Вы находитесь на странице: 1из 8

Food Hydrocolloids 17 (2003) 399–406

www.elsevier.com/locate/foodhyd

Heat-induced gelation of whey proteins observed by rheology, atomic


force microscopy, and Raman scattering spectroscopy
Shinya Ikeda*
Department of Food and Human Health Sciences, Osaka City University, 3-3-138 Sugimoto, Sumiyoshi-ku, Osaka 558 8585, Japan
Received 19 July 2002; revised 21 November 2002; accepted 3 December 2002

Abstract
Rheological and structural transitions during heat-induced gelation of whey proteins were investigated using mechanical spectroscopy,
atomic force microscopy (AFM), and Raman scattering spectroscopy. b-Lactoglobulin aqueous solutions containing 0.1 mol/dm3 NaCl at
pH 7 exhibited solid-like mechanical spectra before gelation. Heating such a solution resulted in the formation of an opaque gel that exhibited
frequency independent tan d values, an indicative of self-similar network structures. Translucent gels were formed in the absence of added
salts at pH 7 and their tan d values were frequency dependent. AFM images of heat-induced gel precursors revealed that these aggregates
were composed of ellipsoidal primary particles, regardless of the concentration of added NaCl, confirming that aggregation occurs in two-
step: the formation of primary aggregates and the subsequent aggregation of the primary aggregates. The size of primary aggregates and the
rate of aggregation increased with increasing NaCl concentrations. Thus, transitions from translucent to opaque gels with increasing ionic
concentrations are likely to be caused mainly by kinetic effects without accompanying fundamental changes in the two-step aggregation
mechanism. Raman scattering spectroscopy allowed discrimination between these two gel types based on secondary structures in denatured
b-lactoglobulin molecules: a decrease in the a-helix structure content was more pronounced in translucent gels, while considerable fractions
of b-sheet structures remained in both types of gels. A significant involvement of hydrophobic interactions in the formation of opaque gels
was suggested by an intense band assigned to hydrophobic side chains.
q 2003 Elsevier Science Ltd. All rights reserved.
Keywords: Globular protein; Heat-denaturation; Sol–gel transition; Viscoelasticity; Atomic force microscopy; Raman scattering spectroscopy

1. Introduction amino acids and potentially functional minor proteins such


as lactoferrin (Sloan, 2002).
Whey protein ingredients such as whey protein concen- Heating an aqueous solution of globular protein induces
trates (WPCs) and isolates (WPIs) are mixtures of globular denaturation and aggregation of protein molecules and a
proteins found in bovine milk. Due to mass production of macroscopic gel can be formed at a sufficiently high protein
whey protein ingredients as byproducts from cheese concentration. A state transition from a sol to a gel is
manufacture, their effective utilization is a continuous prominently reflected in rheological transitions. Thus, a
demand in the industry. Extensive investigation has been number of gel-forming biopolymers including proteins and
conducted primarily on physical functions of whey proteins polysaccharides are utilized in the food industry for
such as gelling, foaming, emulsifying, and stabilizing structuring products and controlling texture. In the case of
abilities since physical properties of food products are key whey proteins, gel network structures and rheological
attributes that determine consumer acceptance (Szczesniak, properties of heat-induced gels can be controlled by varying
1998). Currently, physiological functions of whey proteins gelling conditions such as pH, the ionic strength, species of
are also attracting attention. Whey protein ingredients co-existing ions, and so forth (Clark, 1998). Whey protein
contain reasonable amounts of high-sulfur containing molecules, the isoelectric points of which are near pH 5, are
net negatively charged at neutral pH. If the ionic strength is
sufficiently low, intermolecular electrostatic repulsion is
* Tel.: þ81-6-6605-2862; fax: þ 81-6-6605-3086. dominant. Heat-induced aggregation proceeds relatively
E-mail address: ikeda@life.osaka-cu.ac.jp (S. Ikeda). slowly, leading to the formation of a so-called fine-stranded
0268-005X/03/$ - see front matter q 2003 Elsevier Science Ltd. All rights reserved.
doi:10.1016/S0268-005X(03)00033-X
400 S. Ikeda / Food Hydrocolloids 17 (2003) 399–406

gel, composed of finely stranded nanometer-thick networks, weak background scattering from water (Li-Chan, 1996). In
exhibiting transparent or translucent appearance and this study, Raman scattering spectroscopy was employed in
rubbery texture (Clark, 1998; Ikeda & Foegeding, 1999a; order to detect molecular structural changes accompanying
Ikeda, Foegeding, & Hagiwara, 1999; Ikeda & Morris, heat-induced gelation of whey proteins. Heat-induced gel
2002; Kavanagh, Clark, & Ross-Murphy, 2000b; Langton & precursors were visualized using AFM and their structures
Hermansson, 1992; Stading & Hermansson, 1991). Inter- were investigated. Obtained structural information was
molecular repulsion can be screened by shifting pH toward anticipated to give bases for understanding rheological
the isoelectric points or by increasing ionic strength. Heat- transitions occurring during the formation of varied types of
induced aggregation is accelerated in screened conditions heat-induced whey protein gel networks.
and results in the formation of a white opaque gel composed
of micrometer-sized particulate aggregates (Clark, 1998;
Langton & Hermansson, 1992; Stading & Hermansson, 2. Materials and methods
1991). Gels of this type, often called particulate gels, easily
release liquids and require relatively large stresses to be 2.1. Materials and preparation
deformed. (Ikeda & Foegeding, 1999a; Ikeda et al., 1999;
Stading & Hermansson, 1991). Thus, some phenomenolo- Three times crystallized b-lactoglobulin (a mixture of
gical rules have already been established with respect to the variant A and B, product no. L-0130) was purchased from
relationships between gelling conditions and structural and Sigma Chemicals (St Louis, MO). WPI was a Bi-Pro grade
physical features of resulting gels; however, relatively little product in the form of freeze-dried powders supplied by
knowledge has been accumulated regarding rheological Davisco Foods International (LeSuer, MN). Other chemi-
transitions occurring during gelation (Ikeda, Nishinari, & cals were of reagent grade quality. Protein samples were
Foegeding, 2001b). dissolved in distilled water or 0.1 – 0.3 mol/dm3 NaCl
Electron microscopy has played a key role in classifying aqueous solutions. The protein solutions were adjusted to
network structural types in heat-induced globular protein pH 7, 5.4, or 2 by adding small amounts of hydrochloric acid
gels (Clark, 1998; Kavanagh et al., 2000b; Langton & or sodium hydroxide. The concentration of b-lactoglobulin
Hermansson, 1992; Stading & Hermansson, 1991). Electron was determined spectrophotometrically at 278 nm using the
microscopy imaging, however, needs to be conducted under extinction coefficient e 278 ¼ 0:955 cm2 =mg (Bell &
a vacuum, requiring elaborate sample preparation, such as McKenzie, 1967) whenever needed.
dehydration, fixing, replication, or staining, in order to
preserve the natural state of the specimen. Alternatively, 2.2. Rheological measurements
hydrated biopolymer samples can be directly imaged in
atmospheric pressure with minimal preparation procedures Degassed sample solutions were placed between a cone
using atomic force microscopy (AFM) (Ikeda & Morris, (5 cm in diameter and 0.04 rad in cone angle) and plate test
2002; Ikeda, Morris, & Nishinari, 2001a; Morris, Kirby, & fixture of the instrument (RFSII, Rheometrics, Inc., NJ) pre-
Gunning, 1999). Samples to be visualized by AFM should set at 70 8C and immediately covered with paraffin oil to
be examined on a substrate that is rigid and molecularly flat. prevent water vaporization. Frequency v sweep measure-
Biopolymer samples are usually deposited onto freshly ments of the storage modulus G0 and the loss modulus G00
cleaved mica surfaces and scanned in air or under a liquid were performed during isothermal heating at 70 8C at either
with a probe that maintains constant contact force with the 15 or 30 s intervals in the frequency range of 1– 100 rad/s at
sample surface or constant separation distance between the a strain of 0.03, which was within the linear viscoelastic
sample and the probe (Kirby, Gunning, & Morris, 1996; strain region determined by preliminary experiments. The
Morris et al., 1999). AFM images are generated based on the examined frequency range was restricted in order to
vertical movements of the sample or the probe during complete each sweep within ca. 1 min. Values of the
scanning, depending on the design of the instruments. Thus, complex modulus Gp ðGp2 ¼ G02 þ G002 ) at a frequency of
an AFM image inherently contains quantitative information 1 rad/s and tan d ð¼ G00 =G0 ) at varied frequencies were
on heights of objects in the image. calculated and reported.
While AFM has been confirmed to be capable of
resolving even sub-domain structures in native globular 2.3. Atomic force microscopy
protein molecules (Morris et al., 1999), investigations on
even more detailed molecular structures usually require The sample solutions were heated in sealed Pyrex tubes
spectroscopy. Raman scattering spectroscopy is one of rare immersed in a hot water bath preset at 80 8C for pre-
spectroscopy methods that provide information on peptide specified times and diluted to protein concentrations of ca.
backbone conformations in protein molecules at high 40 – 44 mg/ml. Aliquots (2 ml) of the diluted sample
protein concentrations sufficient to cause gelation (Li-Chan, solutions were immediately spread onto freshly cleaved
1996). Another advantage of Raman scattering spectroscopy mica sheets and imaged under butanol. AFM images were
is that it is directly applicable to aqueous systems due to produced using an East Coast Scientific (Cambridge, UK)
S. Ikeda / Food Hydrocolloids 17 (2003) 399–406 401

manufactured AFM located at the Institute of Food


Research (Norwich, UK). Samples were placed on top of
the piezoelectric scanner, brought into contact with a
V-shaped Nanoprobe cantilever tip (Digital Instruments,
Santa Barbara, CA), and scanned at a preset cantilever
deflection. Topographical images were generated from
vertical movements of the sample during scanning necess-
ary to maintain the preset cantilever deflection. Equivalent
error signal mode images were generated based on the slight
fluctuations of the cantilever deflection around the preset
value during scanning.

2.4. Raman scattering spectroscopy

Heat-induced gels were formed by heating the sample


solutions in sealed containers in a water bath preset at 80 8C
for 60 min, cooled to room temperature, and kept in a 5 8C
cold room overnight. Raman spectra were recorded on a
Raman microscope (System 1000, Renishaw plc, Old Town, Fig. 1. Gp (solid square) and tan d developments in 7% w/w b-lactoglobulin
UK) at room temperature (25 8C) with excitation from the aqueous solution containing 0.1 mol/dm3 NaCl at pH 7 during isothermal
782 nm line of a titanium sapphire crystal laser (Mira model heating at 70 8C. Values of tan d were determined at 1 rad/s (open circle),
2.5 rad/s (solid triangle), 25 rad/s (open square), and 40 rad/s (solid circle).
900-P, Coherent Inc., Santa Clara, CA). A drop (50 ml) of
the sample solution or a gel sample that was cut into a disk
(3 mm in diameter and 2 mm in height) was placed on a with the beginning of increasing Gp values. This point
quartz plate. The laser was focused through a £ 50 was thus considered to be the gelation point. The tan d
objective at a power of 50 mW on the sample surface and value was always less than unity even before the gelation
back-scattering was collected with the same objective and point, suggesting that the sol before gelation behaves as
captured using a Peltier cooled charge-coupled-device if it were a solid against linearly small strains. At the
(CCD) array detector. The accuracy of the wavenumber gelation point, tan d values determined at varied
was checked daily using the 520 cm21 band of silicon. Each frequencies converged and remained frequency indepen-
spectrum was smoothed with the 15 points fifth degree dent until ca. 5000 s. Following equations were satisfied
Savitsky-Golay function using the Grams/386 software when tan d values were independent of the frequency
(Galactic Industries Corporation, Salem, NH).
G0 , G00 , vn ð1Þ
G00 =G0 ¼ tanðnp=2Þ ð2Þ
3. Results and discussion
These equations have been proposed originally for chemi-
3.1. Rheological measurements cally crosslinking polymeric systems and claimed to be
valid only at the gelation threshold (Chambon & Winter,
Certain polysaccharides such as agarose, i- or 1987; Winter & Chambon, 1986). Such a gel at the gelation
k-carrageenan, and gellan gum can form a gel on point, referred to as a critical gel, is known to possess self-
cooling due to the conformational transition from random similar network structures (Muthukumar, 1989). The value
coils to double-helices. Such gelation process can be of the exponent n in Eq. (1) can be related to the fractal
monitored rheologically: the mechanical spectrum that is dimension of the self-similar network and the validity of
the frequency v dependence of the storage G0 and loss self-similarity in a wide length scale is guaranteed only
moduli G00 exhibits features characteristic of dilute or when both equations are satisfied simultaneously (Chambon
semi-dilute solutions before gelation and transforms to a & Winter, 1987; Winter & Chambon, 1986). Therefore, the
so-called gel spectrum during gelation (Hossain, Nemoto, validity of Eqs. (1) and (2) for heat-induced b-lactoglobulin
& Nishinari, 1997). However, this was not the case with gels formed in the presence of 0.1 mol/dm3 NaCl suggests
heat-induced gelation of whey proteins. Fig. 1 shows the that backbone structures that transmit the load on the gel
time evolution of rheological parameters in a 7% w/w were self-similar. This observation seems to be in line with
b-lactoglobulin aqueous solution containing 0.1 mol/dm3 the results of previous light-scattering studies conducted at
NaCl at pH 7, induced by isothermally heating at 70 8C. lower protein concentrations: b-lactoglobulin has been
The tan d value gradually decreased with time at first, shown to form self-similar aggregates at the ionic strength
followed by slight increase after ca. 2000 s heating, and of 0.1 mol/dm3 (Aymard, Gimel, Nicolai, & Durand, 1996;
then started decreasing rapidly at ca. 2400 s, coinciding Gimel, Durand, & Nicolai, 1994). Prolonged heating, longer
402 S. Ikeda / Food Hydrocolloids 17 (2003) 399–406

than ca. 5000 s, expanded differences among tan d values


determined at varied frequencies, suggesting a loss of self-
similarity due to restructuring.
The combined content of b-lactoglobulin and the
second major whey protein a-lactalbumin usually
exceeds 80% w/w of total solids in WPI. Gelation
phenomena of WPI are dominated by b-lactoglobulin
since the heat-stability of a-lactalbumin is even superior
(Kavanagh, Clark, Gosal, & Ross-Murphy, 2000a). Fig. 2
shows the time evolution of rheological parameters
during heating a 7% w/w WPI aqueous solution contain-
ing 0.1 mol/dm3 NaCl at pH 7. Although experimental
conditions were the same as those used for gaining
results shown in Fig. 1, except for using WPI in place
of b-lactoglobulin, dynamic mechanical response was
too low to be detected in an early stage of gelation
(ca. , 1500 s). Slightly before gelation, sufficient mech-
anical response was recovered but solid-like mechanical
Fig. 3. Gp (solid square) and tan d developments in 14% w/w b-
characteristics were found to be lost ðtan d . 1Þ: lactoglobulin aqueous solution at pH 7 during isothermal heating at 70 8C
Frequency independent tan d was observed immediately in the absence of added salt. Values of tan d were determined at 1 rad/s
after the gelation point. However, no data satisfied both (open circle), 2.5 rad/s (solid triangle), 6.3 rad/s (open square), and 40 rad/s
Eqs. (1) and (2) at the same time. Thus, the presence of (solid circle).
minor whey proteins appeared to disturb the formation of
self-similar gel network structures of b-lactoglobulin. values were highly frequency dependent and critical gel-like
Whey protein dispersions usually form translucent gels mechanical characteristics (i.e. Eqs. (1) and (2)) were not
in the absence of added salt at neutral pH. Fig. 3 represents observed in the vicinity of the gelation point. Prolonged
the time evolution of rheological parameters during heating heating after gelation narrowed differences in tan d values
14% w/w b-lactoglobulin without added salt. The tan d determined at varied frequencies. At a late stage of heating,
value remained below unity at most frequencies at an early both Eqs. (1) and (2) were approximately satisfied within the
stage of heating (ca. , 1000 s) but increased beyond unity as range of experimental errors, indicating the formation of
the gelation point is approached, which is an indicative of self-similar gel networks. According to previous findings
the progressing formation of linear aggregates. The tan d based on light-scattering studies, fine-stranded b-lactoglo-
bulin aggregates possess self-similarity in the length scale
lager than the thickness of strands (Aymard, Nicolai,
Durand, & Clark, 1999).
A WPI dispersion without added salt exhibited tan d
values larger than unity before gelation (Fig. 4). The tan d
value appeared to have a maximum before gelation and
become frequency independent of the gelation point. These
mechanical transitions are similar to those typical of
chemically cross-linking polymer systems (Takahashi,
Yokoyama, Masuda, & Takigawa, 1994). Therefore, it
was considered that, in the heated WPI dispersion, linear
aggregation predominated first, resulting in an increase in
tan d; but effects of extensive branching formation led to a
rapid G0 growth overcoming the developments in G00 : It may
be another point of interest that the value of the exponent n
determined at the gelation point was 0.67, identical with an
ideal value predicted based on the percolation theory
(Takahashi et al., 1994).

3.2. Atomic force microscopy


Fig. 2. Gp (solid square) and tan d developments in 7% w/w WPI aqueous
solution containing 0.1 mol/dm3 NaCl at pH 7 during isothermal heating at
70 8C. Values of tan d were determined at 1 rad/s (open circle), 16 rad/s Complicated rheological transitions during heat-induced
(solid triangle), 25 rad/s (open square), and 40 rad/s (solid circle). gelation of whey proteins are considered to reflect structural
S. Ikeda / Food Hydrocolloids 17 (2003) 399–406 403

Fig. 5. Topographical (a) and equivalent error signal mode AFM image (b)
of WPI aggregates formed in 11% w/w WPI aqueous solution at pH 7 by
heating at 80 8C for 60 min in the absence of added salt. Image size is
5 mm £ 5 mm.

in the presence of other whey proteins. It is also likely that at


a high protein concentration the growth of primary
Fig. 4. Gp (solid square) and tan d developments in 14% w/w WPI aqueous aggregates is forced to cease by the secondary aggregation
solution at pH 7 during isothermal heating at 70 8C in the absence of added among growing primary aggregates.
salt. Values of tan d were determined at 1 rad/s (open circle), 2.5 rad/s In the presence of 0.1 mol/dm3 NaCl at pH 7, heating a
(solid triangle), 10 rad/s (open square), and 63 rad/s (solid circle).
10% w/w WPI solution at 80 8C for only a few minutes
normally results in the formation of an opaque gel (Ikeda &
developments in protein aggregates. Previous electron Foegeding, 1999b). Thus, gel precursors were formed by
microscopy images of translucent gels formed at neutral heating a 2% w/w solution at 80 8C for 30 min and imaged
pH without added salt have revealed that the gel networks using AFM (Fig. 6). The heated solution before dilution
are composed of monomer/dimer-thick fine-strands (Lang- appeared to be white opaque, indicating intensive aggrega-
ton & Hermansson, 1992; Stading & Hermansson, 1991). tion despite the much lower protein concentration, com-
On the other hand, scattering studies have revealed that pared to the case without added NaCl. Observed aggregates
aggregation at neutral pH proceeds in two steps (Aymard were once again composed of elementary globular aggre-
et al., 1996; Gimel et al., 1994). The proposed first step is gates. Measured heights of these primary particles were ca.
the formation of primary aggregates that are globular in 18 nm, larger than those formed in the absence of NaCl.
shape and thus designated as globules. Additionally, it has These results suggest that well-known transitions in gel
been found that the second step aggregation among globules network types from fine-stranded to particulate with
can be suppressed in a dilute system and that, in this case, increasing ionic strength at neutral pH do not involve
the size of globules is always ca. 30 nm in diameter, fundamental changes in the two-step aggregation mechan-
insensitive to the ionic strength in the gelling system ism but result from accelerated aggregation.
(0.003 – 0.1 mol/dm3). Therefore, heat-induced gel precur- Translucent fine-stranded gels of WPI can also be formed
sors were formed at relatively high protein concentrations in at acidic pH below the isoelectric points of the proteins.
this study and visualized using AFM.
Fig. 5 shows examples of WPI gel precursors formed in
an 11% w/w WPI solution at pH 7 without added salts by
heating at 80 8C for 60 min. Visualized aggregates were
polydisperse in size and shape although the observed shapes
of the aggregates are not necessarily representatives of their
three-dimensional shapes in a sol due to probable orien-
tation, flow, or extension occurring on spread on mica
surfaces. However, the aggregates were generally composed
of fairly regular globular particles, confirming that the heat-
induced gelation of WPI at neutral pH also occurs in two
steps. An estimated average height of individual primary
aggregates was 11 nm, much shorter than what is expected
Fig. 6. Topographical (a) and equivalent error signal mode AFM image (b)
from the results of previous scattering studies made for pure of WPI aggregates formed in 2% w/w WPI aqueous solution containing
b-lactoglobulin systems. These results may indicate the 0.1 mol/dm3 NaCl at pH 7 by heating at 80 8C for 30 min. Image size is
aggregation of b-lactoglobulin was more or less obstructed 1.5 mm £ 1.5 mm.
404 S. Ikeda / Food Hydrocolloids 17 (2003) 399–406

if each individual aggregate of whey proteins is composed


of individual or mixed species of globular proteins.

3.3. Raman scattering spectroscopy

Recent Fourier transform infrared (FT-IR) spectroscopy


studies on heat-induced gelation of b-lactoglobulin have
concluded that fine-stranded gels are formed by extensively
denatured protein molecules and that the formation of
particulate gels involves only minor modification of the
conformation of the protein (Lefèvre & Subirade, 2000).
Fig. 7. Topographical (a) and equivalent error signal mode AFM image (b) This model seems to be worth revisiting since the b-barrel
of WPI aggregates formed in 2% w/w WPI aqueous solution containing structure in b-lactoglobulin is known to be heat-resistant
0.1 mol/dm3 NaCl at pH 2 by heating at 80 8C for 180 min. Image size is
and the molecular size has been found to increase only ca.
3 mm £ 3 mm.
10% by heating at normal gelling temperatures in dilute
conditions (Panick, Malessa, & Winter, 1999). In the
Fig. 7 shows WPI gel precursors formed at pH 2 in the present study, Raman scattering spectroscopy was
presence of 0.1 mol/dmd3 NaCl by heating at 80 8C for employed to detect secondary structural changes accom-
180 min. NaCl was added in order to avoid possible partial panying heat-induced gelation of b-lactoglobulin. Com-
precipitation of protein molecules (Aymard et al., 1999), but parisons were made between two types of gel networks:
the heated solution remained clear to the eye. Fig. 7 reveals translucent fine-stranded b-lactoglobulin gels formed at pH
that finely stranded aggregates were formed in the heated 7 and 2 without added salt and opaque particulate gels
solution. The width and height of the fibrous aggregates formed at pH 7 with 0.1 or 0.3 mol/dm3 NaCl and at pH 5.4.
were fairly uniform. The measured heights were ca. 10 nm. The intensity of the band around 940 cm21 is considered
Since the radius of native b-lactoglobulin molecules is ca. to be proportional to the a-helix content (Clark, Saunderson,
1.9 nm in an aqueous solution (Aymard et al., 1999), 10 nm- & Suggett, 1981; Li-Chan, 1996). Recorded intensities of
thick strands cannot be just linearly connected monomers of this a-helix band were normalized using 1005 cm21 band
b-lactoglobulin. Previous rheological studies on b-lacto- intensities, according to the conventional procedure
globulin and a-lactalbumin blends (Kavanagh et al., 2000a) (Li-Chan, 1996), and reported in Fig. 8(a). Relatively
have concluded that these whey proteins can form coupled weak intensities of this band is perhaps due to the low a-
gel network at pH 3 and 7. Moreover, it is known in the field helix content of b-lactoglobulin: only about one tenth of the
of biomedical research, where a possible link between the total residues are involved in the a-helix structures
formation of fibrous protein deposits and physiological (Fogolari, Ragona, Zetta 1998). The intensity at pH 2 was
malfunctioning is an important subject, that various globular substantially low even before heating, which is probably
proteins with no structural similarities can form fibrous due to effects of protonation of carboxyl side groups and
aggregates when protein molecules are only partially ionization of amino side groups in the a-helix portion and/or
denatured and slowly aggregate (Guijarro, Sunde, Jones, unfolding of a-helix structures although b-lactoglobulin is
Campbell, & Dobson, 1998). Surface probe microscopy is known as an exceptionally acid-stable protein (Fogolari
anticipated to serve as a tool in future studies for clarifying et al., 1998). The band intensity decreased by heating at pH

Fig. 8. Effects of heat-induced gelation of 15% w/v b-lactoglobulin on Raman scattering peak intensity at 940 cm21 normalized using peak intensity at
1005 cm21 (a), and on peak wavenumber around 1240 cm21 (b) and 1665 cm21 (c). Open and solid columns represent unheated solutions and gels formed by
heating at 80 8C for 60 min, respectively. The numbers below pH values represent the NaCl concentration in mol/dm3.
S. Ikeda / Food Hydrocolloids 17 (2003) 399–406 405

7 with or without 0.1 mol/dm3 NaCl and at pH 2 but was not responsible for the two-step aggregation at neutral pH
totally lost. In the case of particulate gels formed at pH 5.4 (Aymard et al., 1999). Further Raman scattering spec-
or 7 with 0.3 mol/dm3 NaCl, the intensity remained almost troscopy studies are desirable since some Raman scattering
unchanged. The a-helix fraction in b-lactoglobulin is intensity bands are known to reflect conformational states of
known to totally unfold when heated to ca. 70 8C in a dilute disulfide bonds in protein molecules (Li-Chan, 1996).
solution (Qi Holt, Mc Nulty, Clarke, Brownlow, Jones
1997). Therefore, the results shown in Fig. 8(a) may indicate
that heat-resistance of a-helix structures can be improved at 4. Conclusion
a relatively high protein concentration. It is also possible
that non-aggregated and renatured molecules on cooling Heat-induced aggregation of b-lactoglobulin and WPI
contributed to the band intensity of heated samples. was confirmed to be a two-step process at neutral pH,
The Raman scattering intensity peak representing consisting of the formation of granular primary aggregates
b-sheet and disordered structures is expected to appear as the first step and the subsequent aggregation of these
near 1230 –1240 and 1245 –1270 cm21, respectively (Clark primary aggregates, regardless of the ionic concentration.
et al., 1981; Li-Chan, 1996). The peak position at ca. The growth and aggregation of primary particles were found
1242 cm 21 before heating (Fig. 8(b)) is typical of to be concurrent processes, which seems to cause rheolo-
predominant contributions from b-sheet structures and gical transitions during heat-induced gelation of whey
consistent with primarily b-sheet pleated structures of proteins quite different from those of ordinary gelling
b-lactoglobulin (Fogolari et al., 1998). This peak is known systems. Heat-induced gels formed by b-lactoglobulin
to shift to a higher wavenumber when b-sheet structures are alone in the presence of 0.1 mol/dm3 NaCl exhibited
unfolded (Frushour & Koenig, 1975) but remained almost at rheological features of self-similar networks, while the
the same position after heat-induced gelation except for the presence of other whey proteins disturbed aggregation of
case at pH 2 (Fig. 8(b)). Therefore, b-sheet structures b-lactoglobulin more or less, preventing from the formation
appeared not to unfold significantly during heat-induced of self-similar networks from WPI. In the absence of added
gelation. In the formation of fine-stranded gels at pH 2, this salt, an increase in tan d values is more pronounced at the
b-sheet peak shifted to a lower wavenumber, which is an early stage of gelation, which is an indicative of more linear
indicative of strongly hydrogen bonded b-sheets. Similar aggregation of primary particles. Heat-induced aggregation
peak shifts accompanying the formation of fine-stranded at acidic pH below the isoelectric point resulted in the
gels at acidic pH have been observed in the previous FT-IR formation of finely stranded aggregates with only a few
studies (Lefèvre & Subirade, 2000). molecular thickness. Secondary structural changes due to
The peak position around 1665 – 1672 cm21 reflects heat-induced gelation reflected more clearly transitions in
relative amounts of secondary structures: a-helix bands gel network types with an increase in the ionic concentration
exhibit a peak around 1660 cm21 and b-sheet and at neutral pH rather than the shift from two-step aggregation
disordered structure bands have a peak around 1670 cm21 at neutral pH to fine-stranded aggregation at acidic pH.
(Carew, Stanley, Seidel, & Gergely, 1983). Thus, the
general peak shift from ca. 1665 to 1672 cm21 due to heat-
induced gelation (Fig. 8(c)) suggests a general decrease in Acknowledgements
a-helical structures with an increase in disordered and/or b-
sheet structures. No obvious difference was recognized The author is grateful to Professor Katsuyoshi Nishinari
between different gel network types. of Osaka City University, Osaka, Japan and Professor
Particulate gels formed at pH 7 with 0.3 mol/dm3 NaCl E. Allen Foegeding of North Carolina State University,
and at pH 5.4 showed an intense and broad band centered Raleigh, USA for supporting rheological studies, Dr V. J.
around 1345 cm21 (not shown) that is considered to reflect Morris of Institute of Food Research, Norwich, UK for
significant changes in hydrophobic environments around instructing AFM, Professor E. C. Y. Li-Chan of The
aliphatic and aromatic side chains (Li-Chan, 1996). There- University of British Columbia, Vancouver, Canada for
fore, particulate gels are likely to be formed via hydro- instructing Raman scattering spectroscopy, and Professor
phobic interactions among only slightly denatured protein Shuryo Nakai of The University of British Columbia,
molecules. At the secondary structural level, fine-stranded Vancouver, Canada for valuable discussions on biophysical
gels can be characterized by a more prominent loss in functions of proteins.
a-helical structures than particulate gels but b-sheet
structures appeared to be preserved well in both types of
gels. Thus, the shift from two-step aggregation at neutral pH
References
to fine-stranded aggregation at acidic pH appeared not to
involve drastic changes in secondary structures. It is often Aymard, P., Gimel, J. C., Nicolai, T., & Durand, D. (1996). Experimental
pointed out that the involvement of the disulfide exchange evidence for a two-step process in the aggregation of b-lactoglobulin at
reaction that occurs at neutral pH but not at pH 2 should be pH 7. Journal de Chimie Physique, 93, 987 –997.
406 S. Ikeda / Food Hydrocolloids 17 (2003) 399–406

Aymard, P., Nicolai, T., Durand, D., & Clark, A. (1999). Static and Ikeda, S., Nishinari, K., & Foegeding, E. A. (2001b). Mechanical
dynamic scattering of b-lactoglobulin aggregates formed after heat- characterization of network formation during heat-induced gelation of
induced denaturation at pH 2. Macromolecules, 32, 2542– 2552. whey protein dispersions. Biopolymers, 56, 109 –119.
Bell, K., & McKenzie, H. A. (1967). The isolation and properties of bovine Kavanagh, G. M., Clark, A. H., Gosal, W. S., & Ross-Murphy, S. B.
b-lactoglobulin C. Biochimica et Biophysica Acta, 147, 109–122. (2000a). Heat-induced gelation of b-lactoglobulin/a-lactalbumin
Carew, E. B., Stanley, H. E., Seidel, J. C., & Gergely, J. (1983). Studies of blends at pH 3 and pH 7. Macromolecules, 33, 7029–7037.
myosin and its proteolytic fragments by laser Raman spectroscopy. Kavanagh, G. M., Clark, A. H., & Ross-Murphy, S. B. (2000b). Heat-
Biophysical Journal, 44, 219–224. induced gelation of globular proteins: part 3. Molecular studies on low
Chambon, F., & Winter, H. H. (1987). Linear viscoelasticity at the gel point pH b-lactoglobulin gels. International Journal of Biological Macro-
of a crosslinking PDMS with imbalance stoichiometry. Journal of molecules, 28, 41– 50.
Rheology, 31, 683 –697. Kirby, A. R., Gunning, A. P., & Morris, V. J. (1996). Imaging
Clark, A. H. (1998). Gelation of globular proteins. In S. E. Hill, D. A. polysaccharides by atomic force microscopy. Biopolymers, 38,
Ledward, & J. R. Mitchell (Eds.), Functional properties of food 355 –366.
macromolecules (2nd ed.) (pp. 77– 142). Gaithersburg: Aspen Publish- Langton, M., & Hermansson, A.-M. (1992). Fine-stranded and particulate
ers Inc. gels of b-lactoglobulin and whey protein at varying pH. Food
Clark, A. H., Saunderson, D. H. P., & Suggett, A. (1981). Infrared and laser- Hydrocolloids, 5, 523–539.
Raman spectroscopic studies of thermally-induced globular protein Lefèvre, T., & Subirade, M. (2000). Molecular differences in the formation
gels. International Journal of Peptide and Protein Research, 17, and structure of fine-stranded and particulate b-lactoglobulin gels.
353– 364. Biopolymers, 54, 578–586.
Fogolari, F., Ragona, L., Zetta, L., Romagnoli, S., De Kruif, K. G., & Li-Chan, E. C. Y. (1996). The applications of Raman spectroscopy in food
Molinari, H. (1998). Monomeric bovine b-lactoglobulin adopts a b-
science. Trends in Food Science and Technology, 7, 361–370.
barrel fold at pH 2. FEBS Letters, 436, 149 –154.
Morris, V. J., Kirby, A. R., & Gunning, A. P. (1999). Atomic force
Frushour, B. G., & Koenig, J. L. (1975). Raman studies of the crystalline,
microscopy for biologists. London: Imperial College Press.
solution, and alkaline-denatured states of b-lactoglobulin. Biopolymers,
Muthukumar, M. (1989). Screening effect on viscoelasticity near the gel
14, 649 –662.
point. Macromolecules, 22, 4656–4658.
Gimel, J.-C., Durand, D., & Nicolai, T. (1994). Structure and distribution of
Panick, G., Malessa, R., & Winter, R. (1999). Differences between
aggregates formed after heat-induced denaturation of globular proteins.
the pressure- and temperature-induced denaturation and aggregation
Macromolecules, 27, 583–589.
of b-lactoglobulin A, B, and AB monitored by FT-IR
Guijarro, J. I., Sunde, M., Jones, J. A., Campbell, I. D., & Dobson, C. M.
spectroscopy and small-angle X-ray scattering. Biochemistry, 38,
(1998). Amyloid fibril formation by an SH3 domain. Proceedings of the
National Academy of Science of the United States of America, 95, 6512–6519.
4224–4228. Qi, X. L., Holt, C., McNulty, D., Clarke, D. T., Brownlow, S., & Jones,
Hossain, K. S., Nemoto, N., & Nishinari, K. (1997). Dynamic viscoelas- G. R. (1997). Effect of temperature on the secondary structure of b-
ticity of iota carrageenan gelling system near sol–gel transition. Nihon lactoglobulin at pH 6.7, as determined by CD and IR spectroscopy: a
Reoroji Gakkaishi, 25, 135 –142. test of the molten globule hypothesis. Biochemical Journal, 324,
Ikeda, S., & Foegeding, E. A. (1999a). Effects of lecithin on thermally 341 –346.
induced whey protein isolate gels. Food Hydrocolloids, 13, 239 –244. Sloan, A. E. (2002). The top 10 functional food trends: the next generation.
Ikeda, S., & Foegeding, E. A. (1999b). Dynamic viscoelastic properties of Food Technology, 56, 32– 57.
thermally induced whey protein isolate gels with added lecithin. Food Stading, M., & Hermansson, A.-M. (1991). Large deformation properties of
Hydrocolloids, 13, 245–254. b-lactoglobulin gel structures. Food Hydrocolloids, 5, 339–352.
Ikeda, S., Foegeding, E. A., & Hagiwara, T. (1999). Rheological study on Szczesniak, A. S. (1998). Sensory texture profiling—historical and
the fractal nature of the protein gel structure. Langmuir, 15, scientific perspectives. Food Technology, 52, 54–57.
8584–8589. Takahashi, M., Yokoyama, K., Masuda, T., & Takigawa, T. (1994).
Ikeda, S., & Morris, V. J. (2002). Fine-stranded and particulate aggregates Dynamic viscoelasticity and critical exponents in sol–gel transition of
of heat-denatured whey proteins visualized by atomic force an end-linking polymer. Journal of Chemical and Physics, 101,
microscopy. Biomacromolecules, 3, 382– 389. 798 –804.
Ikeda, S., Morris, V. J., & Nishinari, K. (2001a). Microstructure of Winter, H. H., & Chambon, F. (1986). Analysis of linear viscoelasticity of a
aggregated and nonaggregated k-carrageenan helices visualized by crosslinking polymer at the gel point. Journal of Rheology, 30,
atomic force microscopy. Biomacromolecules, 2, 1331–1337. 367 –382.

Вам также может понравиться