Anda di halaman 1dari 10

ARTICLE IN PRESS

Journal of Cereal Science 45 (2007) 275–284


www.elsevier.com/locate/yjcrs

Relationship between the glass transition temperature and the


melt flow behavior for gluten, casein and soya
Carlos Bengoecheaa, Abdessamad Arrachidb,, Antonio Guerreroa,
Sandra E. Hillb, John R. Mitchellb
a
Departamento de Ingenierı´a Quı´mica, Universidad de Sevilla, c/ P. Garcia Gonzalez 1, 41012 Sevilla, Spain
b
Division of Food Sciences, School of Biosciences, University of Nottingham, Sutton Bonington Campus, Loughborough LE12 5RD, UK
Received 29 January 2006; received in revised form 11 June 2006; accepted 25 August 2006

Abstract

The effects of moisture content (25–45% wwb) and temperature (75–120 1C) on the viscosity of gluten, soya and rennet casein systems
was studied using a capillary rheometer. An attempt was made to relate the viscosities to the glass transition temperature measured by
differential scanning calorimetry, dynamic mechanical thermal analysis and the phase transition analyzer. The temperature where the
material flowed was also determined by the latter technique. All three-protein systems showed shear and extension thinning. Over the
shear rate range investigated (1–103 s1), gluten had a substantially lower viscosity than the other two proteins, although the difference
was less pronounced at the highest temperature studied. This low viscosity is reflected by lower values of the glass transition temperature,
the melt flow temperature and the dynamic moduli E0 and E00 in the rubbery state. The results are discussed in terms of the structure and
heat induced changes for the three proteins and their relevance to food processing considered.
r 2006 Elsevier Ltd. All rights reserved.

Keywords: Glass transition; Viscosity; Rheology; Extrusion; Differential scanning calorimetry; Dynamic mechanical thermal analysis; Phase transition
analyzer; Plasticization; Protein

1. Introduction et al., 1991; Singh and Smith, 1999; Zhang et al., 1998). The
variables generally studied are temperature, shear rate and
The rheological behavior of protein ‘‘melts’’ at relatively water content (Colonna et al., 1989; Harper, 1981)
low water contents (15–50% wwb) is important for although time dependent changes such as protein associa-
extrusion processing to produce food products such as tions have sometimes been taken into account (Morgan
texturised vegetable proteins (TVP). In addition, rheologi- et al., 1989; Remsen and Clark, 1978).
cal behavior of gluten at low water contents is one factor More recently there has been an increasing interest in the
influencing baking performance. There have been extensive glassy state in foods (Blanshard and Lillford, 1993). The
studies on the viscosity of biopolymer melts using both glass transition temperature (Tg) has been measured
pressure capillary rheometry and on-line extrusion rhe- extensively for both carbohydrate and protein dominated
ometers (Bhattacharya, 1993; Breuillet et al., 2002; Fujio systems, generally using differential scanning calorimetry
(DSC) or dynamic rheological methods. Of particular
Abbreviations: C–K, Couchman–Karasz; DMTA, dynamic mechanical interest for food applications has been the plasticizing role
thermal analysis; DSC, differential scanning calorimetry; G–T, of water, sugars, polyols and other low molecular weight
Gordon–Taylor; PTA, phase transition analyzer; SPI, soya protein isolate; additives (Kalichevsky et al., 1992a, b, 1993; Morales and
TVP, texturised vegetable protein Kokini, 1997; Pommet et al., 2003; Roos, 1995; Zhang
Corresponding author. PTC NESTLE YORK, Haxby Road, York
YO91 1XY, UK. Tel.: +44 1904 60 31 16; fax: +44 1904 60 48 87.
et al., 2005).
E-mail address: Abdessamad.Arrachid@rdyo.nestle.com A technique that has been recently developed to provide
(A. Arrachid). information both about the glass transition temperature

0733-5210/$ - see front matter r 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jcs.2006.08.011
ARTICLE IN PRESS
276 C. Bengoechea et al. / Journal of Cereal Science 45 (2007) 275–284

and the melt rheology of biopolymer systems, particularly further information about the relationship between protein
within the context of extrusion processing, is the phase type, the glass transition temperature and the melt
transition analyser (PTA) (Plattner et al., 2001). In the rheology.
PTA a piston applies pressure to material initially in a
closed chamber and the temperature is increased. The glass 2. Materials and methods
transition temperature is associated with the point where
the material softens, giving rise to increased movement of 2.1. Materials
the piston. A flow temperature, Tf, can also be determined
by replacing the closed die at the base of the chamber by Rennet casein was obtained from Kerry Foods Ltd
one containing a small hole. Tf is taken as the temperature (High Protein Milk Extract, UK), gluten from RIBA S.A.
where the material flows as evidenced by a continuous (Glutenflor Supervital, Barcelona, Spain), and soya protein
movement of the piston at a constant temperature. For isolate (SPI) from Protein Technologies International
maize these two temperatures have been shown to differ for (SUPRO 500E, Leper, Belgium).
crops grown in different locations and have been related to
extrusion behavior (Plattner et al., 2001). 2.2. Methods
In this study, three protein systems were examined:
gluten, soya and casein. All three are used in extrusion 2.2.1. Sample preparation
processes, and gluten and soya are also important in baked 2.2.1.1. Capillary rheometry. Samples were prepared by
products. Large amounts of information can be found in adding the required amount of water to the powder and
the literature about their structure and their properties in mixing with a Kenwood mixer (KenWood Mixer KMC
extrusion (Bhattacharya and Hanna, 1986; Chen et al., 500, KenWood) to ensure homogeneity. Gluten forms a
1978; Damodaran and Paraf, 1997; Jao et al., 1978; viscoelastic dough when mixed with water; so to prepare a
Kitabatake and Doi, 1992). Gluten is an amorphous homogenous powder it was necessary to freeze the
mixture of polypeptides that can be divided according to hydrated material in liquid nitrogen, prior to milling it to
their functionality into monomeric and polymeric poly- a powder using a Knifeter 1095 Sample Mill (Foss Tecator,
peptides. Monomeric polypeptides consist mainly of Hoganas, Sweden). For all three proteins water content
storage proteins called gliadins. They can aggregate by was measured after each experiment performed with the
non-covalent interactions and contribute to the viscosity capillary rheometer and was found to match the predicted
and extensibility of gluten. The vast majority of polymeric water content.
polypeptides are found in the glutenin fraction. Their
amino acid compositions are similar to the gliadins. They 2.2.1.2. DSC and PTA. Samples were equilibrated over
are additionally stabilized by disulfide bonds, and tend to P2O5 or saturated salt solutions of MgCl2, Mg(NO3)2, KI,
contribute to the elasticity and strain hardening behavior NaCl or KNO3 or distilled water, which produced
of gluten (MacRitchie and Lafiandra, 1997; Weegels et al., equilibrium relative humidities (RH) of 0%, 32.8%,
1996). Soya consists of mainly 7S and 11S globulins. The 52.8%, 68.9%, 75.3%, 93.7% and 100%, respectively.
7S globulin is a trimer with a molecular weight around Samples were equilibrated at room temperature for at least
150–200 kD (Fukushima, 1991). The 11S globulin is seven days.
composed by six subunits and has a reported molecular
weight of 300–400 kD (Fukushima, 1991; Pearson, 1982). 2.2.1.3. DMTA. Samples were prepared by hydrating to
Soya is used to produce TVP (Zhang et al., 2001) and is 17% water at 100% RH overnight and then pressing to a
functionally involved in gelation and emulsification (Utsu- thickness of 0.5–1 mm in a mold under pressure
mi et al., 1997). Casein is the principal protein in bovine 3.1  103 kPa at a temperature between 70 and 90 1C.
milk and consists of four components, as1- (38%), as2- Samples were then cut into 20  8  1 mm strips and stored
(10%), b- (36%) and k-casein (13%), which have molecular over salt solutions of various relative humidities (as for
weights in the range of 19–26 kD and vary in hydro- DSC and PTA: Section 2.2.1.2) to obtain a variety of water
phobicity (Dalgleish, 1997). In milk, they are arranged in a contents. Samples were stored for at least a week before
micelle. In rennet-coagulated casein, micellar aggregation measurements were made. Following equilibration, the
by hydrophobic bonding is thought to precede coagulation. water content was checked for agreement with predicted
Rennet casein performs well in dry spinning processes values and the measured values were used in the
where it is extruded at temperatures of 70 1C and above interpretation of the results. Before the measurement, the
(Visser, 1988). sample was coated with silicone oil (Dow Corning, USA)
One objective of the work described in this paper is to to avoid water loss.
determine if more conventional methods for determining a
glass transition temperature, DSC, dynamic mechanical 2.3. Water content
thermal analysis (DMTA) and melt rheology (pressure
capillary rheometry) are consistent with the information Water contents were obtained by drying to constant
obtained from the PTA. A second objective was to obtain weight at 105 1C.
ARTICLE IN PRESS
C. Bengoechea et al. / Journal of Cereal Science 45 (2007) 275–284 277

2.4. Capillary rheometry heated a second time at 10 1C/min to 180 1C. Finally the
sample was cooled to 20 1C at 50 1C/min. The glass
A Rosand RH7 twin bore capillary rheometer (Rosand transition temperature, Tg, was determined as the tem-
Flowmaster RH7, Bohlin Instruments, UK) was used. The perature midpoint of the heat capacity change observed
sample was driven simultaneously through a capillary die during the second run.
(l ¼ 32 mm, + ¼ 0:5 mm, y ¼ 901) as well as an orifice die
(+ ¼ 0:5 mm, y ¼ 901). The samples (25%, 30%, 35%, 2.6. DSC data fitting
40% and 45% (wwb) of water content) were loaded in the
barrels, allowed to equilibrate for 10 min and extruded at To predict the Tg of mixtures from the Tg’s of the
different temperatures (65, 75, 85, 95, 120 1C). The components the empirical Gordon–Taylor (Gordon and
extrusion was carried out at ram speeds of 2, 5.63, 15.9, Taylor, 1952) (G–T) equation was used:
44.7, 126, 355 and 1000 mm/s. The pressure was recorded
W 1 T g1 þ KW 2 T g2
as a function of the piston speed. Tg ¼ , (7)
The data were fitted to the power law equation as used in W 1 þ KW 2
previous studies (Colonna et al., 1989; Harper, 1981; Singh where W is the mass fraction of each component, Tg is the
and Smith, 1999): glass transition temperature in Kelvin, and K is a system
constant inversely proportional to the plasticizing effect of
s ¼ m_gn , (1)
the diluent (1) on the protein (2). This model is suitable for
where s is the shear stress, g_ is the shear rate, n is the flow bicomponent systems (food component and water or other
behavior index and m is the consistency coefficient plasticizer) and has proven to be useful to fit DSC data for
(Brydson, 1981). starch, maltose, maltodextrins (Roos and Karel, 1991) and
The instrument software obtains the shear viscosity ðZ ¼ for cereal proteins (de Graaf et al., 1993; Kokini et al.,
s=_gc Þ from the Rabinowitsch corrected wall shear rate: 1995; Madeka and Kokini, 1996). For multicomponent
  systems, the Couchman–Karasz (Couchman and Karasz,
3n þ 1
g_ c ¼ g_ , (2) 1978) (C–K) equation has been used:
4n
where g_ the wall shear rate for a Newtonian fluid is given W 1 DC p1 T g1 þ W 2 DC p2 T g2
Tg ¼ . (8)
by W 1 DC p1 þ W 2 DC p2
 
4Q This equation requires the value of DCp of water, which
_g ¼ , (3) is subject to considerable debate (Kalichevsky et al.,
pr3
1992b). G–T equation is equivalent to the C–K equation
Q is the volumetric flow rate and r the capillary radius. if K ¼ DC p2 =DC p1 (4), and DCp is the change in heat
The wall shear stress is given by capacity observed at Tg. The values used for Tg1 and DCp1
ðPL  P0 Þr of water are 134 K (139 1C) and 1.94 J g1 K1, respec-
s¼ , (4) tively (Kalichevsky et al., 1993; Sugisaki et al., 1968).
2LL
Values of DCp2 for the dry protein (casein, soya, gluten)
where PL is the pressure drop across the long die, length
were calculated from Eq. (3), assuming a value of
LL, and P0 the pressure drop across a zero length die
1.94 J g1K1 for DCp1 for water.
obtained from the pressure drop across the orifice die.
The Cogswell method (Cogswell, 1972) is used to obtain
an extensional viscosity ðl ¼ se =Þ, where the extensional 2.7. PTA
stress is given by
The PTA is a closed-chamber capillary ‘‘rheometer’’.
3ðn þ 1Þ One of the main differences between PTA and a capillary
se ¼ P0 (5)
8 rheometer is that the latter extrudes at constant piston
and the extensional strain rate by speed while the PTA performs at constant pressure.
It consists of two sealed chambers, top and bottom,
4 Z_g2
¼ . (6) separated by an interchangeable capillary die. The two
3 ðn þ 1ÞP0 chambers house electric heaters and contain a hollow
cavity that allows a cooling fluid to be used. The pistons,
mounted together as sidebars, are held in a fixed position
2.5. DSC during testing. Air cylinders, mounted to the bottom of the
PTA, maintain constant pressure on the sample. A linear-
Calorimetric measurements were performed using a displacement transducer measures the samples’ deforma-
Perkin Elmer DSC-7 (Perkin Elmer, UK). An empty tion, compaction, and flow relative to initial sample height
stainless steel pan was used in the reference holder. The (Plattner et al., 2001). In this study the measurements of Tg
sample was heated first time at 10 1C/min from 60 to and Tf were carried out at a heating rate of 5 1C/min and an
180 1C. It was then cooled to 60 1C at 50 1C/min and applied pressure of 150 bar.
ARTICLE IN PRESS
278 C. Bengoechea et al. / Journal of Cereal Science 45 (2007) 275–284

2.8. Dynamic mechanical thermal analyzer 10


Casein
9
Soya
The dynamic mechanical thermal analyzer (Rheometric 8

Displacement (mm)
7 Gluten
Scientific, Piscataway, USA) was used with the sample
presented in the single cantilever-bending mode at a 6
frequency of 1 Hz and a strain of 0.03% which was in the 5
linear region. The heating rate used was 3 1C/min. The 4
results from two runs were averaged at each water content. 3
2
1
3. Results
0
0 20 40 60 80 100 120 140 160 180 200
3.1. Glass transition temperature Temperature (C)

Fig. 1 displays the DSC response for the second DSC Fig. 2. Displacement temperature plots from the phase transition analyser
scan for the three proteins at a single, comparable moisture indicating position of glass transition temperature. Moisture contents
were: soya (13.7%, wwb), casein (12.9%, wwb) and gluten (11.7%, wwb).
content. All three proteins showed an endothermic peak
Applied pressure: 150 bar. Heating rate: 5 1C/min.
immediately above the glass transition temperature in the
DSC first scans. This was particularly pronounced for
gluten (first run scan shown in Fig. 1). This endothermic bulk density and the lower degree of compression as the
peak has been reported previously for a number of proteins temperature increases for gluten compared with the other
including gluten subunits (Castelli et al., 2000) and is two proteins. This may suggest that even in the glassy state,
probably due to physical aging (enthalpy relaxation) an gluten is more deformable than the other proteins but such
effect well known for synthetic polymers (Drozdov, 2001; an interpretation needs to be treated with caution as the
Hay, 1993; Toufeili et al., 2002). The midpoint of the heat gluten samples were comminuted in a different way
capacity change from the reheat was taken as the glass (freezing followed by subsequent milling).
transition temperature. This change was less clear for soya From DMTA data, as a function of temperature the
than for casein and gluten. Morales and Kokini (1997) glass transition temperature can be taken as the tempera-
have reported that at, for example, a water content of 15% ture where a decrease occurs in E0 or the maximum in tan d
the glass transition temperature of the 11S globulin is ( ¼ E00 /E0 ) or E00 . In this work, the E00 maximum was taken
about 40 1C higher than the 7S globulin. Since the soya as the glass transition temperature since this falls in
isolate used in this work is a mixture of these two proteins, between the other two temperatures. Fig. 3 shows the
this may explain the breadth of the transition. tan d, E0 and E00 responses for all three proteins at
Fig. 2 shows the piston displacement as a function of comparable water contents. A peak appears in the DMTA
temperature from the PTA. As the material begins to scans around 60 1C independent of water content (Fig. 3).
soften, the piston movement increases more rapidly with Kalichevsky et al. (1992b) also reported this low tempera-
temperature. The glass transition temperature was taken as ture transition. They stated that this may be due to the
the temperature where the derivative of the displacement onset of short range motions, whereas the glass transition is
temperature plot was a maximum. An interesting feature of the onset of main chain motion. Di Gioia et al. (1999)
the displacement temperature response is the higher initial indicated two weak relaxations at 65 and 25 1C in the
DMTA scans attributed to secondary relaxation of
proteins. There is considerable evidence from a range of
Gluten 1st run
techniques for a glassy transition in all polypeptide water
Heat flow - Endothermic

Gluten
systems at a temperature around 200 K (Ringe and Petsko,
2003) and it is possible that this mechanical relaxation is
Casein another manifestation of this. Whereas this mobility
transition may be relevant to the low temperature storage
of biological materials, it does not influence the main
Soya rheological responses of the material. Gluten shows a
higher decrease in the storage modulus 102 Pa and also a
considerably higher tan d peak than found for the other
two proteins in the glass transition region. Our results are
0 20 40 60 80 100 120 140 160 180 consistent with data of Kalichevsky et al. (1993) who
Temperature (C) compared the tan d and E0 responses for gluten, casein,
caseinate, ovalbumin and gelatin around the glass transi-
Fig. 1. DSC response for the three proteins during the reheat. The first
run for gluten only is represented. Moisture contents were: soya (13.7%, tion. They also found a fall in E0 of approximately 102 Pa
wwb), casein (12.9%, wwb) and gluten (11.7%, wwb). Heating rate: 10 1C/ and a tan d maximum close to 0.75 at the glass transition of
min. gluten at a water content of 15%. These data also
ARTICLE IN PRESS
C. Bengoechea et al. / Journal of Cereal Science 45 (2007) 275–284 279

Fig. 3. It is appropriate to add, however, that Pouplin et al.


1.E+10 (1999) reported that the tan d maximum varied with water
Soya
content and also reported an absolute value for E0 which
Casein was 102 Pa lower than shown in Fig. 3 for the sample
1.E+09
Gluten containing 11.7% water.
Fig. 4 compares the Tg values for the three proteins from
E' (Pa)

1.E+08 the different techniques. DMTA gave the highest Tg and


DSC the lowest. Kalichevsky et al. (1992a) also indicated a
1.E+07 higher value for the glass transition temperature obtained
from DMTA at 1 Hz than the glass transition temperature
from DSC, though the differences were smaller than in our
1.E+06
-100 -50 0 50 100 150 200 study. Of particular interest are the values given by the
Temperature (C)

a
1.E+09 200
CASEIN DMTA
Soya 180
DSC
Casein 160

Temperature (C)
PTA
Gluten 140
1.E+08
120
E" (Pa)

100
80
1.E+07 60
40
20
1.E+06 0
-100 -50 0 50 100 150 200 0 5 10 15 20 25 30
Temperature (C) Moisture content (%, wwb)

b
0.9 200
DMTA
Socv
180 SOYA
0.8 DSC
0.7
Casein 160
Temperature (C)

PTA
0.6
Gluten 140
Tan delta

0.5 120
0.4 100
0.3 80
0.2 60
40
0.1
20
0
-100 -50 0 50 100 150 200 0
0 5 10 15 20 25 30
Temperature (C)
Moisture content (%, wwb)
Fig. 3. E0 and E00 and tan d responses from DMTA for the three proteins.
Moisture contents were soya (13.7%, wwb), casein (12.9%, wwb) and
c
200
gluten (11.7%, wwb). Frequency: 1 Hz. Heating rate: 3 1C/min. GLUTEN DMTA
180
DSC
160
Temperature (C)

PTA
140
suggested a much lower glass transition temperature for 120
gluten than obtained for casein and caseinate, as also found 100
in our work. Of the proteins in this investigation, only 80
gelatin gave a comparable E0 decrease to gluten and 60
showed a large tan d peak (0.9). The mechanically 40
measured glass transition temperature for gelatin as judged 20
from the temperature of the tan d peak was higher than 0
0 5 10 15 20 25 30
gluten but lower than the casein systems. This relatively
Moisture content (%, wwb)
large fall in modulus of 102 Pa has also been reported for
dry gluten by Pouplin et al. (1999), who found a value of Fig. 4. Comparison of the effect of the water content on Tg values
the tan d maximum of 0.75 compared with 0.8 shown in measured using DSC, DMTA and PTA for casein, soya and gluten.
ARTICLE IN PRESS
280 C. Bengoechea et al. / Journal of Cereal Science 45 (2007) 275–284

PTA. This measures a temperature where the modulus of 250


Tg CASEIN
the powder decreases and might be expected to relate most
Tf
closely to the temperature where the storage modulus E0 200

Temperature (C)
falls. The lower temperature for the PTA determined Tg
compared with the DMTA value taken from the E00 150
maximum is consistent with this. It should be appreciated
that the glass transition is a kinetic phenomenon which will 100
depend not only on the criteria used to but also on the time
scale of measurement. 50
All three techniques show that gluten has the lowest glass
0
transition temperature at all moisture contents, and casein 0 5 10 15 20 25 30
the highest. Table 1 shows the results from the fitting of the Moisture content (%, wwb)
G–T to the DSC values for Tg. The glass transition
predicted for dry gluten using this approach is in good 250
SOYA Tg
agreement, within experimental error, with the values Tf
200
reported in the literature, (433 K (Hoseney et al., 1986);

Temperature (C)
435 K (Kalichevsky et al., 1992b); 412–448 K (for different
150
subunits of gluten) (Noel et al., 1995; Sartor and Johari,
1996); 423 K (Cherian and Chinachoti, 1996); 460 K 100
(Pouplin et al., 1999); 448 K (Micard et al., 2001); 401 K
(Toufeili et al., 2002). Similar agreement was found with 50
the Tg values for dry casein by Kalichevsky et al. (1992b)
(Tg ¼ 417 K) and Mizuno et al. (1999) (Tg ¼ 408 K)) and 0
for soya (387 (7S)–433 (11S) (Morales and Kokini, 1997); 0 5 10 15 20 25 30
423–473 K (by DMA) (Ogale et al., 2000)). Moisture content (%, wwb)
The very high Tg in the absence of plasticizers compared 250
with most synthetic polymers (Aklonis and MacKnight, Tg GLUTEN
1983; Ferry, 1970) could be explained by the presence of a Tf
200
Temperature (C)

significant incidence of polymer-polymer interactions, by a


high density of hydrogen bonding, or by ionic or 150
hydrophobic interactions (Pouplin et al., 1999).
100

3.2. Flow temperature (Tf)


50

Fig. 5 compares Tf and Tg both measured at a pressure 0


of 150 bar for the three proteins at different moisture 0 5 10 15 20 25 30
contents. It can be seen that gluten which has the lowest Tg, Moisture content (%, wwb)
also starts to flow at much lower temperatures than the
other two proteins. For example, at a water content of 15% Fig. 5. Effect of the moisture content on Tg and the melt flow temperature
Tf for casein, soya and gluten. Tg and Tf were obtained at a pressure of
Tf for gluten is 75 1C compared with 140 1C for soya 150 bar and a heating rate of 5 1C/min.
and 120 1C for casein. Another important feature is the
constant difference in temperature between Tg and Tf for soya can be related to the longer transition observed in the
the three proteins irrespective of the water content. This DSC plot as well as in the DMTA (E0 ).
difference is 37 1C for gluten, 39 1C for casein and
75 1C for soya. This result suggests a strong link between 3.3. Capillary viscosity
the glass to rubber transition and the ability of the proteins
to flow. The higher difference between Tg and Tf found for Fig. 6 shows the results for the shear and extensional rate
dependence of the viscosity from the capillary rheometer at
Table 1 the highest and the lowest water contents (45% (wwb) and
Gordon-Taylor and Couchman-Karasz fitting parameters obtained for 25% (wwb)) and at temperatures of 120 and 75 1C. All the
gluten, soya and casein proteins show shear and extensional thinning behaviors for
Protein Tg (K) K DCp [k ¼ DC p2 =DC p1 ] (J g1 K1) all the conditions studied. The extensional viscosities are
much higher than the shear viscosities with a Trouton ratio
Gluten 434.9 4.9370.61 0.394 of 10–70 at a common deformation rate of vicinity of
Soya 445.0 4.4270.79 0.439
1.101 s1. In all cases, the shear viscosity at the lower shear
Casein 479.4 4.9770.73 0.391
rates is much lower for gluten than it is for the other two
ARTICLE IN PRESS
C. Bengoechea et al. / Journal of Cereal Science 45 (2007) 275–284 281

1.E+09 1.E+10
Extensional and Shear viscosity

Extensional or shear viscosity


1.E+08 1.E+09
1.E+07 1.E+08
1.E+07
1.E+06
1.E+06
1.E+05
(Pa.s)

(Pa.s)
1.E+05
1.E+04
1.E+04
1.E+03
Casein 1.E+03 Casein
1.E+02 1.E+02
Soya Soya
1.E+01 1.E+01
Gluten Gluten
1.E+00 1.E+00
1.E-03 1.E-02 1.E-01 1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 1.E-03 1.E-02 1.E-01 1.E+00 1.E+01 1.E+02 1.E+03 1.E+04
Extensional or shear rate (1/s) Extensional or shear rate (1/s)

1.E+09
1.E+10

Extensional or shear viscosity


Extensional or shear viscosity

1.E+09 1.E+08
1.E+08 1.E+07
1.E+07 1.E+06
1.E+06 1.E+05

(Pa.s)
(Pa.s)

1.E+05 1.E+04
1.E+04
1.E+03
1.E+03 Casein Casein
1.E+02
1.E+02 Soya Soya
1.E+01
1.E+01 Gluten Gluten
1.E+00 1.E+00
1.E-03 1.E-02 1.E-01 1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 1.E-03 1.E-02 1.E-01 1.E+00 1.E+01 1.E+02 1.E+03 1.E+04

Extensional or shear rate (1/s) Extensional or shear rate (1/s)

Fig. 6. Extensional and shear viscosity of the three proteins at different temperatures and moisture contents. Extensional viscosities are represented by the
highest values. Conditions were: (a) 75 1C–25% (wwb); (b) 75 1C–45% (wwb); (c) 120 1C–25% (wwb); (d) 120 1C–45% (wwb).

proteins. The difference is greatest for the conditions Table 2


closest to the glass transition temperature (75 1C and 25% Model fitting parameters for the capillary shear data for gluten, soya and
casein
moisture content) with gluten having a shear viscosity
between one and two orders of magnitude lower than soya Protein K0 n DE (kJ/mol) a R2
and casein. This is consistent with the Tf values measured
with the PTA which at 20% water is about 50 1C lower Gluten 2340 0.45 5.70 4.96 0.78
Soya 7.86 0.34 30.3 5.83 0.91
than the other two proteins. The glass transition tempera-
Casein 0.399 0.27 29.8 8.04 0.95
ture is also lower for gluten than the other two proteins as
previously discussed. Soya always gives the highest values
of extensional viscosities compared with the two other that the constants obtained will be dependent on the
proteins. The difference is highest for the low temperature. time–temperature history and thus should be regarded only
Some general appreciation of the effect of deformation as an empirical representation of the data set. As
rate, moisture content and temperature on viscosity can be mentioned in the introduction, models have been proposed
obtained by fitting an equation of the form (Colonna et al., to take crosslinking into account (Morgan et al., 1989;
1989): Remsen and Clark, 1978), but to include this requires a
much more complex and extensive experimental procedure.
Z ¼ k0 expDE=RT expaMC g_ n1 , (9)
Tables 2 and 3 show the fitted values for the constants in
where Z is the viscosity, k0, a constant, DE, the activation Eq. (9) for the shear and extensional viscosities, respec-
energy, MC, the moisture content of the sample, R, the tively. The values of n show that the materials appear to
universal gas constant, T, the absolute temperature, g_ in show more extension thinning than shear thinning as it can
this case is the shear or extensional rate, n, the flow be clearly seen from Fig. 6. The values also show that, in
behavior index, and a, a constant, to the data obtained at general, the gluten shear viscosity is less dependent on
all five temperatures and moisture contents. Since dena- temperature and moisture content than the other proteins,
turation and time dependent crosslinking reactions will whereas the temperature dependence of the extensional
occur at the higher temperatures it should be appreciated viscosity is much higher for soya.
ARTICLE IN PRESS
282 C. Bengoechea et al. / Journal of Cereal Science 45 (2007) 275–284

Table 3 (Kalichevsky et al., 1993). The relationship between


Model fitting parameters for the capillary extensional data for casein, polysaccharide structure and a calorimetrically determined
gluten and soya
Tg has been extensively discussed by Bizot et al. (1997) who
Protein K0 n DE (kJ/mol) a R2 consider the increase in Tg with various types of chain
immobilization. For a 1-4 linked polysaccharide chains, it
Gluten 22000 0.17 3.02 6.78 0.95 is not possible to observe a glass transition calorimetrically
Soya 0.050 0.16 42.5 7.38 0.94
possibly because of specific inter residue hydrogen bonds
Casein 3790 0.21 14.8 8.69 0.88
(Gidley et al., 1993). Other interactions which contribute to
immobilize the polymer e.g. crystallization, the presence of
4. Discussion ordered secondary structure (e.g., b-sheets or a-helices in
proteins), covalent crosslinks e.g., disulfide bonds, and
Results showed the difficulty to relate viscosity (shear other interchain non-covalent interactions sometimes
and extensional) behavior to glass transition temperature. called hyperentanglements (Morris et al., 1981) would be
The extensional viscosity is related to the pressure drop at expected to increase Tg, and if they are preserved at
the entrance to the capillary of the capillary rheometer and temperatures above Tg, they would reduce the size and
will be influenced by features such as the resistance to increase the width of the glass transition. All these effects
molecular elongation and the extent of molecular entangle- will increase the viscosity.
ment between these molecules. Extensional viscosity does The low shear viscosity of gluten relative to the two
not appear to relate to the glass transition temperature in other proteins studied along with the higher value of tan d
the way that shear viscosity does. It might be suggested and low moduli in the rubbery region, suggests that the
that a higher then expected Trouton ratio for gluten could immobilizing interactions are lower in this protein system
be related to the strain hardening properties of gluten in than in most other biopolymers i.e., it is closer to an ideal
extension (Breuillet et al., 2002), whereas in comparison synthetic polymer. The reason for this is not clear. The
with casein, soya possess more ordered globular structures greater hydrophobicity as shown by the lack of water
resisting orientation. solubility may be a factor. Another possibility could be
Of particular interest is the very low shear viscosity differences in the content of proline+hydroxyproline
found for the gluten melt compared with the other two residues in the proteins. These residues are not compatible
proteins. The zero shear viscosity is related to the stress with most ordered secondary structures (a-helix, b-sheet)
relaxation modulus G(t) by the expression: and, therefore, will be another factor contributing to an
Z 1 expanded disordered structure rather than one that is
Z¼ tGðtÞd ln t, (10) compact and ordered (Mohammed et al., 2000). Gluten
0 and gelatin, which also show a large fall (102 Pa) in the
where the frequency dependence of G0 mirrors the time storage modulus at the glass transition (Kalichevsky et al.,
dependence of the stress relaxation modulus G(t) (Ferry, 1993), both have a higher praline/hydroxyproline content
1970) (high frequencies will be equivalent to short times than soya and casein (Mohammed et al., 2000).
and the extensional modulus EðtÞ ¼ 3GðtÞ). If it is assumed It should be appreciated that the range of factors
that time temperature superposition holds, then a low contributing to the immobilization of the polypeptide
viscosity relates both to a low Tg, a low value of G* or G0 in chain makes a detailed interpretation difficult. For
the rubbery state, a narrow transition and a narrow example, Mizuno et al. (2000) demonstrated an increase
rubbery region in terms of time or temperature. For a in Tg on crosslinking casein with microbial transglutami-
completely amorphous high molecular weight polymer nase as would be expected, but a counter-intuitive decrease
with increasing temperature, there will be a glass to rubber for soya. The latter was attributed to a reduction in b sheet
transition which is evidenced by a fall in the elastic secondary structure associated with the crosslinking reac-
modulus of about 103 Pa to a value governed by the density tion. An important complicating factor is that heating will
of non-specific and ill defined chain entanglements result in additional crosslinking. This is indicated by
(Aklonis and MacKnight, 1983; Ferry, 1970). In addition exotherms observed in DSC studies of gluten at low water
to the fall in modulus, the glass transition is evidenced by a contents and discussed in detail by Sartor and Johari
peak in tan d. The height of the tan d peak relates (1996). This behavior of gluten protein could be explained
quantitatively to the volume fraction of the relaxing phase by disulfide/thiol exchange and thiol oxidation that take
and for a completely amorphous polymer the height of the place upon heating, leading to an increase of the covalent
peak is 1.0, and the breadth of this transition is typically coupling by disulfide bonds of gluten proteins (Morel et al.,
10–20 1C (Wetton et al., 1986). The extensive studies of 2000; Schofield et al., 1983; Weegels and Hamer, 1998).
water plasticised biopolymers have invariably shown Crosslinks involving tyrosine have also been suggested
smaller and broader transitions than would be expected (Tilley et al., 2001).
for an ideal high molecular weight synthetic polymer. For Protein crosslinking, which in general would be expected
example, the fall in modulus is often only an order of to increase with temperatue and water content, will
magnitude and the height of the tan d peak is less than 0.5 increase the viscosity of the protein melt.
ARTICLE IN PRESS
C. Bengoechea et al. / Journal of Cereal Science 45 (2007) 275–284 283

The relative contributions of the glass transition and Chen, A.H., Jao, Y.C., Larkin, J.W., Goldstein, W.E., 1978. Rheological
crosslinking reactions can be understood qualitatively from model of soy dough in extrusion. Journal of Food Process Engineering
Eq. (10). If the time dependence of the relaxation modulus 2, 337–342.
Cherian, G., Chinachoti, P., 1996. H and O Nuclear magnetic resonance
is considered, crosslinking will increase both the width and study of water in gluten in the glassy and rubbery state. Cereal
height of the rubbery plateau and an increase in the glass Chemistry 73, 618–624.
transition temperature will increase the width of the glassy Cogswell, F.N., 1972. Measuring extensional rheology of polymer melts.
zone. Both these effects will increase viscosity, which is an Transactions of the Society of Rheology 16, 383–403.
integral of the stress relaxation modulus over time. The Colonna, P., Tayeb, J., Mercier, C., 1989. Extrusion cooking of starch and
starchy products. In: Mercier, C., Linko, P., Harper, J.M. (Eds.),
observation that under all conditions gluten has a lower Extrusion Cooking. American Association of Cereal Chemists, St.
shear viscosity than the other two proteins, particularly at Paul, MN, p. 247.
low shear rates where Eq. (10) is likely to be most relevant, Couchman, P.R., Karasz, F.E., 1978. A classical thermodynamic
suggests that what dominates its low viscosity is the low Tg. discussion of the effect of composition on glass transition tempera-
The low values for DE and a for gluten compared with the tures. Macromolecules 11, 117–119.
Dalgleish, D.G., 1997. Structure-function relationship of caseins. In:
other two proteins is probably due to crosslinking reactions Damodaran, S., Paraf, A. (Eds.), Food Proteins and their Applica-
becoming more dominant at higher temperatures and tions. Marcel Dekker Inc., New York, Basel, Hong Kong,
water contents resulting in a lower decrease in viscosity pp. 199–223.
with increasing temperature and water content. Cross- Damodaran, S., Paraf, A., 1997. Food Proteins and their Applications.
linking could explain the increase in the modulus (E0 ) seen Marcel Dekker Inc., New York, Basel, Hong Kong, p. 681.
de Graaf, E., Madeka, H., Cocero, A.M., Kokini, J.L., 1993. Determina-
for casein and gluten at high temperatures (Fig. 3) though tion of the effect of moisture on gliadin glass transition using
water loss might also be a factor. mechanical spectrometry and differential scanning calorimetry. Bio-
technology Progress 9, 210–213.
Di Gioia, L., Cuq, B., Guilbert, S., 1999. Thermal properties of corn
5. Conclusions gluten meal and its proteic components. International Journal of
Biological Macromolecules 24, 341–350.
The PTA gives values for the glass transition tempera- Drozdov, A.D., 2001. Physical aging in amorphous polymers: comparison
ture and also the temperature of flow which are consistent of observations in calorimetric and mechanical tests. European
with the information obtained from DSC, DMTA and Polymer Journal 37, 1379–1389.
Ferry, J.D., 1970. Viscoelastic Properties of Polymers. Wiley, New York,
capillary rheometry. Gluten gives a much lower shear London, Sydney, Toronto.
viscosity than soya and casein which is partly related to a Fujio, Y., Hayashi, N., Hayakawa, I., 1991. Effect of moisture content on
lower Tg and the involvement of more of the protein in the flow behaviour of molten soy-protein isolate under an elevated
mobilization occurring above Tg. This is shown not only by temperature. International Journal of Food Science and Technology
transition techniques (DSC, DMTA, capillary rheometry) 26, 45–51.
Fukushima, D., 1991. Structures of plant storage proteins and their
but also by the measurements of Tg and Tf using the PTA.
functions. Food Reviews International 7, 353–381.
This may be expected to have implications not only for Gidley, M.J., Cooke, D., Ward-Smith, S., 1993. Low moisture poly-
extrusion processing but for the performance of gluten in saccharide systems: thermal and spectroscopic aspects. In: Blanshard,
baked products. J.M.V., Lillford, P.J. (Eds.), The Glassy State in Foods. Nottingham
University Press, Nottingham, pp. 303–315.
Gordon, M., Taylor, J.S., 1952. Ideal copolymers and the 2nd-order
References transitions of synthetic rubbers.1. Non-crystalline copolymers. Journal
of Applied Chemistry 2, 493–500.
Aklonis, J.J., MacKnight, W.J., 1983. Introduction to polymer viscoelas- Harper, J.M., 1981. Extrusion in foods. CRC, Boca Raton, USA.
ticity. Wiley, New-York. Hay, J.N., 1993. Physical ageing in polymer blends. In: Blanshard, J.M.V.,
Bhattacharya, M., 1993. Slit rheometer studies of wheat-flour dough. Lillford, P.J. (Eds.), The Glassy State in Foods. Nottingham
Journal of Texture Studies 24, 391–409. University Press, Nottingham, pp. 269–279.
Bhattacharya, M., Hanna, M.A., 1986. Viscosity modelling of dough in Hoseney, R.C., Zeleznak, K.J., Lai, L.S., 1986. Wheat gluten: a glassy
extrusion. Journal of Food Technology 21, 167–174. polymer. Cereal Chemistry 63, 285–286.
Bizot, H., Le Bail, P., Leroux, B., Davy, J., Roger, P., Buleon, A., 1997. Jao, Y.C., Chen, A.H., Lewandowski, D., Irwin, W.E., 1978. Engineering
Calorimetric evaluation of the glass transition in hydrated linear and analysis of soy dough rheology in extrusion. Journal of Food Process
branched polyanhydroglucose compounds. Carbohydrate Polymers Engineering 2, 97–112.
32, 33–50. Kalichevsky, M.T., Jaroszkiewicz, E.M., Ablett, S., Blanshard, J.M.V.,
Blanshard, J.M.V., Lillford, P.J., 1993. The Glassy State in Foods. Lillford, P.J., 1992a. The glass transition of amylopectin measured by
Nottingham University Press, Nottingham. DSC, DMTA and NMR. Carbohydrate Polymers 18, 77–88.
Breuillet, C., Yildiz, E., Cuq, B., Kokini, J.L., 2002. Study of the Kalichevsky, M.T., Jaroszkiewicz, E.M., Blanshard, J.M.V., 1992b. Glass
anomalous capillary Bagley factor behavior of three types of wheat transition of gluten. 1. Gluten and gluten sugar mixtures. International
flour doughs at two moisture contents. Journal of Texture Studies 33, Journal of Biological Macromolecules 14, 257–266.
315–340. Kalichevsky, M.T., Blanshard, J.M.V., Marsh, R.D.L., 1993. Application
Brydson, J.A., 1981. Flow Properties of Polymer Melts. George Godwin, of mechanical spectroscopy to the study of glassy biopolymers and
London. related systems. In: Blanshard, J.M.V., Lillford, P.J. (Eds.), The
Castelli, F., Gilbert, S.M., Caruso, S., Maccarone, D.F., Fisichella, S., Glassy State in Foods. Nottingham University Press, Nottingham,
2000. Thermoanalytical characterisation of high molecular weight pp. 133–156.
glutenin subunits. Water effect on their glass transition. Thermo- Kitabatake, N., Doi, E., 1992. Denaturation and texturization of food
chimica Acta 346, 153–160. protein by extrusion cooking. In: Kokini, J.L., Ho, C.-T., Karwe,
ARTICLE IN PRESS
284 C. Bengoechea et al. / Journal of Cereal Science 45 (2007) 275–284

M.V. (Eds.), Food Extrusion Science and Technology. Marcel Dekker Remsen, C.H., Clark, J.P., 1978. A viscosity model for a cooking dough.
Inc., New York, Basel, Hong Kong, pp. 361–372. Journal of Food Process Engineering 2, 39–64.
Kokini, J.L., Cocero, A.M., Madeka, H., 1995. State diagrams help Ringe, D., Petsko, G.A., 2003. The ‘‘glass transition’’ in protein dynamics:
predict rheology of cereal proteins. Food Technology 49 (74), 76–82. what it is, why it occurs, and how to exploit it. Biophysical Chemistry
MacRitchie, F., Lafiandra, D., 1997. Structure-function relationships of 105, 667–680.
wheat proteins. In: Damodaran, S., Paraf, A. (Eds.), Food Protein and Roos, Y.H., 1995. Characterization of food polymers using state
their Applications. Marcel Dekker Inc., New York, Basel, Hong diagrams. Journal of Food Engineering 24, 339–360.
Kong, pp. 293–324. Roos, Y.H., Karel, M., 1991. Water and molecular weight effects on the
Madeka, H., Kokini, J.L., 1996. Effect of glass transition and crosslinking glass transition in amorphous carbohydrates and carbohydrate
on rheological properties of zein: development of a preliminary state solutions. Journal of Food Science 56, 1676–1681.
diagram. Cereal Chemistry 73, 433–438. Sartor, G., Johari, G.P., 1996. Polymerization of a vegetable protein,
Micard, V., Morel, M.-H., Bonicel, J., Guilbert, S., 2001. Thermal wheat gluten, and the glass-softening transition of its dry and reacted
properties of raw and processed wheat gluten in relation with protein state. Journal of Physical Chemistry 100, 19692–19701.
aggregation. Polymer 42, 477–485. Schofield, J.D., Bottomley, R.C., Timms, M.F., Booth, M.R., 1983. The
Mizuno, A., Mitsuiki, M., Motoki, M., 1999. Glass transition temperature effect of heat on wheat gluten and the involvement of sulfhydryl-
of casein as affected by transglutaminase. Journal of Food Science 64, disulphide interchange reactions. Journal of Cereal Science 1, 241–253.
769–799. Singh, N., Smith, A.C., 1999. Rheological behaviour of different cereals
Mizuno, A., Mitsuiki, M., Motoki, M., Ebisawa, K., Suzuki, E., 2000. using capillary rheometry. Journal of Food Engineering 39, 203–209.
Relationship between the glass transition of soy protein and molecular Sugisaki, M., Suga, H., Seki, S., 1968. Calorimetric study of glassy state. 4.
structure. Journal of Agricultural and Food Chemistry 48, 3292–3297. Heat capacities of glassy water and cubic ice. Bulletin of the Chemical
Mohammed, Z.H., Hill, S.E., Mitchell, J.R., 2000. Covalent crosslinking Society of Japan 41, 2591.
in heated protein systems. Journal of Food Science 65, 221–226. Tilley, K.A., Benjamin, R.E., Bargorogoza, R.E., Okot-Kotber, B.M.,
Morales, A., Kokini, J.L., 1997. Glass transition of soy globulins using Prakash, O., Kwen, H., 2001. Tyrosine cross-links molecular basis of
differential scanning calorimetry and mechanical spectrometry. Bio- gluten structure and function. Journal of Agricultural and Food
technology Progress 13, 624–629. Chemistry 49, 2627–2632.
Morel, M.H., Bonicel, J., Micard, V., Guilbert, S., 2000. Protein Toufeili, I., Lambert, I.A., Kokini, J.L., 2002. Effect of glass transition
insolubilization and thiol oxidation in sulfite-treated wheat gluten and cross-linking on rheological properties of gluten: development of a
films during aging at various temperatures and relative humidities. preliminary state diagram. Cereal Chemistry 79, 138–142.
Journal of Agricultural and Food Chemistry 48, 186–192. Utsumi, S., Matsumura, Y., Mori, T., 1997. Structure-function relation-
Morgan, R.G., Steffe, J.F., Ofoli, Y., 1989. A generalized viscosity model ships of soy proteins. In: Damodaran, S., Paraf, A. (Eds.), Food
for extrusion of protein doughs. Journal of Food Process Engineering Proteins and their Applications. Marcel Dekker Inc., New York,
11, 55–78. Basel, Hong Kong, pp. 257–292.
Morris, E.R., Cutler, A.N., Ross-Murphy, S.B., Rees, D.A., Price, J., Visser, J., 1988. Dry spinning of mil proteins. In: Blanshard, J.M.V.,
1981. Concentration and shear rate dependence of viscosity in random Mitchell, J.R. (Eds.), Food Structure—its Creation and Evaluation.
coil polysaccharide solutions. Carbohydrate Polymers 1, 5–21. Butterworths, London, pp. 197–218.
Noel, T.R., Parker, R., Ring, S.G., Tatham, A.S., 1995. The glass- Weegels, P.L., Hamer, R.J., 1998. Temperature-induced changes of wheat
transition behavior of wheat gluten proteins. International Journal of products. In: Hamer, R.J., Hoseney, R.C. (Eds.), Interactions: the
Biological Macromolecules 17, 81–85. Keys to Cereal Quality. American Association of Cereal Chemist Inc.,
Ogale, A.A., Cunningham, P., Dawson, P.L., Acton, J.C., 2000. St Paul, pp. 95–130.
Viscoelastic, thermal and microstructural characterisation of soy Weegels, P.L., Hamer, R.J., Schofield, J.D., 1996. Functional properties of
protein isolate films. Journal of Food Science 65, 672–679. wheat glutenin. Journal of Cereal Science 23, 1–17.
Pearson, A.M., 1982. Soy proteins. In: Hudson, B.J.W. (Ed.), Develop- Wetton, R.E., Morton, M.R., Rowe, A.M., 1986. Comparison of DMTA
ments in Food Proteins. Elsevier Applied Science, London, pp. 67–108. and DETA. American Laboratory 18, 96.
Plattner, B., Strahm, B., Rausch, K., 2001. ASAE Annual International Zhang, W., Hoseney, R.C., Faubion, J.M., 1998. Capillary rheometer of
Meeting, Sacramento, California, USA; ASAE, St. Joseph, MI; Vol. corn endosperm: glass transition, flow properties, and melting of
ASAE Meeting Paper Number 01-6067. starch. Cereal Chemistry 75, 863–867.
Pommet, M., Redl, A., Morel, M.H., Guilbert, S., 2003. Study of wheat Zhang, J., Mungara, P., Jane, J., 2001. Mechanical and thermal properties
gluten plasticization with fatty acids. Polymer 44, 115–122. of extruded soy protein sheets. Polymer 42, 2569–2578.
Pouplin, M., Redl, A., Gontard, N., 1999. Glass transition of wheat gluten Zhang, X.Q., Burgar, I., Do, M.D., Lourbakos, E., 2005. Intermolecular
pasticized with water, glycerol or sorbitol. Journal of Agricultural and interactions and phase structures of plasticized wheat proteins
Food Chemistry 47, 538–543. materials. Biomacromolecules 6, 1661–1671.

Anda mungkin juga menyukai