You are on page 1of 8

Food Hydrocolloids 28 (2012) 325e332

Contents lists available at SciVerse ScienceDirect

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

Molecular forces involved in heat-induced pea protein gelation: Effects


of various reagents on the rheological properties of salt-extracted pea protein gels
Xiang Dong Sun, Susan D. Arntfield*
Department of Food Science, University of Manitoba, Winnipeg, MB, Canada R3T 2N2

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

Article history: The molecular forces involved in the gelation of heat-induced pea protein gel were studied by moni-
Received 1 October 2010 toring changes in gelation properties in the presence of different chemicals. At 0.3 M concentration,
Accepted 21 December 2011 sodium thiocyanate (NaSCN) and sodium chloride (NaCl) showed more chaotropic characteristic and
enhanced the gel stiffness, whereas sodium sulfate (Na2SO4) and sodium acetate (CH3COONa) stabilized
Keywords: protein structure as noted by increasing denaturation temperatures (Td) resulting in reduced storage
Pea protein
moduli (G0 ). To determine the involvement of non-covalent bonds in pea protein gelation, guanidine
Gelation
hydrochloride (GuHCl), propylene glycol (PG), and urea were employed. The significant decrease in G0 of
Molecular forces
Reagents
pea protein gels with the addition of 3 M GuHCl and 5 M urea indicated that hydrophobic interactions
Storage modulus and hydrogen bonds are probably involved in pea protein gel formation. The increase in G0 with
increasing PG concentration (5e20%), demonstrated hydrogen bonds and electrostatic interaction
involvement. No significant influence was observed on G0 with addition of different concentrations of b-
mercaptoethanol (2-ME), low levels of dithiothreitol (DTT), and up to 25 mM N-ethylmaleimide (NEM),
which indicated that disulfide bonds are not required for gel formation, but data at higher DTT and NEM
concentrations and slow cooling rates showed a minor contribution by disulfide bonds. Reheating and
recooling demonstrated that gel strengthening during the cooling phase was thermally reversible but not
all the hydrogen bonds disrupted in the reheating stage were recovered when recooled.
Ó 2012 Elsevier Ltd. All rights reserved.

1. Introduction protein isolates were extracted by pH adjustment and were


recovered by isoelectric precipitation. The contributions of
The gel forming ability of pea proteins upon heating is an constituent proteins to gel formation and the molecular forces
important functional property, which affects their utilization in involved in gelation have been studied for protein isolates (Bora,
foods. Although a number of studies have focused on the effects of Brekke, & Powders, 1994), purified vicilin (O’Kane, Happe,
plant proteins especially soy proteins on gelation properties of Vereijken, Gruppen, & van Boekel, 2004a) and purified legumin
comminuted meat products, limited studies (Sanjeewa, (O’Kane, Happe, Vereijken, Gruppen, & van Boekel, 2004b; O’Kane,
Wanasundara, Pietrasik, & Shand, 2010; Serdaroglu, Yildiz-Turp, & Vereijken, Gruppen, & van Boekel, 2005) again using materials
Abrodimov, 2005; Su, Bowers, & Zayas, 2000; Pietrasik & Janz, 2010; recovered by isoelectric precipitation. From these studies, infor-
Verma, Ledward, & Lawrie, 1984) were conducted to incorporate mation is available on the gelation of legumin, vicilin and pea
pea protein into comminuted meat products. Their results showed protein isolates that have been recovered by pH manipulation and
that pea protein or pea flour had a weakening effect on the texture isoelectric precipitation.
of meat products. To further the adoption of pea proteins as func- Dry pea seeds contain approximately 20e27% protein (Wang &
tional additives in foods, it is essential to understand their gelation Daun, 2004) of which 65e70% are the salt extractable globular
mechanism. Heat-induced gels have been prepared from pea storage proteins legumin, vicilin and convicilin (Schroeder, 1982).
protein isolates (Shand, Ya, Pietrasik, & Wanasundara, 2007) and Composition of globular proteins varies among pea genotypes and
pea protein isolates containing transglutaminase (Shand, Ya, the legumin/vicilin ratio has been shown to fluctuate between 0.2
Pietrasik, & Wanasundara, 2008). In these applications the pea and 1.5 (Casey, Charman, Wright, Bacon, & Guldager, 1982) with
vicilin being the major protein for most cultivars. Subunit compo-
sition of the globular proteins within a given cultivar, are also
* Corresponding author. Tel.: þ1 204 474 9866; fax: þ1 204 474 7630. variable. Legumin is the hexamer (6 subunits) of disulfide linked
E-mail address: arntfie@cc.umanitoba.ca (S.D. Arntfield). basic and acidic subunits which fit together as two trigonal

0268-005X/$ e see front matter Ó 2012 Elsevier Ltd. All rights reserved.
doi:10.1016/j.foodhyd.2011.12.014
326 X.D. Sun, S.D. Arntfield / Food Hydrocolloids 28 (2012) 325e332

antiprisms. Heterogeneity of these legumin polypeptides is a result interacting through hydrogen bonds that might otherwise interact
of the production of legumin precursors from a number of gene with the solvent surrounding the molecule.
families: four/five acidic and five/six basic polypeptides have been Sulfhydryl/disulfide interchange has been proposed to be
identified. Variable amino acid compositions have also been re- involved in soy protein gelation based on the reaction of the gel (a
ported. Approximately two cysteine and three methionine residues loss of the gel integrity) to several reagents: b-mercaptoethanol (b-
per w60 kDa subunit were found for the acidic and basic subunits, ME or 2-ME) (Briggs & Wolf, 1957; Catsimpoolas & Meyer, 1970;
respectively by Casey and Short (1981), while Croy, Gatehouse, Utsumi, Damodaran, & Kinsella, 1984; Utsumi & Kinsella, 1985;
Tyler, and Boulter (1980) reported seven and four. Pea vicilin is Wolf & Briggs, 1958; Wolf, 1993), dithiothreitol (DTT) (McKlem,
a trimer made of 3 subunits of w50 kDa (Gatehouse, Croy, Morton, 2002; Utsumi & Kinsella, 1985; Wolf, 1993), and N-ethyl-
Tyler, & Boulter, 1981). Smaller polypeptides which have been maleimide (NEM) (Briggs & Wolf, 1957; Catsimpoolas & Meyer,
associated with isolated vicilin are reported to be due to post 1970; Shimada & Cheftel, 1988; Utsumi & Kinsella, 1985; Wang &
translational proteolysis (Gatehouse, Lycett, Delauney, Croy, & Damodaran, 1990; Wolf & Briggs, 1958). When electrostatic forces
Boulter, 1983). Fragments of 19 and 30 kDa or 33 and 16 or are involved in gel formation of whole plasma protein, gel strength
12.5 kDa are produced depending on the site of cleavage during is affected by pH and salts (Hickson, Dill, Morgan, Suter, &
proteolysis (Gatehouse, Lycett, Croy, & Boulter, 1982). O’Kane et al. Carpenter, 1980; Hermansson, 1982a, b; O’Riordan, Kinsella,
(2004a) separated vicilin into two fractions and named them as Mulvihill, & Morrissey, 1988, O’Riordan, Mulvihill, Kinsella, &
vicilin 1 and vicilin 2 and observed that on SDS-PAGE, vicilin 2 Morrissey, 1988). Therefore, involvement of electrostatic interac-
contained a third globulin protein, convicilin (w70 kDa). They tions can be determined by the effect of salts and pH.
concluded that convicilin was not a separate, third globulin of pea More experimental data are needed to better understand the
and should be denoted as the a-subunit of the salt extractable pea importance of various forces in network formation and to ascertain
protein vicilin. This is in contrast to prior work by Gatehouse et al. their impact on rheological properties of pea protein gels, partic-
(1981) who suggested convicilin was a distinct protein (280 kD) ularly gels formed by pea proteins that have been isolated by salt
made up of 4e10 kD subunits. Pea vicilin is reported to contain no extraction and precipitation by dilution. As no pH adjustments are
cysteine and few methionine residues (Croy et al., 1980; Thanh & used in this isolation method there should be no changes in protein
Shibasaki, 1976). structure due to pH manipulation. Involvement of hydrophobic
Protein gelation is the cross-linking of its polypeptide chains to interactions and hydrogen bonds during gelation of isoelectrically
form a three-dimentional network. Cross linking of proteins is precipitated pea protein have been reported previously (O’Kane
caused by different molecular forces and may involve hydrogen et al., 2004b), but no work has been done using various salts to
bonds, ionic attractions, disulphide bonds, hydrophobic associa- investigate the role of electrostatic interactions. The objective of
tions or a combination of the above (Otte, Schumacher, Ipsen, Ju, & our current research was to evaluate the effects of various salts
Qvist, 1999). Some reagents have been used to investigate molec- (NaSCN, Na2SO4, CH3COONa, NaCl), chemicals that target non-
ular forces (Table 1). covalent interactions (propylene glycol [PG], DTT, urea, GuHCl),
The molecular forces involved in the gel network are dependent chemicals that target covalent interactions (NEM, DTT and 2-ME) as
upon the protein, and protein structure which can be influenced by well as reheating and recooling on the properties of pea protein
the method used for protein isolation (Shimada & Matsushita, 1980; gels prepared form a salt extracted pea protein isolate. This allowed
Utsumi & Kinsella, 1985). The involvement of different interaction us to elucidate the molecular forces involved in gel network
forces in formation and structure of protein gels can be deduced formation and maintenance.
from effects of pH, salts, reducing agents and dissociating agents
(Clark, Judge, Richards, Stubbs, & Suggett, 1981; Mulvihill, Rector, & 2. Materials and methods
Ensella, 1990; Utsumi & Kinsella, 1985). To determine which
molecular forces are involved in the formation of heat-induced pea 2.1. Commercial pea flour and salt-extracted protein isolate (PPIs)
protein gel, different chemical reagents can be employed.
Hydrogen bonds and hydrophobic interactions in protein can be Commercial pea flours were kindly donated by Nutri-Pea Ltd
destabilized by urea and guanidine hydrochloride (GuHCl). Urea is (Portage la Prairie, MB, Canada) and Best Cooking Pulse, Inc.
usually used to denature proteins, but the mechanism is not (Portage la Prairie, MB, Canada). The source of the flour had to be
completely understood. Urea denatures a protein molecule through changed during the experiment; the source of the flour has been
preferential adsorption with charged protein solutes, dehydrating noted in the results. The flours were made by milling dehulled
the molecules and causing repulsion between proteins, stabilizing Canadian yellow pea, and protein content was greater than 25%.
the unfolded form (Wallqvist, Covell, & Thirumalai, 1998). Conse- The pea protein isolate (PPI) was prepared from this flour using
quently, urea probably interferes with both hydrophobic interac- a salt-extraction method that has been described previously (Sun &
tions and hydrogen bonding by dehydrating protein molecules and Arntfield, 2010). After freeze drying (Genesis SQ Freeze Dryer,
Gardiner, NY, U.S.A.), the PPI contained 81.9% protein as determined
Table 1 by Kjeldahl method using an N to protein conversion factor of 5.7
Effect of various reagents on molecular forces exist in protein. (AACC, 1982).
Non-covalent bonds Covalent bond References
2.2. Rheology
Electrostatic Hydrophobic Hydrogen Disulfide
interaction interaction bond bond
All samples were prepared with 14.5% (w/v) pea protein isolate
DTT Disrupt Rüegg and
at its natural pH. Control was prepared with distilled water without
Rudinger (1977)
GuHCl Disrupt Disrupt Tanford (1968) any salt. For the salt series, 0.3 M solutions of Na2SO4 (Fisher
2-ME Disrupt Scientific, Fair Lawn, New Jersey, USA), CH3COONa (Mallinckrodt,
NEM Disrupt Creighton (1993) Inc., Pointe-Claire, Quebec, Canada), NaCl (Fisher Scientific, Ottawa,
PG Promote Disrupt Promote Tanford (1962) Canada) and NaSCN (Fisher Scientific, Fair Lawn, New Jersey, USA)
Urea Disrupt Disrupt Gordon & Jencks
(1963)
were prepared with distilled water. While pH values were not
adjusted, the pH values have been measured for the mixtures with
X.D. Sun, S.D. Arntfield / Food Hydrocolloids 28 (2012) 325e332 327

the various salt solutions. GuHCl (electrophoresis grade; Fisher hydrophobic core of the protein (Zhang & Cremer, 2006). At low ionic
Scientific, Fair Lawn, New Jersey, USA), urea (Fisher Scientific, strengths, salts are believed to primarily influence electrostatic
Nepean, Ontario, Canada), DTT (Sigma Chemical Company, St. Louis, interactions by interacting with charged groups on the proteins,
USA), NEM (Sigma Chemical Company, St. Louis, USA), PG (Sigma- whereas at higher concentrations the ion specific effects, or lyotropic
aldrich, Inc., St. Louis, USA), NaSCN and 2-ME (MP Biochemicals Inc., effects, become prominent (Damodaran & Kinsella, 1981). At these
Solon, Ohio, USA) were dissolved in 0.3M NaCl solution to produce higher concentrations, NaSCN has been shown to be a destabilizing
the desired concentrations. Again, pH was not adjusted for these salt while NaCl stabilizes protein structure (Damodaran & Kinsella,
samples. For most denaturants, a slight increase in pH was ex- 1981; von Hippel & Schleich, 1969). The data in this study support
pected, as was seen with the salts. The buffering capacity of the the possibility of a lyotropic effect at 0.3 M salt as the order of
protein would limit the increase. To achieve complete suspension, increase in denaturation temperature corresponded to the position
samples were mixed by a Vortex-Genie Mixer (Scientific Industries of these salts in the lyotropic series. Melander and Horváth (1977)
Inc., Bohemia, N.Y., USA) for 1 min prior to loading to a TA 2000 indicated that the property of a salt that affected hydrophobic
rheometer (TA Instruments, Newcastle, Del. USA). Rheological tests interactions in proteins was determined by its molal surface tension
were performed as previously described (Sun & Arntfield, 2010). increment (s) independently of the salt concentration. They also
pointed out that s formed the basis of a natural lyotropic series.
2.3. Differential scanning calorimetry (DSC) Within this series, NaSCN is considered to be a destabilizing salt,
with a strong ability to bind to proteins. Not only does this change
The thermal properties of salt extracted pea protein isolate the protein charge, but by creating an excessive negative charge, the
suspensions were examined using a DSC Q2000 and high volume SCN facilitates unfolding of the globular proteins at lower
pans as described previously (Sun & Arntfield, 2010). Thermal temperatures, thus explaining the low denaturation temperature
curves were obtained using 10e20 mL of sample at a concentration observed with this salt. In contrast, SO2
4 anions are considered to be
of 10% and a heating rate of 10  C/min with an empty pan as stabilizers of protein structure (Damodaran & Kinsella, 1981). This is
reference. The sample was heated over a temperature range of reflected in the increased denaturation temperature. It would
30e120  C in a standard DSC cell that had been calibrated with both appear that non-polar groups are further buried within the protein
indium and sapphire standards. Each sample was analyzed in structure resulting in higher temperatures to promote unfolding.
duplicate. This affect was attributed to the salts rather than the pH shift as
previous work has shown not significant change in the denaturation
2.4. Statistical analysis temperature for PPI at the pH values seen with these salts (Sun &
Arntfield, 2011).
All data were analyzed for significant differences, with Although the inclusion of salts increased the thermal denatur-
minimum significance set at the 5% level (P < 0.05), using Tukey’s ation temperatures, with the exception of NaCl the DH values for
test by GraphPad InStat software version 3.06 (GraphPad Software the pea protein in the 0.3 M salts were lower than that of the 10%
Inc. La Jolla, CA, USA). pea protein dispersed in water (no salt) at its natural pH (5.65) and
this could be explained by the protein structure stabilizing effect of
NaCl. For the NaSCN, changes in protein conformation due to the
3. Results and discussion
increased net charge would explain the lower DH value. Confor-
mational changes due to stabilization may also account for the
3.1. Effect of sodium salts
lower DH values for the Na2SO4 and CH3COONa. Damodaran and
Kinsella (1981) indicated that although higher denaturation
The thermal denaturation data for the pea protein isolate in the
temperature values reflect resistance to thermal denaturation, the
0.3 M salts indicated that the denaturation temperature increased
tertiary and quaternary structures of the stabilized protein might
when salts were present with the greatest increase for Na2SO4, fol-
not be the same as that of the native protein, and a lower DH value
lowed by CH3COONa, NaCl and finally NaSCN (Table 2). The effects of
can result.
salts on protein structure involve three mechanisms: electrostatic
In the presence of the 0.3 M Na2SO4, CH3COONa, NaCl, and
shielding effects, non-specific charge neutralization and direct
NaSCN, rheological data for the PPIs showed improved structure
ionemacromolecule interactions (Zhang & Cremer, 2006). It is
development for all four treatments in comparison to the water
believed that the ion specific effects arise from changes in the
control (Table 2). Poor solubility for the water sample may account
for the lack of network formation. Among the four treatments, the
Table 2 SCN anion showed the greatest increase in G0 values followed by
Effect of 0.3 M sodium salt on the thermal denaturation of 10% PPIs at natural pH Cl, CH3COO, and SO2 4 . It appears that the ability of the pea
(5.65) and rheological properties of 14.5% PPIs gels (pH 5.65) at 1 Hz sweep
protein to form gels in the presence of 0.3 M sodium salts was, in
frequency.
part, related to the thermal denaturation temperature and there-
Salt Td ( C) ΔH G0 (Pa) Tan d fore also dependent upon the lyotropic influence of these salts. For
(J/g protein)
CH3COO, and SO2 4 , the denaturation temperature was increased
Control - pH 5.7 86.2  0.1a 15.8  0.0c 0.35  0.2a 1.36  0.6b to values in excess of 95  C, the temperature used for gel formation.
Sodium sulfate 104.6  0.0d 11.6  0.9ab 248.5  62.9a 0.18  0.0a
(Na2SO4) - pH 5.9
As a result, the degree of protein unfolding was limited, thus
Sodium acetate 98.0  0.1c 13.5  0.4b 1198  197a 0.17  0.0a reducing the potential for network formation. The high G0 value
(CH3COONa) - pH with SCN resulted from the lower denaturation temperature and
6.3 exposure of reactive groups as well as electrostatic shielding effects
Sodium chloride 94.3  0.3b 17.8  0.2c 4516  188b 0.17  0.0a
which minimized charge repulsion within the protein. While these
(NaCl) - pH 5.7
Sodium thiocyanate 93.6  0.0b 10.7  0.3a 14560  18667c 0.16  0.0a salts affected gel stiffness differently they did not impact tan d of
(NaSCN) - pH 6.0 the networks formed as the tan d was not significantly affected by
Values represent means  SD of duplicates.
the salt used (Table 2).
a-d
Column values followed by the same superscript letter are not significantly As NaSCN is a chaotropic salt which destabilizes proteins in
different (p < 0.05). PPIs was extracted from Nutri-Pea Ltd. sample. solution and promotes protein solubility, the ability to form
328 X.D. Sun, S.D. Arntfield / Food Hydrocolloids 28 (2012) 325e332

a strong network was interesting. Although SCN is a destabilizing Table 3


anion, the Td value at a concentration of 0.3 M was still high Effect of guanidine hydrochloride (GuHCl) concentration on the rheological
properties of PPIs gels at 1 Hz and thermal denaturation of PPIs. The concentration of
(93.6  C). A similar result has been reported for the 11S soy globulin PPIs dispersion was 14.5% at natural pH 5.65 (all samples contained 0.3 M NaCl).
(Damodaran, 1988). In a study of chaotropic anions on the release of
insoluble membrane proteins, Hincha (1998) observed that the GuHCl G0 (Pa) Tan d Td ( C) ΔH
concentration (J/g protein)
presence of chaotropic salts reduced the energy barrier for the (M)
dissociation of proteins from their binding sites on the membrane.
0 4516  188a 0.1678  0.00a 86.2  0.1a 15.81  0.0a
In the present study, this reduction in the energy at the relatively 0.3 619.5  113.8b 0.108  0.00a 93.04  0.09b 10.92  0.26b
low NaSCN concentration led to enhanced gel stiffness. Higher 1.0 203  51.6c 0.249  0.02a 93.24  0.23b 9.08  0.13c
NaSCN concentrations were examined to further investigate the 3.0 0.08  0.04c 15.09  5.86b _ _
impact of this chaotropic salt. With 1 M NaSCN, the gel stiffness (G0 ) Values represent means  SD of duplicates.
a-c
was reduced significantly, although the tan d was not affected, but Column values followed by the same superscript letter are not significantly
with 3 M NaSCN, gel formation was inhibited (Fig. 1). At the higher different (p < 0.05).
concentrations, the binding of SCN resulted in a change in overall
No denaturation was observed at 3.0M GuHCl. PPIs was extracted from Best Cooking
Pulse, Inc. sample.
protein charges such that the structure was destabilized
(Damodaran & Kinsella, 1981; von Hippel & Schleich, 1969) and gel
formation was inhibited. a concentration of 3.0 M GuHCl, gel formation was inhibited. This
evidence supports the need for hydrogen and ionic bonds in gel
formation.
3.2. Effects of various reagents on non-covalent bonds

3.2.2. Effect of PG
Various molecular forces include hydrogen bond, hydrophobic
PG may disrupt hydrophobic forces and enhance hydrogen
and electrostatic interactions. They tend to be disrupted by most
bonds and electrostatic interactions by lowering the dielectric
chaotropic agents such as detergents, urea, or GuHCl (Sood &
constant of solvent, and reducing the energy barrier to
Slattery, 2003). To investigate these non-covalent bonds that
proteineprotein interaction enough to enable structure formation
contribute to the gel formation, PPIs was dispersed into various
(Utsumi & Kinsella, 1985). Our results showed that the gel stiffness
reagents (GuHCl, PG, and Urea) solution containing 0.3 M NaCl prior
of pea protein gradually increased with the increasing amount of
to heat treatment.
added PG but the tan d value did not change (Fig. 2) which might
reflect the effect of PG on enhancement of the hydrogen bonds, and
3.2.1. Effect of GuHCl
increasing G00 by promoting proteineprotein interactions. This
GuHCl is a strong ionic denaturing agent (Tanford, 1968), which
would suggest that hydrogen bonds and electrostatic interaction
weakens hydrophobic interactions and inhibits hydrogen and ionic
play a prominent role in determining the stiffness of pea protein
bonds. Tanford (1968) concluded that GuHCl gave the most
gel. These results are consistent with the suggestion that electro-
extensively unfolded state, in which the protein molecules are
static interactions are important in the formation of elastic gels, and
devoid of their native conformation and behave as random coils.
suggest that hydrogen bonding complements electrostatic inter-
GuHCl is a more effective denaturant than urea, unfolding proteins
actions in pea protein gels.
at two to three times lower concentrations than urea (Greene &
Pace, 1974), and GuHC1 is chemically stable, while urea slowly
3.2.3. Effect of urea
decomposes to form cyanate and ammonia.
Zou, Habermann-Rottinghaus, and Murphy (1998) indicated
As expected, the addition of GuHCl to the pea protein isolate
that urea binds to amide groups through hydrogen bonds,
used in this study resulted in protein denaturation, as evidenced by
decreasing the hydrophobic effect through dehydration of the
gradual decrease in DH up to 1.0 M GuHCl and no measureable
protein molecule, and pointing out that hydrophobic groups and
structure change with 3 M GuHCl (Table 3). The higher Td values
hydrophilic groups are involved in the denaturation caused by urea.
with 0.3 and 1.0 M GuHCl are an indication that the more stable
Walstra (2003) also reported that the denaturing effect of urea is
structural components were retained, despite the lower enthalpy
caused by a dehydration of peptide bonds which were bound by
values. While the Td values at 0.3 and 1.0 M GuHCl were low enough
urea also weakens hydrophobic interactions. In Fig. 3, G0 was
for the protein to unfold during gel preparation, there was a gradual
decrease in the G0 values with higher levels of GuHCl. At

Fig. 1. Effect of different concentration NaSCN on gelation properties of 14.5% pea Fig. 2. Effect of different concentration PG on gelation properties of 14.5% pea protein
protein isolate dispersion contains 0.3 M NaCl, at pH 5.65. isolate dispersion contain 0.3 M NaCl, at pH 5.65.
X.D. Sun, S.D. Arntfield / Food Hydrocolloids 28 (2012) 325e332 329

Fig. 3. Effect of different concentration urea on gelation properties of 14.5% pea protein
isolate dispersion contain 0.3 M NaCl, at pH 5.65.
Fig. 4. Effect of different concentration DTT on gelation properties of 14.5% pea protein
isolate dispersion contain 0.3 M NaCl, at pH 5.65.
unaffected when the urea concentration was raised to 2 M, but
decreased dramatically at 5 M and no gel was formed at 8 M. Urea
bonds to pea protein gel network formation and maintenance.
denatured pea protein severely by breaking down hydrogen bonds
Results were similar to those with DTT in that 2-ME had no effect
and hydrophobic interactions, preventing network formation. This
on the G0 value for the pea protein gels and tan d values were higher
reduction in G0 of pea protein gel confirms the involvement of
in the presence of 2-ME (Fig. 5) further supporting the observation
hydrogen bonds and/or hydrophobic interactions in gel networks.
that disulfide bonds do not play a major role in pea protein gel
Tanford (1968) indicated that urea is a strong denaturing agent
formation. O’Kane et al. (2004b) also concluded that under normal
which can induce an extensively unfolded state, in which the
heating conditions, gel formation for pea legumin does not require
protein molecule behaves like a random coil. The denatured states
disulfide bonds.
obtained by other denaturants such as LiClO4, 2-ME, or DTT are
“intermediate” states between native and urea denatured states
3.3.3. Effect of NEM
(Tanford, 1968).
NEM is a reagent which can react with sulfhydryl groups to form
Hydrogen bonds, hydrophobic interactions, and electrostatic
a stable alkyl derivative; preventing the formation of disulfide
interactions are all believed to be important in the overall balance
bonds between protein molecules (Creighton, 1993). Our results
of attractive and repulsive forces contributing to network
showed that NEM concentrations of 50 mM and less had no
formation.
significant impact on gel rheological characteristics. With 100 mM
NEM, the G0 values were greater than without NEM or at lower NEM
3.3. Effect on covalent bonds concentrations and tan d values were also higher than the control
(Fig. 6). Hua, Cui, Wang, Mine, and Poysa (2005) investigated the
Of the two major pea proteins, the vicilin contains no cysteine effect of NEM on the gel forming ability of different varieties of
residues (Croy et al., 1980; Thanh & Shibasaki, 1976) and would not soybean and obtained the similar results. A similar phenomenon
be involved in disulphide bond formation. Although there is vari- involving myofibrillar protein was attributed to cleaving of inter-
ability in the legumin protein it is generally thought to contain 2e7 and intra-molecular disulfide bonds thus facilitating protein
cysteine residues per 60 KDa subunit (Casey & Short, 1981; Croy unfolding and increasing the exposure of reactive groups involved
et al., 1980; O’Kane et al., 2004b) and the 60 KDa subunits each in hydrogen bonding and ionic and hydrophobic interactions
contain a basic and an acidic subunit held together by 1 or more (Ustunol, Xiong, Means, & Decker, 1992). This may also be the case
disulfide bonds. As a result, it is expected that some sulphydryl for pea proteins at an NEM concentration of 100 mM, where
groups would be available for disulfide bond formation during a stronger less elastic gel is formed.
gelation.

3.3.1. Effect of DL-dithiothreitol (DTT)


DTT is often used to reduce the disulfide bonds of proteins and
to prevent intra- and intermolecular disulfide bonds from forming
between cysteine residues (Rüegg & Rudinger, 1977). Although DTT
is a disulfide reducing agent, our results showed that its addition to
pea protein dispersions at low concentrations (0.05e0.15 M) did
not disrupt protein gel stiffness, as there was no significant change
in G0 values with increasing levels of DTT (Fig. 4). However, at DTT
concentrations of 0.05 and 0.1 M, the tan d values were significantly
higher than the control, suggesting a decrease in the relative
elasticity of the networks. Thus disulfide bonds do not appear to be
necessary for pea protein gel formation, but they have a minor
effect on gel characteristics.

3.3.2. Effect of 2-mercaptoethanol (b-mercaptoethanol, 2-ME)


By competing for sulfhydryl group, mercaptoethanol can reduce
disulfide bonds of protein (Wang & Damodaran, 1990). As a result, Fig. 5. Effect of different concentration 2-ME on gelation properties of 14.5% pea
the effect of adding 2-ME reflects the contribution of disulfide protein isolate dispersion contain 0.3 M NaCl, at pH 5.65.
330 X.D. Sun, S.D. Arntfield / Food Hydrocolloids 28 (2012) 325e332

Fig. 8. Effect of reheating and recooling on gelation properties of G0 of 14.5% (w/v) pea
Fig. 6. Effect of different concentration NEM on gelation properties of 14.5% pea protein isolate dispersed in 0.3 M NaCl at natural pH 5.65.
protein isolate dispersion contain 0.3 M NaCl, at pH 5.65.

interactions are endothermic and as a result, are stronger at high


It was noted that when the sample was cooled slowly (0.5  C/ temperatures and weaker at low temperature (opposite to that for
min), the effect of added NEM was more pronounced. G0 values for hydrogen bonds) (Damodaran, 1996). Therefore, structure forma-
a sample with 20 mM NEM were greater than those of without tion at these high temperatures probably involves hydrophobic
NEM, whereas no difference could be seen when cooling at 2  C/ interactions.
min (Fig. 7). O’Kane et al. (2004b) indicated that the main forces A reheating and recooling process was used to investigate the
involved in the formation of gel structure of pea legumin protein contribution of hydrogen bonds to the formation of pea protein gel.
are hydrogen bonds, hydrophobic interaction, and disulfide bonds Upon reheating, the G0 values steadily declined as the temperature
at slower cooling rates (1  C/min heating and 0.2  C/min cooling). rose to 95  C, with a rate of decline almost the same as the rate of
Our result was consistent with their observation. They explained increase in G0 during cooling and reached approximately the same
that disulfide bonds became involved in the gel network since the level as G0 of pea protein dispersion which was heated to 95  C
slower cooling rates provide time to react and contribute to gel (Fig. 8). This indicated that changes in gel structure during the
stiffness. cooling phase were largely thermally reversible. Since hydrogen
Overall, under normal heating and cooling (1  C or higher) bonds are weakened with increasing temperature, the thermal
disulfide bonds do not contribute to gel formation and have only reversibility of the curves provided further evidence for the
a minor impact on gel characteristics. This is consistent with the contribution of hydrogen bonds to gel stiffening (Fig. 8). Structures
results of O’Kane et al. (2004b), who indicated that disulfide bonds obtained during the heating phase or the beginning of the cooling
had minimumg involvement in the network formation of isolated phase did not appear to be reversible.
pea legumin proteins. During the final recooling phase, hydrogen bonds are again
believed to be important in gel stiffening. Since hydrogen bonding
3.4. Effect of reheating and recooling is favored at low temperatures, the gradual increase in G0 values
during the cooling phase is again attributed to this molecular force.
Heating and cooling curves monitor structure development of However, G0 values could not reach the same level as obtained
pea protein isolate during processing. Rheological measurement during the first cooling stage, probably because during recooling,
indicated that structure development begins at temperatures of not all the hydrogen bonds disrupted in the reheating stage could
about 85e87  C during the heating phase (Fig. 8). Hydrophobic recover.
In light of the fact that hydrophobic interactions are weakened
and hydrogen bonds strengthened at lower temperatures, the
100000 gradual increase in G0 values during cooling and recooling indicates
hydrogen bonds played a more important role than hydrophobic
interactions in the stage of pea protein gel formation.
10000 Results from heating pea protein pretreated with DTT and 2-ME
Storage modulus (G')

indicated that disulfide bonds were not necessary for gel forma-
tion; however, inclusion of NEM at higher level, plus data obtained
1000 at a cooling rate of 0.5  C/min with the addition of NEM (Fig. 7)
provided evidence that under some conditions disulfide bonds
2 °C/min heating and cooling could contribute to the resulting structure. Léger and Arntfield
2 °C/min heating and cooling with 20 mM NEM (1993) indicated that if a gel fails to melt upon heating this is
100
2 °C/min heating and 0.5 °C/min cooling
usually indicative of the presence of covalent bonds. This situation
2 °C/min heating and 0.5 °C/min cooling with 20 mM NEM
was seen for pea protein gel reheated to 95  C (Fig. 8) as a weak gel
10 structure was retained. In this case, disulfide bonds were probably
20 30 40 50 60 70 80 90 100 involved in the stabilization of the gel structure.
Temperature (°C)
4. Conclusion
Fig. 7. Effect of the addition of 20 mM NEM on the development of storage modulus
(G0 ) during heating and cooling phase of 14.5%, 0.3 M NaCl, pea protein dispersion.
Heating phase of different treatments is essentially identical with each other thus not As it is difficult to attribute the inclusion of a given reagent with
shown. changes to one specific type of interaction, a range of chemical
X.D. Sun, S.D. Arntfield / Food Hydrocolloids 28 (2012) 325e332 331

modifiers have been used in this study. In addition, changes in pH Hincha, D. K. (1998). Release of two peripheral proteins from chloroplast thylakoid
membranes in the presence of a hofmeister series of chaotropic anions. Archives
and ionic strength can affect gelation properties as has been re-
of Biochemistry and Biophysics, 358, 385e390.
ported previously (Sun & Arntfield, 2011). Considering the overall von Hippel, P. H., & Schleich, T. (1969). Effects of neutral salts on the structure and
results, there is good evidence that for pea protein isolated by salt conformational stability of macromolecules in solution. In S. N. Timasheff, &
extraction and precipitated by dilution, non-covalent bonds play G. D. Fasman (Eds.), Structure and stability of biological macromolecules (pp. 417).
New York: Marcel Dekker.
a key role in gel formation, and disulfide bonds were not required to Hua, Y., Cui, S. W., Wang, Q., Mine, Y., & Poysa, V. (2005). Heat induced gelling
form a gel. A minor role for disulfide bonds, under some conditions, properties of soy protein isolates prepared from different defatted soybean
was supported by reduced gel stiffness when using slow heating flours. Food Research International, 38, 377e385.
Léger, L. W., & Arntfield, S. D. (1993). Thermal gelation of the 12S canola globulin.
and cooling rates in the presence of NEM and incomplete melting Journal of the American Oil Chemists’ Society, 70, 853e861.
upon reheating the gel. This role was shown to affect gel elasticity McKlem, L. K. (2002). Investigation of molecular forces involved in gelation of
but not gel stiffness. While electrostatic and hydrophobic interac- commercially prepared soy protein isolates. MS thesis, North Carolina State
University.
tions contributed to the initial structure development during gel Melander, W., & Horváth, C. (1977). Salt effects on hydrophobic interactions in
formation, hydrogen bonds appeared to be responsible for stiff- precipitation and chromatography of proteins: an interpretation of the lyo-
ening the gel during cooling. tropic series. Archives of Biochemistry and Biophysics, 183, 200e215.
Mulvihill, D. M., Rector, D., & Ensella, J. E. (1990). Effect of structuring and
destructuring anionic ions on the rheological properties of thermally induced
b-lactoglobulin gels. Food Hydrocolloids, 4, 267e276.
Acknowledgments O’Kane, F. E., Happe, R. P., Vereijken, J. M., Gruppen, H., & van Boekel, M. A. J. S.
(2004a). Characterization of pea vicillin. 2. Consequences of compositional
heterogeneity on heat-induced gelation behaviour. Journal of Agricultural and
The support of the Natural Sciences and Engineering Research Food Chemistry, 52, 3149e3154.
Council of Canada (NSERC) is gratefully acknowledged. O’Kane, F. E., Happe, R. P., Vereijken, J. M., Gruppen, H., & van Boekel, M. A. J. S.
(2004b). Heat-induced gelation of pea legumin: comparison with soybean
glycinin. Journal of Agricultural and Food Chemistry, 52, 5071e5078.
O’Kane, F. E., Vereijken, J. M., Gruppen, H., & van Boekel, M. A. J. S. (2005). Gelation
References behaviour of protein isolates extracted from 5 cultivars of Pisum sativum L.
Journal of Food Science, 70, C132eC137.
AACC. (1982). Approved methods of the American association of cereal chemists (8th O’Riordan, D., Kinsella, J. E., Mulvihill, D. M., & Morrissey, P. A. (1988). Gelation of
ed.). St. Paul, MN. plasma proteins. Food Chemistry, 33, 203e214.
Bora, P. S., Brekke, C. J., & Powders, J. R. (1994). Heat induced gelation of pea (Pisum O’Riordan, D., Mulvihill, D. M., Kinsella, J. E., & Morrissey, P. A. (1988). The effect of
sativum) mixed globulins, vicilin and legumin. Journal of Food Science, 59, salts on the rheological properties of plasma protein gels. Food Chemistry, 34,
594e596. 1e11.
Briggs, D. R., & Wolf, W. J. (1957). Studies on the cold-insoluble fraction of the water- Otte, J., Schumacher, E., Ipsen, R., Ju, Z., & Qvist, K. B. (1999). Protease induced
extractable soybean proteins. I. Polymerization of the 11S component through gelation of unheated and heated whey proteins: effects of pH, temperature and
reactions of sulfhydryl groups to form disulfide bonds. Archives of Biochemistry concentration of proteins, enzyme and salts. International Dairy Journal, 9,
and Biophysics, 72, 127e144. 801e812.
Casey, R., Charman, J. E., Wright, D. J., Bacon, J. R., & Guldager, P. (1982). Quantitative Pietrasik, Z., & Janz, J. A. M. (2010). Utilization of pea flour, starch-rich and fiber-rich
variability in Pisum seed globulins: its assessment and significance. Plant Foods fractions in low fat bologna. Food Research International, 43, 602e608.
for Human Nutrition, 31, 333e346. Rüegg, U. T., & Rudinger, J. (1977). Cleavage of disulfide bonds in proteins. Methods
Casey, R., & Short, M. N. (1981). Variation in the amino acid composition of legumin in Enzymology, 47, 111e116.
from Pisum. Phytochemistry, 20, 21e23. Sanjeewa, W. G. T., Wanasundara, J. P. D., Pietrasik, Z., & Shand, P. J. (2010). Char-
Catsimpoolas, N., & Meyer, E. W. (1970). Gelation phenomena of soybean globulins. acterization of chickpea (Cicer arietinum L.) flours and application inlow-fat
I. Proteineprotein interactions. Cereal Chemistry, 47, 559e570. pork bologna as a model system. Food Research International, 43, 617e626.
Clark, A. H., Judge, F. J., Richards, B., Stubbs, J., & Suggett, A. (1981). Electron Schroeder, H. E. (1982). Quantitative studies on the cotyledonary proteins in the
microscopy of network structure in thermally induced globular protein gels. genus Pisum. Journal of the Science of Food and Agriculture, 33, 623e633.
International Journal of Peptide and Protein Research, 17, 380e392. Serdaroglu, M., Yildiz-Turp, G., & Abrodimov, K. (2005). Quality of low-fat meatballs
Creighton, T. E. (1993). Protein structure and molecular properties (2nd ed.). New containing legume flours as extenders. Meat Science, 70, 99e105.
York, USA: W.H. Freeman & Co. Shand, P. J., Ya, H., Pietrasik, Z., & Wanasundara, P. K. J. P. D. (2007). Physicochemical
Croy, R. R. D., Gatehouse, J. A., Tyler, M., & Boulter, D. (1980). The purification and and textural properties of heat-induced pea protein isolate gels. Food Chemistry,
characterisation of a third storage protein (convicilin) from the seeds of pea 102, 1119e1130.
(Pisum sativum L.). Biochemical Journal, 191, 509e516. Shand, P. J., Ya, H., Pietrasik, Z., & Wanasundara, P. K. J. P. D. (2008). Trans-
Damodaran, S. (1988). Refolding of thermally unfolded soy proteins during the glutaminase treatment of pea proteins: effect on physicochemical and rheo-
cooling regime of the gelation process: effect on gelation. Journal of Agricultural logical properties of heat-induced protein gels. Food Chemistry, 107, 692e699.
and Food Chemistry, 36, 262e269. Shimada, K., & Cheftel, J. C. (1988). Determination of sulfhydryl groups and disulfide
Damodaran, S. (1996). Amino acids, peptides, and proteins (Chapter 6). In bonds in heat-induced gels of soy protein isolate. Journal of Agricultural and
O. R. Fennema (Ed.), Food chemistry (3rd ed.). (pp. 350) New York, USA: Marcel Food Chemistry, 36, 147e153.
Dekker, Inc. Shimada, K., & Matsushita, S. (1980). Relationship between thermocoagulation of
Damodaran, S., & Kinsella, J. E. (1981). The effect of neutral salts on the stability of proteins and amino acid composition. Journal of Agricultural and Food Chemistry,
macromolecules. Journal of Biological Chemistry, 256, 3394e3398. 28, 413e417.
Gatehouse, J. A., Croy, R. R. D., Morton, H., Tyler, M., & Boulter, D. (1981). Charac- Sood, S. M., & Slattery, C. W. (2003). The use of lithium chloride to study human
terization and subunit structures of the vicilin storage proteins of pea (Pisum milk micelles. Journal of Dairy Science, 86, 78e85.
sativum L.). European Journal of Biochemistry, 118, 627e633. Su, Y. K., Bowers, J. A., & Zayas, J. F. (2000). Physical characteristics and micro-
Gatehouse, J. A., Lycett, G. W., Croy, R. R. D., & Boulter, D. (1982). The post- structure of reduced-fat frankfurters as affected by salt and emulsified fats
translational proteolysis of the subunits of vicilin from pea (Pisum sativum L.). stabilized with nonmeat proteins. Journal of Food Science, 65, 123e128.
Biochemical Journal, 207, 629e632. Sun, X. D., & Arntfield, S. D. (2010). Gelation properties of salt extracted pea protein
Gatehouse, J. A., Lycett, G. W., Delauney, J., Croy, R. R. D., & Boulter, D. (1983). induced by heat treatment. Food Research International, 43, 509e515.
Sequence specificity of the post-translational proteolytic cleavage of vicilin, Sun, X. D., & Arntfield, S. D. (2011). Dynamic oscillatory rheological measurement
a seed storage protein of pea (Pisum sativum L.). Biochemical Journal, 212, and thermal properties of pea protein extracted by salt method: effect of pH
427e432. and NaCl. Journal of Food Engineering, 105, 577e582.
Gordon, J. A., & Jencks, W. P. (1963). The relationship of structure to the effective- Tanford, C. (1962). Contributions of hydrophobic interactions to the stability of the
ness of denaturing agents for proteins. Biochemistry, 2, 47e57. globular conformation of proteins. Journal of the American Chemical Society, 84,
Greene, R. F., Jr., & Pace, C. N. (1974). Urea and guanidine hydrochloride denatur- 4240e4247.
ation of ribonuclease, lysozyme, a-chymotrypsin, and b-lactoglobulin. Journal of Tanford, C. (1968). Protein denaturation. Advances in Protein Chemistry, 23, 121e282.
Biological Chemistry, 249, 5388e5393. Thanh, V. H., & Shibasaki, K. (1976). Heterogeneity of beta-conglycinin. Biochimica et
Hermansson, A. M. (1982a). Gel characteristics-compression and penetration of Biophysica Acta, 439, 326e338.
blood plasma gels. Journal of Food Science, 47, 1960e1964. Ustunol, Z., Xiong, Y. L., Means, W. J., & Decker, E. A. (1992). Forces involved in mixed
Hermansson, A. M. (1982b). Gel characteristics-structure as related to texture and pork myofibrillar protein and calcium alginate gels. Journal of Agricultural and
water binding of blood plasma gels. Journal of Food Science, 47, 1965e1972. Food Chemistry, 40, 577e580.
Hickson, D. W., Dill, C. W., Morgan, R. G., Suter, D. A., & Carpenter, Z. L. (1980). Utsumi, S., Damodaran, S., & Kinsella, J. E. (1984). Heat-induced interactions
A comparison of heat-induced gel strength of bovine plasma and egg albumen between soybean proteins: preferential association of 11S basic subunits and
proteins. Journal of Animal Science, 51, 69e73. b subunits of 7S. Journal of Agricultural and Food Chemistry, 32, 1406e1412.
332 X.D. Sun, S.D. Arntfield / Food Hydrocolloids 28 (2012) 325e332

Utsumi, S., & Kinsella, J. E. (1985). Forces involved in soy protein gelation: effects of 1efd063dbc685ffa5040f927cdb94f9b/The-Chemical-Composition-and-
various reagents on the formation, hardness and solubility of heat-induced gels Nutritive-Value-of-Canadian-Pulses.pdf Retrieved on Oct. 13, 2010.
made from 7S, 11S and soy isolate. Journal of Food Science, 50, 1278e1282. Wolf, W. J. (1993). Sulfhydryl content of glycinin: effect of reducing agents. Journal
Verma, M. M., Ledward, D. A., & Lawrie, R. A. (1984). Utilization of chickpea flour in of Agricultural and Food Chemistry, 41, 168e176.
sausages. Meat Science, 11, 109e121. Wolf, W. J., & Briggs, D. R. (1958). Studies on the cold-insoluble fraction of the
Wallqvist, A., Covell, D. G., & Thirumalai, D. (1998). Hydrophobic interactions in water-extractable soybean proteins. II. Factors influencing conformation
aqueous urea solutions with implications for the mechanism of protein dena- changes in the 11S component. Archives of Biochemistry and Biophysics, 76,
turation. Journal of American Chemist Society, 120, 427e428. 377e393.
Walstra, P. (2003). Physical chemistry of foods. New York: Marcel Dekker, Inc. Zhang, Y., & Cremer, P. S. (2006). Interactions between macromolecules and
Wang, C. H., & Damodaran, S. (1990). Thermal gelation of globular proteins: weight- ions: the Hofmeister series. Current Opinion in Chemical Biology, 10,
average molecular weight dependence of gel strength. Journal of Agricultural 658e663.
and Food Chemistry, 38, 1157e1164. Zou, Q., Habermann-Rottinghaus, S. M., & Murphy, K. P. (1998). Urea effects on
Wang, N., & Daun, J. K. (2004). The chemical composition and nutritive protein stability: hydrogen bonding and the hydrophobic effect. Proteins:
value of Canadian pulses. http://www.pulsecanada.com/uploads/1e/fd/ Structure, Function, and Genetics, 31, 107e115.

You might also like