You are on page 1of 7

Relationships of Hydrophobicity to Emulsifying Properties

of Heat Denatured Proteins


LEANDROS P. VOUTSINAS, ELAINE CHEUNG, and SHURYO NAKAI

ABSTRACT point to evidence, however, that emulsifying properties and


Effects of heating on the emulsifying properties of selectedfood solubility are not well correlated (Aoki et al., 1980; McWat-
proteins and the protein surfacehydrophobicity (So) asa predictor ters and Cherry, 1975; McWatters and Holmes, 1979a,
of thesepropertieswere investigated.The emulsifyingpropertiesof 1979b; Smith et al., 1973; Wang and Kinsella, 1976).
the proteins studied were differently affected by heating. Heat- The protein hydrophobicity has been lately receiving
denaturation was not always accompaniedby loss of emulsifying much attention since the hydrophobic interactions are
properties, but, on the contrary, in some casesresulted in great considered to play important roles in the functional prop-
improvement.The emulsifying properties could welI be predicted erties of food proteins (Kato et al., 1981 ; Kinsella, 1979).
solely on the basisof So level but not on the basisof solubihty level, Keshavarz and Nakai (1979) reported a significant correla-
which indicated that So is a very important property determining
protein functionality. However,the emulsifying activity, emulsion tion between surface hydrophobicity (determined by
stability and fat binding capacity of the proteins studied could be hydrophobic chromatography and hydrophobic partition
explained and more accurately predicted using So and solubihty techniques) and interfacial tension of the proteins studied.
together. Kato and Nakai (1980) subsequently reported that the
surface kydrophobicity (determined fluorometrically)
showed significant correlations with interfacial tension and
INTRODUCTION emulsifying activity of the proteins studied. Their results
TO BE USEFUL and successful in food applications, pro- suggest that the emulsification of oil with protein can be
teins in addition to providing essential amino acids, should explained using the concept of protein hydrophobicity.
ideally possessseveral desirable characteristics referred to as Nakai et al. (1980b) also reported that the effective (sur-
functional properties (Wang and Kinsella, 1976). Moreover, face) hydrophobicity showed good correlations with inter-
according to Johnson (1970) the functional and physical facial tension and emulsifying activity of the plant proteins
properties, rather than the nutritional value, of protein in studied. It is noteworthy that these authors observed a
protein-containing products will largely determine their closer correlation of emulsification capacity with hydro-
acceptability as ingredients in prepared foods. phobicity than with solubility.
The ability of protein to aid the formation and stabiliza- While many factors influence the performance of pro-
tion of emulsions is critical for many applications in teins in food systems, heat treatment is one of the most
chopped, comminuted meats, cake batters, coffee whiteners, important and is very often used during the processing of
milk, mayonnaise, salad dressings, and frozen desserts. In protein products. This study was instigated by the observa-
these products varying emulsifying and stabilizing capacities tion that the emulsifying capacity of soy protein was not
are required because of the differing composition and adversely affected even by texturization which caused a loss
stresses to which these products are subjected (Kinsella, in solubihty. The objectives of this paper, therefore, were to
1979). Moreover, the ability of proteins to bind fats is a determine the effect of heating on the emulsifying proper-
very important functional property for such applications ties of selected food proteins, and to assessthe value of
as meat replacers and extenders, principally because it surface hydrophobicity as a predictor of the emulsifying
enhances flavor retention and reputedly improves mouth- properties of these proteins.
feel (Kinsella, 1976). Soy proteins have been added to
ground meats to promote fat absoption or fat binding, and
thus decrease cooking losses and maintain dimensional MATERIALS&METHODS
stability in the cooked product (Wolf and Cowan, 1975).
To evaluate the emulsifying properties of a protein its Materials
solubility profile is usually determined, because of its use- Bovine serum albumin (#A-4503), p-lactoglobulin (#L6879
fulness as an excellent index of protein functionality (Kin- from milk) and ovalbumin (#A-5503) were ail purchasedfrom
sella, 1976). Good solubility can markedly expand poten- SigmaChemicals(St. Louis, MO). Soy protein isolatewasobtained
tial applications of a protein (Kinsella, 1976). Denaturation, from General Mills, Inc., (Minneapolis,MN). Promine-Dwas pur-
chasedfrom Central Soya Co. (Chicago,IL). Pea protein isolate
on the other hand, implicates losses of protein functional (a) (M412-046),Century cultivar field pea, was obtainedfrom POS
properties and is usually measured as a loss of solubility Pilot Plant Corp., Univ. of Saskatchewan(Saskatoon,Sask.).Pea
(Nakai and Powrie, 198 1). Generally, surfactant properties protein isolate (b), Pro-PulseWlOO, was obtained from Griffith
are related to the aqueous solubility of proteins (Kinsella, Laboratories Ltd. (Scarborough,Ont.). Vital wheat gluten, Whet-
1976). A positive correlation between solubility and the pro 75%, was suppliedby Industrial Grain ProductsLtd. (Thunder
ability of a protein to emulsify and stabilize an emulsion Bay, Ont.). Canola Protein isolate was preparedby the method of
has been reported in many studies (Crenwelge et al., 1974; Nakai et al. (198Oa).The protein content of isolatebatches(a) and
Pearson et al., 196.5; Swift and Sulzbacher, 1963; Volkert (b) was 84% and 88%, respectively.Sunflower protein isolatewas
and Klein, 1979; Yasumatsu et al., 1972). Many authors preparedby the method of Nakai et al. (1980a).Wheyprotein con-
centrate (75%) was obtained from Sodispro Technol. (St. Hya-
&the, Que.). Gelatin, Bloom 300, was purchasedfrom United
StatesBiochemicalCorp. (Cleveland,OH).
Whole caseinwaspreparedfrom skim milk by acid precipitation.
Authors Voutsinas, Cheung and Nakai are affiliated with the Dept.
Myosin was extracted from the pectoralis superficialisand pro-
of Food Science, The Univ. of British Columbia, Vancouver, British
fundus musclesof a freshly slaughteredmale chicken (broiler 11 -
Columbia, Canada V6T 2A2.
12 wk old) with cold KCl-potassium phosphatebuffer, pH 6.5
accordingto the method of Perry (1955).
26-JOURNAL OF FOOD SCIENCE-Volume 48 (1983)
Preparation of protein samples Protein (surface) hydrophobicity
For hydrophobicity, solubility, emulsifying activity index, Protein surface hydrophobicity was fluorometrically deter-
emulsion stability index and fat binding capacity determinations, mined according to the method of Kato and Nakai (1980) after
the following protein samples were prepared. To increase the ac- slight modification. Each protein sample (2 ml) was serially diluted
curacy of regression analysis, as many different extents of denatura- with 0.01 M phosphate buffer, pH 7.4, to obtain protein concen-
tion as possible were attempted thus resulting in samples of differ- trations ranging from 0.00156% to 0.05%. Two sets of protein
ent solubility. For protein samples which needed constant pH to samples were prepared. Ten liters of cis-parinaric acid solution were
reproduce the same degree of denaturation, the weakest phosphate added to the first set of tubes. The parinaric acid-protein conjugate
buffer (O.OlM) was used. For proteins in which the solubility was was then excited at 325 nm and the relative fluorescence intensity
difficult to change by heating, CaClz was used to disperse the pro- was measured at 420 nm in an Aminco-Bowman spectrofluoro-
teins. meter, using a slit width of 0.5 mm. The fluorescence intensity read-
Bovine serum albumin (BSA). A 1% solution in distilled water. ing was standardized by adjusting the fluorometer reading to an
Heat denatured: the pH was adjusted to 4.0 and then the solution arbitrary value of 7.4 when 10 liters of cis-parinaric acid solution
was heated at 100°C for 5 min and homogenized in an Omni-Mixer was added to 2 ml of decane. Then, the fluorescence readings of the
(Ivan Sorvall, Norwalk, CT) for 1 min at speed setting 1 (lowest protein samples were taken. The net fluorescence intensity at each
speed). protein concentration was determined by subtracting the fluores-
p-Lactoglobulin. A 1% solution in distilled water. Heat dena- cence intensity of each sample without cis-parinaric acid from the
tured: the pH of the solution was adjusted to 1.0, heated at 100°C fluorescence intensity of the corresponding sample containing
for 15 min and then homogenized for 1 min at speed setting 1. cis-parinaric acid. The initial slope (So) of the fluorescence intensity
Soy protein isolate. A 1% aqueous dispersion, stirred on magnetic vs protein concentration plot was used as an index of the protein
stirrer for 5 min, and then, homogenized for 1 min at speed setting surface hydrophobicity. The initial slope was determined by linear
5. Samples l-4: pH 5.5, heated at 100°C for 0.25,0.5,1.0, and 2.0 regression analysis using a Monroe (Orange, NJ) 1880 programmable
min, respectively. Sample 5: pH 5.5, autoclaved at 121°C for 15 calculator.
min. Sample 6: pH 7.2, autoclaved at 121°C for 15 min.
Promine-D. A 1% dispersion in O.OlM phosphate buffer, pH 7.4, Solubility index
stirred magnetically for 5 min and homogenized for 1 min at speed
setting 5. Protein samples (l%, w/v, in O.OlM phosphate buffer pH 7.4)
Ovalbumin. A 1% aqueous solution. Samples l-4: pH 5.6, were dispersed by stirring with a magnetic stirrer for 5 min and then
heated at 80°C for 1.5, 2.0, 2.5, and 3.0 min, ;espectively.-Sample blended in a Sorval Omnimixer at speed setting 5 for 1 min. The pH
5: pH 5.6, heated at 100°C for 5 min. Sample 6: pH 1.0, heated of each dispersion was adjusted to 7.4 by adding 1N NaOH. A por-
at 100°C for 15 min. tion of each protein suspension was then centrifuged at 27,000 x g
Pea protein isolate (a). A 1% aqueous dispersion, stirred mag- for 30 min. Aliquots of the suspension and the supernatant after
netically for 5 min, and then, homogenized for 1 min at setting 5. centrifugation were diluted and the protein contents were deter-
Samples l-4: pH 5.8, heated at 80°C for 1, 2,4 and 7 min, respec- mined by the Phenol-Biuret method (Brewer et al., 1974)..The per-
tively, and then, homogenized for 10 sec. cent solubility index (s) was taken as the ratio of the protein con-
Pea protein isolate (b). A 1% dispersion in O.OlM phosphate buf- tent of the supernatant to that of the suspension.
fer, pH 7.4, was magnetically stirred 5 min, and then, homogenized
for 1 min at speed setting 5. Fat binding capacity (FBC)
Vital gluten. Samples l-5: 1% dispersions in 0.5, 1.0,1.5, 1.75 The FBC was determined according to the method of Voutsinas
and 2.0 N acetic acid, respectively. Then, the samples were mag- and Nakai (1981).
netically stirred for 5 mln, homogenized 15 set at speed setting 1,
and heated at 100°C for 30 min. Emulsifying activity index (EAI)
Canola protein isolate (a). A 1% aqueous dispersion, was stirred
EAI was determined by the tubidimetric technique of Pearse
for 5 min, and then, homogenized for 1 min at speed setting 5.
and Kinsella (1978). To prepare emulsion, each protein sample
Samples l-4: pH 5.5, heated at 100°C for 0.5, 1.0, 1.5 and 2.0
min, respectively, and then homogenized for 5 set at speed setting was diluted to a concentration of 0.5% with O.OlM phosphate buf-
1. Sample 5: pH 7.2, autoclaved at 121°C for 10 min. fer, pH 7.4. Two ml of corn oil and 6 ml of the diluted protein dis-
persion were homogenized together in an Omni-mixer with a micro-
Canola protein isolate (b). A 1% dispersion in O.OlM phosphate
buffer, pH 7.4, was stirred for 5 min, and then, homogenized for attachment (Ivan Sorvall, Inc., Norwalk, CT) at speed setting 1 for
1 min at speed setting 5. 1 min at 20°C.
Sunflower protein isolate. A 1% dispersion in O.OlM phosphate
buffer, pH 7.4, was stirred for 5 min, and then, homogenized for Emulsion stability index (EN)
1 min at speed setting 5. ES1 was determined by a modification of the method of Pearse
Whey protein. A 1% solution in 0.03M CaC12. Samples l-3: and Kinsella (1978) as follows: the emulsion, prepared as described
pH 6.0, heated at 80°C for 4, 5 and 15 min, respectively. Sample above (EAI), was held at room temperature and aliquots (0.1 ml)
4: pH 5.8, heated at 80°C for 15 min. were taken directly from the bottom of the container, containing
Whole casein. A 1% solution in O.OlM CaClT, pH 7.4. Samples the emulsion, at different time intervals and diluted to 50 ml
1 and 2: 1% solution in O.OlM CaCl,, pH 7.4, a&claved at 12i’C (500 x dilution) with O.OlM phosphate buffer, pH 7.4, containing
for 5 and 20 min, respectively. Sample 3: 1% solution in 0.02M 0.1% sodium dodecyl sulfate (SDS). The absorbance of diluted
CaC12,pH 7.4, autoclaved at lil°C fo; 20 min. emulsions at 500 nm was then iecorhed with a Beckman DB spec-
Gelatin. A 1% dispersion in O.OlM phosphate buffer, pH 7.4, trophotometer. The half-life (min) of the absorbance decay with
stirred 5 min, and then, homogenized for 1 min at speed setting 5. time, determined graphically, was used as ESI.
Heat denatured: 0.5% dispersion in O.OlM phosphate buffer, pH
7.4, heated at 75°C for 2.5 min with stirring. Viscosity
Myosin. To form a stock solution for determination of hydro- Viscosity measurements of 0.5% protein dispersion at 20°C were
phobicity, solubility, and other functional properties, the original made using a Brookfield Synchro-Lectric viscometer, Model LVT
gel was diluted twice with 0.3M NaCl in O.OlM phosphate buffer, fitted with a UL adapter, at 60 rpm.
pH 7.4. The control sample was unheated and samples 1 and 2 were
heated at 75°C for 1 and 5 min, respectively. Statistical analysis
Prior to analysis, all heated and unheated protein samples-
except Promine-D, pea protein (b), canola protein (b), sunflower Simple and multiple regression analyses were used to determine
protein, gelatin, and myosin-were’ dialysed against O.OlM phos- the relationships between hydrophobicity, solubility and emulsify-
phate buffer, pH 7.4, containing 0.02% sodium azide. After hydro- ing properties of the protein samples. These analyses were carried
phobicity, solubility and emulsifying properties were determined, the out by using a Monroe 1880 programmable calculator. In addition,
protein samples were freeze-dried and subsequently used for fat backwards stepwise multiple regression analyses and surface visuali-
binding capacity (FBC) determinations. zation plotting were carried out using an Amdahl470 V/8 computer
at the University of British Columbia. Five independent variables

Volume 48 (1983hJOlJRNAL OF FOOD SCIENCE-27


HYDROPHOBICITY AND EMULSIFYING PROPERTIES. ..

were used in the initial equation in the backwards stepwise proce- properties. It is evident that, for most proteins under a
dure including surface hydrophobicity (S,,), solubility index (s), given set of conditions, protein solubility decreased as heat-
interaction of So and s, and quadratic powers of So and s. Depend- ing time increased due to the progressive denaturation of
ent variables included EAI, ESI, and FBC. The models for the the protein. Gelatin, as expected, was completely solubil-
prediction of the emulsifying properties were selected on the basis ized on heating due to the rupture of hydrogen bonds
of the statistical s$nificance of F-probabilities of the partial regres-
sion coefficients. which are responsible for its insolubility (Blanshard, 1970).
As protein denaturation progressed, as seen by the decrease
RESULTS & DISCUSSION in protein solubility, the hydrophobicity usually increased.
This is due to the gradual exposure of hydrophobic amino
Effect of heat treatment on emulsifying properties acid residues of native proteins which are usually buried in
the interior of the molecules (Tanford, 1973) as a result of
Table 1 shows the relationships of hydrophobicity and the protein unfolding. In the case of whey protein, it can
solubility index of selected proteins with their emulsifying be seen, that excessive heating resulted in a decrease of

Table l-Relationships between protein hydrophobicity. solubility index, emulsifying activity, emulsion stability and fat binding capacity of
various proteinsa

Solubility
ESlb (min)
Hydrophobicity index (s/o) EAI
Protein EC-J) (s) (m*/g) I II FBC (%I

Bovine serum albumin -control 325 100.0 148 108.50 91.50 25.0
Bovine serum albumin - heated 304 26.8 140 90.00 109.00 54.4
P-Lactoglobulin -control 426 100.0 96 27.20 37.00 4.2
fi-Lactoglobulin - heated 192 6.4 51 25.30 27.50 16.5
Soy isolate - control 95 26.4 42 6.65 25.20 90.0
-sample 1 (pH 5.5, 100°C. 0.25 min) 97 26.4 51 7.30 26.00 86.3
-sample 2 (pH 5.5, 100°C. 0.5 min) 131 24.0 56 5.55 17.00 81.7
- sample 3 (pH 5.5, lOO’C, 1 .O min) 150 14.2 48 1 .oo 5.13 76.0
- sample 4 (pH 5.5, lOO’C, 2.0 min) 144 4.2 58 0.98 14.03 73.5
- sample 5 (pH 5.5, 121°C. 15 mini 160 8.2 50 0.76 11.20 68.4
. - sample6 (pH 7.2, 121’C. 15 min) 128 77.2 76 112.50 100.00 54.8
Promine-D 39 29.1 65 1.10 5.80 85.3
Ovalbumin -control 6 100.0 24 0.56 0.70 42.7
- sample 1 (pH 5.6, 80°C. 1.5 min) 38 90.6 34 0.63 1.10 45.9
- sample 2 (pH 5.6,BO”C. 2.0 mini 87 75.8 35 1.10 1.80 57.5
- sample 3 tpH 5.6,BO”C. 2.5 min) 95 70.6 45 1.08 1.70 66.9
- sample 4 (pH 5.6, BO’C, 3.0 min) 142 48.9 49 5.85 8.40 82.5
- sample 5 (pH 5.6, lOO’C, 5 min) 269 10.6 78 24.30 138.00 102.6
-sample6 (pH 1.0, lOO’C, 15min) 296 46.0 136 12.60 108.00 101.3
Pea isoiate (a) -control 66 42.6 61 3.67 15.50 66.4
- sample 1 (BO”C, 1 min) 77 34.7 66 0.60 0.42 53.9
- sample 2 (80°C. 2 min) 104 29.5 35 0.56 0.31 51.6
- sample 3 (80°C. 4 min) 100 25.2 37 0.85 0.16 50.4
-sample 4 (80°C. 7 min) 121 20.2 25 1.45 0.16 57.5
Pea isolate tb) 59 58.0 54 1.60 17.60 78.9
Gluten -sample 1 (0.5 N acetic, lOO”C, 30 min) 65 71.6 70 0.36 17.60 38.0
-sample 2 (1.0 N acetic, 100°C. 30 min) 59 79.0 69 0.55 25.00 41.8
-sample 3 (1.5 N acetic, lOO”C, 30 min) 65 85.6 82 0.28 23.80 34.8
- sample 4 (1.75 N acetic, lOO’C, 30 min) 60 89.6 74 0.80 26.60 39.9
-sample 5 (2.0 N acetic, lOO”C, 30 min) 57 93.3 69 0.78 23.30 32.2
Canola isolate (a) -control 65 28.3 60 0.54 3.76 33.9
- sample 1 tpH 5.5, 100°C. 0.5 min) 75 15.9 63 0.49 2.80 29.1
- sample 2 (pH 5.5, lOO”C, 1 .O min) 77 10.5 49 0.17 1.27 29.2
-sample 3 (pH 5.5, 100°C. 1.5 min) 86 3.6 41 0.12 0.45 35.1
-sample 4 (pH 5.5, 100°C. 2.0 min) 78 2.9 40 0.14 0.47 42.5
- sample 5 (pH 7.2, 121°C. 10 min) 101 21.2 50 2.20 8.40 35.7
Canola isolate (b) 55 44.0 66 0.25 3.35 56.0
Sunflower isolate 47 31 .o 75 1.40 5.30 105.8
Whey protein - control 182 88.7 87 50.30 87.50 74.5
- sample 1 (pH 6.0,8O”C, 4 min) 211 75.2 98 61.30 102.00 75.2
- sample 2 (pH 6.0,8O”C, 5 min) 164 65.5 89 63.00 104.00 72.5
- sample 3 IpH 6.0,80°C, 15 min) 128 50.4 82 48.30 104.50 100.8
- sample 4 (pH 5.8,8O”C, 15 min) 132 61.6 85 46.30 108.00 99.4
Casein - control 28 100.0 58 1.70 39.30 11.3
-sample 1 tO.OlM CaClg, 121°C. 5 min) 30 76.2 56 9.50 35.00 18.1
-sample 2 (0.01 M CaC12, 121°C. 20 min) 21 71.5 49 14.75 24.30 16.7
-sample 3 (0.02M CaClg, 121°C, 20 min) 23 70.2 50 12.50 16.80 14.0
Gelatin -control 5 15.3 46 4.65 3.10 19.1
-heated (75OC, 2.5 min) 6 100.0 59 10.20 9.40 __-_
Myosin -control 14= 1 oo.o= 43= 36.00’ _--_
- sample 1 (75°C. 1 min) 44c 50.9c 5oc 22.40’ ----
- sample 2 (75’C, 5 min) 54c 16.5’ 48’ 1o.ooc ---_

E average of duplicate determination


I: 0.1 M NaCl added; II: NaCl not added
’ 0.3 M NaCl added

28-JOURNAL OF FOOD SCIENCE-Volume 48 (7983)


hydrophobicity, due probably to participation of some emulsifying properties (EAI and ESI) of ovalbumin and
of the exposed hydrophobic groups in hydrophobic inter- lysozyme by heat denaturation was also reported by Kato
actions. For casein, heating did not result in any substan- and Nakai (1980) and Kato et al. (1981).
tial change of its hydrophobicity and this was expected, Another example of protein denaturation resulting in
since casein exists in a random coil conformation (Morr, improvement of functionality was reported by Aoki et al.
1979). The results in Table 1 also indicate that for samples (1981), who determined the effect of alcohol modifica-
having the same solubility, the more hydrophobic the tion of soy protein on its emulsion stabilizing properties.
protein, the greater are its emulsifying properties, The soy protein was denatured with 50% alcohol by treat-
Looking, specifically, at the results of soy protein in ment at 35OC for 2 hr. The emulsion stabilizing properties
Table 1, it can be seen that, the EAI of all heated samples of soy protein modified with ethanol or n-propanol de-
were slightly greater than that of the control. The ESI, creased with increasing solubility, whereas the emulsion
however, was initially slightly increased by heating, but stabilizing properties of the unmodified soy protein in-
subsequently started to decrease as the solubility progres- creased. Aoki et al. (1981) attributed the improved emul-
sively decreased; that is, solubility became an increasingly sion stability brought about by the alcohol modification of
important factor controlling this property. The higher EAI soy protein to the perturbation and unfolding of the hy-
values of heated soy protein samples were probably due to drophobic interior structure of the native soy protein by
their increased hydrophobicity values. On the other hand, alcohol, and to the resulting increase in the exposed hydro-
the observed decrease in ES1 upon heating may be due to phobic amino acid residues. It should be noted, that the
a decreasein solubility as a result of aggregation. increased emulsion stabilizing properties of soy protein be-
The proteins studied here can be divided into four tween pH 2 and 7, achieved through alcohol modification
groups according to the effect of heating on their emulsi- by Aoki et al. (19 8 1), is very important because soy protein
fying properties. The first group includes BSA, gluten, and can be expected to play a significant role in the stabiliza-
whey protein whose EAI was not substantially affected tion of a wide range of food emulsions, all meat emul-
by heating, whereas ES1 was improved by heating. The sions falling well within these pH limits.
second group includes soy protein and myosin whose EAI
was slightly improved by heating, whereas the ES1 was Statistical analysis
decreased. The third group includes @-lactoglobulin, pea, Regression equations for predicting the emulsifying and
canola, and casein, whose emulsifying properties were ad- fat binding properties from hydrophobicity and solubility
versely affected by heating. Finally the fourth group in-
data of heat denatured proteins in Table 1 are presented in
cludes ovalbumin and gelatin whose emulsifying properties
Table 3. Simple linear regression analyses showed that the
were markedly improved upon heating. It is evident, there-
coefficients of determination (r*) between S, and EAI,
fore, that heating does not have uniform effects on the
and between Se and ES1 (detetmined without NaCl) were
emulsifying properties of different proteins.
highly significant (P < 0.001). No significant correlation
The improvement of emulsifying properties of gelatin was found, however, between solubility and EAI or ESI.
upon heating might have been due to the increase in solu-
bility since its hydrophobicity was not changed by heat- Moreover, no correlation was found between So or solu-
bility and FBC. Although significant correlations were
ing. Kato and Nakai (1980), and Kato et al. (1981) re- found between S, and ES1 as well as solubility and ES1
ported that the emulsifying properties (EAI and ESI) of
determined in the presence of O.lM NaCl, such correla-
ovalbumin and lysozyme were markedly improved upon tions are not considered meaningful since the effect of
heating. They also found that this improvement was cor-
O.lM NaCl on solubility and S, were not determined. It
related with the higher hydrophobicity of heat denatured
is reasonable that 0.1 M NaCl may positively or negatively
protein samples as compared to those of the native pro-
affect the solubility of a protein. Sodium chloride was
teins. Furthermore, their results indicated that the more
used in the ES1 test, to obtain a general idea on the salt
hydrophobic the proteins, the greater the decreasein inter-
sensitivity of the emulsions formed by different proteins.
facial tensions and the improvement in emulsifying proper-
ties. During the handling of the ovalbumin samples we Relatively high salt concentrations are used in some foods,
e.g. meat products. As seen in Table 1, O.lM NaCl usually
noticed that the apparent viscosities of heat-denatured
has a detrimental effect upon the emulsion stability. NaCl
samples increased. This is evident in Table 2, which shows
exerts its negative effect on emulsion stability probably by
the apparent viscosity changes of some proteins upon heat-
ing. It is suggested, therefore, that higher viscosity appears
to be another factor contributing to the great improvement Table 2 - Effect of heating on the apparent viscosity of some
of emulsifying properties of heat denatured ovalbumin proteinsa measured at 20°C
samples. Apparent viscosity
As shown in Table 1, the FBC of different proteins is Proteina 1Pa.s x 1 O-3)
differently affected by heating. While the FBC of Canola
Whey - control 1.14
protein is relatively unaffected by heating, the FBC’s of - sample 1 IpH 6.0, 80°C, 4 min) 1 .I4
BSA, /I-lactoglobulin, ovalbumin, whey protein, and casein - sample 2 (pH 6.0, 80°C, 5 min) 1.14
are positively affected, and the FBC’s of soy and pea pro- - sample 3 (pH 6.0, 80°C, 15 min) 1.16
teins are adversely affected by heating, - sample 4 (pH 5.8, 80°C, 15 min) 1.20
Protein denaturation usually decreases solubility then
adversely affects protein functionality. However, as shown BAS - control 1.13
in this study, emulsifying and fat binding properties of - heated (pH 4.0, IOO’C, 5 min) 1.24
some proteins (ovalbumin, whey protein) can be improved
by denaturation. According to Morr (1979) denaturation of Ovalbumin - control 1 .I0
- sample 1 (pH 5.6,80°C, 1.5 min) 1 .I3
the whey protein molecule, if produced at the proper stage - sample 2 (pH 5.6, 80°C, 2.0 min) 1.13
of the protein concentrate isolation/utilization process, can - sample 3 (pH 5.6, 8O’C. 2.5 min) 1.16
improve the functionality. The improvement in functional- - sample 4 (pH 5.6,80°C, 3.0 min) 1 .I 7
ity is probably due to an unfolding of the molecule to ex- - sample 5 (pH 5.6,lOO”C. 5 min) 1 .I8
pose hydrophogic amino acid residues, thus making the - sample 6 (pH 1 .O, 100°C. 15 min) 1.90
protein more amphiphilic and capable of orienting at the
oil-water interface (Morr, 1979). A great improvement of a0.5% protein in 0.01 M phosphate buffer, pH 7.4

Volume 48 /1983)-JOURNAL OF FOOD SCIENCE-29


I HYDROPHOBICITY AND EMULSIFYING PROPERTIES. . .

reducing the surface charge and by withdrawing surface al properties studied. As shown in Fig. 1, regardless of the
water of oil droplets. solubility, as hydrophobicity increases the EAI is initially
As seen in Table 3, multiple regression analysesbetween increased and then slightly declined. Although at low and
S,, solubility and EAI or ESI, show highly significant corre- medium hydrophobicity values, increasing solubility levels
lations (R* values are 0.542 and 0.434, P < 0.001, respec- increase the EAI, at very high S, values, solubility does not
tively). appear to play an important role in the emulsifying capac-
Since the Student’s t-values of partial regression coeffi- ity of proteins.
cients are highly significant, both hydrophobicity and solu- The ES1 of the proteins is significantly affected by So, s,
bility significantly affect the emulsifying properties of the the interaction of S, and s, and the s2 (Table 3). The R*
proteins. As shown in Table 3, the backwards stepwise value of the model is 0.584, indicating that 58.4% of the
multiple regression analysis gives a highly significant corre- variability in ES1 could be accounted for by these 4 vari-
lation between S,, solubility and FBC of the 48 protein ables. As shown in Fig. 2, regardless of the solublity, as
samples in Table 1. Comparing the regression models in S, increases the ES1 is initially increased and then de-
Table 3, it is apparent that the models obtained by step- creased. At low and medium So values, increasing solubility
wise regression analysis have higher R* values and lower index levels increases ESI. However, at very high S,, values,
S.E. (standard error of the estimate) values than those of solubility index does not appear to be important for the
the linear regression models. Our discussion, therefore, emulsion stability.
will be confined within the models calculated by stepwise The FBC of the proteins studied was significantly af-
regression analysis. fected by Se, s, and the squares of these two variables
The EAI of the proteins studied was significantly affected (Table 3). The R* of the model was 0.473 (P < O.OOl),
by So, solubility and the square of solubility (Table 3). indicating that 47.3% of the variability in FBC could be
The R* value for EAI was 0.583, indicating that 58.3% of accounted for by the four variables. The statistical signifi-
the variability in EAI of the protein studied could be cance of quadratic powers of S, and solubility in the FBC
accounted for by the three independent variables S,, s model in Table 3 indicates that as the value of S, or s
.and s*. In this model, the statistical significance of the increases, the FBC value increases first then declines (i.e.
square of solubility index indicates that as the value of solu- the response of FBC to increasing levels of S, or solubil-
bility index increases, the effect of that variable declines, ity can be depicted as a curvilinear graph).
i.e., the EAI response to increasing levels of solubility may The finding of this study that with increasing S, the
be depicted as curvilinear rather than linear. Response emulsifying properties initially increase and then decrease
surface plots (contours) were generated by the computer can be explained by taking into account the fact that the
to visualize the effects of S6 and solubility on the function- emulsifying properties of proteins ultimately depend on the

Table 3 - Regression models for prediction of emulsifying and fat binding properties of heat denatured proteins.
I
Dependent Regression Variable Regression
variable analysis description coefficient F-ratio Fprobability t-valuea p-value

EAI Simple Constnat 41.162


(n=52; r2=0.464, linear so 0.202
P<O.OOl ;
S.E.b=18.95)

EAI Multiple Constant 29.283


(n=52: R2~0.542. linear so 0.207 - 7.202X” 0.698
P<O.OOl ; S 0.226 - 2.887” 0.280
s.E.b=I 7.78)

EAI Backwards Constant 16.877 4.531 0.039


(n=52; R2~0.583, multiple SO 0.214 58.77 0.000 0.720
P<O.OOI ; nonlinear S 0.932 7.83 0.008 1.151
s.E.b=17.401 s2 -0.007 4.73 0.035 - -0.894

ESI Simple Constant 0.880


(n=49; r2=0.377, linear so 0.274
P<0.001;
S.E.b=30.98)

ESI Multiple Constant -14.308


(n=49; R2=0.434, linear so 0.278 5.613*** 0.623
P<O.OOl ; S 0.295 2.153 0.239
s.E.b=29.53)

ESI Backwards Constant -69.463 18.02 0.000


(n=49; R2=0.584, multiple so 0.565 34.31 0.000 1.268
P<O.OOI ; nonlinear 2.034 13.62 0.001 - I.651
S.E.b=26.70) i xs -0.004 10.08 0.003 - -0.779
52 -0.012 5.92 0.019 -1.044

FBC Backwards Constant 4.895 0.16 0.691 -


(n=48; R2=0.473. multiple so 0.451 13.84 0.001 I .445
P<O.OOl ; nonlinear s 1.398 13.84 0.001 I .602
S.E.b=20.97) sg -0.001 10.15 0.003 - -1.270
S2 -0.014 11.79 0.001 - -1 .J45

a*: ~03.05; **: P<O.OI; ***: P<O.OOI


btandard error of estimate

30-JOURNAL OF FOOD SCIENCE-Volume 48 (1983)


suitable balance between the hydrophile and lipophile, and migrate to, adsorb at, unfold, and rearrange at an interface.
do not necessarily increase as the proteins become more Therefore, solubility in the aqueous phase, is closely re-
lipophilic (Aoki et al., 198 1). Excessive denaturation of the lated to surface activity of proteins (Kinsella, 1979). Many
soy protein by n-propanol resulted in a lower emulsion sta- authors, on the other hand, have reported that high levels
bility as compared to the moderate denaturation by ethanol of solubility were not necessarily associated with maximum
(Aoki et al. 1981). There may be many factors, e.g., molec- emulsifying properties (Aoki et al., 1980; McWatters and
ular size, molecular flexibility and charge, besides the Cherry, 1975; McWatters and Holmes, 1979a, 1979b;
balance of hydrophile and lipophile which participate in Smith et al., 1973; Wang and Kinsella, 1976). Flint and
determining the emulsifying properties of proteins. Johnson (1981) investigated a film formation by soy pro-
Wolf and Cowan (1975) reported that in ground meat tein (isolate) at an oil-water interface for the pH range
products fat absorption or binding appeared to be involved l-10. Definite film formation was observed at all pH
in formation and stabilization of an emulsion. Thus, they values below the isoelectric point (4.6) of the protein
suggested that fat adsorption may simply be another and up to pH 6.5. At pH 5.4, despite the low solubility
aspect of emulsification. This helps to explain the obser- strong film formation occurred. However, at pH’s above 5.4
vation in this study that hydrophobicity has a curvilinear the strength of films formed gradually decreased until at an
effect also on FBC, as the concept of hydrophile-lipophile upper limit pH of 7.5, the presence of an interfacial layer
balance can be utilized again. Voutsinas and Nakai (1981) could barely be seen. The marked pH dependence of
already reported that with increasing S, the FBC was ini- film formation on the alkaline side of the isoelectric point
tially increased and then decreased. was attributed to the fact that with increasing pH the pro-
Pearson et al. (1965) reported that only that fraction of tein becomes more soluble in the aqueous phase and con-
protein which is soluble can function as an effective emulsi- sequently less likely to be brought out of solution (coagu-
fying agent. hloreover, Franzen and Kinsella (1976) sug- lated) at the phase boundary. The ability of soy protein to
gested that, as a protein becomes more soluble, it forms form a film even at very low pH’s, where the solubility of
layers around the fat droplet to facilitate association with protein is high, suggestedthat this phenomenon may not be
the aqueous phase. Granular, insoluble proteins, however, linked solely to solubility but also to the availability of
separate from the oil phase or just float on the oil surface lipophilic groups for binding at the oil-water interface.
where they remain inert and contribute little toward emul- Changes in the protein conformation may occur at acid
sification. Similarly, soluble proteins enclose the fat glob- pH’s which enhance the combination of the protein and oil
ules and render the emulsion more stable to heat treatment. leading to the formation of interfacial film described
Atso,, according to Bull (1972), the surface activity of a (Flint and Johnson, 1981). -Continued on nextpage
protein is a function of the ease with which the protein can

52 65 79, _

9 78 148 217 286 355 425

Fig. I-Emulsifying activity index response surface contour as a function of hydrophobicity (So) and solubility index (s).

7
9 78 148 217 286 355 425

SO
Fig. %Emukion stability index response surface contour as a function of hydrophobiciv (So) and solubijiw index (s).

Volume 48 (1983)-JOURNAL OF FOOD SCIENCE-31


HYDROPHOBICITY AND EMULSIFYING PROPERTIES. ..

Considering the above contradictory situation, the re- McWatters, K.H. and Cherry, J.P. 1975. Functional properties of
ueanut paste as affected by moist heat treatment of full-fat pea-
sults of this study showed that high solubility had a nega- nuts. J. Food Sci. 40: 1205..
tive effect on FBC in agreement with the results of Vout- McWatters. K.H. and Holmes, M.R. 1979a. Influence of pH and salt
sinas and Nakai (1982). They attributed the low FBC of concentration on nitrogen solubility and emulsification properties
of soy flour. J. Food Sci. 44: 770.
soluble proteins to their conformation (mainly helical) McWatters, K.H. and Holmes, M.R. 1979b. Influence of moist heat
which may not permit their binding sites to be sterically on solubility and emulsification properties of soy and peanut
flours. J. Food Sci. 44: 774.
available for interaction with oil or to the limited accessof Morr, C.V. 1979. Conformation and functionality of milk proteins.
oil (hydrocarbon chains) to the protein binding sites due to In “Functionality and Protein Structure,” Ed. A. Pour-El, Amer.
a possible barrier around them formed by the excessive Chem. Sot. SvmDosium Series 92. Washington. DC.
Nakai, S., Ho,-L.; Tung, M.A., aid Quinn, J.g. 1980a. Solubiliza-
number of protein polar groups. tion of rapeseed, SOY and sunflower protein isolates by surfactant
In conclusion, the emulsifying properties of proteins are and proteinase treatments. Can. Inst. Food Sci. Technol. J. 13: 14.
Nakai. S., Ho, L., Helbig. N., Kato, A.. and Tung, M.A. 1980b.
influenced by both hydrophobicity and solubility. Al- Relationship between hydrophobicity and emulsifying properties
though important, solubility alone cannot fully explain the of some plant proteins. Can. Inst. Food Sci. Technol. J. 13: 23.
Nakai, S. and Powrie, W.D. 1981. Modification of proteins for func-
emulsifying properties especially when the proteins are tional and nutritional improvements. In “Cereals: A Renewable
heat denatured. Resource ” Ed. Y. Pomeranz and L. Munck, Amer. Assoc. Cereal
Chem., S;. Paul, MN.
Pearse, K.N. and Kinsella, J.E. 1978. Emulsifying properties of pro-
REFERENCES teins: evaluation of a turbidimetric technique. J. Agric. Food
Chem. 26: 716.
Aoki, H., Taneyama, 0.. and Inami, M. 1980. Emulsifying proper- Pearson, A.M., Spooner, M.E., Hegarty, G.R.. and Brat&r, L.J.
ties of SOY protein: Characteristics of 7s and 11s proteins. J. 1965. Emulsifying capacity and stability of soy sodium pro-
Food Sci. 45: 534. teinate, potassium caseinate and non-fat dry milk. Food Tech-
Aoki, H., Taneyama, 0.. Orimo, N.. and Kitagawa, I. 1981. Effect nol. 19: 1841.
of lipophilization of soy protein on its emulsion stabilizing proper- Perry. S.V. 1955. Myosin adenosinetriphosphatase. In “Methods
ties. J. Food Sci. 46: 1192. in Enzymology,” Ed. S.P. Colowick and N.O. Kaplan, Vol. 2.
Blanshard, J.M.V. 1970. Stabilizers -their structure and properties. Academic Press, New York, NY.
J. Sci. Food Agric. 21: 393. Smith, G.C., Juhn, H., Carpenter, Z.L., Mattil, K.F., and Carter,
Brewer, J.M., Pesce, A.J., and Ashworth, R.B. 1974. “Experimental C.M. 1973. Efficacy of protein additives of emulsion stabilizers in
Techniques in Biochemistry.” Prentice-Hall, Inc.. Englewood frankfurters. J. Food Sci. 38: 849.
Cliffs. NJ. Swift, C.E. and Sulzbacher, W.L. 1963. Cornminuted meat emul-
Bull. H.B. 1972. Adsorbed surface films of egg albumin. J. Colloid sions: factors affecting meat proteins as emulsion stabilizers.
Interfacial Sci. 41: 305. Food Technol. 17: 224.
Crenwelge, D.D., Dill, C.W., Tybor, P.T.. and Landmann, W.A. Tanford, C. 1973. “The Hydrophobic Effect: Formation of Mice&s
1974. A comparison of the emulsification capacities of some pro- and Biological Membranes.” John Wiley & Sons, New York, NY.
tein concentrates. J. Food Sci. 39: 175. Volkert, M.A. and Klein, B.P. 1979. Protein dispersibility and emul-
Flint, F.O. and Johnson, R.F.P. 1981. A study of film formation sion characteristics of flour soy products. J. Food Sci. 44: 93.
by soy protein isolate. J. Food Sci. 46: 1351. Voutsinas. L.P. and Nakai, S. 1982. A simple turbidimetric method
Franzen, K.L. and Kinsella, J.E. 1976. Functional properties of suc- for determining the fat binding capacity of proteins. J. Agric.
cinylated and acetylated leaf protein. J. A&c. Food Chem. Food Chem. (In press).
24: 914. Wang,-J.C. and Kinselia, J.E. 1976. Functional properties of novel
Johnson, D.W. 1970. Functional properties of oilseed proteins. J. proteins: alfalfa leaf proteins. J. Food Sci. 41: 286.
Amer. Oil Chem. Sot. 47: 402. Wolf. W.J. and Cowan. J.C. 1975. “Sovbeans as a Food Source.”
Kato, A. and Nakai. S. 1980. Hydrophobicity determined by a CR’C Press, Inc., Clewland, OH. -
fluorescence probe method and its correlation with surface proper- Yasumatsu, K., Sawada, K., Moritaka, S., Misaki, M., Toda, J..
ties of proteins. Biochim. Biphys. Acta. 624: 13. Wada, T., and Ishii. K. 1972. Whipping and emulsifying properties
Kate, A., Tsutsui, N., Kobayashi, K., and Nakai, S. 1981. Effects of of soybean products. Agric. Biol. Chem. 36: 719.
partial denaturation on surface properties of ovalbumin and Ms received 5117182; revised 8/g/82; accepted g/6/82.
lysozyme. Agr. Biol. Chem. 45: 2755.
Keshavarz, E. and Nakai. S. 1979. The relationship between hydro- A part of this work was presented at the 41st Annual Meeting of
phobicity and interfacial tension of proteins. Biochim. Biophys. the Institute of Food Technologists, Atlanta, GA, June 7-10, 1981.
Acta. 575: 269. This research was conducted according to the terms of contract
Kinsella, J.E. 1976. Functional properties of proteins in foods: a no. OSU80-00026 for Agriculture Canada. (Scientific authority:
survey. Crit. Rev. Food Sci. Nutr. 7: 219. Dr. V. Harwalkar). The authors are grateful for the encouragement
Kinsella, J.E. 1979. Functional properties of soy proteins. J. Amer. given by late Dr. J.R. Quinn.
Oil Chem. Sot. 56: 242.

ELECTRICAL STIMULATION OF RABBITMUSCLE.. . From page 22

Locker, R.H., Davey, C.L., Nottingham, P.M.. Haughey, D.P., and trical stimulation on beef, pork. lamb and goat meat. Proc. 26th
Law, N.H. 1975. New concepts in meat processing. Adv. Food Eur. Meet. Meat Res. Work., Colorado Springs, CO, P. 19.
Res. 21: 157. Spudich, J.A. and Watt, S. 1971. The regulation of rabbit skeletal
Lowey, S. and Cohen, C. 1962. Studies on the structure of myosin. muscle contraction. J. Biol. Chem. 246: 4866.
J. Mol. Biol. 4: 293. Taylor, A.A., Shaw, B.G. and MacDougall, D.B. 1980. Hot debonlng
Mannherz, H.G., Brehme, H.. and Lamp, U. 1975. Depolymerization beef with and without electrical stimulation. Meat Sci. 5: 109.
of F-actin to G-actin and its repolymerization in the presence of Tonomura, Y., Tokura, S., Sekiya, K., and Imamura, K. 1961.
analogs of adenosine triphosphate. Eur. J. Biochem. 60: 109. Influence of solvent composition on the molecular shape and
Olson. D.G., Parrish, F.C., Dayton, W.R.. and Gall, D.E. 1977. Effect enzymic activity of myosin A. Arch. Biochem. Biophys. 95: 229.
of postmortem storage and calcium activated factor on the myo- Weber, K. and Osborn, M. 1969. The reliability of molecular weight
fibrillar proteins of bovine skeletal muscle. J. Food Sci. 42: 117. determinations by dodecyl sulfate-polyacrylamide gel electro-
Rizzino, A.A.. Barouch. W.W., Eisenberg. E. and Moos, C. 1970. phoresis. J. Biol. Chem. 244: 4406.
Actin-heavy meromyosin binding: determination of binding West, R.L. 1979. Effects of high temperature conditioning on
stoichiometry from ATPase kinetic measurements. Biochemistry muscle tissue. Food Technol. 33(4): 41.
9: 2402. Whiting, R.C., Strange, E.D., Miller, A.J., Benedict. R.C., Mozersky,
Savell, J.W., Smith, G.C.. Dutson, T.R., Carpenter, Z.L., and Suter. S.M., and Swift, C.E. 1981. Effects of electrical stimulation on the
D.A. 1977. Effect of electrical stimulation on palatability of beef. functional properties of lamb muscle. J. Food Sci. 46: 484.
lamb and goat meat. J. Food Sci. 42: 702. Will, P.A., Henrickson, R.L., Morrison, R.D., and Odell, G.V. 1979.
Shaw. F.D. and Walker, D.J. 1977. Effect of low voltage stimulation Effect of electrical stimulation on ATP depletion and sarcomere
of beef carcasses on muscle PH. J. Food Sci. 42: 1140. length in delay-chilled bovine muscle. J. Food Sci. 44: 1646.
Smith, G.C., Savell, J.W., Dutson. T.R., Hostetler, R.L., Terrel, MS received 6112182; revised g/30/82; accepted 10/6/82.
R.N., Murphey. C.E., and Carpenter. Z.L. 1980. Effects of elec-

~&JOURNAL OF FOOD SCIENCE-Volume 48 (1983)

You might also like