Rheology of Protein Dispersions

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 29

RHEOLOGY OF PROTEIN DISPERSIONS

MARVIN A. TUNG
Food Science Department
University of British Columbia
Vancouver, British Columbia, V6T 1 W5, Canada
(Manuscript received December 27, 1977)

ABSTRACT

Utilization of proteins as new food products or as ingredients in


formulated foods depends on the protein's ability to provide desired
functional performance, in particular the ability to impart structure and
texture. This review emphasizes the relevance of rheological testing in
clarifying the physico-chemical nature of the protein which permits
prediction o f functional behavior under a variety of conditions, as
experienced in complex food systems. Emphasis has been placed on
studies of protein (soy, egg albumen, caseinate, single cell, whey, gluten,
rapeseed, gelatin) solutions and gels in which fundamental rheological
principles have been applied. Appropriate rheological techniques are
discussed as are factors which influence flow behavior since these pro-
vide the key to protein modification to impart desired functionality
and hence effective utilization.

INTRODUCTION

Proteins in various forms derived from conventional and nonconven-


tional sources are increasingly being utilized by the food industry. Their
attractiveness lies in their nutritional value and functional properties
which permit their use as ingredients t o impart desirable structure,
texture, flavor and color characteristics to formulated food products.
As specific functional characteristics of the proteins are required for
each food system, rheologmd properties of protein dispersions are of
considerable pertinence in the manufacture of protein foods, providing
criteria for protein application. Hermansson and akesson (1975a,b) and
Hermansson (1975a) obtained good correlations between flow behav-
ior-related functional properties of added proteins (soy protein isolate,
caseinate and whey protein concentrate) and water-binding properties
of a model meat system, as well as texture changes in commercial
meatballs. Sabater de Sabaths and Soler Nolla (1976)found the flow

Journal o f Texture Studies 9 (1978)3-31. All Rights Reserved


@Copyright 1978 b y Food & Nutrition Press, Inc., Westport, Connecticut 3
4 MARVIN A. TUNG

characteristics of emulsions to be highly correlated (r 2 0.98) with


sausage texture. Similarly, Hawkins (1974) concluded that a knowledge
of the rheological behavior of comminuted meat mixes was invaluable
in achieving a desired target meat product. Recently, Hsu
et al. (1977a,b) quantified the nutritional and functional properties of
protein ingredients and used computer techniques to determine formu-
lations which would provide model foods with specific nutritional and
functional characteristics.
Flow properties of protein dispersions are governed by composition,
as well as molecular shape, size and charge, which are influenced by
environmental conditions such as temperature, concentration, pH, ionic
strength and previous processing or treatment history. Rheological
parameters can be useful indices of structural changes in proteins. More-
over, information on the relationships between environmental condi-
tions and flow properties of various proteins can be used to modify
rheological behavior so as to impart the desired functionality.
Knowledge of flow properties of protein dispersions is also of prac-
tical significance in optimal process design for unit operations such as
pumping, mixing, heating, cooling, spray drying, and others. Rheolog-
ical measurements can be used for quality control prior to and during
the manufacturing process, as well as for the final product.
Rheological properties of protein dispersions are of vital importance
in spinning for texturization. Viscosity, related to the chain length or
degree of polymerization of the molecules and thus related to fiber
formation, is a critical factor. The highest viscosity short of gel forma-
tion provides the best spinning conditions (Thomson 1946). However,
viscosity alone may not be the critical factor, rather several rheological
properties as a whole are responsible for the spinnability of protein
dispersions (Huang and Rha 1971). It is necessary for a spinning dope
to have some minimum internal resistance to deformation in order to
remain as a distinct and continuous filament without complete disper-
sion or breakage after emerging from the spinnerette (Huang and
Rha 1971).

NON-NEWTONIAN FLOW

Since rheology is the study of the manner in which materials respond


to an applied stress or strain, the conditions of testing are of prime
importance. Most food materials are complex systems in which rheolog-
ical properties exhibit stress and/or strain dependence. For quality
control purposes the test conditions are relatively unimportant provid-
ing the same conditions are used in all tests. However, as pointed out by
RHEOLOGY OF PROTEIN DISPERSIONS 5

Hermansson (1975b), much of the published rheological data on


protein dispersions is of limited relevance since most of the data consist
of single point measurements and give no indication of flow behavior at
differing rates of shear or over time. In addition, the term “apparent
viscosity” has been misused in several studies since this term applies to
a value obtained when a non-Newtonian fluid is subjected to a constant
rate of shear. The instrument used in such studies (Brookfield Synchro-
Lectric viscometer) was equipped with T-bar or disk-type fixtures; the
rate of shear, thus, varied across the surface of the rotating fixture.
Equipped in this manner, such an instrument would be useful for the
measurement of Newtonian liquids in which viscosity is independent of
shear rate, but cannot provide meaningful information in terms of flow
behavior of more complex rheological systems. Data obtained in this
manner for non-Newtonian fluids should be referred to as “Brookfield
consistency”.

Shear Rate Thinning


Most protein dispersions exhibit pseudoplastic flow behavior in
which the fluid exhibits shear thinning over a wide range of shear rates.
Examples are: egg albumen (Tung etal. 1971), soy protein (Circle
et al. 1964; Hermansson 1975b), caseinate and whey protein concen-
trate (Hermansson 1975b), calcium co-precipitates with casein (Hayes
et d. 1969), rapeseed protein (Gill and Tung 1976), and single cell
protein concentrate (Huang and Rha 1971).
In polymer dispersions, such as protein systems, molecules are
randomly oriented and entangled when the fluid is at rest. This condi-
tion of random structure follows the tendency of systems in nature to
maximize entropy. The material behavior when sheared may be
imagined to change over different shearing conditions. As the shear rate
is increased, the asymmetric dispersed molecules tend to align them-
selves with the shear planes so that frictional resistance is reduced. In
this manner the random structure at ultralow shear rates gives way to
shear-oriented structure at higher shear rates. The progressively decreas-
ing resistance to flow at higher shear rates is observed as a decreasing
apparent viscosity. For any given pseudoplastic system it is reasonable
to expect that at some high shear rate, the macromolecules will be fully
aligned along the laminar shear planes and that no further streamlining
will be possible; this will be the limit of shear rate thinning for the
fluid. As the shear rate is further increased, shear stress will be propor-
tional to the shear rate.
Pseudoplastic behavior of highly solvated dispersed systems may be
explained by a slightly different mechanism. When undisturbed, the
6 MARVIN A. TUNG

dispersed molecules or particles influence several adjacent solvent


molecules by reducing their mobility and forming a solvated layer
structure. At very low shear rates there is little effect on the layered
structure and interactions would be constant as the aggregates are of
constant size. However, higher shear rates would progressively remove
the solvated layers, giving a reduced aggregate size and hence, a lower
apparent viscosity through the intermediate shear range. A t some high
shear rate the solvated layers would be completely removed, resulting in
a constant apparent viscosity at very high shear rates.
The pseudoplastic flow behavior of liquid egg albumen is shown in
Fig. 1. Measurements were made using a Haake Rotovisko RV1
viscometer equipped with narrow-gapped coaxial cylinder spindles and
cups (Tung et al. 1971). Equilibrium flow behavior data were fitted by
a form of the well known power-law model:

where q is the apparent viscosity, f is the shear rate and m and n


parameters are the consistency coefficient and the flow behavior index,
respectively. It is of interest to note that an increase in temperature
resulted in lower apparent viscosities and that egg albumen displayed
greater pseudoplasticity, as indicated by smaller flow behavior indices,
at higher temperatures (0.757, 0.704, 0.674 and 0.667 for 10, 20, 30
and 40°C, respectively).

Time Dependent Thinning


Some fluids exhibit time dependent flow that may produce decreas-
ing or increasing effects on shear stress and apparent viscosity with
time, at a constant shear rate and temperature. In some cases, this
change is reversible and the fluid will recover its original condition with
time at rest, or the change brought about by shearing may be irrevers-
ible. A thixotropic system will experience a reversible decrease in shear
stress and apparent viscosity at a constant shear rate and temperature.
Fluids of this type are thought to consist of asymmetric dispersed
particles or molecules that interact through adhesive or weak secondary
bonding forces to form a network or aggregated structure at rest. When
continuous flow is imposed on the system at constant shear rate, some
of the interparticle bonds break with time, thereby offering less resis-
tance to flow, and the shear stress relaxes to some constant value over a
period of time. Since the shear rate is constant during this stress relaxa-
tion period, this may also be interpreted as an apparent viscosity decay.
When the shearing treatment is stopped, the dispersed particles can
RHEOLOGY OF PROTEIN DISPERSIONS 7

3 -
I
500 1000 1500 2000 2500 3000
Sheor rote. I-’

FIG. 1. EFFECT OF TEMPERATURE ON THE FLOW BEHAVIOR OF


LIQUID EGG ALBUMEN (DATA FROM TUNG ETAL. 1 9 7 1 )

once again form the network, or aggregated structure, as Brownian


motion gradually restores the particles to positions where interparticle
adhesion can occur. Over a sufficient period of time the original density
of linkages is achieved, thus the process is reversible.
The major difference between the structural shear rate sensitivity of
thixotropic and pseudoplastic systems is in the time required to reach a
stable shear-induced condition, and to restore the original structure
when shear stresses are removed. This change and reversal is instan-
taneous in pseudoplastic fluids, whereas a relatively long time is re-
quired in thixotropic fluids. Hermansson (1975b) found that disper-
sions of soy protein isolate and whey protein concentrate (WPC) were
thixotropic.
In some fluids, the time dependent loss of structure is not recover-
able and fluids that experience a permanent loss of structure are called
rheodestructive. The loss of structure is envisaged in the manner
described earlier for thixotropy; however, there is no tendency to
reform the structural elements when shearing is stopped. Not all time
dependent shear thinning systems fall into one or other of the purely
thixotropic or totally rheodestructive categories. In some materials,
partial recovery of structure is noted. An example of this behavior is
shown in Fig. 2 for liquid egg albumen. The apparent viscosity at the
begmning of the curve determined 32 hr after the original test is higher
8 MARVIN A. TUNG

30 -

n
u 20-

-
A Inltiol test

I l After 32 hours

I 1 1 1
OO 10 20 30 40 50

FIG. 2. APPARENT VISCOSITY DECAY IN LIQUID EGG ALBUMEN


AT 10°C AND A SHEAR RATE OF 3140 s' TESTED BEFORE AND
AFTER 32 HR STORAGE AT 4'C (DATA FROM TUNG ET A L . 1971 )

than the apparent viscosity at equilibrium in the initial test. Thus, there
is an appreciable recovery of structure and the structural breakdown
mechanism in egg albumen at a constant shear rate is a combined thixo-
tropic-rheodestructive process (Tung et al. 1971).
An understanding of the rheological principles and the causes under-
lying the response of a food system to applied forces is necessary for
maximizing the benefits of rheology as applied to quality assurance,
process evaluation and product development. It is particularly im-
portant that the conditions of rheological testing reflect the parameters
(shear rate, pressure, temperature, time) encountered in product
application. Appropriate methods provide a complete rheological
description and, thus, a better understanding of the protein nature at
the molecular level. This, in turn, permits prediction of functional
behavior under a variety of conditions, as experienced in food systems,
while pointing the way to chemical or physical modifications which will
result in desired rheological properties.

WATER RELATIONS

Hermansson (1972, 1975b) and Hermansson and akesson (1975a)


RHEOLOGY OF PROTEIN DISPERSIONS 9

related the solubility and swelling properties of a soy protein isolate


(Promine-D from Central Soya Co., Inc., Chicago, IL), caseinate and
whey protein concentrate (WPC) t o their flow properties. Highly
soluble nonswelling proteins had low viscosity, which is the case for
many globular proteins such as those in WPC. Soluble proteins with
high initial swelling, such as caseinate, showed a highly concentration
dependent viscosity, probably reflecting the content of partially
solvated particles. Promine-D with its lower solubility but high limited
swelling exhibited high viscosity at relatively low concentration. When
relating these parameters to moisture loss in model meat systems,
Hermansson and Akesson (1975a) found solubility positively related to
moisture loss, whereas swelling and viscosity were negatively related.
Thus, Promine-D provided superior water binding properties in a food
system.
Fleming e t al. (1974) related the Brookfield consistency and water
absorption characteristics of slumes of sunflower and soybean flours,
concentrates and isolates. Water uptakes of the sunflower proteins were
lower than those of the soy proteins as was the consistency for disper-
sions of sunflower protein isolates.

CONCENTRATION DEPENDENCE

Environmental factors which affect protein structure and inter-


actions will influence the viscosity of their solutions. In extremely
dilute dispersions the total viscosity effect is the sum of the effects
caused by each of the individual suspended particles. However, as their
concentration is increased the disturbances of solvent flow produced by
the suspended particles are n o longer independent since flow patterns
overlap, and aggregation and solvent immobilization occur (Frisch and
Simha 1956).
The apparent viscosity of protein dispersions has been found to rise,
in most cases exponentially in line with rheological volume concentra-
tion theory (Frisch and Simha 1956), with increasing protein concentra-
tion; for example, soy proteins (Circle e t a l . 1964; Ehninger and
Pratt 1974; Hermansson 1975b), succinylated fish myofibrillar protein
( G r o n i n g e r 1 9 7 3 ) , single-cell protein concentrate (Huang and
Rha 1971), sunflower proteins (Fleming et al. 1974), caseinate (Hayes
a n d Muller 1961; Hermansson 1975b), whey protein concentrate
(Hermansson 1975b), rapeseed protein (Gill and Tung 1976) and
gelatin (Ward and Saunders 1958).
The effect of protein concentration on the flow properties of
10 MARVIN A. TUNG

Promine-D, caseinate and WPC at pH 7 in distilled water was studied by


Hermansson (1975b). She evaluated flow parameters for the power-law:

u=m+
and power-law plastic:

u = uy + m’+n’ (3)

constitutive equations, where u is the shear stress, uy is the yield stress,


+ is the shear rate and m, m’, n and n’ are parameters. Results of these
experiments are shown in Table 1. With increasing concentration of
Promine-D there was increasing deviation from Newtonian behavior and
at 8% concentration a yield value was apparent. The yield value in-
creased considerably as the protein concentration increased further.
Caseinate possessed strikingly different properties. Yield values were
not detected even at high concentrations and below 12%the dispersions
were almost Newtonian and of low viscosity. Above 12%, they were
slightly pseudoplastic and the consistency coefficient increased greatly
with increasing concentration. WPC was characterized as being of low
viscosity over a broad concentration range. The flow was almost
Newtonian in the range 4-1276, pseudoplastic in the range 14-1696,
while at higher concentrations yield values could be estimated and the
flow was plastic.
Viscosities of the three protein dispersions at a shear rate of 100 s-’
are presented in Fig. 3. These values were derived from the data of
Table 1 by calculating the shear stress (a) at ?; = 100 s-’ and defining
the apparent viscosity (77) as o/f at that point. From this figure it can
be seen that the apparent viscosities of Promine-D were higher than
those of caseinate and WPC over a broad concentration range. The
change in slope of the apparent viscosity curve at 10%protein concen-
tration reflects the large increase in the yield value relative to the vis-
cosity term at higher concentration. Hermansson (1975b) interpreted
these flow characteristics of Promine-D as strongly indicative of the
presence of protein-protein interactions leading to the formation of a
protein network at higher concentrations. Processing had greatly
influenced the rheological behavior as a ‘native’ soy protein isolate
(prepared under mild conditions on a pilot plant scale) was less pseudo-
plastic, had lower consistency coefficients and no yield values
(Hermansson 1975b).
The caseinate flow characteristics indicated that the increase in the
apparent viscosity was more likely due to protein-water interactions
( h y d r a t i o n ) t h a n t o strong protein-protein interaction forces
RHEOLOGY OF PROTEIN DISPERSIONS 11

Table 1. Effects of protein concentration on power-law and power-law


plastic flow model parameters for protein dispersions in distilled water at
25°C (data from Hermansson 1975b)
I
Protein Conc. OY -
m or m i
?LPe (%, d.b.) (dyne cm 2 , (dyne s" c m 2 ) nor n

Promine-D 4 0 0.11 0.91


6 0 0.33 0.85
8 7 1.5 0.76
10 48 7.8 0.68
12 194 11.0 0.67
14 597 25.0 0.64
16 1271 62.0 0.52
Caseinate 4 0 0,036 0.92
6 0 0.036 0.99
8 0 0.062 1.01
10 0 0.23 0.94
12 0 1.1 0.90
14 0 9.1 0.83
16 0 89.0 0.82
18 0 632.0 0.82
20 0 2862.0 0.90
WPC 4 0 0.027 0.94
6 0 0.030 0.97
8 0 0.034 0.98
10 0 0.059 0.95
12 0 0.079 0.93
14 0 0.24 0.87
16 0 0.78 0.77
18 8 1.1 0.77
20 21 3.9 0.67

(Hermansson 1975b). Although WPC dispersions exhibited the lowest


apparent viscosities in the whole concentration range (Fig. 3), the WPC
flow properties were more similar to those of Promine-D than to those
of caseinate.

pH DEPENDENCE

The effect of pH on the flow properties of protein dispersions is


marked (Table 2 and Fig. 4). Promine-D yield values and apparent
viscosities increased with increasing pH from 5.9 to 10.9
(Hermansson 1975b). This corresponds with the increasing solubility of
12 MARVIN A. TUNG

A 8 16 20
'2 0
Concentration. L

FIG. 3. EFFECT OF CONCENTRATION ON APPARENT VISCOSITY


OF PROTEIN DISPERSIONS IN DISTILLED WATER (T = 25OC,
+
APPARENT VISCOSITY DEFINED AS u/+ AT = 100 S' , DERIVED
FROM DATA OF HERMANSSON 1975b)

Table 2. Effects of pH on power-law and power-law plastic flow


model parameters for 12%protein dispersions in distilled water at 25°C
(data from Hermansson 1975b)
I
Protein OY
m or m
-
5Pe PH (dyne cm ) (dyne s" cm-' ) n o r n'

Promine-D 5.9 65 2.7 0.65


8.0 310 19.0 0.61
9.0 466 29.0 0.58
10.0 506 30.0 0.59
10.9 885 38.0 0.59
Caseinate 6.0 0 1.0 0.90
8.O 0 5.8 0.88
9.1 0 9.7 0.86
9.8 0 10.0 0.87
10.4 0 6.8 0.88
11.2 0 1.8 0.81
11.8 0 1.0 0.85
WPC 6.0 0 0.12 0.92
~n
v.w
n
v
A i C
U.10
n no
U.34

10.0 0 0.44 0.81


RHEOLOGY O F PROTEIN DISPERSIONS 13

1000

-A

:100
*
u
0
a
P
4

w PC
10 -

6 7 8 9 10 11 12
DH

FIG. 4 . EFFECT OF pH ON APPARENT VISCOSITY OF PROTEIN


DISPERSIONS IN DISTILLED .WATER (T = 25’C, APPARENT VIS-
COSITY DEFINED AS o/f AT y = 100 < I , DERIVED FROM DATA OF
HERMANSSON 1975b)

soy proteins as the pH diverges from their isoelectric point (“pH 4.5)
(van Megan 1974).Circle et al. (1964)had found only a slight increase
in Promine-D Brookfield consistency at pH 8 and 9 as compared to
pH 7. Hutton and Campbell (1977)reported an increase in Brookfield
consistency of Promosoy but a decrease for Promine-D as pH was
increased from 5 t o 7. It should be noted that both Circle et al. (1964)
and Hutton and Campbell (1977) used the T-spindles thus possibly
confounding their results.
Fleming et al. (1974),noting that the consistency of soy flours
increases markedly during alkaline protein isolation and that the pH
treatment may contribute to the high consistency of the final isolate,
adjusted soy and sunflower flour and concentrate to pH 1 2 and then
back to pH 6. This pH-activation process improved water absorption
properties and increased Brookfield consistency of most soy and sun-
flower products.

Soybean Proteins
The viscosity of soy protein dispersions appears to be at a maximum
point near pH 12.3 (Ishino and Okamoto 1975). Above 14.5% protein
concentration, pH 12 solutions gel (Kelly and Pressey 1966).Ishino and
14 MARVIN A. TUNC

Okamoto found that molecular interaction of alkali denatured soy


proteins occurred above 8% concentration at pH 12.3. Under these
conditions, a remarkable increase in solution Brookfield consistency, or
gelation, was observed after dialysis against phosphate buffer, pH 7.2,
apparently as a result of intermolecular bonding. Proteins treated below
pH 11.0 and at more than 8% concentration did not increase in con-
sistency after dialysis suggesting that soybean proteins change their
conformation above pH 11.0.Nondialyzed alkali denatured soy protein
had high Brookfield consistency below pH 10.4 which decreased
rapidly b e t d n pH 10.4 and 10.9 (Fig. 5). Between pH 11.2 and 12.2,
the Brookfield consistency increased slightly and then decreased again.

7 8 9 10 11 12 13
PH

FIG. 5. EFFECT OF pH ON BROOKFIELD CONSISTENCY OF 10.4%


ALKALI-DENATURED SOY PROTEIN DISPERSIONS AT 25OC BE-
FORE AND AFTER DIALYSIS AGAINST PHOSPHATE BUFFER,
pH 7.2 (DATA FROM ISHINO AND OKAMOTO 1975)

Correlating chemical and viscometric d a t a , Ishino and


Okamoto (1975) suggested that the high viscosity below pH 10.4 is
based on aggregation and on increased hydration of proteins while the
increase in viscosity between pH 11.2 and 12.2 may be caused by the
proteins unfolding, dissociating into subunits and interacting. Above
pH 12.9,there may be a destruction of exposed hydrophobic regions of
the protein. Kelly and Pressey (1966)found that the increased con-
sistency of soy protein exposed to high pH levels w a s accompanied by a
RHEOLOGY OF PROTEIN DISPERSIONS 15

rapid shift in the sedimentation constants of the 2, 7, 11 and 15s


ultracentrifuge components to essentially that of 3s.
The striking effects of elevated pH on soy protein dispersion vis-
cosity permits their utilization in fiber spinning. The protein slurries are
adjusted to pH 10-12 with alkali and the high viscosity spinning dope is
pumped through spinnerettes into an acid-salt bath where the proteins
coagulate into continuous filaments (Kinsella 1976). As a result of
alkali denaturation and the pressure and shearing forces of the spinning
process, the polypeptide chains unfold, uncoil and realign along the
long axes to form fibers by means of hydrogen, ionic and disulfide
bonds (Kelly and F'ressey 1966;Huang and Rha 1974).

Caseinate, Whey and Gelatin


The apparent viscosity of caseinate increases with pH from 6.0 to a
maximum a t 9.8-10.0 (Fig. 4) where it rapidly decreases
(Hermansson 197513). The greater effect of pH on the apparent vis-
cosity of caseinate as compared to Promine-D can be correlated with its
similarly marked effect on caseinate swelling (Hermansson 1972).
Hayes and Muller (1961)found the maximum apparent viscosity of
alkaline caseins at a slightly lower pH, in contrast to the results of Pun
et al. (1972) where the maximum was nearer pH 11. These latter
authors suggest that the increase in apparent viscosity is the result of
formation of increasing amounts of the largely hydrated caseinate ions
which begin to aggregate to form hydrated casein micelles.
Hermansson (1975b)found that the apparent viscosity of the WPC
dispersions increased only slightly with pH and the viscosity remained
low in the range studied (pH 6-10). Gelatin dispersions appear to have
maximum viscosities at pH 3 and 11 (Wardand Saunders 1958).

Single-Cell Proteins
In a study of single-cell protein concentrate (SCP)from Torula yeast,
Huang and Rha (1971)found concentrations of 10-25% protein, with
15-2076 being optimum, suitable for spinning. Maximum apparent
viscosity in the SCP dope was induced around pH 9 (Table 3). AU SCP
dopes were pseudoplastic except the dispersion of 15% concentration
and pH 9 which was essentially Newtonian. The authors noted that the
pH 12 dispersions developed gel-like aggregates at 20 and 25% concen-
tration and thus showed time dependent flow;results were presented
for comparative purposes only. Samples containing 10 and 15% SCP
had little or no yield values. For 20 and 25% protein dispersions, plastic
flow was evident with the yield values, consistency coefficients and
16 MARVIN A. TUNG

Table 3. Effects of protein concentration on power-law and power-law plastic


flow model parameters and apparent viscosity (f = 10 S 1 ) for single-cell protein
dispersions at 60°C (data from Huang and Rha 1971)

Protein aY -' m or m' I V'10d


(%I PH (dyne cm ) (dyne s" cm-2 ) n or n ICP)

10 8 0 0.12 0.67 5.6


9 0 0.082 0.73 4.4
10 0 0.11 0.65 4.9
12 0 0.033 0.85 2.3
15 Ba 0.7 1.8 0.50 64.0
ga 0.7(0)b O.ga(1.4) l.l(O.97) 130.0
10 0 0.24 0.78 14.0
12 0 0.17 0.53 5.8
20 Ba 3 2.6 0.67 150
9 7 21.0 0.61 930
10 1 0.34 0.81 32
12c 11 7.0 0.54 350
25 8 42 65.0 0.74 400
9 73 67.0 0.81 4400
1oa 24 17.0 0.60 920
12ac 49 59.0 0.68 540
a Air bubble entrapment
b Powerlaw parameters in parentheses
Gel-like aggregate formatiof
Defined as 014 at $ = 10 s-

flow behavior indices sensitive to pH changes. Huang and Rha (1971)


explain the increase in apparent viscosity up to pH 9 as probably being
the result of increased solubility, while at pH 10 depolymerization of
the peptide chains may have caused the apparent viscosity to decrease.

EFFECTS OF OTHER FACTORS

Ionic Strength
Ionic strength also has a marked effect on flow properties of protein
dispersions. Groninger (1973) found that sodium chloride greatly
reduced the consistency of succinylated fish myofibrillar protein dis-
persions (Table 4). Fleming et al. (1974)found that the effect of 5%
sodium chloride on the Brookfield consistency of 15% dispersions of
soy and sunflower proteins depended on the various products (flours,
concentrates, isolates) used. The viscosity of gelatin solutions decreased
RHEOLOGY OF PROTEIN DISPERSIONS 17

Table 4. Effect of sodium chloride on the


consistency of succinylated fish protein dis-
persionsa at 25OC (data from Groninger 1973)

Protein Salt Consistencyb


(%I (%) (CP)

1 0 1700
0.1 600
0.2 130
2 0 5500
0.1 2000
0.2 1150

aApproxirnately 509b of the +amino groups of


the Protein were reacted with anhydride
bMeasured with a Brookfield LVT at 80 rprn

in the presence of added sodium chloride (Ward and Saunders 1958).


Using 0.2, 0.5 and 1.OM NaCI, Hermansson (1975b) found that salt
increased the apparent viscosity of caseinate, had little effect on WPC,
and decreased the apparent viscosity of Promine-D dispersions up to
0.5M sodium chloride (Table 5). Interestingly, the rheological param-
eters for Promine-D showed a reversal in the trend between 0.5 and
1.OM NaCl. This reversal may correspond with the critical salt concen-
tration of 0.7M found necessary to solubilize soy proteins at pH 4.5
(van Megan 1974). Ehninger and F’ratt (1974) found that the effects of
0.1 and 0.2M NaCl on Promine-D and Promine-R consistency depended
on the protein concentration and the pH level between pH 5.5 and 6.5.
Promine-D swelling and solubility also decreased with increasing ionic
strength (Hermansson 1972; Hermansson and Akesson 1975b). The
change in flow properties may be attributed to the presence of less
swollen, more rigid aggregates and less solvated protein molecules
(Hermansson 1975b).

Other Bonding Modifiers


Many other agents affect the rheological properties of protein dis-
persions. Of particular interest are the effects of agents which disrupt
the quaternary structure by interfering with the protein’s ability to
form disulfide linkages or hydrogen bonds. Shibasaki etal. (1969)
studied soy protein systems and the effects of several factors on re-
duced viscosity (qred):
18 MARVIN A. TUNG

where q is the viscosity of the dispersion, q, is the solvent viscosity and


C is the protein concentration. They found that the reduced viscosity
of soy protein fractions increased linearly with urea concentration and
with time. However, initially there was a decrease in qred with
2-4Murea which the authors correlated with dissociation of the
protein into subunits using starch gel electrophoresis. At longer times or
with greater urea concentrations (6-8M),decreases in reduced viscosity
as a result of dissociation were overtaken by increasing qred resulting
from unfolding of the proteins. Addition of 2-mercaptoethanol rapidly
decreased qred. Results with alternate hydrogen bond and disulfide
bond cleaving agents (guanidine hydrochloride and cysteine, respec-
tively) were the same as those for urea and 2-mercaptoethanol.

Table 5. Effects of sodium chloride concentration on power-lawand power-law


plastic flow model parameters and apparent viscosity (9 = 100 8 )for protein
dispersions at 25OC (data from Hermansson 1975b)
I
Protein Conc. NaCl Conc. =Y - m or m , rlq1ooa
Type (% d.b.) (M) (dyne cm 2 , (dyne 8" cm-2) n or n (cp)

Promine-D 10 0 48 6.1 0.68 188


0.2 0 0.73 0.81 30
0.5 0 0.39 0.86 20
1.o 0 0.72 0.80 29
16 0 1270 40.0 0.64 2030
0.2 239 29.O 0.60 700
0.5 132 23.0 0.58 465
1.o 153 22.0 0.60 500
Caseinate 12 0 0 1.1 0.90 69
0.2 0 2.5 0.86 130
0.5 0 4.5 0.86 235
1.0 0 14.0 0.82 610
WPC 20 0 21 3.9 0.67 106
0.2 25 3.4 0.67 99
0.5 32 3.2 0.68 105
1.o 26 3.7 0.66 103

aDefined as a/+ at = 100 s-'

Kelley and Pressey (1966)found that factors which influence sulf-


hydryl groups have a marked effect on the Brabender viscosity of soy
protein dispersions. Mixing in an oxygen atmosphere resulted in a
twofold increase in Brabender consistency compared with mixing in
nitrogen. Similarly, low concentrations of bromate and iodate
RHEOLOGY OF PROTEIN DISPERSIONS 19

( 5 mmoles/800 ml solution) resulted in significant consistency in-


creases. The disulfide bond cleaving agents sodium sulfite and thio-
glycolic acid at very low concentrations reduced the maximum
Brabender consistency.

VISCOELASTICITY

Of the many physical treatments applied to protein dispersions


perhaps the application of heat is the most common and certainly one
of the more important. The capacity of proteins to form gels and
provide a structural matrix for holding water, flavors and other ingre-
dients is a critical functional attribute.
In the denaturing of a protein, the polypeptide chains unfold ex-
posing non-polar groups which associate in a three dimensional network
(Ferry 1948). More recently, Tombs (1974) suggested that globular
proteins form gels as a result of aggregation of the denatured protein to
form strands followed by the interaction of the strands to form a gel
mesh. These interactions which form the linkages may be electrostatic,
hydrophobic, covalent or hydrogen bonds. Thermoreversible gels have a
preponderance of intermolecular hydrogen bonds, whereas gels with
covalent bonds or a few disulfide linkages per polymer chain may be
thermoirreversible.
Most studies of protein gels have used viscometric or consistency
measurements only, or else have estimated the parameter generally
known as “gel strength”. However, all polymer dispersions (including
gels), multiphase systems and tissue structures, hence most fluid and
semi-solid foods, are viscoelastic in nature. That is, they possess solid-
like elastic and fluid-like viscous behavior simultaneously. This dual
nature is particularly relevant to material response in situations of
unsteady motion, that is, when stresses and strains are changing with
time, as in cutting, mashing, chewing, mixing, pouring, spreading or
pumping. Moreover, in steady shear flow, a viscoelastic fluid will
exhibit an unbalanced stress acting in the direction normal to the plane
of flow. The fluid will rise around a rotating spindle and this occurrence
is known as the Weissenberg effect. Another consequence of this
normal force is “die swell” which is an increase in diameter of a mate-
rial extruded under pressure through a tube. Normal forces increase
with increasing tangential shear. This lateral expansion is of importance
in extrusion processing.
A complete rheological description of a viscoelastic material requires
measurement of a number of parameters over several decades of time.
20 MARVIN A. TUNG

Empirical test methods are useful for quality control purposes although
results may depend on the test employed. A series of gels, even of the
same type, will not necessarily be ranked in the same order of
“strengths” by a rupture or a deformation test (Mitchell 1976).
Hermansson and Akesson (1975b) found that results of an empirical
test of gel strength did not reflect evident differences in gel character.
Elucidation of the fundamental nature of the gel structure requires
rheological test methods that will measure viscoelastic behavior. From
the measurement of viscoelastic properties, information can be ob-
tained about the nature and rates of the configurational rearrange-
ments, and the disposition and interaction of the macromolecules in
their short-range and long-range interrelations (Ferry 1970). Informa-
tion on structural aspects of a system at a molecular level may be used
to predict behavior on a macroscopic scale, such as mechanical behavior
in processing and utilization (Stanley and Tung 1976).
As viscoelasticity is imparted by the three-dimensionally interlinked
network of dispersed polymer molecules held together by relatively
weak van der Waal’s forces in fluid foods or the much stronger forces
found in solids, rheological evaluation must be made under conditions
which, as near as possible, are non-destructive in nature, that is, with
very low stress or strain, so that alteration of internal structure is
minimized. Methods usually employed include creep compliance-time
studies, stress relaxation at constant strain, and dynamic testing.

CREEP COMPLIANCE TESTS

In creep compliance testing a small constant stress is applied to a


sample and the resulting strain is followed with time. When the stress is
applied, the sample initially exhibits a purely elastic response so that if
the stress is removed during that period, the sample will fully recover its
original shape. However, after this initial very short time, the rate of
strain decreases as the weakest linkages in the internal structure of the
sample rupture. Progressively stronger linkages rupture until the rate of
strain is constant and flow occurs. Plotting the data as strain against
time (Fig. 6 ) permits the determination of a number of rheological
parameters. In stress relaxation tests, a predetermined strain is suddenly
imposed and kept constant while the relaxation of stress with time is
measured.

Agar, Egg Albumen and Soy Isolate Gels


Isozaki et al. (1976)used creep compliance and stress relaxation to
RHEOLOGY OF PROTEIN DISPERSIONS 21

I rLoad removed

Recovery curve

Creep curve

I
0
Time

FIG. 6. IDEALIZED CREEP AND RECOVERY CURVES FOR A


VISCOELASTIC MATERIAL

study the viscoelasticity of agar, egg albumen and soy isolate gels.
Measurements were made with a parallel plate viscoelastometer and a
Rheolometer. Creep and relaxation curves of all gels could be approx-
imated by sixelement Voigt and Maxwell models respectively. Linear
plots on logarithmic coordinates were found for the relationships
between E,, the Young’s modulus of the Hookean body of the Voigt
model, as well as EM the Young’s modulus of the first Maxwell body
of the Maxwell model, and protein concentration (C). The following
relationships were determined using the line of best fit:

for 1.0-1.8% agar; E, = 1.8 x 105C2*0, EM, = 0.87 X 105C1.9,


for 10-15% egg albumen; E, = 1.3 X lO2C3ao, EM = 0.21 X 102C3.4,
for 18-21% soy protein; E, = 1 . 4 X 10-2C5.7, E M , = 0.24 X 10-3C6*9.

Thus, in general, the value of the exponent for concentration is approx-


imately 2 for agar, 3 for egg albumen and 6-7 for soy protein. While
conceding that it is possible for the differences in values of the
exponents to be the result of differences in concentration levels
employed for the different protein gels, the authors feel that since all
gels show E, values of 1to 5 X 106 dyne cm-2, and EM values of 0.5 ,
to 3.0 X 106 dyne cm-2, these results suggest differences in the gel
structure between the samples.
22 MARVIN A. TUNG

DYNAMIC TESTS
In dynamic testing small deformations and short time spans are used,
conditions that will not alter the structure of a material and will satisfy
requirements of linear viscoelasticity theory based on infinitesimal
strains and strain rates where the ratio of stress to strain is a function of
time (or frequency) alone, and not of stress magnitude (Ferry 1970). If
a sinusoidally varying strain is imposed on a sample which responds in a
linear manner, a sinusoidally varying stress will result. For an elastic
material, the stress will be in phase with the strain, while for a viscous
material, the stress will be in phase with the strain rate which is 90” out
of phase in advance of the strain. When a viscoelastic material is sub-
jected to sinusoidally oscillating strain, the stress is neither exactly in
phase nor 90° out of phase but is intermediate in response with some of
the energy input stored and recovered in each cycle and some dissipated
as heat. The extent of the phase shift will reflect the extent of viscous
and elastic natures of the material.
From the curves developed for any given specimen, the elastic and
viscous components known as the dynamic shear storage modulus (G’)
and the dynamic shear loss modulus (G”) can be calculated. The loss
tangent is expressed as tan 6 = G”/G’, while the dynamic viscosity is
q’ = G“/w, where o is the oscillatory frequency.

Hydrated Wheat Gluten


Cumming and Tung (1975) determined the dynamic shear behavior
of rehydrated commercial wheat gluten subjected to small deformations
under oscillatory shear using the cone and plate geometry in a
Weissenberg rheogoniometer. Most traditional methods of rheological
testing applied to flour dough involved large deformations and, while of
great practical significance in specific processes, are of limited value in
clarifying the viscoelastic nature of gluten which is related to its unique
functional properties in baking.
Hydrated wheat gluten had an interesting rheological nature in
dynamic shear (Cumming and Tung1975). The storage modulus or
elastic component(G’) was very sensitive to the moisture content.
Tan 6 was greatly influenced by oscillatory frequency, increasing with
w , which implied increasing viscous behavior. However, even at large o,
G‘ was still about twice the value of G”, illustrating the relatively elastic
character of gluten. Working of the specimen also was found to affect
G’ and G”, resulting in an increase in magnitude of both parameters, a
characteristic practically observed in dough mixing operations as work
hardening. Working did not affect the loss tangent, tan 6, indicating no
RHEOLOCY OF PROTEIN DISPERSIONS 23

influence on the proportional contribution of storage and loss moduli.


The findings of Cumming and Tung (1975) were compatible with, and
lent support to, a published theory for the mechanism of the visco-
elastic response of wheat gluten (Bemardin and Kasarda 1973).
A later report (Cumming and Tung 1977) detailed modification of
the rheological properties of rehydrated gluten by manipulating
protein, starch, lipid and moisture levels, concluding that careful
control of these factors combined with various energy input systems
could be expected to yield improved, as well as new and novel, proc-
esses and products.

Rapeseed Protein Dispersions


Gill and Tung (1976, 1978) studied the rheology of 125 rapeseed
protein dispersions in steady and oscillatory shear. Significant increases
in apparent viscosity were observed when samples containing as little as
1%protein were heated, while gelation was obtained at 4.5% protein.
Highest apparent viscosities were found with gels prepared at pH 10
(Fig. 7). Gels at pH 6 demonstrated anomalously high apparent viscosi-
ties in view of their relatively amorphous microstructure as revealed by
scanning electron microscopy (Gill and Tung1978). The authors
speculated that at pH 6, the gel’s high apparent viscosity may be the
result of protein insolubility and aggregation rather than of the three
dimensional structural integrity. No gel formation was observed at
pH 2. Reductive alkylation to methylate free e-amino groups prevented
gel formation (Fig. 7), but the authors found no differences between
the content of e-amino groups of lysine in modified and unmodified
protein. They concluded that modification of some other functional
group essential for gelation had occurred in preference to e-amino.
Rapeseed protein gels adjusted to 0.5 and 1.OM NaCl demonstrated
dramatically higher apparent viscosities (Gill and Tung 1978). If ionic
bonds were of major importance in 125 rapeseed protein gels, increased
ionic strength would have been expected to result in reduced gel
strength as a consequence of ion competition for the interacting
functional groups of the protein. However, since Gillberg and
Tomell (1976) found that NaCl increased protein solubility, more
effective overlapping of functional groups may explain the higher
apparent viscosities. Addition of 1M urea as a hydrogen bond disrupting
agent, or of 0.15M dithiothreitol as a disulfide bond disrupting agent,
did not prevent gelation and had only small effects on apparent vis-
cosity (Gill and Tung 1978).
The effects of oscillation frequency on the loss tangents of pH 9.2
rapeseed protein gels containing various additives are illustrated in
24 MARVIN A. TUNG

10’-
\

FIG. 7. EFFECT OF pH AND REDUCTIVE ALKYLATION


ON THE FLOW BEHAVIOR OF 4.5% RAPESEED PRO-
TEIN DISPERSIONS AT 23OC (DATA FROM GILL AND
TUNG 1978)

Fig. 8. The small inflection points observed at low frequency may be


the result of entanglement coupling (Ferry 1970) in which extended
linear fragments interact in a specific frequency range such that an
increase in elastic behavior is observed. Although the addition of NaCl
increased the absolute strength of the pH 9.2 gel system, its effect on
elasticity was minimal. The urea treated sample appeared to show a
slight relative increase in elastic component with increasing o and
exhibited the largest proportional elastic response overall. The sample
modified by reductive alkylation demonstrated a greater dependency
on frequency of oscillation. Interestingly, the 0.1 5M dithiothreitol
treated gel was unable to store proportionally as much energy in oscilla-
tory shear as the other gels; its loss tangent was, thus, considerably
higher. The dithiothreitol gel then, while comparably viscous in relation
to a control gel, was considerably less elastic, suggesting that disulfide
bonds do play a relevant role in matrix formation. Gill and Tung (1978)
concluded that gel formation in this rapeseed protein is a complex
phenomenon involving many types of bonds. It is particularly worthy
RHEOLOGY OF PROTEIN DISPERSIONS 25

o+lM NoCl
r+0.5M NaCl
D + 0.15 M d i t h i o t h r e i t o I
no additives
A modified by
reductive olkyla tion

I 0.01
, I
0.1
I
1
4
10
Frequency. HZ

FIG. 8. EFFECT OF VARIOUS ADDITIVES ON THE DYNAMIC!


SHEAR LOSS TANGENT OF 4.5%RAPESEED PROTEIN DISPERSIONS
AT 23OC (DATA FROM GILL AND TUNG 1978)

of note that the dynamic experiments yielded information on matrix


formation not evident in steady shear experiments.

Gelatin Gels
Dynamic testing has also been applied to the investigation of the
aging or gel formation process of gelatin gels. Nijenhuis (1974)mea-
sured the storage and loss moduli at a number of frequencies as a
function of aging time using a coaxial cylinder dynamic rheometer. The
reduced (to a single reference temperature) storage moduli of the
gelatin gels plotted against the logarithm of aging time yielded nearly
straight lines with almost constant slopes, their positions dependent on
the aging temperature but not on the frequency of oscillation. Since
there was no apparent frequency dependence, Nijenhuis (1974)con-
cluded that the applied frequency range lay completely within the
rubbery region of the gel. This was in contrast to results for polyvinyl-
chloride (PVC)in di-(a-ethylhexyl)phthalate where the storage moduli
were frequency dependent even after 100 hr of aging (Nijenhuis 1974).
At high frequency there was an increase in the PVC storage modulus
characterizing the onset of the rubber-glass transition. Moreover, the
slopes of the G' us. w plots on logarithmic coordinates differed for
varying aging time. These differences between the dynamic properties
26 MARVIN A. TUNG

suggested that the aging or gel formation of gelatin and polyvinyl-


chloride are fundamentally different (Nijenhuis 1974).

SOYBEAN PROTEIN GELS

Although few studies of soybean protein gels have employed


fundamental techniques, much information has been gained in experi-
ments in which consistency or “gel strength” has been measured. The
gelation of soy protein dispersions can be caused by heating and cooling
or by the addition of calcium.
Circle et al. (1964) used PromineD to study the gelation of soy
protein. Gel stability under stress conditions of heating was found to
depend primarily on the protein concentration for any given tempera-
ture and time. For high protein concentrations, 10 min at 70°C was
sufficient to derive rigid gels. Brookfield consistency of heated dis-
persions (pH7, 10% protein, 100°C, 30min) was increased by salt
addition (NaCl, NaN03, NaHP04 H2 0)in contrast to their effects on
unheated dispersions. The reducing agents, sodium sulfite and cysteine,
markedly reduced Brookfield consistencies of unheated dispersions and
heated gels, their thinning action presumed to be the result of cleavage
of disulfide bonds. Even sodium sulfite, however, was ineffective in
preventing gelation when protein concentration was high. Two other
reducing agents, sodium hypophosphite and sodium nitrite, were found
ineffective.
Catsimpoolas and Meyer (1970, 1971a,b) attempted to characterize
the mechanism of soybean globulin gelation and the molecular forces
involved. They found that a minimum protein concentration of 8%was
required to obtain a self supporting gel network at pH 7 and room
temperature. The heating process activated the protein sol to the progel
state characterized by a significantly higher Brookfield consistency.
Catsimpoolas and Meyer (1970) found that this step was irreversible
and likely resulted from the disruption of the quaternary structure of
soybean globulins, which has been demonstrated by a number of
workers (Wolf and Tamura1969; Cumming etal. 1973; Hashizume
etal. 1975). The gel was obtained by cooling the progel and an addi-
tional increase in consistency was observed as the globulins reaggregated
into a network structure stabilized by hydrogen and ionic bonds; this
transition was reversible (Catsimpoolas and Meyer 1970). The gels were
found to be more sensitive to excess heat at pH extremes, while highest
Brookfield consistency values were obtained at pH 7-9. Ehninger and
Pratt (1974), however, found that gels of Promine-D and Promine-R at
lower protein concentrations (<9%) had higher consistencies at
RHEOLOGY OF PROTEIN DISPERSIONS 27

pH < 6.0 than at pH 6.5. Hutton and Campbell (1977) confirmed that
Promine-D heated dispersions had a maximum Brookfield consistency
at pH 5-6 but the Promosoy 100 maximum occurred at the highest pH
studied, pH 7.0. Aoki (1965) reported that strong gels were formed
only at pH > 5. Obviously, pH effects were dependent on purity and
pretreatment of the soy protein studied.
Catsimpoolas and Meyer (1970) found that the effects of NaCl
(0.2-2M) on gel consistency were somewhat anomalous. At tempera-
tures above 70°C, the consistencies of the progel and gel decreased with
increasing NaCl concentration, but below this temperature higher
Brookfield consistency was favored by a higher concentration of salts.
In agreement with these results are those of Hermansson (1972) who
found that Promine-D gels (10% protein, heated at 75°C) exhibited
rapidly decreasing Brookfield consistency with increasing ionic strength
(0-0.8M NaCl), effects which correlated well with swelling properties.
At acidic pH (4.5-6.5), 0.2M NaCl decreased the Brookfield consis-
tency and stability of soy dispersions heated to about 90°C and cooled
(Ehninger and Pratt 1974).
Excess heat (125°C) and chemical modification (by disulfide or
hydrogen bond cleaving agents, for example) converted the progel to a
metasol which did not gel on cooling (Catsimpoolas and Meyer 1970).
Aoki and Sakuri (1968) reported that gelation was prevented by the
addition of reducing agents having disulfide bond cleaving ability and,
although to a lesser extent, by oxidizing agents. However, Catsimpoolas
and Meyer(1970) found that while low concentrations (0.1%) of
mercaptoethanol inhibited gelation, high concentrations (10%)
enhanced it, while another reagent capable of blocking sulfhydryl
groups (0.1% N-ethylmaleimide) had no effect on gelation. Similarly,
Circle e t a l . (1964) reported that cysteine at the 0.05% level inhibited
gelation but at the 0.5% concentration w a s considerably less effective.
Thus, the role of disulfide bonds in soy protein gelation is unclear.
Another anomaly was reported by Aoki and Sakurai (1968) who found
that urea and guanidine hydrochloride generally prevented gel forma-
tion but at low concentrations these hydrogen bond disrupting agents
promoted gel formation of rather brittle gel precursors.
Dispersions of soy protein in water miscible solvents formed gels of
higher Brookfield consistency than did water dispersions, effects which
were associated with the solvent's ability to unfold proteins
(Catsimpoolas and Meyer 1971a). Aliphatic chain length of either a
water miscible solvent or a triglyceride affected gelation (Catsimpoolas
and Meyer 1971a,b). Sucrose and dextrose also influenced Brookfield
consistency of soy protein gels, their effects being dependent on
protein concentration and pH (Ehninger and Pratt 1974).
28 MARVIN A. TUNG

SUMMARY

A complete coverage of the literature on rheology of protein dis-


persions has not been attempted in this presentation. Rather, the
relevance of rheology has been emphasized with examples cited in
which rheological testing has been used to clarify intinsic or functional
properties of proteins. Emphasis, too, has been placed on factors
which influence flow properties of protein dispersions, since these
provide the key for modification to impart desired functionality. Some
progress has been made in applying these concepts to food fabrication
but much remains to be accomplished in the application of funda-
mental physicochemical information to protein modification and
process parameter optimization, prerequisites to effective protein utili-
zation. Particular attention has been paid to soy protein dispersions
which, although widely studied, have been the subject of few funda-
mental rheological investigations. It is hoped that the interest of food
rheologists will be stimulated in order that more fundamental studies
will be pursued.

ACKNOWLEDGMENTS

The assistance of Maureen R. Garland in the preparation of this


review is sincerely appreciated. This work was supported in part by the
National Research Council of Canada.

REFERENCES

AOKI, H. 1965. The gelation of soybean protein. 11. The fundamental factors
affecting gelation. Nippon Nogai Kagaku Kaishi 39(7), 270-276.
AOKI, H. and SAKURAI, M. 1968. The gelation of soybean protein. IV. The
effects of reducing agents, oxidizing agents and protein denaturants. Nippon
Nogai Kagaku Kaishi 42(9), 544-552.
BERNARDIN, J. E. and KASARDA, D. D. 1973. The microstructure of wheat
protein fibrils. Cereal Chem. 50,735-745.
CATSIMPOOLAS, N. and MEYER, E. W. 1970. Gelation phenomena of soybean
globulins. I. Protein-protein interactions. Cereal Chem. 47, 559-570.
CATSIMPOOLAS, N. and MEYER, E. W. 1971a. Gelation phenomena of soybean
globulins, 11. Protein-water miscible solvent interactions, Cereal Chem. 48,
150-158.
CATSIMPOOLAS, N. and MEYER, E. W. 1971b. Gelation phenomena of soybean
globulins. 111. Protein-lipid interactions. Cereal Chem. 48,159-167.
CIRCLE, S. J., MEYER, E. W. and WHITNEY, R. W. 1964. Rheology of soy
protein dispersions. Effect of heat and other factors on gelation. Cereal
Chem. 41,157-172.
RHEOLOGY OF PROTEIN DISPERSIONS 29

CUMMING, D. B., STANLEY, D. W. and deMAN, J. M. 1973. Fate of water soluble


soy protein during thermoplastic extrusion. J. Food Sci. 38,320-323.
CUMMING, D. B. and TUNG, M. A. 1975. Dynamic shear behavior of commercial
wheat gluten. Can. Inst. Food Sci. Technol. J. 8,206-210.
CUMMING, D. B. and TUNG, M. A. 1977. Modification of the ultrastructure and
rheology of rehydrated commercial wheat gluten. Can. Inst. Food Sci.
Technol. J. 10,109-119.
EHNINGER, J. N. and PRATT, D. E. 1974. Some factors influencing gelation and
stability of soy protein dispersions. J. Food Sci. 39,892-896.
FERRY, J. D. 1948.Protein gels. Adv. Protein Chem. 4, 1-78.
FERRY, J. D. 1970. Viscoelastic Properties of Polymers, 2nd e d . , John Wiley and
Sons, New York.
FLEMING, S. E., SOSULSKI, F. W., KILARA, A. and HUMBERT, E. S. 1974.
Viscosity and water absorption characteristics of slurries of sunflower and
soybean flours, coccentrates and isolates. J. Food Sci. 39,188-191.
FRISCH, H. L. and SIMHA, R. 1956. The viscosity of colloidal suspensions and
macromolecular solutions. In Rheology Theory and Applications, Vol. 1,
(F. R. Eirich, ed.), pp. 525-613,Academic Press, New York.
GILL, T. A. and TUNG, M. A. 1976. Rheological, chemical and microstructural
studies of rapeseed protein dispersions. Can. Inst. Food Sci. Technol. J. 9,
75-79.
GILL, T. A. and TUNG, M. A. 1978. Thermally induced gelation of the 12s
rapeseed glycoprotein. Accepted by J. Food Sci.
GILLBERG, L. and TORNELL, B. 1976. Preparation of rapeseed protein isolates.
Precipitation of rapeseed proteins in the presence of polyacids. J. Food Sci. 41,
1070-1075.
GRONINGER, H. S., JR. ,1973.Preparation and properties of succinylated fish
myofibrillar protein. J. Agr. Food Chem. 21,978-981.
HASHIZUME, K., NAKAMURA, N. and WATANABE, T. 1975.Influence of ionic
strength on conformation changes of soybean proteins caused by heating, and
relationship of its conformation changes to gel formation. Agr. Biol. Chem. 39,
1339-1347.
HAWKINS, A. E. 1974. The contribution of rheology to an improved meat
product. Dechema-Monogr. 77,321-335.
HAYES, J. F. and MULLER, L. L. 1961. Factors affecting the viscosity of solu-
tions of acid-precipitated caseins. Aust. J. Dairy Technol. 16,265-269.
HAYES, J. F., MULLER, L. L. and FRASER, P. 1969. Studies on co-precipitates
of milk proteins. Part 5 - Investigations on viscosity of co-precipitates in dis-
persions of high concentration. Aust. J. Dairy Technol. 24,75-78.
HERMANSSON, A.-M. 1972. Functional properties of proteins for foods-
swelling. Lebensm.-Wiss. u. Technol. 5( l),24-29.
HERMANSSON, A.-M. 1975a. Functional properties of added proteins correlated
with properties o f meat systems. Effect on texture of a meat product. J. Food
Sci. 40,611-614.
HERMANSSON, A.-M. 1975b. Functional properties of proteins for foods-flow
properties. J. Texture Studies 5,425-439.
HERMANSSON, A.-M. and AKESSON, C. 1975a. Functional properties of added
proteins correlated with properties of meat systems. Effect of concentration and
temperature on water-binding properties of model meat systems. J. Food
Sci. 40,595-602.
HERMANSSON, A.-M. and AKESSON, C. 197513. Functional properties of added
30 MARVIN A. TUNG

proteins correlated with properties of meat systems. Effect of salt on water-


binding properties of model meat systems. J. Food Sci. 40,603-610.
HSU, H. W., SATTERLEE, L. D. and KENDRICK, J. G. 1977a.Computer blending
predetermines properties of protein foods. Part I. Experimental design. Food
Prod. Dev. 11 (7),52-58.
HSU, H. W., SATTERLEE, L. D. and KENDRICK, J. G. 1977b.Computer blending
prec' +erminesproperties of protein foods. Part 11. Results and discussion. Food
Proa. 3ev. 11 (8),70-78.
HUANG, F. and RHA, C. K. 1971. Rheological properties of single-cell protein
c o n c e n t r a t e : dope formation and its flow behavior. J. Food Sci. 36,
1131-1 134.
HUANG, F. and RHA, C. 1974. Protein structures and protein fibers - a review.
Polymer Eng. Sci. 14(2), 81-91.
HUTTON, C. W. and CAMPBELL, A. M. 1977. Functional properties of a soy
concentrate and a soy isolate in simple systems and in a food system. Emulsion
properties, thickening function and fat absorption. J. Food Sci. 42,457-460.
ISHINO, K. and OKAMOTO, S. 1975. Molecular interaction in alkali denatured
soybean proteins. Cereal Chem. 52,9-21.
ISOZAKI, H., AKABANE, H. and NAKAHAMA, N. 1976. Viscoelasticity of hydro-
gels of agar-agar. Analysis of creep and stress relaxation. J. Agr. Chem. SOC.
Japan 50(6), 265-272.
KELLEY, J. J. and PRESSEY, R. 1966. Studies with soybean protein and fiber
formation. Cereal Chem. 43,195-206.
KINSELLA, J. E. 1976.Functional properties of proteins in foods: a survey. CRC
Crit. Rev. Food Sci. Nutr. 7(3), 219-280.
MITCHELL, J. R. 1976. Rheology of gels. J. Texture Studies 7,313-339.
NIJENHUIS, K. te. 1974. Investigation into the aging process of gel systems by the
measurement of their dynamic moduli. Dechema-Monogr. 77,177-186.
PURI, B. R., MOHINDROO, U. and MALIK, R. C. 1972. Studies in physico-
chemical properties of caseins. Part 111. Viscosities of casein solutions in differ-
ent alkalies. J. Ind. Chem. SOC. 49,855-861.
SABATER de SABAThS, A. and SOLER NOLLA, G. 1976. Rheology of emulsions
used in meat industry. Effect on sausage texture. Anales de Bromato-
logia 28(1),81-98.
SHIBASAKI, K., KIMURA, Y. and O'KUBO, K. 1969. Food chemical studies on
soybean proteins. Part V. The viscotic behavior of urea denatured proteins.
Nippon Shokuhin Kogyo Gakkai-shi 16(7), 298-303.
STANLEY, D. W. and TUNG, M. A. 1976. Microstructure of food and its relation
t o texture. In Rheology and Texture in Food Quality, (J.M.deMan,
P. W. Voisey, V. F. Rasper and D. W. Stanley, eds.), pp. 28-78, Avi Publishing
Co., Inc., Westport, Conn.
THOMSON, R. H. K. 1946. The extrusion of filaments from solutions of vegetable
globulin. SOC. Dyers and Colourists, Symposium on Fibrous Proteins,
p. 173-177,Chorley and Pickersgill Ltd., Leeds, England.
TOMBS, M. P. 1974. Gelation of globular proteins. Faraday Discuss. Chem.
SOC.57,158-164.
TUNG, M. A., WATSON, E. L. and RICHARDS, J. F. 1971. Rheology of egg
albumen. Trans. ASAE 14(1),17-19.
van MEGAN, W. H. 1974.Solubility behavior of soybean globulins as a function of
pH and ionic strength. J. Agr. Food Chem. 22(1),126-129.
RHEOLOGY OF PROTEIN DISPERSIONS 31

WARD, A. G. and SAUNDERS, P. R. 1958. The rheology of gelatin. In Rheology


Theory and Applications, Vol. 11, (F. R. Eirich, ed.), pp. 313-362, Academic
Press, New York.
WOLF, W. J. and TAMURA, T. 1969. Heat denaturation of soybean 11s protein.
Cereal Chem. 46,331-344.

You might also like