study (0.35 wt% (ref. 58) reduced. Moreover, in this study, the role of both biopolymers
versus 4.8  103 wt%) which may indicate that the size of the in network formation induced by the electrostatic attractive
cross-linking agent may be important for XG hydrogel forma- interaction was reported for the rst time, blg may be regarded
tion. Recently, some authors have reported the formation of XG as a crosslinking agent at pH4, i.e., there may exist at least two
gels by annealing48 and freezing and thawing XG solutions.61 XG molecules that share the same blg molecules to support
These authors interpreted the gelation of XG as the result of electrostatic interactions, as reported by Girard, Turgeon et al.;23
intermolecular interactions during processing. Fogolari, Ragona et al.63 that blg may likely interact with nega-
Based on the previously discussed CLSM, the rheological tively charged macromolecules through more than one amino
proles and earlier reports in the literature, the gel formation acid zone. Therefore, the linkage of XG chains though blg seems
mechanism is proposed as follows (see Fig. 12). In aqueous to be identical to a polymerization process, which is supported
solution, xanthan gum exhibits pseudoplastic behavior due to by the lamentous nature of the gel structures (Fig. 9 and 11).
the presence of large-scale molecular assemblies.48 The assem- This is contrary to the aggregate nature reported in the previous
blies of xanthan gum molecules occur by an end-to-end asso- study.12 Due to the reduced mobility of XG chains in the system,
ciation52 or by a side–side association.62 each XG chain may be considered a nucleus. Furthermore, as
As the pH decreases, the repulsive interaction between XG pH is reduced, the proteins will diffuse to the XG chain and
chains decreases, while the attractive interaction between blg aggregate, due to attractive interactions that support the
and XG increases. The aggregation of proteins through the formation of soluble complexes, followed by the formation of
polysaccharide chains and the decrease in pH may enhance the interpolymeric complexes. The growth of soluble complexes by
intermolecular interaction between XG chains through blg. the aggregation of blg through XG chains indicates that there is
When the pH approaches the Ip of the protein and is still higher an increase in the size of structural domains, which is consis-
than pH4, the interaction between negatively charged groups on tent with the increase in the scattered light intensity observed in
XG chains with positively charged patches on the blg results in the previous study during gelation.12 Understanding the
the formation of soluble complexes, followed by the formation mechanism of this type of structuration is important because it
of interpolymer complexes as the pH decreases further. The could be representative of other protein–polysaccharide asso-
interpolymer complexes may also be soluble up to the point ciative behaviors found in various biological systems (protein–
where large-scale interactions are forged. The association that hyaluronic acid,64,65 chitosan–protein,66 etc.) or because it could
occurs as electrostatic repulsion decreases results in a sol–gel be used to develop materials for tissue-engineering applica-
transition at the point of gelation. At high blg–XG ratios, larger tions67 and in the control of the rheological properties of some
strands were observed in the gel structures (Fig. 9c and d); thus, food products. For example, the production of very low
the gels may have multiple layers of proteins that aggregate concentrations of exopolysaccharide by lactic acid bacteria
along XG chains. It should be noted that the gels obtained in signicantly affects the properties of fermented product gel and
this study are pH-reversible. The gels melt rapidly when sus- water-retention capacity.68,69 Protein-anionic polysaccharides
pended in phosphate buffer at pH 7 above the Ip of the protein may become as important as cationic–anionic biopolymers in
(results not shown); however, they can gel again when the pH materials development.70 Moreover, the gels formed in this
decreases below the Ip of the protein, which affirms that the study are pH-reversible hydrogels due to the electrostatic nature
junction zones are electrostatic in nature. This proposition is of the junction zones. This reversibility is attractive for the
slightly different from that reported previously.12 In both design of physical gels geared toward biomedical applications19
propositions, the gelation process is stated to initiate with the or controlled release of bioactive molecules.71

Fig. 12 Schematic of the interaction and gel formation process between blg (circles) and XG (rod sharp); rectangle presents aggregation zones of blg on XG chains,
arrow presents electrostatic cross-linking zones of XG chains by blg.

This journal is ª The Royal Society of Chemistry 2013 Soft Matter, 2013, 9, 3063–3073 | 3071
View Article Online

Soft Matter Paper

Conclusions 14 S. L. Turgeon, M. Beaulieu, C. Schmitt and C. Sanchez, Curr.


Opin. Colloid Interface Sci., 2003, 8(4–5), 401–414.
The gelation kinetics and structural evolution of electrostatic 15 V. Tolstoguzov, Food Hydrocolloids, 2003, 17, 1–23.
blg–XG gels are followed by rheological measurement and 16 K. Ako, D. Durand and T. Nicolai, So Matter, 2011, 7, 2507–
confocal scanning microscopy. The blg aggregated along the XG 2516.
chains that initially provided a tenuous network for gel struc- 17 N. W. H. Cheetham and E. N. M. Mashimba, Carbohydr.
turation and could be regarded as a crosslinking agent. The Polym., 1988, 9, 195–212.
gelation process followed three stages: induction stage, gelation 18 M. Tako, T. Teruya, Y. Tamaki and K. Ohkawa, Colloid Polym.
stage, and quasiequilibrium stage, corresponding to the Sci., 2010, 288, 1161–1166.
formation of soluble complexes, the formation of junction 19 C. Tsitsilianis, So Matter, 2010, 6, 2372–2388.
zones followed by the formation of interpolymer complexes, 20 Y. Zhao, H. Yokoi, M. Tanaka, T. Kinoshita and T. Tan,
and the stabilization of the gel at the nal pH, respectively. The Biomacromolecules, 2008, 9, 1511–1518.
gel strength mainly depended on the XG content. The gel 21 H. Yan, H. Frielinghaus, A. Nykanen, J. Ruokolainen,
structure stabilized at an electrical charge equivalence pH A. Saiani and A. F. Miller, So Matter, 2008, 4, 1313–1325.
during the acidication and did not change aer this pH (EEP), 22 E. H. C. Bromley, M. R. H. Krebs and A. M. Donald, Faraday
and this EEP value depends on the biopolymer ratio. The blg– Discuss., 2005, 128, 13–27.
XG ratio was observed to be an important factor in controlling 23 M. Girard, S. L. Turgeon and S. F. Gauthier, J. Agric. Food
the gelation process, the gel strength, and the gel structure, and Chem., 2003, 51, 6043–6049.
an optimal ratio for electrostatic interaction was estimated. 24 G. Sworn, in Food Stabilisers, Thickeners and Gelling Agents,
This novel hydrogel could be used for encapsulation of active Wiley-Blackwell, 2009, pp. 325–342.
molecules. 25 T. A. Camesano and K. J. Wilkinson, Biomacromolecules,
2001, 2, 1184–1191.
Acknowledgements 26 I. T. Norton, D. M. Goodall, S. A. Frangou, E. R. Morris and
D. A. Rees, J. Mol. Biol., 1984, 175, 371–394.
ACDI/PCBF is acknowledged for its fellowship for Xuan Thang 27 R. Mezzenga, P. Schurtenberger, A. Burbidge and M. Michel,
Le along with Diane Gagnon and Sandra I. Laneuville for their Nat. Mater., 2005, 4, 729–740.
technical assistance and fruitful discussion. This work has been 28 K. L. Spiller, S. J. Laurencin, D. Charlton, S. A. Maher and
funded by NSERC discovery grant. A. M. Lowman, Acta Biomater., 2008, 4, 17–25.
29 R. Lutz, A. Aserin, Y. Portnoy, M. Gottlieb and N. Garti,
Notes and references Colloids Surf., B, 2009, 69, 43–50.
30 F. Weinbreck, R. de Vries, P. Schrooyen and C. G. de Kruif,
1 J. L. Drury and D. J. Mooney, Biomaterials, 2003, 24, 4337– Biomacromolecules, 2003, 4, 293–303.
4351. 31 Y. L. Tan, A. Y. H. Singh and Y. Hemar, J. Texture Stud., 2007,
2 J.-Y. Xiong, J. Narayanan, X.-Y. Liu, T. K. Chong, S. B. Chen 38, 404–422.
and T.-S. Chung, J. Phys. Chem. B, 2005, 109, 5638–5643. 32 S. Laneuville, C. Sanchez, S. Turgeon, J. Hardy and P. Paquin,
3 L. van den Berg, Y. Rosenberg, M. A. J. S. van Boekel, in Food Colloids: Interactions, Microstructure and Processing,
M. Rosenberg and F. van de Velde, Food Hydrocolloids, ed. E. Dickinson, The Royal Society of Chemistry,
2009, 23, 1288–1298. Cambridge, 2005, pp. 443–465.
4 H. Firoozmand, B. S. Murray and E. Dickinson, Langmuir, 33 J. M. Park, B. B. Muhoberac, P. L. Dubin and J. Xia,
2007, 23, 4646–4650. Macromolecules, 1992, 25, 290–295.
5 L. Donato, C. Garnier, B. Novales, S. Durand and 34 K. W. Mattison, I. J. Brittain and P. L. Dubin, Biotechnol.
J.-L. Doublier, Biomacromolecules, 2005, 6, 374–385. Prog., 1995, 11, 632–637.
6 I. Capron, T. Nicolai and D. Durand, Food Hydrocolloids, 35 J. Xia, P. L. Dubin and H. Dautzenberg, Langmuir, 1993, 9,
1999, 13, 1–5. 2015–2019.
7 S. L. Turgeon, M. Beaulieu, C. Schmitt and C. Sanchez, Curr. 36 J. Xia and P. L. Dubin, in Macromolecular Complexes in
Opin. Colloid Interface Sci., 2003, 8, 401–414. Chemistry and Biology, ed. P. Dubin, J. Bock, R. Davis,
8 A. N. Gupta, H. B. Bohidar and V. K. Aswal, J. Phys. Chem. B, Donald N. Schulz and C. Thies, Springer-Verlag, 1994, pp.
2007, 111, 10137–10145. 247–271.
9 P. Cairns, M. J. Miles, V. J. Morris and G. J. Brownsey, 37 H. H. Winter and C. Francois, J. Rheol., 1986, 30, 367–382.
Carbohydr. Res., 1987, 160, 411–423. 38 S. B. Ross-Murphy, Rheol. Acta, 1991, 30, 401–411.
10 S. L. Turgeon and M. Beaulieu, Food Hydrocolloids, 2001, 15, 39 J. A. Lucey, M. Tamehana, H. Singh and P. A. Munro, Food
583–591. Res. Int., 1998, 31, 147–155.
11 S. de Jong, H. J. Klok and F. van de Velde, Food Hydrocolloids, 40 A. Tobitani and S. B. Ross-Murphy, Macromolecules, 1997, 30,
2009, 23, 755–764. 4845–4854.
12 S. I. Laneuville, S. L. Turgeon, C. Sanchez and P. Paquin, 41 X. Thang Le and S. L. Turgeon, in Gums and Stabilisers for the
Langmuir, 2006, 22, 7351–7357. Food Industry 16, The Royal Society of Chemistry, 2012,
13 M. F. Butler, Biomacromolecules, 2002, 3, 676–683. pp. 141–149.

3072 | Soft Matter, 2013, 9, 3063–3073 This journal is ª The Royal Society of Chemistry 2013
View Article Online

Paper Soft Matter

42 J.-L. Doublier and L. Choplin, Carbohydr. Res., 1989, 193, 57 J. Fujiwara, T. Iwanami, M. Takahashi, R. Tanaka,
215–226. T. Hatakeyama and H. Hatakeyama, Thermochim. Acta,
43 M. J. Wilkins, M. C. Davies, D. E. Jackson, C. J. Roberts and 2000, 352–353, 241–246.
S. J. B. Tendler, J. Microsc., 1993, 172, 215–221. 58 H. Nolte, S. John, O. Smidsrød and B. T. Stokke, Carbohydr.
44 A. Patrick Gunning, A. R. Kirby and V. J. Morris, Polym., 1992, 18, 243–251.
Ultramicroscopy, 1996, 63, 1–3. 59 T. Lund, O. Smidsrød, B. Torger Stokke and A. Elgsaeter,
45 G. Sworn and Monsanto, in Handbook of Hydrocolloids, ed. P. Carbohydr. Polym., 1988, 8, 245–256.
A. W. G. O. Phillips, CRC, Boca Raton, 2000, pp. 103–115. 60 Z. H. Mohammed, A. Haque, R. K. Richardson and
46 A. B. Rodd, D. E. Dunstan and D. V. Boger, Carbohydr. E. R. Morris, Carbohydr. Polym., 2007, 70, 38–45.
Polym., 2000, 42, 159–174. 61 P. Giannouli and E. R. Morris, Food Hydrocolloids, 2003, 17,
47 M. Milas, M. Rinaudo, M. Knipper and J. L. Schuppiser, 495–501.
Macromolecules, 1990, 23, 2506–2511. 62 V. J. Morris, D. Franklin and K. I'Anson, Carbohydr. Res.,
48 M. Iijima, M. Shinozaki, T. Hatakeyama, M. Takahashi and 1983, 121, 13–30.
H. Hatakeyama, Carbohydr. Polym., 2007, 68, 701–707. 63 F. Fogolari, L. Ragona, S. Licciardi, S. Romagnoli,
49 I. Capron, S. Alexandre and G. Muller, Polymer, 1998, 39, R. Michelutti, R. Ugolini and H. Molinari, Proteins: Struct.,
5725–5730. Funct., Genet., 2000, 39, 317–330.
50 T. Kobori, A. Matsumoto and S. Sugiyama, Carbohydr. 64 K. Kakehi, M. Kinoshita and S.-i. Yasueda, J. Chromatogr., B:
Polym., 2009, 75, 719–723. Anal. Technol. Biomed. Life Sci., 2003, 797, 347–355.
51 S. Laneuville Ballester, S. Turgeon, C. Sanchez and 65 X. Hu, Q. Lu, L. Sun, P. Cebe, X. Wang, X. Zhang and
P. Paquin, WO Patent 2007014447, 2007. D. L. Kaplan, Biomacromolecules, 2010, 11, 3178–
52 G. Holzwarth and E. B. Prestridge, Science, 1977, 197, 757– 3188.
759. 66 M. R. Kasimova, A. n. Velázquez-Campoy and H. M. Nielsen,
53 J. G. Southwick, A. M. Jamieson and J. Blackwell, Biomacromolecules, 2011, 12, 2534–2543.
Macromolecules, 1981, 14, 1728–1732. 67 S. Van Vlierberghe, P. Dubruel and E. Schacht,
54 T. S. Bjørn, S. Olav and E. Arnljot, Biopolymers, 1989, 28, 617– Biomacromolecules, 2011, 12, 1387–1408.
637. 68 P. Duboc and B. Mollet, Int. Dairy J., 2001, 11, 759–768.
55 A. R. Kirby, A. Patrick Gunning and V. J. Morris, Carbohydr. 69 A. N. Hassan, J. Dairy Sci., 2008, 91, 1282–1298.
Res., 1995, 267, 161–166. 70 M. Rinaudo, Polym. Int., 2008, 57, 397–430.
56 E. Pelletier, C. Viebke, J. Meadows and P. A. Williams, 71 Y. Zhang, L. Tao, S. Li and Y. Wei, Biomacromolecules, 2011,
Biopolymers, 2001, 59, 339–346. 12, 2894–2901.

This journal is ª The Royal Society of Chemistry 2013 Soft Matter, 2013, 9, 3063–3073 | 3073

You might also like