Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Food Hydrocolloids 63 (2017) 130e138

Contents lists available at ScienceDirect

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

Whey proteinepectin soluble complexes for beverage applications


Ty B. Wagoner, E. Allen Foegeding*
Department of Food, Bioprocessing and Nutrition Sciences, Box 7624, North Carolina State University, Raleigh, NC 27695-7624, USA

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

Article history: There is a strong interest in the consumption of beverages containing whey proteins due to implications
Received 2 May 2016 in health outcomes such as increased satiety and metabolic regulation. However, low thermal stability
Received in revised form limits the conditions under which whey protein beverages can be formulated. Studies have shown that at
8 August 2016
a narrow pH range near the protein isoelectric points, whey proteins and polysaccharides self assemble
Accepted 22 August 2016
into soluble complexes (SCs) that exhibit unique functionality. This study investigated the formation and
Available online 24 August 2016
thermal stability of SCs under conditions relevant to beverage applications. Complexes were formed at
pH 5 using whey protein isolate (WPI; 1e6% w/w) and high-methoxyl pectin (HMP; 0.125e0.75% w/w)
Keywords:
Whey protein
and then heat-set at 85  C for 25 min. Hydrodynamic properties, particle size distribution, zepotential,
Polysaccharides and dispersion viscosity were evaluated before and after heat-setting. Mean particle diameter ranged
Rheology from 300 to 715 nm for unheated SCs, and 230 nm to 1 mm for heat-set SCs. Heat-setting SCs led to a
Complexes significant (p < 0.05) reduction in intrinsic viscosity from 93.6 mL/g to 79.5 mL/g, suggesting confor-
Beverages mational changes that favor a smaller hydrodynamic size. Dispersions of SCs exhibited decreased
Thermal stability apparent viscosity, consistent with the lower intrinsic viscosity and smaller particle size. Heat-set SCs (4%
WPI, 0.5% HMP) remained as sub-micron particles (d ¼ 303e829 nm) after pH adjustment (pH 4e7) and
thermal processing (142  C for 6 s), indicating that WPI and HMP can be heat-set into complexes with
enhanced colloidal stability in beverage applications.
© 2016 Elsevier Ltd. All rights reserved.

1. Introduction (pI) of whey proteins, where protein precipitation, microphase


separation, or gelation may occur after heating (Ako, Nicolai,
There is strong interest in the consumption of whey proteins Durand, & Brotons, 2009; Langton & Hermansson, 1992).
due to potential health benefits, including increased satiety, One method of influencing whey protein thermal stability is via
metabolic regulation, and as part of a weight loss intervention the formation of whey protein and polysaccharide complexes
€rck, 2007; Zafar,
program (Hector et al., 2015; Nilsson, Holst, & Bjo (Turgeon, Schmitt, & Sanchez, 2007). Complexation occurs as a
Waslien, AlRaefaei, Alrashidi, & AlMahmoud, 2013). One partic- result of electrostatic interactions between anionic polysaccharides
ular category of interest is protein beverages, such as meal- and positively charged surface patches (notably eNHþ 3 groups) on
replacement beverages or recovery sports drinks. Beverages the proteins (Cooper, Dubin, Kayitmazer, & Turksen, 2005;
require thermal processing for safety and shelf stability, but whey Tolstoguzov, 2003). This “charge patch” mechanism may also be
proteins are notoriously fastidious to heat, particularly at pH or influenced by charge regulation and ion-dipole interactions at pH
ionic strength where electrostatic stabilization is diminished. near the pI (da Silva & Jo €nsson, 2009). Depending on the electro-
Currently, stable beverages are formulated under acidic or neutral static conditions, the complexes will remain as small, soluble
conditions, where astringency and off-flavors respectively are of complexes (SCs) or undergo associative or segregative phase sepa-
concern (Beecher, Drake, Luck, & Foegeding, 2008; Evans, ration (Cooper et al., 2005; de Kruif, Weinbreck & de Vries, 2004;
Zulewska, Newbold, Drake, & Barbano, 2010). To mitigate senso- Tolstoguzov, 2003; Turgeon et al., 2007; Weinbreck, Wientjes,
rial defects, whey protein beverages would ideally be formulated at Nieuwenhuijse, Robijn, & Kruif, 2004).
pH 4e6. However, this pH range encompasses the isoelectric point Soluble complexes have previously been shown to exhibit
functionality in a number of food applications (Champagne &
Fustier, 2007; Liu, Xu, & Guo, 2007; Schmitt & Turgeon, 2011).
This represents a unique opportunity in beverage applications,
* Corresponding author.
where to our knowledge the use of soluble complexes has not been
E-mail address: allen_foegeding@ncsu.edu (E.A. Foegeding).

http://dx.doi.org/10.1016/j.foodhyd.2016.08.027
0268-005X/© 2016 Elsevier Ltd. All rights reserved.
T.B. Wagoner, E.A. Foegeding / Food Hydrocolloids 63 (2017) 130e138 131

explored. Previous research indicates that SCs are formed over the whey proteins as determined by the manufacturer was 51.2% belg,
pH range 4e6 (Gente s, St-Gelais, & Turgeon, 2010; Jones & 21.3% aelactalbumin, 24.3% GMP, 1.6% immunoglobulins, 0.8%
McClements, 2011; Krzeminski, Prell, Weiss, & Hinrichs, 2014; bovine serum albumin and 0.2% lactoferrin. Mineral composition,
Salminen & Weiss, 2013). Their physical characteristics can be as determined by inductively coupled plasma atomic emission
optimized via judicious selection of pH, polymer ratio, ionic spectroscopy, was 14.4% N, 0.3% P, 0.4% K, 0.6% Ca, 1.1% S, and 0.2%
strength, ionic species, polysaccharide charge density, and cosol- Na. High-methoxyl pectin (71.8% esterified as reported by the
vents (Chanasattru, Jones, Decker, & McClements, 2009; Girard, manufacturer, <1% ash) was provided by CP Kelco (Atlanta, GA,
Turgeon, & Gauthier, 2002; Hirt & Jones, 2014; Murphy, Cho, USA). Chemical reagents were of analytical grade and supplied by
Farkas, & Jones, 2015; Salminen & Weiss, 2013; Sperber, Cohen Sigma-Aldrich (St. Louis, MO, USA). All reported protein concen-
Stuart, Schols, Voragen, & Norde, 2009). Moreover, SCs exhibit trations account for ingredient purity and are expressed as con-
stability to a heat-setting step above protein denaturation tem- centration of protein, not concentration of WPI powder.
peratures and under conditions normally associated with low sta-
bility, such as pH near the pI or high salt (Jones & McClements, 2.2. Preparation of solutions and formation of soluble complexes
2008; Krzeminski et al., 2014). The heat-setting step directs pro-
tein aggregation and complexation towards the formation of par- Stock protein solutions (10% w/w protein) were formed by sol-
ticles of a discrete size ranging from 225 to 540 nm depending on ubilizing WPI in deionized water (>17 MU) by stirring at 300 rpm
formation conditions (Jones & McClements, 2010; Salminen & for 3 h at ambient temperature (23 ± 1  C), then stored at 5  C
Weiss, 2013). However, most of these studies formed complexes overnight. Stock HMP solutions (2% w/w) were formed by stirring
at low protein concentrations e typically 0.5% or below (as reported HMP in deionized water at 600 rpm, heating to 70  C to hydrate,
in review by Jones & McClements, 2011). In order to meet FDA re- and stirring at 400 rpm for 6 h at ambient temperature (23 ± 1  C).
quirements for “high” or “excellent source of protein” claims, Stock solutions were diluted to final concentrations with deionized
beverages would need to contain 10 g protein per serving, or ~4% water and adjusted to appropriate pH using 1 M citric acid or 1 M
protein per 250 mL (Code of Federal Regulations, 2016 title 21, sec NaOH. Immediately prior to the formation of complexes, pectin and
101.54). WPI solutions were adjusted to pH 7. The SCs were formed at an 8:1
To be commercially viable, SCs need to be prepared with com- mass ratio of protein to pectin. This ratio was identified in a pre-
mon food ingredients and economically feasible processing oper- liminary study (data not shown) as forming the smallest particles,
ations. In this study, whey protein isolate (WPI) and high-methoxyl and agrees with the results of Jones, Decker, and McClements
pectin (HMP) were selected due to current widespread usage in (2009).
beverages. Whey protein isolate is a value-added ingredient con- The SCs were formed by first combining the pH 7 stock solutions
taining 90% protein, with the mixture and amount of individual at an 8:1 mass ratio of protein to pectin and diluting to the desired
proteins dependent on the fractionation technique (Huffman & concentration with deionized water. The pH was then adjusted to 5
Harper, 1999). Whey protein isolate derived from sweet whey is with 1 M citric acid to form SCs. Heat-set SCs were formed by
comprised of ~50% belg, ~15% aelactalbumin, ~20% glyco- heating solutions in an 85  C water bath for 25 min. After heating,
macropeptide (GMP), and small amounts of bovine serum albumin, solutions were cooled in an ice slurry and stored overnight at 5  C
lactoferrin, and immunoglobulins (Farrell et al., 2004). Pectin is an before analysis.
anionic polysaccharide derived from acid extraction of plant ma-
terials, most commonly apple pulp or citrus peel. All pectins consist
2.3. Intrinsic viscosity
of a linear backbone of repeating Degalacturonic acid monomers
linked by ae1e4 glycosidic bonds. The presence of neutral sugar
Intrinsic viscosity of polymer solutions was measured using a
side chains and branching are also present to some extent and vary
Cannon-Fenske (Cannon Instruments, State College, PA) capillary
by pectin source. The pKa of carboxyl moieties ranges from 2.9 to
viscometer per the method described by Vardhanabhuti and
3.3 depending on pectin source, and ~70% are naturally methylated
Foegeding (1999) with several modifications. Viscometers were
(Whistler & BeMiller, 1993). Pectin with 50% methylation is
immersed in a temperature controlled water bath set at 40 ± 0.2  C.
classified as HMP. Pectin is commonly used to stabilize acidic
Deionized water was used to dilute the stock protein solution to
casein-based dairy beverages where it adsorbs to the surface of
five concentrations between 0.01 and 0.16% w/w. The pH was
casein micelles, providing steric and electrostatic stabilization
adjusted, if needed, after dilution of the SCs; however, the ionic
(Lopes da Silva & Rao, 2006; Pereyra, Schmidt, & Wicker, 1997;
strength was not adjusted in order to keep the protein:salt ratio
Tromp, de Kruif, van Eijk, & Rolin, 2004).
constant. Pectin was prepared and diluted in 50 mM NaCl to five
The primary requirements for beverage applications are high
concentrations between 0.0125 and 0.2% w/w. For all measure-
protein concentration, stability to thermal processing, appropriate
ments, 7 mL of sample were added to the viscometer and allowed to
viscosity and sensorial attributes, and post-processing shelf sta-
equilibrate for 10 min. Two measurements were performed 10 min
bility. Therefore, the specific goals of this study were to determine
apart and averaged for each data point. Viscometers were rinsed
the WPI and HMP concentrations over which SCs could be formed
three times with deionized water and once with acetone before
and evaluate properties relevant to their use in beverages,
being completely dried between runs. Specific viscosity (hsp) and
including size distribution, hydrodynamic properties, surface
relative viscosity (hrel) were calculated using Eqs. (1) and (2),
charge, and flow behavior. Additionally, the use of heat-set com-
respectively, where t is the efflux time of the sample solution and t0
plexes in a model beverage was explored.
is the efflux time of the solvent (Kragh, 1961).
2. Materials and methods t  t0
hsp ¼ (1)
t0
2.1. Materials
t
Whey protein isolate (WPI) was provided by Glanbia Nutri- hrel ¼ (2)
t0
tionals (Twin Falls, ID, USA). Protein content was 92.13% as deter-
mined by Kjeldahl analysis (N  6.38). The amount of individual Intrinsic viscosity [h] was determined by plotting hsp/c and hrel/
132 T.B. Wagoner, E.A. Foegeding / Food Hydrocolloids 63 (2017) 130e138

c as a function of polymer concentration via extrapolation of the the cube of particle radius. zePotential was calculated using phase
best-fit line to zero polymer concentration on the basis of the analysis light scattering. A voltage of 3 V was applied at 10 Hz, and
Huggins (Eq. (3)) and Kraemer (Eq. (4)) equations, where kH is the five measurements were averaged over 20 s periods in order to
Huggins constant for a given solvent, kK is the Kramer constant for a calculate electrophoretic mobility. Electrophoretic mobility is
given solvent, c is polymer concentration, and [h] is intrinsic vis- related to zepotential by the Smoluchowski equation.
cosity (Tanford, 1961).
2.6. Apparent viscosity
hsp
¼ ½h þ kH ½h2 c (3)
c Viscosity flow profiles were developed via shear rate sweeps at
ambient temperature (23 ± 1  C) using a controlled stress rheom-
lnðhrel Þ
¼ ½hc þ kK ½h2 c (4) eter (Anton Paar MCR 302, Graz, Austria) under controlled rate
c mode. A double gap concentric cylinder attachment was filled with
3.8 mL sample and pre-sheared for 15 s at 10 s1. The sample was
This technique is valid for solutions with hrel values up to ~ 2. All hrel
held at zero shear for 10 s before a linear shear rate ramp was
values were less than 2 with the exception of 0.16% SCs, which had
applied from 0.1 to 200 s1. Flow profiles were modeled using the
hrel values less than 2.8. These values fit the model per the Huggins
Power Law model (Eq. (7)) for fluids where K represents the flow
and Kramer equations and were included in subsequent
consistency coefficient (Pa sn), n represents the dimensionless flow
calculations.
behavior index, h represents apparent viscosity (Pa s), and g_ rep-
resents shear rate (s1).
2.4. Laser diffraction particle size analysis
h ¼ K g_ n1 (7)
The size distribution of SCs was determined using laser
diffraction on a Mastersizer 3000 with Hydro EV sample interface
(Malvern Instruments Ltd., Worcestershire, UK). Size calculations
were made using the Mie scattering model for spherical particles. 2.7. Ultra high temperature (UHT) thermal processing
Water was used as a dispersion solvent with a refractive index of
1.333, and a refractive index of 1.541 and absorbance of 0.001 was Soluble complexes were prepared and heat-set at pH 5 as
used for SCs. Sample solutions were added to reach an obscuration described above. The pH was then adjusted to 3, 4, 6 or 7 using 1 M
level of 10 ± 1% and stirred at 2000 rpm for the duration of the test. citric acid or 1 NaOH. Solutions were heat-treated per the method
The instrument measured the volume fraction in each of 100 previously described by Wagoner, Ward, and Foegeding (2015).
separate size classes ranging from 0.01 to 3000 mm. Particle size Briefly, 2 mL samples were equilibrated to 23  C in capped boro-
results are expressed as 50th and 90th percentile values (D50 and silicate test tubes and processed in a 150  C oil bath under manual
D90) and volume mean diameter D[4,3] (Eq. (5)), where ni is the agitation. Solution temperature reached 141  C in 62 s and was then
number of particles of diameter di. Reporting various size classes held at 141  C for 6 s to mimic thermal processing commonly used
allowed for statistical analysis among treatments. for protein beverages. After the hold time, test tubes were cooled in
P an ice slurry for 1 min and allowed to equilibrate overnight at 5  C
ni d4i before analysis.
D½4;3 ¼ P (5)
ni d3i
2.8. Experimental design

All data are from three independent experimental replications


2.5. Dynamic light scattering (DLS) particle size analysis and unless stated otherwise. Where applicable, mean values are
zepotential given ± standard deviation. Statistical analysis was performed us-
ing the analysis of variance procedure in JMP v12.0 (SAS Institute,
Intensity-weighted z-average diameter and zepotential were Cary, NC, USA). Statistical significance was indicated at p < 0.05.
measured at ambient temperature (23 ± 1  C) simultaneously using Differences among means were determined using Tukey's HSD post
a Wyatt Mo €biuz (Wyatt Technology, Santa Barbara, CA). Samples
hoc comparison test.
were diluted to a protein concentration of 1 mg/mL to reduce
multiple scattering effects. The pH was adjusted, if needed, after
3. Results and discussion
dilution. To calculate particle diameter using DLS, fluctuations due
to Brownian motion were measured over 30 s intervals and quan-
3.1. Intrinsic viscosity
tified per a second order correlation function with a regularization
fit in the accompanying software (Dynamics v7.3.1, Wyatt Tech-
Intrinsic viscosity, [h], represents the contribution to viscosity
nology). The calculated diffusion coefficient is related to the hy-
by a dispersed polymer and is proportional to hydrodynamic vol-
drodynamic diameter per the Stokes-Einstein equation. Mass
ume (Lopes da Silva & Rao, 2006). The [h] of HMP and WPI were
concentration (cmass) was used to approximate the mass percentage
measured independently for comparative purposes, and for SCs to
of particles in distinct size ranges based on the measured scattering
evaluate hydrodynamic changes after heat-setting. All pectin so-
intensity (Eq. (6)), where I is the intensity of scattered light, M is
lutions were formulated at 50 mM NaCl due to the polyelectrolyte
molar mass, P is a form factor, r is particle radius, and q is the
behavior of pectin; hred increases at dilute concentrations in the
angular dependence of scattered light.
absence of added salt due to long-range electrostatic repulsion
Iðr; qÞ (Pals & Hermans, 1952). Ionic strength of the remaining samples
cmass f (6) was not adjusted. The WPI was adjusted to pH 6 due to precipita-
MðrÞ  Pðr; qÞ
tion at pH 5.
The Rayleigh-Gans approximation for Rayleigh spheres was The [h] of HMP (461 mL/g) falls in the established range for
used as a form factor, which assumes molar mass is proportional to citrus-derived HMP (Table 1) (Fishman, Gillespie, Sondney, & El-
T.B. Wagoner, E.A. Foegeding / Food Hydrocolloids 63 (2017) 130e138 133

Table 1 increase in protein concentration from 1 to 6% did not affect the


Intrinsic viscosity of high-methoxyl pectin (HMP), whey protein isolate (WPI), and shape of the size distribution, although the D50 and D[4,3] means
4% WPIeHMP soluble complexes before and after heat-setting.
trended towards a larger diameter (Table 2). After heat-setting, the
Sample pH [h] (mL/g)a SCs were generally smaller in D50, D90, and D[4,3] means at all
HMP (50 mM NaCl) 5.0 461.0 ± 8.8 a protein concentrations. As with the unheated SCs, a concentration
WPI 6.0b 5.1 ± 0.6 d effect on particle diameter was observed.
SCs 5.0 93.6 ± 1.6 b The z-average diameter of unheated SCs ranged from 222 nm at
Heat-set SCsc 5.0 79.5 ± 4.2 c
1% protein to 353 nm at 6% WPI (Table 3). Unheated SCs also tended
Letters indicate significant differences among means within a column as determined to be more variable in particle size (i.e., larger standard deviation),
by Tukey's HSD at a ¼ 0.05.
which was similarly observed as greater polydispersity (i.e., wider
a
Reported [h] means are averages based on the extrapolation of hrel/c and hsp/c to
zero concentration.
peak width) via laser diffraction. As the size population was poly-
b
WPI was evaluated at pH 6.0 due to flocculation at pH 5.0. disperse, particle size distributions were separated into three size
c
Solutions were heat-set at 85  C for 25 min. classes: 1e10 nm, 10e100 nm, and >100 nm. Of the unheated SCs
larger than 100 nm, mean diameter increased with protein con-
centration from 311 nm at 1% protein to 648 nm at 6% protein.
Atawy, 1991). An [h] of 5.1 mL/g for WPI is similar to other reported The z-average diameter of heat-set SCs increased from 257 nm
values for WPI (Vardhanabhuti & Foegeding, 1999). This is greater at 1% protein to 1046 nm at 6% protein (Table 3). Moreover, as
than reported values for belactoglobulin (3.7 mL/g) and the Ein- protein concentration increased the mass percentage decreased in
stein hard sphere model for globular proteins (2.5 mL/g) (Ba ez,
the smaller size classes and increased in the >100 nm size class.
Moro, Ballerini, Busti, & Delorenzi, 2011; Ross-Murphy, 1994). The This suggests an increased extent of aggregation at higher protein
slight deviation from belg values can be attributed to the combi- concentration. This trend was previously reported for belg and
nation of proteins in WPI, notably the large percentage of GMP in HMP complexes formed and heat-set at pH 4.75, where z-average
this sweet whey-derived ingredient. The [h] of unheated diameter increased from 350 nm (1.0% belg, 0.5% HMP) to 550 nm
WPIeHMP SCs was 93.6 mL/g. After heat-setting, SCs had a (2.0% belg, 1.0% HMP) (Jones & McClements, 2010). Slight differ-
significantly lower [h] of 79.5 mL/g, suggesting a reduced hydro- ences can be attributed to differences in polymer type and forma-
dynamic volume after heat-setting. One explanation for the tion conditions, although these results generally agree with our DLS
reduced intrinsic viscosity is the formation of more tightly packed results. Compared to heat-set SCs, unheated SCs had more mass in
(i.e., dense) complexes, which would be accompanied by a reduc- the 1e10 nm size class (43e56%). After heat-setting, only 1e10% of
tion in particle size and hydrodynamic volume. Similar behavior total mass fell in the 1e10 nm size class, with over 88% of mass in
has been reported for heated whey protein and pectin complexes, the >100 nm size class.
where under certain conditions the mean particle size was lower A reduction in mean diameter after heat-setting would support
after heating above protein denaturation temperatures the decreased hydrodynamic volume observed from intrinsic vis-
(Kazmierski, Wicker, & Corredig, 2003; Krzeminski et al., 2014). cosity, whereas the increase in mean diameter observed using DLS
would support a greater degree of protein aggregation. In the case
3.2. Particle size distribution of unheated SCs, D[4,3] means were approximately twice as large as
z-average diameter at all protein concentrations. This could be due
Intrinsic viscosity provided an estimate of hydrodynamic vol- to assumptions on real and imaginary refractive indices of Mie
ume relative to mass but did not directly provide information about scattering theory. Conversely, the z-average diameter of heat-set
particle size. In beverage applications, colloidal stability is most SCs was always larger than D[4,3] and the disparity increases with
important and is governed by the sedimentation rate of dispersed protein concentration. The formation of more linear aggregates
particles. Although Stokes’ law is an ideal model representing an during heat-setting would favor a larger calculated spherical
isolated sphere at infinite dilution, it can be used as a simple sta- diameter.
bility model, where colloidal stability is predicted by small particle One explanation for a transition towards a smaller diameter
size, minimal density difference between the particle and solvent, after heat-setting is the core shell theory of structural rearrange-
or high continuous phase viscosity (Dickinson, 1992). ment (Jones & McClements, 2011; Peinado, Lesmes, Andre s, &
To evaluate the size distribution of SCs, a combination of tech- McClements, 2010; Santipanichwong, Suphantharika, Weiss, &
niques e laser diffraction and dynamic light scattering (DLS) e McClements, 2008). Protein denaturation at elevated tempera-
were used. These complementary techniques indirectly provide tures leads to weakened protein-polysaccharide electrostatic in-
size information based on scattering angle (laser diffraction) or teractions, disruption of hydrogen bonding, and complex
diffusion of particles due to Brownian motion (DLS). Additionally, dissociation. The dissociated denatured proteins aggregate via
each technique favors a particular size class and implies certain nucleation and growth to form microgels in the absence of poly-
assumptions; thus, a combination of the two techniques is an saccharides (Bromley, Krebs, & Donald, 2006; Jones & McClements,
appropriate way to characterize the samples (Shekunov, 2010). As cationic patches reform on the aggregate exterior, elec-
Chattopadhyay, Tong, & Chow, 2007). It is worth noting that un- trostatic interactions cause the anionic polysaccharide to attach on
der our experimental conditions, both techniques assume a the surface, limiting the extent of protein aggregation (Jones &
spherical shape, thus giving a size approximation of an equivalent McClements, 2010). Thus, this reformed particle has a proposed
hard sphere. Particle size distribution was described using number- protein aggregate “core” and polysaccharide “shell.” The anionic
based D50 and D90 values, and D[4,3] volume mean diameter. D[4,3] is surface charge provides electrostatic stabilization. Additionally,
a volume-weighted average and can be skewed by a small number extension of the polysaccharide into the bulk solvent may provide
of large particles, as they occupy a greater volume than small par- steric stabilization in the same way pectin stabilizes micellar casein
ticles; conversely, D50 and D90 values are not skewed by the pres- in acidic milk beverages (Tromp et al., 2004).
ence of a few very large particles to the same extent (Xu, 2000).
Per laser diffraction, unheated WPIeHMP SCs had a monomodal 3.3. zePotential
distribution with peaks ranging from 200 to 400 nm and a small,
trailing “shoulder” of particles in the 5e10 mm range (Fig. 1). An The core shell theory would suggest greater surface charge (i.e.,
134 T.B. Wagoner, E.A. Foegeding / Food Hydrocolloids 63 (2017) 130e138

Fig. 1. Size distribution of unheated and heat-set WPIeHMP soluble complexes formed at pH 5. Percentage indicates protein concentration on a mass basis, at an 8:1 mass ratio of
protein to polysaccharide. Heat-set complexes were heated 25 min at 85  C.

Table 2 to 42.9 ± 0.4 at 6% WPI. These values are slightly more negative
Size characteristics of WPIeHMP soluble complexes (SCs) based on protein con- than the 32 mV reported for SCs (0.1% belg, 0.05% HMP) formed
centration as determined by laser diffraction.
and heat-set at pH 4.75 (Jones, Lesmes, Dubin, & McClements,
Protein (%) D50 (nm) D90 (nm) D[4,3] (nm) 2010). Unheated and heat-set SCs both exceeded the j30j mV
Unheated SCs 1.0 210 ± 54 bcd 1150 ± 432 ab 557 ± 241 ab threshold commonly used as a predictor of colloidal stability. The
4.0 271 ± 33 ab 1350 ± 234 a 664 ± 166 a greater magnitude of surface charge after heating agrees with the
5.0 243 ± 18 abc 1183 ± 140 ab 653 ± 248 a aforementioned core shell theory of structural rearrangement
6.0 300 ± 27 a 1436 ± 155 a 715 ± 100 a
during heating and would suggest that e at least in terms of surface
Heat-set SCs 1.0 172 ± 8 cd 493 ± 20 d 230 ± 4 c charge e SCs would exhibit greater colloidal stability after heat-
± ± ±
4.0 164 2 d 666 9 cd 273 2 c
setting. Furthermore, a greater zepotential magnitude after heat-
5.0 190 ± 3 cd 813 ± 15 bc 329 ± 7 bc
6.0 170 ± 4 cd 895 ± 14 cd 367 ± 6 bc
setting supports the core shell theory of stabilization during heat-
setting.
All complexes were formed at pH 5. Heat-set SCs were heated for 25 min at 85  C.
In regards to both size and zepotential, it is important to note
D50 and D90 represent the diameter that 50% or 90% of particles are smaller than.
D[4,3] values represent the volume mean average. Values are given ± standard de- that much of the research on complexation and heat-setting of
viation. Letters indicate significant differences among means within a column as whey protein/pectin complexes involves belg because it is a well-
determined by Tukey's HSD at a ¼ 0.05.

Table 3
Average particle size and mass percent of unheated and heat-set soluble complexes in various size classes.

Protein (%) z-Average diameter (nm) 1e10 nm 10e100 nm >100 nm

Mean diameter (nm) Mass (%) Mean diameter (nm) Mass (%) Mean diameter (nm) Mass (%)

1%, unheated 222 f 8.5 42.6 b 38.5 ab 0.4 311 e 57 c


4%, unheated 286 def 7.5 52.2 a 67.7 a 0.7 524 d 47 d
5%, unheated 316 de 8.2 51.3 a 61.2 ab 0.3 671 c 48 d
6%, unheated 353 d 6.1 56.3 a 58.5 ab 0.5 648 c 43 d

1%, heat-set 257 ef 11.3 9.9 c 23.3 b 1.4 303 e 89 b


4%, heat-set 585 c 6.2 6.3 cd 32.3 ab 0.5 692 c 93 ab
5%, heat-set 876 b 8.1 4.7 cd 900 b 95 ab
6%, heat-set 1046 a 3.6 1.1 d 1053 a 99 a

All complexes were formed at pH 5. Heat-set SCs were heated for 25 min at 85  C. Z-average diameter represents the intensity-weighted mean diameter. Mean diameter and
mass percentage are reported for three size classes: 1e10 nm, 10e100 nm, and >100 nm. Letters indicate significant differences among means within a column as determined
by Tukey's HSD at a ¼ 0.05.

electrostatic stabilization), which in addition to particle size is an characterized model protein for mechanistic studies (Verheul,
indicator of colloidal stability. Surface charge is commonly esti- Roefs, & de Kruif, 1998). However, from an application perspec-
mated via zepotential, where a high magnitude of charge e either tive, WPI is a more appropriate ingredient for beverages. In this
positive or negative e is desired to attenuate flocculation. A j30j mV study, the WPI was derived from membrane processing and con-
threshold is often used as a guideline for predicting colloidal sta- tained 24.3% GMP. The SCs were formed at pH 5, which falls nears
bility (Jacobs, Kayser, & Müller, 2000). the pI of belg (pH 5.3) and aelactalbumin (pH 4.8) (Swaisgood,
The zepotential of unheated SCs ranged from 33.6 ± 3.8 at 1% 1982). However, GMP has a pI < 3.8 and would thus possess a net
protein to 39.4 ± 0.5 at 6% protein, suggesting a concentration negative charge at pH 5 (Bernal & Jelen, 1985). With a negative
effect (Fig. 2). Heat-set SCs had a greater magnitude (i.e., more charge, GMP would either, theoretically, interact with cationic
negative) zepotential ranging from 38.9 ± 1.1 at 1% WPI protein patches during structural rearrangement or remain
T.B. Wagoner, E.A. Foegeding / Food Hydrocolloids 63 (2017) 130e138 135

Fig. 2. zePotential of whey protein isolate e high methoxyl pectin soluble complexes Fig. 3. Rheological profiles of dispersions of whey protein isolateehigh methoxyl
before and after heat-setting. Complexes were formed at pH 5, and protein concen- pectin soluble complexes. Solutions were formed at pH 5 and are either unheated
tration was varied at a protein to pectin ratio of 8:1 w/w. Letters indicate significant (solid line) or heat-set at 85  C for 25 min (dotted line). Percentage indicates protein
differences among means as determined by Tukey’s HSD at a ¼ 0.05. concentration at a constant 8:1 ratio of protein to pectin. Lines are drawn according to
the power law parameters presented in Table 2.

dispersed in the bulk continuous phase.


in viscosity is in congruence with a smaller [h] and D[4,3] of heat-set
SCs, indicating that the viscosity change could simply be due to the
3.4. Apparent Viscosity smaller hydrodynamic size/volume of the particles. An alternative
hypothesis is that unbound pectin dispersed in the continuous
Variations in the size and shape of SCs would result in different phase becomes bound in particles (Jones & McClements, 2010).
flow patterns of SC dispersions. Flow behavior is particularly Heat-set SCs exhibited more negative zepotential, suggesting that a
important due to the impact of viscosity on processing consider- greater amount of pectin adsorbed to the surface of the particles.
ations and consumer preferences (Thompson, Drake, Lopetcharat, Thus, less free pectin would decrease continuous phase viscosity.
& Yates, 2004). Rheological behavior of protein-polysaccharide This would also explain the lower [h] of heat-set SCs as the po-
coacervates and emulsions stabilized by SCs has been explored, tential contribution to bulk viscosity is lower when pectin is
but fluid dispersions of protein-polysaccharide complexes have not adsorbed to the surface. Future studies comparing the amounts of
been well-characterized (Benichou, Aserin, Lutz, & Garti, 2007; free/bound protein and pectin would be needed to test the veracity
Neirynck et al., 2007; Stone & Nickerson, 2012; Wang, Lee, Wang, of this hypothesis.
& Huang, 2007; Weinbreck et al., 2004). Solutions of both unheated and heat-set SCs trended towards
All samples showed at least some degree of shear-thinning higher K and an increased degree of pseudoplasticity with
behavior and were thus modeled with Power Law parameters increasing protein concentration (Table 4, Fig. 3), which is expected
(Table 4). Shear-thinning behavior is likely due to disruption of based on increased dispersed phase volume. Solutions containing
interparticle linkages or to orientation of particles in the direction heat-set SCs at 5 and 6% protein exhibited lower K values than
of the shear field (Chen, Wen, Janmey, Crocker, & Yodh, 2010). At 1% unheated SCs. These values differ from other reported n (0.56) and
protein, both unheated and heat-set SCs were almost Newtonian in K (110 mPa sn) values of unheated SCs of WPI and enzymatically-
nature (n > 0.98) with nearly identical flow profiles (Fig. 3). At 4e6% modified HMP formed at pH 6 (Lutz, Aserin, Portnoy, Gottlieb, &
protein, the heat-setting treatment resulted in a reduction in vis- Garti, 2009). The higher viscosity and greater degree of pseudo-
cosity over the measured shear rate range (0.1e200 s1). The drop plasticity found by Lutz et al. (2009) could be attributed to varia-
tions in protein composition, pectin modification, pH, or to
differences in complex formation technique.
Table 4
Consistency and flow behavior coefficients of heat-set WPIeHMP SC dispersions.
3.5. Application for soluble complexes in beverages
Protein (%) K (mPa sn) n R2

unheated SCs 1 1.90 f 0.98 a 0.98 As mentioned previously, one application of interest for heat-set
4 20.3 e 0.91 bc 0.99 SCs is in beverages with the benefit of increased thermal stability
5 39.5 c 0.89 cd 0.99 near the pI of whey proteins, where thermal processing produces
6 79.0 a 0.85 e 0.99
precipitates or gels (Wagoner et al., 2015). If thermal processing
heat-set SCs 1 1.92 f 0.99 a 0.95 produces sols, the primary concern for this application would be
4 16.8 e 0.92 b 0.99
colloidal stability, which would be predicted by a small particle size
5 31.9 d 0.87 de 0.99
6 68.0 b 0.86 e 0.98 per Stokes’ Law. This could be accomplished by combining heat-
setting and commercial sterility in one heating process, or using
Power law parameters K (consistency coefficient) and n (flow behavior coefficient)
of solutions containing unheated and heat-set WPIeHMP SCs as affected by protein
heat-set SCs as an ingredient prior to final formulation and thermal
concentration and heat-setting. Letters indicate significant differences among heating (Wagoner, Vardhanabhuti, & Foegeding, 2016). For the
means within a column as determined by Tukey's HSD at a ¼ 0.05. former option, small particles are only formed at a narrow pH range
136 T.B. Wagoner, E.A. Foegeding / Food Hydrocolloids 63 (2017) 130e138

close to the protein pI (Turgeon et al., 2007). With whey proteins Table 5
and pectin, this corresponds to pH 4.5e5.25 due to strong elec- Particle size characteristics of heat-set WPI-HMP soluble complexes after pH
adjustment and UHT thermal processing.
trostatic attraction and the tendency for polymers to remain
interacting during heating (Jones & McClements, 2011). For the pH D50 (nm) D90 (nm) D[4,3] (nm)
latter option, heat-setting may be a method of expanding the op- Control 164 ± 2c 666 ± 9 c 273 ± 2 b
timum range of pH and ionic strength (Jones & McClements, 2008; 3 423 ± 9a 8520 ± 1140 a 2860 ± 627 a
Krzeminski et al., 2014). Additionally, the ratio of protein to poly- 4 173 ± 7c 746 ± 51 c 303 ± 18 b
6 293 ± 61 b 1940 ± 560 b 712 ± 147 b
saccharide could be altered, which has been shown to alter func-
7 359 ± 74 ab 2310 ± 510 b 829 ± 158 b
tionality of SCs in emulsions (Li, Fang, Al-Assaf, Phillips, & Jiang,
Complexes were formed at pH 5 with 4% WPI and 8:1 ratio of protein to pectin,
2012). To evaluate the effects of pH on thermal stability of particles,
adjusted to pH 3, 4, 6 or 7, and thermally processed via UHT conditions (6 s at
dispersions of heat-set SCs (4% protein) formed at pH 5 were 141  C). The control represents SCs that were heat-set at pH 5 but not UHT pro-
treated as model beverages and adjusted to pH 3, 4, 6, or 7 prior to a cessed. D50 and D90 represent the diameter that 50% or 90% of particles are smaller
UHT thermal process similar to commercial sterilization of near- than. D[4,3] values represent the volume mean average. Values are given ± standard
neutral pH protein beverages. deviation. Letters indicate significant differences among means within a column as
determined by Tukey's HSD at a ¼ 0.05.
The particle size distribution of the model beverages after UHT
thermal processing is shown in Fig. 4. The control represents the
heat-set SCs at pH 5 that were not UHT processed. The size distri- suggests aggregation of primary particles concomitant with the
bution of the pH 4 model beverage was similar to that of the un- lower magnitude of surface charge at pH 3 (Jones et al., 2009, 2010).
heat processed control, with a D[4,3] of 303 nm compared to Galacturonic acid has a pKa near pH 3, so the surface of the SCs
273 nm for the control, suggesting a limited loss of stability would be less negatively charged at pH z pKa per the core shell
(Table 5). Stability at pH 4 is significant to beverage processing, as model (Whistler & BeMiller, 1993). This loss of electrostatic repul-
beverages at pH values of 4.6 require a less severe heat process to sion would induce flocculation and/or secondary aggregation of SCs
produce commercial sterility. At pH 6 and 7, the size distribution during UHT processing. Additionally, GMP has a pI < 3.8 (Bernal &
shifted towards larger sizes, with D[4,3] values 2e3 times greater Jelen, 1985), and protonation at acidic pH could facilitate aggrega-
than the control. Moreover, the pH 7 model beverage trended to- tion among SCs. The formation of particles >1 mm indicates that SCs
wards slightly higher particle sizes than the beverage at pH 6, would not provide any benefit at pH 3, especially when considering
although the means were not statistically distinct. Sedimentation whey proteins alone and in beverages typically exhibit high ther-
was not observed in these two treatments over the course of several mal stability in this pH range (Wagoner et al., 2015).
weeks, suggesting at least moderate stability; however, this would
need to be more rigorously evaluated. Previous reports indicate 4. Conclusions
that heat-set SCs (belg and HMP) are stable to pH adjustments
from pH 3e7 (Jones, Decker, & McClements, 2010). This would This study has shown that whey protein isolate and high-
indicate that the change in size distribution observed in this study methoxyl pectin can be used to generate soluble complexes at
is due to the UHT process, not the pH adjustment step. concentrations up to 6% protein, which exceed FDA requirements
The beverage adjusted to pH 3, conversely, shifted towards a for a “good source” label claim. The average size of soluble com-
bimodal distribution with peaks near 400 nm and 9 mm plexes increased as protein concentration was increased. Soluble
(D[4,3] ¼ 2.86 mm) and sedimentation occurred rapidly. This complexes can be heat-set at temperatures above the protein
denaturation temperature. Heat-setting was not detrimental to
stability of the soluble complexes, and moreover, was associated
with a transition to a narrower size distribution, reduced hydro-
dynamic volume, greater magnitude of surface charge, and a
reduction in apparent viscosity of the dispersion. These trends
would favor colloidal stability via a slower sedimentation rate per
Stokes’ Law and strong electrostatic repulsion. A model beverage at
pH 4 containing heat-set SCs at 4% w/v protein can be UHT pro-
cessed without a significant impact on particle size.

Acknowledgements

Support from the North Carolina Agricultural Research Service


and Glanbia Nutritionals is gratefully acknowledged. The use of
trade names in this publication does not imply endorsement by the
North Carolina Agricultural Search Service of the products named
nor criticism of similar ones not mentioned. The authors are
grateful for whey protein isolate and high-methoxyl pectin pro-
vided by Glanbia Nutritionals and CP Kelco, respectively. Assistance
and comments from Chris Daubert are gratefully appreciated.

References

Ako, K., Nicolai, T., Durand, D., & Brotons, G. (2009). Micro-phase separation ex-
plains the abrupt structural change of denatured globular protein gels on
Fig. 4. Size distribution of heat-set whey protein isolate-high methoxyl pectin soluble varying the ionic strength or the pH. Soft Matter, 5(20), 4033e4041. http://
complexes (formed at pH 5 with 4% WPI and 8:1 ratio of protein to pectin). Solutions dx.doi.org/10.1039/B906860K.
were adjusted to pH 3, 4, 6 or 7 and thermally processed via UHT conditions (6 s at ez, G. D., Moro, A., Ballerini, G. A., Busti, P. A., & Delorenzi, N. J. (2011). Comparison
Ba
141  C). The control represents SCs that were heat-set at pH 5 but not UHT processed. between structural changes of heat-treated and transglutaminase cross-linked
T.B. Wagoner, E.A. Foegeding / Food Hydrocolloids 63 (2017) 130e138 137

beta-lactoglobulin and their effects on foaming properties. Food Hydrocolloids, proteinepolysaccharide complexes. Advances in Colloid and Interface Science,
25(7), 1758e1765. http://dx.doi.org/10.1016/j.foodhyd.2011.02.033. 167(1e2), 49e62. http://dx.doi.org/10.1016/j.cis.2010.10.006.
Beecher, J. W., Drake, M. A., Luck, P. J., & Foegeding, E. A. (2008). Factors regulating Kazmierski, M., Wicker, L., & Corredig, M. (2003). Interactions of b-lactoglobulin and
astringency of whey protein beverages. Journal of Dairy Science, 91(7), high-methoxyl pectins in acidified systems. Journal of Food Science, 68(5),
2553e2560. http://dx.doi.org/10.3168/jds.2008-1083. 1673e1679. http://dx.doi.org/10.1111/j.1365-2621.2003.tb12312.x.
Benichou, A., Aserin, A., Lutz, R., & Garti, N. (2007). Formation and characterization Kragh, A. M. (1961). Viscosity. In P. Alexander, & R. J. Block (Eds.), Analytical methods
of amphiphilic conjugates of whey protein isolate (WPI)/xanthan to improve of protein chemistry (Vol. 3). New York: Pergamon Press.
surface activity. Food Hydrocolloids, 21(3), 379e391. http://dx.doi.org/10.1016/ de Kruif, C. G., Weinbreck, F., & de Vries, R. (2004). Complex coacervation of proteins
j.foodhyd.2006.04.013. and anionic polysaccharides. Current Opinion in Colloid & Interface Science, 9(5),
Bernal, V., & Jelen, P. (1985). Thermal stability of whey proteins e A calorimetric 340e349. http://dx.doi.org/10.1016/j.cocis.2004.09.006.
study. Journal of Dairy Science, 68(11), 2847e2852. http://dx.doi.org/10.3168/ Krzeminski, A., Prell, K. A., Weiss, J., & Hinrichs, J. (2014). Environmental response of
jds.S0022-0302(85)81177-2. pectin-stabilized whey protein aggregates. Food Hydrocolloids, 35, 332e340.
Bromley, E. H. C., Krebs, M. R. H., & Donald, A. M. (2006). Mechanisms of structure http://dx.doi.org/10.1016/j.foodhyd.2013.06.014.
formation in particulate gels of b-lactoglobulin formed near the isoelectric Langton, M., & Hermansson, A. M. (1992). Fine-stranded and particulate gels of b-
point. The European Physical Journal E, 21(2), 145e152. http://dx.doi.org/10.1140/ lactoglobulin and whey protein at varying pH. Food Hydrocolloids, 5(6),
epje/i2006-10055-7. 523e539. http://dx.doi.org/10.1016/S0268-005X(09)80122-7.
Champagne, C. P., & Fustier, P. (2007). Microencapsulation for the improved delivery Li, X., Fang, Y., Al-Assaf, S., Phillips, G. O., & Jiang, F. (2012). Complexation of bovine
of bioactive compounds into foods. Current Opinion in Biotechnology, 18(2), serum albumin and sugar beet pectin: Stabilising oil-in-water emulsions.
184e190. http://dx.doi.org/10.1016/j.copbio.2007.03.001. Journal of Colloid and Interface Science, 388(1), 103e111. http://dx.doi.org/
Chanasattru, W., Jones, O. G., Decker, E. A., & McClements, D. J. (2009). Impact of 10.1016/j.jcis.2012.08.018.
cosolvents on formation and properties of biopolymer nanoparticles formed by Liu, H., Xu, X. M., & Guo, S. D. (2007). Rheological, texture and sensory properties of
heat treatment of b-lactoglobulinepectin complexes. Food Hydrocolloids, 23(8), low-fat mayonnaise with different fat mimetics. LWT - Food Science and Tech-
2450e2457. http://dx.doi.org/10.1016/j.foodhyd.2009.07.003. nology, 40(6), 946e954. http://dx.doi.org/10.1016/j.lwt.2006.11.007.
Chen, D. T. N., Wen, Q., Janmey, P. A., Crocker, J. C., & Yodh, A. G. (2010). Rheology of Lopes da Silva, J. A., & Rao, M. A. (2006). Pectins: Structure, functionality, and uses.
soft materials. Annual Review of Condensed Matter Physics, 1(1), 301e322. http:// In Food polysaccharides and their applications (2nd ed.). Boca Raton: CRC Press.
dx.doi.org/10.1146/annurev-conmatphys-070909-104120. Lutz, R., Aserin, A., Portnoy, Y., Gottlieb, M., & Garti, N. (2009). On the confocal
Code of Federal Regulations. (2016). Food Labeling. title 21, sec. 101.54 http://www. images and the rheology of whey protein isolated and modified pectins asso-
ecfr.gov Accessed 08.12.15. ciated complex. Colloids and Surfaces B: Biointerfaces, 69(1), 43e50. http://
Cooper, C. L., Dubin, P. L., Kayitmazer, A. B., & Turksen, S. (2005). Poly- dx.doi.org/10.1016/j.colsurfb.2008.10.011.
electrolyteeprotein complexes. Current Opinion in Colloid & Interface Science, Murphy, R., Cho, Y. H., Farkas, B., & Jones, O. G. (2015). Control of thermal fabrication
10(1e2), 52e78. http://dx.doi.org/10.1016/j.cocis.2005.05.007. and size of b-lactoglobulin-based microgels and their potential applications.
Dickinson, E. (1992). An introduction to food colloids. Oxford University Press. Journal of Colloid and Interface Science, 447, 182e190. http://dx.doi.org/10.1016/
Evans, J., Zulewska, J., Newbold, M., Drake, M. A., & Barbano, D. M. (2010). Com- j.jcis.2014.09.067.
parison of composition and sensory properties of 80% whey protein and milk Neirynck, N., van der Meeren, P., Lukaszewicz-Lausecker, M., Cocquyt, J.,
serum protein concentrates. Journal of Dairy Science, 93(5), 1824e1843. http:// Verbeken, D., & Dewettinck, K. (2007). Influence of pH and biopolymer ratio on
dx.doi.org/10.3168/jds.2009-2723. whey proteinepectin interactions in aqueous solutions and in O/W emulsions.
Farrell, H. M., Jr., Jimenez-Flores, R., Bleck, G. T., Brown, E. M., Butler, J. E., Colloids and Surfaces A: Physicochemical and Engineering Aspects, 298(1e2),
Creamer, L. K., et al. (2004). Nomenclature of the proteins of cows' milkdSixth 99e107. http://dx.doi.org/10.1016/j.colsurfa.2006.12.001.
revision. Journal of Dairy Science, 87(6), 1641e1674. http://dx.doi.org/10.3168/ Nilsson, M., Holst, J. J., & Bjo €rck, I. M. (2007). Metabolic effects of amino acid
jds.S0022-0302(04)73319-6. mixtures and whey protein in healthy subjects: Studies using glucose-
Fishman, M. L., Gillespie, D. T., Sondney, S. M., & El-Atawy, Y. S. (1991). Intrinsic equivalent drinks. The American Journal of Clinical Nutrition, 85(4), 996e1004.
viscosity and molecular weight of pectin components. Carbohydrate Research, Pals, D. T. F., & Hermans, J. J. (1952). Sodium salts of pectin and of carboxy methyl
215(1), 91e104. http://dx.doi.org/10.1016/0008-6215(91)84010-C. cellulose in aqueous sodium chloride. II Osmotic pressures. Recueil Des Travaux
Gente s, M. C., St-Gelais, D., & Turgeon, S. L. (2010). Stabilization of whey protein Chimiques Des Pays-Bas, 71(5), 458e467. http://dx.doi.org/10.1002/
isolatepectin complexes by heat. Journal of Agricultural and Food Chemistry, recl.19520710505.
58(11), 7051e7058. http://dx.doi.org/10.1021/jf100957b. Peinado, I., Lesmes, U., Andre s, A., & McClements, D. J. (2010). Fabrication and
Girard, M., Turgeon, S. L., & Gauthier, S. F. (2002). Interbiopolymer complexing morphological characterization of biopolymer particles formed by electrostatic
between b-lactoglobulin and low- and high-methylated pectin measured by complexation of heat treated lactoferrin and anionic polysaccharides. Langmuir,
potentiometric titration and ultrafiltration. Food Hydrocolloids, 16(6), 585e591. 26(12), 9827e9834. http://dx.doi.org/10.1021/la1001013.
http://dx.doi.org/10.1016/S0268-005X(02)00020-6. Pereyra, R., Schmidt, K. A., & Wicker, L. (1997). Interaction and stabilization of
Hector, A. J., Marcotte, G. R., Churchward-Venne, T. A., Murphy, C. H., Breen, L., von acidified casein dispersions with low and high methoxyl pectins. Journal of
Allmen, M., et al. (2015). Whey protein supplementation preserves postprandial Agricultural and Food Chemistry, 45(9), 3448e3451. http://dx.doi.org/10.1021/
myofibrillar protein synthesis during short-term energy restriction in over- jf970198f.
weight and obese adults. The Journal of Nutrition, 145(2), 246e252. http:// Ross-Murphy, S. B. (1994). Physical techniques for the study of food biopolymers (1st
dx.doi.org/10.3945/jn.114.200832. ed.). London: Blackie Academic & Professional.
Hirt, S., & Jones, O. G. (2014). Effects of chloride, thiocyanate and sulphate salts on b- Salminen, H., & Weiss, J. (2013). Effect of pectin type on association and pH stability
lactoglobulinepectin associative complexes. International Journal of Food Sci- of whey proteindpectin complexes. Food Biophysics, 9(1), 29e38. http://
ence & Technology, 49(11), 2391e2398. http://dx.doi.org/10.1111/ijfs.12560. dx.doi.org/10.1007/s11483-013-9314-3.
Huffman, L. M., & Harper, J. (1999). Maximizing the value of milk through separation Santipanichwong, R., Suphantharika, M., Weiss, J., & McClements, D. J. (2008). Core-
technologies. Journal of Dairy Science, 82(10), 2238e2244. http://dx.doi.org/ shell biopolymer nanoparticles produced by electrostatic deposition of beet
10.3168/jds.S0022-0302(99)75471-8. pectin onto heat-denatured beta-lactoglobulin aggregates. Journal of Food Sci-
Jacobs, C., Kayser, O., & Müller, R. H. (2000). Nanosuspensions as a new approach for ence, 73(6), N23eN30.
the formulation for the poorly soluble drug tarazepide. International Journal of Schmitt, C., & Turgeon, S. L. (2011). Protein/polysaccharide complexes and co-
Pharmaceutics, 196(2), 161e164. http://dx.doi.org/10.1016/S0378-5173(99) acervates in food systems. Advances in Colloid and Interface Science, 167(1e2),
00412-3. 63e70. http://dx.doi.org/10.1016/j.cis.2010.10.001.
Jones, O. G., Decker, E. A., & McClements, D. J. (2009). Formation of biopolymer Shekunov, B. Y., Chattopadhyay, P., Tong, H. H. Y., & Chow, A. H. L. (2007). Particle
particles by thermal treatment of b-lactoglobulinepectin complexes. Food Hy- size analysis in pharmaceutics: Principles, methods and applications. Pharma-
drocolloids, 23(5), 1312e1321. http://dx.doi.org/10.1016/j.foodhyd.2008.11.013. ceutical Research, 24(2), 203e227. http://dx.doi.org/10.1007/s11095-006-9146-7.
Jones, O., Decker, E. A., & McClements, D. J. (2010). Thermal analysis of b-lacto- € nsson, B. (2009). Polyelectrolyteeprotein complexation driven
da Silva, F. L. B., & Jo
globulin complexes with pectins or carrageenan for production of stable by charge regulation. Soft Matter, 5(15), 2862e2868. http://dx.doi.org/10.1039/
biopolymer particles. Food Hydrocolloids, 24(2e3), 239e248. http://dx.doi.org/ B902039J.
10.1016/j.foodhyd.2009.10.001. Sperber, B. L. H. M., Cohen Stuart, M. A., Schols, H. A., Voragen, A. G. J., & Norde, W.
Jones, O. G., Lesmes, U., Dubin, P., & McClements, D. J. (2010). Effect of poly- (2009). Binding of b-lactolobulin to pectins varying in their overall and local
saccharide charge on formation and properties of biopolymer nanoparticles charge density. Biomacromolecules, 10(12), 3246e3252. http://dx.doi.org/
created by heat treatment of b-lactoglobulinepectin complexes. Food Hydro- 10.1021/bm900812x.
colloids, 24(4), 374e383. http://dx.doi.org/10.1016/j.foodhyd.2009.11.003. Stone, A. K., & Nickerson, M. T. (2012). Formation and functionality of whey protein
Jones, O. G., & McClements, D. J. (2008). Stability of biopolymer particles formed by isolatee(kappa-, iota-, and lambda-type) carrageenan electrostatic complexes.
heat treatment of b-lactoglobulin/beet pectin electrostatic complexes. Food Food Hydrocolloids, 27(2), 271e277. http://dx.doi.org/10.1016/
Biophysics, 3(2), 191e197. http://dx.doi.org/10.1007/s11483-008-9068-5. j.foodhyd.2011.08.006.
Jones, O. G., & McClements, D. J. (2010). Biopolymer nanoparticles from heat-treated Swaisgood, H. E. (1982). Chemistry of milk protein. In Developments in dairy
electrostatic proteinepolysaccharide complexes: Factors affecting particle chemistry (Vol. 1, pp. 1e60). London, United Kingdom: Applied Science.
characteristics. Journal of Food Science, 75(2), N36eN43. http://dx.doi.org/ Tanford, C. (1961). Physical chemistry of macromolecules. Wiley & Sons.
10.1111/j.1750-3841.2009.01512.x. Thompson, J. L., Drake, M. A., Lopetcharat, K., & Yates, M. D. (2004). Preference
Jones, O. G., & McClements, D. J. (2011). Recent progress in biopolymer nanoparticle mapping of commercial chocolate milks. Journal of Food Science, 69(9),
and microparticle formation by heat-treating electrostatic S406eS413. http://dx.doi.org/10.1111/j.1365-2621.2004.tb09958.x.
138 T.B. Wagoner, E.A. Foegeding / Food Hydrocolloids 63 (2017) 130e138

Tolstoguzov, V. (2003). Some thermodynamic considerations in food formulation. Wagoner, T. B., Ward, L., & Foegeding, E. A. (2015). Using state diagrams for pre-
Food Hydrocolloids, 17(1), 1e23. http://dx.doi.org/10.1016/S0268-005X(01) dicting colloidal stability of whey protein beverages. Journal of Agricultural and
00111-4. Food Chemistry, 63(17), 4335e4344. http://dx.doi.org/10.1021/acs.jafc.5b00633.
Tromp, R. H., de Kruif, C. G., van Eijk, M., & Rolin, C. (2004). On the mechanism of Wang, X., Lee, J., Wang, Y. W., & Huang, Q. (2007). Composition and rheological
stabilisation of acidified milk drinks by pectin. Food Hydrocolloids, 18(4), properties of beta-lactoglobulin/pectin coacervates: Effects of salt concentra-
565e572. http://dx.doi.org/10.1016/j.foodhyd.2003.09.005. tion and initial protein/polysaccharide ratio. Biomacromolecules, 8(3), 992e997.
Turgeon, S. L., Schmitt, C., & Sanchez, C. (2007). Proteinepolysaccharide complexes http://dx.doi.org/10.1021/bm060902d.
and coacervates. Current Opinion in Colloid & Interface Science, 12(4e5), Weinbreck, F., Wientjes, R. H. W., Nieuwenhuijse, H., Robijn, G. W., & de Kruif, C. G.
166e178. http://dx.doi.org/10.1016/j.cocis.2007.07.007. (2004). Rheological properties of whey protein/gum arabic coacervates. Journal
Vardhanabhuti, B., & Foegeding, E. A. (1999). Rheological properties and charac- of Rheology, 48(6), 1215e1228. http://dx.doi.org/10.1122/1.1795191.
terization of polymerized whey protein isolates. Journal of Agricultural and Food Whistler, R. L., & BeMiller. (1993). Industrial gums: Polysaccharides and their de-
Chemistry, 47(9), 3649e3655. http://dx.doi.org/10.1021/jf981376n. rivatives. Academic Press.
Verheul, M., Roefs, S. P. F. M., & de Kruif, K. G. (1998). Kinetics of heat-induced Xu, R. (2000). Particle characterization: Light scattering methods. Dordrecht: Kluwer
aggregation of b-lactoglobulin. Journal of Agricultural and Food Chemistry, Academic.
46(3), 896e903. http://dx.doi.org/10.1021/jf970751t. Zafar, T. A., Waslien, C., AlRaefaei, A., Alrashidi, N., & Al Mahmoud, E. (2013). Whey
Wagoner, T., Vardhanabhuti, B., & Foegeding, E. A. (2016). Designing whey pro- protein sweetened beverages reduce glycemic and appetite responses and food
teinepolysaccharide particles for colloidal stability. Annual Review of Food Sci- intake in young females. Nutrition Research, 33(4), 303e310. http://dx.doi.org/
ence and Technology, 7(1), 93e116. http://dx.doi.org/10.1146/annurev-food- 10.1016/j.nutres.2013.01.008.
041715-033315.

You might also like