You are on page 1of 9

Food Research International 82 (2016) 95–103

Contents lists available at ScienceDirect

Food Research International

journal homepage: www.elsevier.com/locate/foodres

Improved thermal gelation of oat protein with the formation of


controlled phase-separated networks using dextrin and
carrageenan polysaccharides
Talina Vanessa Nieto Nieto, Yixiang Wang, Lech Ozimek, Lingyun Chen ⁎
Department of Agricultural, Food Nutritional Science, University of Alberta, Edmonton T6G 2P5, Canada

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

Article history: The thermal gelation of oat protein (OP) was investigated in the presence of polysaccharides at different pHs.
Received 19 November 2015 The compressive stress dramatically increased in these phase-separated protein–polysaccharides gels due to
Received in revised form 23 January 2016 an apparent increase in protein concentration. The polysaccharide structure significantly affected the degree of
Accepted 24 January 2016
phase-separation and gel mechanical properties. The observed two-fold increase in gel compressive stress can
Available online 28 January 2016
be attributed to strong repulsive forces caused by carrageenan molecules. These resulted in a greater degree of
Keywords:
phase-separation with the formation of carrageenan rich domains embedded in the protein phase, and a highly
Oat protein ordered protein network, stabilized by hydrogen and hydrophobic interactions. In the case of OP-dextrin gels, the
Polysaccharide rate of phase separation was slower than the rate of protein aggregation, thus the dextrin particles were uniform-
Carrageenan ly distributed within the protein network. This research contributes to the basic understanding required for
Gelation designing textures for novel plant-based protein products.
Phase-separation and gel strength © 2016 Elsevier Ltd. All rights reserved.

1. Introduction charge density of the polysaccharide, as well as processing conditions


such as pH and ionic strength, have a direct influence on the gel
Alternative protein sources are gaining importance as consumers are structure, and consequently also the gelling properties (De Jong, Klok,
incorporating more plant-based proteins in their diets and in some & Van de Velde, 2009). By modulating protein–polysaccharide inter-
cases, opting for a full replacement of animal protein. Currently soy actions, a synergistic effect can be achieved. To date, most of the
and wheat proteins dominate this market. As the plant-based protein research has focused on improving gelling properties of dairy protein–
ingredient sector grows, there is the opportunity to meet the increasing polysaccharide mixtures (Çakır et al., 2012; Croguennoc, Nicolai,
market requirements for novel functionalities and sensory attributes of Durand, & Clark, 2001; Weinbreck, Nieuwenhuijse, Robijn, & de Kruif,
plant-protein sources. Our previous work revealed that oat protein 2003). Research efforts on plant-protein gels have included soy (Baeza
could form polymer-like gels with percolating networks, opening sever- et al., 2002; Pires Vilela et al., 2011), pea (Liu, Elmer, Low, &
al potential gelling applications in various foods (Nieto-Nieto, Wang, Nickerson, 2010) and canola (Uruakpa & Arntfield, 2006). Yet the
Ozimek, & Chen, 2015). knowledge acquired is still very limited.
Previous studies have demonstrated that blends of proteins and The objective of this work is to systematically study how the
polysaccharides can create gels with different microstructure, which presence of different polysaccharides may impact the microstructure
can be applied to control or improve sensorial attributes in food prod- of oat protein gels and their subsequent gelling properties. Three poly-
ucts by providing texture and water retention (Baeza, Carp, Pérez, & saccharides were selected, including dextrin, carrageenan and chitosan.
Pilosof, 2002; Çakır et al., 2012; Pires Vilela, Cavallieri, & Lopes da Dextrin is a low molecular weight carbohydrate (~3 kDa) produced by
Cunha, 2011; Uruakpa & Arntfield, 2006). Coacervate networks the partial acid hydrolysis of starch, thus exhibiting the α-(1 → 4)-Glc
represent one continuous network formed when the attractive structure of amylose and the α-(1 → 4)- and α-(1 → 4,6)-Glc branched
protein–polysaccharide interactions are favorable. Phase-separated net- structure of amylopectin, but with a lower polymerization (Secundo &
works are formed when protein–polysaccharide interactions are segre- Guerrieri, 2005, Silva et al., 2014). Carrageenan is a gel-forming polysac-
gative, resulting in a bi-continuous phase. Interpenetrating networks charide derived from a number of seaweeds, and is a sulfated
are formed when no interactions are established between polymers; polygalactan with 15 to 40% ester-sulfate content, which makes it
here each polymer forms an independent network (Morris, 1986). The an anionic polysaccharide. It has an average relative molecular
weight well above 100 kDa and is formed by alternate units of D -
⁎ Corresponding author. galactose and 3,6-anhydro-galactose (3,6-AG) joined by α-1,3 and
E-mail address: lingyun.chen@ualberta.ca (L. Chen). β-1,4-glycosidic linkage (Necas & Bartosikova, 2013). Chitosan is a

http://dx.doi.org/10.1016/j.foodres.2016.01.027
0963-9969/© 2016 Elsevier Ltd. All rights reserved.
96 T.V. Nieto Nieto et al. / Food Research International 82 (2016) 95–103

linear-cationic polysaccharide derived from partial N-deacetylation Additionally, the water holding capacity (WHC) of the gels prepared
of chitin and is normally obtained from crustacean shells with the was evaluated. For this, a gel sample (0.9–1.2 g) was placed into a
structure α(1 → 4)-linked 2-amino-2deoxy-β-D-glucan (Shahidi, Vivaspin 20 centrifugal filter unit (GE Healthcare Bio-Sciences AB,
Arachchi, & Jeon, 1999). The impact of adding different polysaccha- Uppsala, Sweden) and centrifuged at 453 × g for 10 min at room
ride over the bulk properties of the gels was evaluated at different temperature. The weight of the gel was recorded before (Wi) and after
pHs. The oat protein gel formation mechanism in the presence centrifugation to the nearest 0.0001 mg, considering ΔWtot as the differ-
of polysaccharide is discussed with special emphasis on phase- ence in water content of the sample, before and after centrifugation, the
separated networks. percentage of water loss was expressed as:

2. Materials and methods Wi  ΔWtot


%WHC ¼  100:
Wi
2.1. Materials

Using naked oat grains (Avena nuda) from Wedge Farms Ltd., 2.5. Rheological measurements
Manitoba, Canada (protein content = 16.6% ± 0.64), oat protein isolate
(OP) was extracted according to our previous work (Nieto-Nieto, Wang, The rheological measurements were done with a TA Discovery HR-3
Ozimek, & Chen, 2014) and the protein content was determined to be rheometer (TA instruments, New Castle, DE, USA). To compare the
89.57% ± 0.73 using a Leco nitrogen analyzer (FP-428, Leco Corporation, effects of different molecular interactions on the formation of OP and
St Joseph, MI) using a protein calculation factor of 6.25. Urea, 2- OP-polysaccharide gels, we used the methodology reported by Yang,
mercaptoethanol, sodium dodecyl sulfate, Rhodamine B, Fluorescein Wang, Vasanthan, & Chen, 2014. Gels were prepared as previously de-
5-isothiocyanate (FITC), carrageenan (molecular weight of 300– scribed in the gel preparation section and cut into 1.0 ± 0.1 cm (height)
400 kDa and composed predominantly by κ-carrageenan as reported sections. The resulting gel disk were submerged for 48 h in solutions of
by supplier) and chitosan from shrimp shells (molecular weight of 2-mercapthoethanol (2-ME) (0.2 M), urea (6 M) and sodium dodecyl
150 kDa and 75–85% deacetylation degree) were obtained from sulfate (SDS) (1% w/v), which could disrupt disulfide bonds, hydrogen
Sigma-Aldrich Canada (Oakville, ON, Canada). Food grade dextrin bonds and hydrophobic interactions respectively. Since the storage
Capsul 2730 (molecular weight N 10 kDa) was obtained from Ingredion modulus is sensitive to changes on the gel network, a frequency
Canada Inc. (Mississagua, ON, Canada). sweep test was done to evaluate the dependence of G' to frequency
(0.1–100 rad/s) on gels compressed to keep 80% of its original height.
2.2. Gel preparation A gap of 8.0 ± 0.8 mm was thus used to make sure the gel structure
was well maintained during the test. To study the changes in viscoelas-
Dextrin, chitosan and carrageenan were selected for this study. Gels tic properties as a function of temperature, OP and OP-polysaccharide
were prepared by heating the protein–polysaccharide suspensions at initial suspensions were subject to a temperature ramp from 25 to
pH 3 and 7 adjusted with 0.1 N NaOH or HCl. Protein–polysaccharide 95 °C, hold at 95 °C for 5 min, and then cooled down to 25 °C at a rate
suspensions were prepared by dispersing protein and polysaccharide of 1.5 °C/min. Sample conditioning took place before and after each
powder in deionized water and the mixture was stirred at room tem- temperature ramp for a period of 3 min. For these measurements, ap-
perature overnight using a stirring hotplate (Fisher Scientific, Markham, proximately 1 mL sample was loaded in the bottom plate of the parallel
ON, Canada). The concentration of OP in the mixtures was kept constant plate geometry; the upper plate was lower to the appropriate geometry
at 15% (w/v) and the concentration of polysaccharide was 0.5% (w/v). gap (1 mm). To avoid evaporation during heating a solvent trap was
These values were determined based on preliminary trials, which used and a thin layer of silicone oil was applied. The temperature of
allowed gel formation. Samples were labeled as OP, OP-DE, OP-CH and the bottom plate was controlled with a Peltier system. All rheological
OP-CA, representing gels containing oat protein, oat protein-dextrin, measurements were done within a predetermined linear viscoelastic
oat protein-chitosan and oat protein-carrageenan respectively. Test region, which was determined in preliminary experiments, setting the
tubes containing the suspension were tightly closed and placed in an strain value at 0.05%.
oil bath at 110 °C for 30 min. Once the heat treatment was completed,
the tubes were cooled in an ice bath and stored in the refrigerator 2.6. Turbidity and particle size measurements
overnight.
The turbidity of the samples was measured before and after heating
2.3. Scanning electron microscopy (SEM) (110 °C, 30 min) in order to follow the protein aggregation process.
Changes in turbidity at 600 nm were recorded at a protein concentra-
Morphological observation of the gels was carried out with a Phillips tion of 0.1% at pH 3 and 7 on a Jasco V-530 UV/VIS spectrophotometer
XL-30 scanning electron microscope (FEI Company, Oregon, USA) at an (Jasco Corporation, CA, USA). The absorbance was measured using
acceleration voltage of 20 kV. The samples were frozen in liquid nitro- quartz cuvettes with a 10 mm path length.
gen and freeze-dried before observations. The cross-section and A Zetasizer Nano ZS ZEN1600 system (Malvern Instruments, U.K.)
surfaces of the dry gels were sputtered with gold and platinum, was used to study the change of particle size distribution caused
observed and photographed. by the heat treatment (110 °C for 30 min). Samples were immediately
cooled in an ice bath to room temperature, after completing the heating
2.4. Gel properties period. Samples were then diluted to a total concentration of 0.1% (w/v)
and passed through a 0.45 μm pore size filter prior to measure-
The compressive stress of the gels was evaluated using an Instron ment. Number-based particle size distribution was measured to identify
5967 universal testing machine (Instron Corp., Norwood, MA, USA). the total number of particles of a given size (Van der Meeren,
Gels were dismounted from test tubes and cut into cylindrical pieces Dewettinck, & Saveyn, 2004).
(∼ 10 mm height and ∼ 14 mm diameter). A two cycle compression
test using a 50 N load cell was performed at room temperature at a 2.7. Surface hydrophobicity (Ho)
rate of 1 mm/min and 50% compression. The compressive stress was
calculated as the peak compression force in the 1st bite cycle, divided Protein surface hydrophobicity was determined using the apolar
by the initial cross-section area of the gel sample. fluorescent dye, ANS (Alizadeh-Pasdar & Li-Chan, 2000). Protein and
T.V. Nieto Nieto et al. / Food Research International 82 (2016) 95–103 97

protein–polysaccharide suspensions (1% w/v) were prepared and 3. Results and discussion
heated for 30 min at 110 °C followed by cooling. Samples were diluted
in phosphate buffer to obtain five final concentrations ranging 3.1. Effect of polysaccharide
from 0.002 to 0.020% (w/v). 20 μL of ANS solution (8.0 mM in 0.1 M
phosphate buffer, pH 7) were added to 4 mL sample. The relative Normally, globular proteins form fine stranded gels at pH values that
fluorescence intensity (RFI) was determined using a Fluorescence spec- diverge from the isoelectric point (pI) and at low ionic strength. Never-
trophotometer Sprectramax M3 (Molecular devices, CA, USA) set at theless, SEM images of oat protein gels showed a percolating network
390 and 470 as excitation and emission wavelengths, respectively, structure similar to that observed in polymer gels. Heating likely causes
with a constant excitation and emission slit of 5 nm. The initial slope the oat globulin hexamers to dissociate first into trimmers and then into
of the RFI versus protein concentration plot was calculated by linear highly reactive monomers (Zhao, Mine, & Ma, 2004). As monomers
regression analysis and used as an index of Ho. were exposed during heating, the creation of a side-by-side array of
polypeptide chains promoted intermolecular interactions and conse-
quently formed strong percolating networks (Lefevre & Subirade,
2.8. Confocal laser scanning microscopy (CSLM) 2000). Fig. 1 shows the microstructure of oat protein gels in the
presence of different polysaccharides at pH 3 and 7. At pH 3, those
CSLM was used to observe the distribution of the polysaccharide gels prepared with dextrin and chitosan (OP-DE and OP-CH) (Fig. 1 A
within the protein network. A laser scanning confocal microscope and E) show networks consisting of hollow cells formed between verti-
Zeiss LSM710 (Carl Zeiss Microscopy, Jena, Germany) was used with cal walls. The cells appeared to be interconnected in OP-DE gels, where-
a 10 × objective. Dextrin and carrageenan were labeled covalently as long walls forming large and empty compartments were observed in
with FITC. Specifically, 0.5 g of the polysaccharide and 10 mL of DMSO OP-CH gels. In contrast, gels prepared with carrageenan (OP-CA)
were stirred overnight, and then 7 mg of FITC were added to the (Fig. 1 C) show a condensed and compact structure. At pH 7, gels with
polysaccharide-DMSO mixture. The reaction mixture was protected thick walls and highly interconnected percolating networks were
from the light, heated at 90 °C for 2 h and dialyzed extensively against formed by OP-DE and OP-CA (Fig. 1 B and D). Whereas, a particulate
distilled water in the dark and freeze-dried. Rhodamine B was used structure was observed for OP-CH (Fig. 1 F) with spherical aggregates
for non-covalent labeling of oat protein. A 15% protein suspension was distributed among the gel network. It was revealed in our previous
prepared, and 40 μL of Rhodamine B (5 mg/mL) were added to 1 mL work that as the heating process is taking place, oat protein hexamers
of protein suspension. The mixture was stirred for 2 h at room temper- dissociate into monomers, which then become reactive units that par-
ature, dialyzed against distilled water in the dark and freeze-dried. Once ticipate in the assembly of this unique configuration at a near molecular
protein and polysaccharides were labeled, OP-polysaccharide suspen- level (Nieto-Nieto et al., 2015).
sions were prepared as described previously in the gel preparation To explain the effect of the type of polysaccharide on the gelling
section. A thin slice of gel was cut with scalpel and place in a slide. The properties of oat protein, it is necessary to consider the nature of inter-
fluorescent images were analyzed at wavelength of 488 nm and actions involved. Using protein and anionic polysaccharide as an exam-
516 nm. Images were processed with ZEN 2009 LE software (Carl ple, at pH N pI, repulsion exists between protein and polysaccharide,
Zeiss AG, Oberkochen, Germany). leading to the formation of phase-separated networks; whereas at
pH b pI, electrostatic attraction exists between protein and polysaccha-
ride, resulting in the formation of coacervate networks. Oat protein has
2.9. Statistical analysis a pI of ~5 (Guan, Yao, Chen, Shan, & Zhang, 2007; Wu, Sexson, Cluskey, &
Inglett, 1977), thus electrostatic attractive forces are expected in the OP-
The results were evaluated by analysis of variance (ANOVA) and the CA system at pH 3 and OP-CH system at pH 7. These attractive forces
minimum significance test (Tukey's test) was set at the 5% level could restrict protein unfolding and neutralize protein surface to
(p b 0.05) to compare all means (GraphPad Software, La Jolla, CA, promote random aggregation, leading to the formation of particulate
USA). All experiments were performed at least in three independent tri- networks. OP-CH at pH 3 and OP-CA at pH 7 were likely to form
als and the results were reported as mean ± standard deviation. phase-separated networks due to repulsive interactions between

Fig. 1. SEM images of the cross section of oat protein gels prepared at 110 °C. A scale bar represents 5 μm. A and B correspond to OP-DE gels prepared at pH 3 and 7, respectively. C and D
correspond to OP-CA gels prepared at pH 3 and 7, respectively. E and F correspond to OP-CH gels prepared at pH 3 and 7, respectively.
98 T.V. Nieto Nieto et al. / Food Research International 82 (2016) 95–103

protein and polysaccharide. Thus, the observed changes in microstruc- reactivity of cysteine (Ikeda & Morris, 2002), which may partially
ture could be related to the effect that the polysaccharide has on the explain the formation of weaker and less interconnected networks. In
formation of the protein gel network. The balance between attractive the case of carrageenan and dextrin, the polysaccharide functionality
and repulsive forces is useful to explain why gel networks prepared could be lost after heating under acidic conditions, since these polysac-
at pH 3 seemed to be less interconnected than those prepared at pH 7. charides are susceptible to depolymerisation by acid hydrolysis (Necas
In the case of oat protein, previous research suggests that, the optimum & Bartosikova, 2013; Silva et al., 2014). Thus, a strong synergistic effect
balance is reached at pH 5–7, whereas at a higher or lower pH, dispro- was not observed at an acidic pH. The improvement of the gel strength
portionate electrostatic repulsive forces may prevent the formation of is associated with the expected phase-separation process produced by
highly interconnected networks (Nieto-Nieto et al., 2014). The gels the thermodynamic incompatibility between protein and polysaccha-
prepared with dextrin (OP-DE) showed similar microstructure to that rides. In this process, polysaccharide molecules are preferentially
of oat protein alone at both pH 3 and 7. Since dextrin is a neutral poly- excluded from the protein surface. As the phase-separation process
saccharide, it is not expected to produce any significant change on the advances, protein rich regions as well as polysaccharide rich regions
surface charge of the protein molecules, thus could be present as a co- are formed. The water content is higher in the phase of the more hydro-
solute. Nevertheless, a phase-separated network could still be formed philic component with a greater excluded volume effect. This causes the
due to the thermodynamic incompatibility between dextrin and oat formation of a less concentrated phase, rich in polysaccharide and a
protein during heating. In our previous work, oat protein-inulin gels more concentrated phase, rich in protein (Tolstoguzov, 2007). An
with micro phase-separated network were formed (Nieto-Nieto et al., increase in the apparent protein concentration could explain the
2015). During heating the protein may have started to unfold, increas- improved gel strength in our study, since the rigidity of a protein gel is
ing the thermodynamic incompatibility between oat protein and inulin. determined by the concentration of protein incorporated into the gel
This excluded volume effect caused the system to separate into a network (Tseng, Xiong, & Boatright, 2008). The strength improvement
protein-rich phase and an inulin-rich phase (Turgeon, Beaulieu, was not observed for samples with chitosan under segregative con-
Schmitt, & Sanchez, 2003). ditions due to the brittle character of the samples, nonetheless
Fig. 2 shows the compressive stress and the WHC of oat protein gels fragmented gels did not show water release due to compression indicat-
in the presence of different polysaccharides. Compressive stress is ing that water was held within the network. In contrast, gels prepared
defined as the capacity of the material to withstand a given deforma- under conditions favoring associative interactions showed reduced
tion. At pH 3, oat protein had a compressive stress value of 2.99 kPa. gel strength values. Here the charge neutralization produced by
This value increased to 4.07 and 3.61 kPa in the presence of dextrin the polysaccharide, restricted the protein unfolding and increased
and carrageenan, respectively. The compressive stress decreased to the protein aggregation, leading to particulate networks. The restricted
2.66 kPa with addition of chitosan. At pH 7, the addition of carrageenan protein unfolding associated with weak electrostatic repulsion,
significantly increased (p b 0.05) the compressive stress value of oat prevents the exposure of reactive amino acid residues, which are nor-
protein gels from 13.21 to 25.95 kPa. The gel strength was also im- mally buried at the interior of the protein molecules. Thus the number
proved (16.87 kPa) with addition of dextrin, but decreased in the pres- of possible interactions is reduced, which affects the mechanical
ence of chitosan (11.61 kPa). Stronger gels were formed at neutral strength of the gel.
conditions since at this pH the repulsive and attractive interactions The WHC is a key property of gels, and low values often result in dry
were equilibrated. On the other hand, at pH 3 the attractive interactions products and low texture stability. The oat protein based gels showed
were not able to overcome the strong electrostatic repulsion among good WHC values at pH 3 (86.24–87.84%) and pH 7 (71.28–79.17%).
positively charged amino acid residues, leading to weaker gels. Disulfide WHC values were maintained or improved with carrageenan at pH 7,
bonds are unlikely to be formed under acidic conditions due to low chitosan at pH 3 and dextrin at both pH 3 and 7. In phase-separated
gels, polysaccharides tended to separate into the serum phase and
increase the viscosity of the serum phase rendering it less prone to
syneresis (Laneuville & Turgeon, 2014). Conversely, a dramatic decrease
of WHC was observed with addition of carrageenan at pH 3 and chitosan
at pH 7. Specifically, these samples showed particulate microstructures
in previous SEM observations. In particulate gels, protein aggregation
was promoted and protein-water interactions were restricted, leading
to reduced WHC (Foegeding, 2007).
The improvement of the gelling properties by mixing with polysac-
charides is highly related to the type of interactions that are promoted
in the system. When gels were prepared under conditions that favored
segregative interactions, gel properties were strengthened without
compromising the WHC. In contrast, gels formed under associative
interactions showed reduced WHC values and reduced gel strength. A
similar observation was made in a whey protein-xanthan gum system
(Bertrand & Turgeon, 2007). In this case, xanthan gum had a synergistic
effect when gels were prepared under conditions that favored phase
separation by segregative interactions, and an antagonistic effect
when gels were prepare under conditions that promoted attractive
interactions (Bertrand & Turgeon, 2007). Carrageenan had a strong syn-
ergistic effect on oat protein gelling properties at pH 7, as the compres-
sive stress almost doubled. The addition of dextrin slightly improved the
gel strength at both acidic and neutral conditions. Dextrin was not able
to form a gel on its own, whereas carrageenan formed a very weak, al-
most liquid gel. This indicates that the improvement of the compressive
Fig. 2. Gel properties of OP and OP-polysaccharide gels prepared at pH 3 and 7. Statistical
analysis was performed separately for each pH, which is indicated by the use of primed
stress arises from synergistic protein–polysaccharide interactions. Thus,
letters (a and a' for compressive stress) (A and A' for water holding capacity). Values OP-DE and OP-CA were selected to further study the influence of the
within the same pH not sharing a common superscript differ (p b 0.05). type of polysaccharide on the gel formation mechanism.
T.V. Nieto Nieto et al. / Food Research International 82 (2016) 95–103 99

3.2. Protein interactions reduced at pH values lower than 7 due to the low reactivity of cysteine
(Ikeda & Morris, 2002). Therefore, at pH 3, only the involvement of
While electrostatic attractions are the main driving force in protein– hydrogen bonding was evaluated. All samples treated with urea showed
polysaccharide complex formation, other interactions, such as hydrogen greatly reduced G' values and the reduction levels in OP-DE and OP-CA
bonds and hydrophobic interactions may, contribute to the stability of a were higher than gels with OP alone (Fig. 3). This indicates that the ad-
protein–polysaccharide system (McClements, 2006). To elucidate the dition of the polysaccharide may increase the number of hydrogen
type of interactions involved in the development of the oat protein– bonds in the gels. Hydrogen bonds may take place between the C = O
polysaccharide gels, rheological properties were evaluated by a and N–H groups of the protein and hydroxyl groups of polysaccharides
frequency test in the presence of dissociating reagents including urea, (Secundo & Guerrieri, 2005). Hydrogen bonds are also believed to
2-ME and SDS. These are known to disrupt hydrogen bonds, disulfide participate in the secondary stabilization of complexes formed by
bonds and hydrophobic interactions, respectively. The gels prepared at electrostatic attractive forces (Stone & Nickerson, 2012) such as the
pH 3 and treated with SDS shrank and hardened. This was due to strong one formed by OP-CA at this pH. Hydrophobic bonds may also contrib-
interactions that develop between positively charged protein and nega- ute to gel formation, since urea is able to disrupt hydrophobic forces
tively charged SDS to form highly condensed networks with reduced (Sun & Arntfield, 2012).
protein-water interactions. Thus, readings obtained for these samples Fig. 3 also shows the response of G' to the variation on frequency of
were considered not reliable and were not included. As mentioned oat protein derived gels prepared at pH 7. All OP gels treated with SDS,
earlier, the participation of disulfide bonds in gels prepared at pH 3 urea and 2-ME showed greatly reduced G' values of OP gels. This
was not expected, since the ability of disulfide bond formation is indicates that hydrogen bonds, disulfide bonds and hydrophobic forces,

Fig. 3. Storage modulus (G') of oat protein derived gels prepared at pH 3 and 7 as a function of angular frequency.
100 T.V. Nieto Nieto et al. / Food Research International 82 (2016) 95–103

all play essential roles in the stabilization of the gel network. In the case hydrophobic interactions are expected to participate in the secondary
of OP-DE, a similar effect was observed, however the G' was slightly less support of OP-CA complexes. After heating, the sample turbidity in-
affected in the presence of 2-ME. In contrast, gels prepared with OP-CA creased due to formation of more and larger aggregates, which impaired
were not affected by 2-ME, but were disturbed by SDS and urea. Thus the light transmission due to the light scattering of particles. At pH 7,
hydrogen bonds and hydrophobic forces constituted the main support OP-DE also showed higher turbidity than OP. In contrast, the turbidity
of the OP-DE and OP-CA gel networks at neutral pH. Heat is expected for OP-CA was lower than both OP and OP-DE. Strong repulsive forces
to cause the dissociation of oat globulin hexamers by breaking disulfide between the negatively charged carrageenan and the negatively
bonds linking the acidic and basic subunits. This allows the re-formation charged protein prevented aggregation. Even after heating at pH 7,
of disulfide bonds during the heating process and explains the involve- OP-CA exhibited low turbidity. At this pH, both carrageenan and protein
ment of disulfide bonds in OP gel formation. In the case of gels contain- carry a relatively higher negative charge, which produce a strong
ing carrageenan, the development of new disulfide bonds may be enough electrostatic repulsion to prevent the formation of large aggre-
hindered, probably due to the fact that strong repulsion hampered the gates (Nicolai & Durand, 2013). The particle size distribution of the pro-
required proximity for the SH/S-S- interchange. Previous studies dem- tein aggregates formed in the presence of carrageenan during heating
onstrated that the presence of anionic polysaccharides could protect was also monitored at pH 7 and was compared to oat protein alone.
BSA against extensive aggregation because the formation of intermolec- These results are shown in Fig. 4. OP and OP-CA systems had a peak
ular disulfide bridges was prevented, during or after high-pressure value of 33 nm and 28 nm in diameter, respectively. After heating, the
treatment (Galazka, Smith, Ledward, & Dickinson, 1999). main peak observed for OP was 68 nm. However, the peak value of
OP-CA was 16 nm, suggesting that heating did not form larger aggre-
3.3. Turbidity and particle size gates. OP-DE showed a similar behavior to oat protein alone (results
not shown). In the presence of carrageenan, strong repulsive forces
The changes in turbidity were recorded before and after heating OP limit the size of the aggregates formed, so that rapid random aggrega-
and OP-polysaccharide suspensions to study how the progression of the tion is impaired. The overall reaction rate of heat-induced gelation can
heat-induced aggregation was affected by the presence of the polysac- be determined by both the unfolding and the aggregation reaction
charides. Before heating the turbidity of OP at pH 3 showed the lowest (Banerjee & Bhattacharya, 2012). Under high repulsion conditions,
value among the tested samples (Fig. 4). protein molecules rapidly unfold, and slowly aggregate, since the partic-
OP-DE and OP-CA showed higher turbidity. This may be associated ipation of attractive forces is limited (Totosaus, Montejano, Salazar, &
to hydrogen bonds taking place between the protein and hydroxyl Guerrero, 2002). The combination of a high unfolding rate with a low
groups of dextrin (Secundo & Guerrieri, 2005). For OP-CA, the increased aggregation rate allows the establishment of structures with greater
turbidity is primarily due to electrostatic attraction between the nega- degree of order (Totosaus et al., 2002). Therefore, the addition of carra-
tively charged carrageenan and the positively charged protein. This geenan facilitates the formation of a highly organized oat protein net-
caused charge neutralization and lead to an increase in particle size work. A highly ordered structure implies that the ensemble of reactive
and light scattering efficiency of the protein–polysaccharide complexes units take place in an intimate lateral sequence, which later on, allows
(Jones & McClements, 2011). Moreover, hydrogen bonding and the contribution of other intermolecular interactions (hydrogen bonds

Fig. 4. Changes in turbidity and in the particle size distribution (number based) of unheated and heated OP and OP-polysaccharide suspensions. Values not sharing a common superscript
differ (p b 0.05).
T.V. Nieto Nieto et al. / Food Research International 82 (2016) 95–103 101

and hydrophobic forces) as supporting secondary structure, culminat- phase separation refers to a phase-separated microstructure in which
ing in stronger gels (Lefevre & Subirade, 2000). This may also explain the protein phase contains a small number of large polysaccharide
the two-fold increase in compressive stress with the incorporation of insertions, whereas a lower degree of phase separation is related to a
a small amount of carrageenan at pH 7. Thus is possible to consider large number of small polysaccharide insertions. Macroscopic phase
that the improvement of the mechanical properties is mostly controlled separation was observed in gels prepared with added carrageenan,
by the protein component (Kontogiorgos, Ritzoulis, Biliaderis, & since large polysaccharide insertions in the micrometer-range were
Kasapis, 2006), which is mainly related to an apparent increase in the observed.
protein concentration due to phase separation (Tseng et al., 2008).
3.6. Rheological measurements
3.4. Surface hydrophobicity
Phase separation is promoted by heating (Turgeon et al., 2003) since
Hydrophobic forces play an important role to stabilize oat protein protein denaturation and aggregation intensifies the system incompat-
gels. Changes in surface hydrophobicity after heating produced by addi- ibility by increasing molecular volume or size (Tolstoguzov, 2007). As
tion of different types of polysaccharide were investigated. Normally, the protein–polysaccharide system is heated, protein aggregation and
globular proteins are composed of charged side groups at their surface phase separation are taking place at the same time with different kinet-
with the hydrophobic groups buried at the interior of the structure. ics (Croguennoc et al., 2001). For this reason, the rheological changes
After heating, proteins can unfold exposing buried hydrophobic groups during the whole gel formation process were studied. As shown in
that provide additional sites for crosslinking and thus altering the Fig. 7, the OP-DE system has similar behavior to that of oat protein
surface characteristics. Heating in the presence of dextrin had no alone. During the heating stage, both G' values initially decreased with
significant impact on the surface hydrophobicity, as no strong electro- increasing temperature, then increased sharply at 70 °C. During cooling,
static interactions could be developed between oat protein and dextrin G' still increased until 88 °C. From this temperature and until the end of
(Fig. 5). In contrast, heating in the presence of carrageenan slightly the cooling stage (25 °C), G' slightly decreased. The final G' values were
increased the surface hydrophobicity, suggesting the exposure of around 16 kPa for OP and 18 kPa for OP-DE. The development of G' with
hydrophobic residues due to strong repulsive forces in the OP-CA sus- increasing temperature was different in the presence of carrageenan. A
pension (Mleko, Pikus, & Li-Chan, 1997; Uruakpa & Arntfield, 2006). higher initial G' value was observed for OP-CA at 25 °C. A sharp increase
These exposed hydrophobic groups can function as prospective occurred at around 80 °C and until the end of the heating cycle, then G'
crosslinking points and promote protein intermolecular interactions as continued to increase from 95 to 80 °C and during the whole cooling
the aggregation process advances. Thus, the resulting hydrophobic in- stage, but with a significantly reduced incremental rate. The final G'
teractions are part of the secondary support of gel structure. The larger value reached 22 kPa. This indicates that aggregation started at an
repulsive forces in the OP-CA system could control the rate of protein earlier stage for OP and OP-DE as the sharp increase of G' took place at
aggregation and expose hydrophobic groups during heating, which in a lower temperature (~63 °C) than that of OP-CA (~93 °C). The gel for-
turn produced a highly ordered structure (Bryant & McClements, mation process of carrageenan alone (0.05%w/v) was also monitored
1998) with a higher level of crosslinking, leading to a greater mechani- (data not shown) and had an initial G' value of 4.3 Pa, which is signifi-
cal strength. cantly lower than OP-CA. During the cooling from 30 °C to 25 °C, the
G' value sharply increased from ~4 to 339 Pa, which is also significantly
3.5. Confocal laser scanning microscopy (CSLM) inferior to OP-CA. This further supports previous observations that car-
rageenan produced a synergistic effect during oat protein gel formation.
Confocal images (Fig. 6) confirm the formation of phase-separated Carrageenan alone cannot form gels of such strength at low concentra-
networks in the OP-DE and OP-CA gels at pH 7 with dextrin and carra- tion. Strong electrostatic repulsive forces between negatively charged
geenan. Samples showed polysaccharide insertions occupying the void protein molecules and negatively charged carrageenan lead to thermo-
spaces of the protein network. More dextrin insertions were observed dynamic incompatibility. This means that protein–protein interactions
with smaller sizes (~ 20 μm), whereas carrageenan insertions were and carrageenan-carrageenan interactions were more thermodynami-
larger in size but fewer in number. This indicates that a different degree cally favorable than protein-carrageenan interactions. Thus, the system
of phase separation occurred in these two systems. A higher degree of separates into two phases, each with higher concentration of one of the
components and depleted of the other (Tolstoguzov, 2007).
The nonionic polysaccharide dextrin could also show thermody-
namic incompatibility with the negatively charged oat protein at
neutral pH during heating. Phase separation between oat protein and
inulin, another neutral polysaccharide, resulted in formation of inulin
nanoparticles homogeneously distributed inside the protein network
(Nieto-Nieto et al., 2015). Thermodynamic incompatibility not only
arises from electrostatic repulsion, but it can also be related to molecular
differences of size and shape. Before heating, the entropy of mixing al-
lows a homogenous system. Nevertheless with increasing temperature,
the size of the aggregates increases, and the entropy of mixing is re-
duced as well as the stability of the system leading to phase separation
(Schmitt, Sanchez, Desobry-Banon, & Hardy, 1998). For a better under-
standing of the formation of phase-separated networks, it is necessary
to consider that two events are taking place simultaneously. One is
the phase separation process and the second event is the gelation pro-
cess. An increase in temperature stimulates both processes. The extent
of the phase separation and thus the heterogeneity of the gels depend
on the rate of phase separation compared to the rate of aggregation
(Verbeken, Neirinck, Van Der Meeren, & Dewettinck, 2005). In the
Fig. 5. Surface hydrophobicity of OP and OP-polysaccharide suspensions. Values not case of OP-DE, the rate of phase separation was slower than the rate at
sharing a common superscript differ (p b 0.05). which OP forms gel, because no obvious change in the development
102 T.V. Nieto Nieto et al. / Food Research International 82 (2016) 95–103

Fig. 6. Confocal microscopic photographs of OP and OP-polysaccharide gels formed at pH 7. Scale bar represents 200 μm.

profile of G' was observed during heating compared to OP gel alone. 4. Conclusion
Thus, dextrin was more uniformly distributed within the OP network.
Whereas OP-CA showed a higher rate of phase separation and slower Oat protein gelling properties could be significantly improved by
rate of aggregation, since the G' value of OP-CA increased at a later forming phase-separated networks in the presence of carrageenan and
time and a higher temperature, allowing the formation of carrageenan dextrin. Different levels of phase separation occurred in the OP-DE
rich domains embedded in the protein network. As discussed previous- and OP-CA gels and the heterogeneity of these gels depended on the
ly, such a slow aggregation also permitted the formation of a highly rate of phase separation compared to the rate of protein aggregation.
ordered structure, stabilized by hydrogen and hydrophobic interactions, In the case of OP-DE, the rate of OP gel formation was faster than the
and resulting in the greater gel strength. The charge density played an rate of phase separation, thus the dextrin particles were uniformly
important role to determine the extent of phase separation, since it di- distributed within the protein network. Conversely, OP-CA gels showed
rectly impacts the strength of the electrostatic repulsive forces between a greater level of phase separation due to a slower aggregation process.
protein and polysaccharide, as well as the degree of protein unfolding. This created the formation of larger carrageenan-rich domains embed-
The difference in molecular weight between dextrin and carrageenan ded in the protein network. The degree of phase separation was the
may also contribute to such a different degree of phase separation by af- key to producing gels with different strengths, and may ultimately be
fecting the volume occupied by the polysaccharide. This consequently related to distinctive textures. Strong repulsive forces caused by the
impacts the apparent protein concentration in the protein rich phase addition of carrageenan may promote the creation of a highly ordered
for gel formation. The excluded volume effect is significantly greater network structure. Then this ensemble of reactive protein molecules
for both molecules with a larger size and for non-spherical macromole- in this network may take place in an intimate lateral sequence, permit-
cules such as linear polysaccharide (Tolstoguzov, 2003). Since dextrin ting the development of hydrogen and hydrophobic interactions to
had a much lower molecular weight than carrageenan, the space occu- further strengthen the protein network. Moreover, the improvement
pied by carrageenan is much greater. The thermodynamic incompatibil- of the gel's mechanical properties is mainly controlled by the protein
ity of protein and polysaccharides arises from differences of their component, associated to an apparent increase of the protein concen-
molecular properties such as shape, size or charge (De Jong et al., tration in the gel network. Addition of chitosan did not have a favorable
2009). Hence, the different characteristics possessed by OP-DE and effect on the gel properties.
OP-CA gels may be related to the difference in the extent of phase
separation.
Acknowledgments

The authors are grateful to the Nature Science and Engineering


Research Council of Canada (NSERC), the Alberta Crop Industry Devel-
opment Fund Ltd. (ACIDF), the Alberta innovates Bio Solution (Al Bio)
and the Alberta Barley commission for financial support as well as the
Canada Foundation for Innovation (CFI) for equipment support. Lingyun
Chen would like to thank the Natural Science and Engineering Research
Council of Canada (NSERC)-Canada Research Chairs Program for its
financial support. Talina Vanessa Nieto-Nieto thanks the National
Council of Science and Technology (CONACyT) for the financial support
provided during her PhD program.

References
Alizadeh-Pasdar, N., & Li-Chan, E.C. (2000). Comparison of protein surface hydrophobicity
measured at various pH values using three different fluorescent probes. Journal of
Agricultural and Food Chemistry, 48, 328–334.
Baeza, R.I., Carp, D.J., Pérez, O.E., & Pilosof, A.M.R. (2002). Regular article: κ –carrageenan–pro-
tein interactions: Effect of proteins on polysaccharide gelling and textural properties.
LWT - Food Science and Technology, 35, 741–747.
Banerjee, S., & Bhattacharya, S. (2012). Food gels: Gelling process and new applications.
Fig. 7. Storage modulus (G') of OP, OP-DE and OP-CA during heat-induced gelation at pH 7. Critical Reviews in Food Science and Nutrition, 52, 334–346.
Doted line represents the heating temperature profile. Samples were heated from 25 to Bertrand, M., & Turgeon, S.L. (2007). Improved gelling properties of whey protein isolate
95 °C, hold at 95 °C for 5 min, and then cooled down to 25 °C at a rate of 1.5 °C/min. by addition of xanthan gum. Food Hydrocolloids, 21, 159–166.
T.V. Nieto Nieto et al. / Food Research International 82 (2016) 95–103 103

Bryant, C.M., & McClements, D.J. (1998). Molecular basis of protein functionality with spe- Pires Vilela, J.A., Cavallieri, Â. L.F., & Lopes da Cunha, R. (2011). The influence of gelation
cial consideration of cold-set gels derived from heat-denatured whey. Trends in Food rate on the physical properties/structure of salt-induced gels of soy protein isolate–
Science & Technology, 9, 143–151. gellan gum. Food Hydrocolloids, 25, 1710–1718.
Çakır, E., Daubert, C.R., Drake, M.A., Vinyard, C.J., Essick, G., & Foegeding, E.A. (2012). The Schmitt, C., Sanchez, C., Desobry-Banon, S., & Hardy, J. (1998). Structure and
effect of microstructure on the sensory perception and textural characteristics of technofunctional properties of protein–polysaccharide complexes: A review. Critical
whey protein/κ-carrageenan mixed gels. Food Hydrocolloids, 26, 33–43. Reviews in Food Science and Nutrition, 38, 689–753.
Croguennoc, P., Nicolai, T., Durand, D., & Clark, A. (2001). Phase separation and association Secundo, F., & Guerrieri, N. (2005). ATR-FT/IR study on the interactions between gliadins
of globular protein aggregates in the presence of polysaccharides: 2. Heated mixtures and dextrin and their effects on protein secondary structure. Journal of Agricultural
of native β-lactoglobulin and κ-carrageenan. Langmuir, 17, 4380–4385. and Food Chemistry, 53, 1757–1764.
De Jong, S., Klok, H.J., & Van de Velde, F. (2009). The mechanism behind microstructure Shahidi, F., Arachchi, J.K.V., & Jeon, Y. (1999). Food applications of chitin and chitosans.
formation in mixed whey protein–polysaccharide cold-set gels. Food Hydrocolloids, Trends in Food Science & Technology, 10, 37–51.
23, 755–764. Silva, D.M., Nunes, C., Pereira, I., Moreira, A.S.P., Domingues, M.R.M., Coimbra, M.A., &
Foegeding, E.A. (2007). Rheology and sensory texture of biopolymer gels. Current Opinion Gama, F.M. (2014). Structural analysis of dextrins and characterization of dextrin-
in Colloid & Interface Science, 12, 242–250. based biomedical hydrogels. Carbohydrate Polymers, 114, 458–466.
Galazka, V., Smith, D., Ledward, D., & Dickinson, E. (1999). Interactions of ovalbumin with Stone, A.K., & Nickerson, M.T. (2012). Formation and functionality of whey protein
sulphated polysaccharides: effects of pH, ionic strength, heat and high pressure treat- isolate–(kappa-, iota-, and lambda-type) carrageenan electrostatic complexes. Food
ment. Food Hydrocolloids, 13, 81–88. Hydrocolloids, 27, 271–277.
Guan, X., Yao, H., Chen, Z., Shan, L., & Zhang, M. (2007). Some functional properties of oat Sun, X.D., & Arntfield, S.D. (2012). Molecular forces involved in heat-induced pea protein
bran protein concentrate modified by trypsin. Food Chemistry, 101, 163–170. gelation: Effects of various reagents on the rheological properties of salt-extracted
Ikeda, S., & Morris, V.J. (2002). Fine-stranded and particulate aggregates of heat-denatured pea protein gels. Food Hydrocolloids, 28, 325–332.
whey proteins visualized by atomic force microscopy. Biomacromolecules, 3, 382–389. Tolstoguzov, V. (2003). Some thermodynamic considerations in food formulation. Food
Jones, O.G., & McClements, D.J. (2011). Recent progress in biopolymer nanoparticle and Hydrocolloids, 17, 1–23.
microparticle formation by heat-treating electrostatic protein–polysaccharide com- Tolstoguzov, V. (2007). Ingredient interactions in complex foods: aggregation and phase
plexes. Advances in Colloid and Interface Science, 167, 49–62. separation. In D. McClements (Ed.), Understanding and Controlling the Microstructure
Kontogiorgos, V., Ritzoulis, C., Biliaderis, C.G., & Kasapis, S. (2006). Effect of barley of Complex Foods (pp. 185–206). Woodhead Publishing in Food Science, Technology
bbeta N − glucan concentration on the microstructural and mechanical behav- and Nutrition.
iour of acid-set sodium caseinate gels. Food Hydrocolloids, 20, 749–756. Totosaus, A., Montejano, J.G., Salazar, J.A., & Guerrero, I. (2002). A review of physical and
Laneuville, S.I., & Turgeon, S.L. (2014). Microstructure and stability of skim milk acid gels chemical protein-gel induction. International Journal of Food Science & Technology, 37,
containing an anionic bacterial exopolysaccharide and commercial polysaccharides. 589–601.
International Dairy Journal, 37, 5–15. Tseng, Y.-., Xiong, Y.L., & Boatright, W.L. (2008). Effects of inulin/oligofructose on the ther-
Lefevre, T., & Subirade, M. (2000). Molecular differences in the formation and structure of mal stability and acid-induced gelation of soy proteins. Journal of Food Science, 73,
fine-stranded and particulate β-lactoglobulin gels. Biopolymers, 54, 578–586. E44–E50.
Liu, S., Elmer, C., Low, N.H., & Nickerson, M.T. (2010). Effect of pH on the functional behav- Turgeon, S., Beaulieu, M., Schmitt, C., & Sanchez, C. (2003). Protein–polysaccharide inter-
iour of pea protein isolate–gum Arabic complexes. Food Research International, 43, actions: phase-ordering kinetics, thermodynamic and structural aspects. Current
489–495. Opinion in Colloid & Interface Science, 8, 401–414.
McClements, D.J. (2006). Non-covalent interactions between proteins and polysaccha- Uruakpa, F., & Arntfield, S. (2006). Surface hydrophobicity of commercial canola proteins
rides. Biotechnology Advances, 24, 621–625. mixed with κ-carrageenan or guar gum. Food Chemistry, 95, 255–263.
Mleko, S., Pikus, S., & Li-Chan, E. (1997). Interactions of kappa-carrageenan with whey Van der Meeren, P., Dewettinck, K., & Saveyn, H. (2004). Particle size analysis. In L. Nollet
proteins in gels formed at different pH. Food Research International (Ottawa, Ont.) (Ed.), Handbook of food analysis: Methods and instruments in applied food analysis, Vol.
Food Research International, 30, 427–433. 138. (pp. 1805–1807). Marcel Dekker AG.
Morris, V. (1986). Multicomponent gels. In G. Phillips, D. Wedloc, & P. Williams (Eds.), Verbeken, D., Neirinck, N., Van Der Meeren, P., & Dewettinck, K. (2005). Influence of
Gums and Stabilisers for the Food Industry, Vol. 3. (pp. 87–99). London: Elsevier κ-carrageenan on the thermal gelation of salt-soluble meat proteins. Meat
Applied Science. Science, 70, 161–166.
Necas, J., & Bartosikova, L. (2013). Carrageenan: a review. Veterinarni Medicina, 58, Weinbreck, F., Nieuwenhuijse, H., Robijn, G.W., & de Kruif, C.G. (2003). Complex forma-
187–205. tion of whey proteins: exocellular polysaccharide EPS B40. Langmuir, 19, 9404–9410.
Nicolai, T., & Durand, D. (2013). Controlled food protein aggregation for new functionality. Wu, Y.V., Sexson, K.R., Cluskey, J.E., & Inglett, G.E. (1977). Protein isolate from high protein
Current Opinion in Colloid & Interface Science, 18, 249–256. oats: preparation, composition and properties. Journal of Food Science, 42, 1383–1386.
Nieto-Nieto, T.V., Wang, Y.X., Ozimek, L., & Chen, L. (2014). Effects of partial hydrolysis on Yang, C., Wang, Y., Vasanthan, T., & Chen, L. (2014). Impacts of pH and heating tempera-
structure and gelling properties of oat globular proteins. Food Research International, ture on formation mechanisms and properties of thermally induced canola protein
55, 418–425. gels. Food Hydrocolloids, 40, 225–236.
Nieto-Nieto, T.V., Wang, Y.X., Ozimek, L., & Chen, L. (2015). Inulin at low concentration Zhao, Y., Mine, Y., & Ma, C. (2004). Study of thermal aggregation of oat globulin by laser
significantly improves the gelling properties of oat protein — A molecular mechanism light scattering. Journal of Agricultural and Food Chemistry, 52, 3089–3096.
study. Food Hydrocolloids, 50, 116–127.

You might also like