You are on page 1of 8

Journal of Food Engineering 127 (2014) 67–74

Contents lists available at ScienceDirect

Journal of Food Engineering


journal homepage: www.elsevier.com/locate/jfoodeng

High moisture extrusion cooking of pea protein isolates: Raw material


characteristics, extruder responses, and texture properties
Raffael Osen ⇑, Simone Toelstede, Florian Wild, Peter Eisner, Ute Schweiggert-Weisz
Fraunhofer Institute for Process Engineering and Packaging IVV, Giggenhauser Straße 35, 85354 Freising, Germany

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

Article history: The production of palatable meat analogues using high moisture extrusion cooking is a complex process
Received 10 May 2013 that depends on both the properties of the protein ingredients and the extrusion conditions. Three com-
Received in revised form 22 November 2013 mercial pea protein isolates were compared in order to investigate which protein properties affect extru-
Accepted 25 November 2013
der responses and product texture properties. The comparison revealed that although their basic
Available online 4 December 2013
chemical compositions were similar their functional properties affected the viscosity of the protein mass
during the initial heating phase of the extrusion process. The product texture properties depended on the
Keywords:
cooking temperature and were basically similar among the proteins, although considerably different
Pea protein isolate
Functional properties
energy input was observed during texturization. Our findings show that pea protein isolates are valuable
High moisture extrusion cooking raw materials for the development of fibrous whole-muscle meat alternatives, opening up a wide range
Texture properties of products for different consumer requirements.
Extruder responses Ó 2013 Elsevier Ltd. All rights reserved.

1. Introduction long cooling die by varying the moisture, temperature, pressure


and shear respectively (Noguchi, 1990). The combination of these
By 2050 the world’s population is projected to reach 9 billion process parameters results in molecular transformation and chem-
and 70% more food will be required (Aiking, 2011). Between ical reaction of the protein molecules which contribute to stabiliza-
1950 and 2000 meat production increased from 45 to 229 mil- tion of the three-dimensional network formed after the extrusion
lion tons and this is expected to increase to 465 million tons by step (Chen et al., 2011; Liu and Hsieh, 2007).
2050 (Steinfeld et al., 2006). The inherently inefficient conversion One key feature of high moisture meat analogues is their fibrous
of plant protein into animal protein as a result of animal metabo- structure which resembles muscle meat. In order to investigate
lism makes meat production responsible for a disproportionate textured protein products, several methods for textural profiling
share of environmental pressures such as land use, freshwater have been used. Texture profile analysis, a common method that
depletion, global warming and biodiversity loss (Steinfeld et al., determines the compression force of a probe, has been reported
2006). A promising solution to reduce the impact of meat produc- to relate little to fiber formation under high moisture extrusion
tion on the environment could be offered by partial replacement of conditions (Lin et al., 2000). Another approach is to measure the
meat protein with plant protein products in the human diet (Smil, force and deformation at rupture upon stretching of extrudates.
2000). Due to high variability of the results, Thiebaud et al. (1996) sug-
Since the 1960s extrusion cooking has been employed to pro- gested texture determination by shearing. This resembles the sen-
duce meat analogues using common starches and proteins as raw sation when food is first cut by the front incisors when introduced
materials. The traditional extruded meat analogues produced by into the mouth and is a common objective method for evaluating
low moisture extrusion (< 35%) have a sponge-like texture and beef tenderness (Caine et al., 2003; de Huidobro et al., 2005). Eval-
require rehydration prior to consumption (Guy, 2001). These uation of the cutting strength in longitudinal and transverse direc-
products are used as meat extenders or ground meat substitutes. tions, with respect to the flow direction in the cooling die, has been
However, they fail to mimic the appearance and texture of fibrous used to assess the quality of fiber formation (Chen et al., 2010;
whole-muscle meat. Fang et al., 2013).
One promising technology for obtaining fibrous meat-like struc- The effects of the protein ingredients and extrusion conditions
tures from plant proteins is the high moisture extrusion cooking on final product texture are reflected by their influence on extruder
(HMEC) process. The proteins are plasticized and texturized in a responses such as motor torque, die pressure, and specific mechan-
ical energy (SME). Upon thermal energy input, macromolecular
⇑ Corresponding author. Tel.: +49 8161 491450; fax: +49 8161 491444. transformations influence the rheological properties of the
E-mail address: raffael.osen@ivv.fraunhofer.de (R. Osen).
protein–water melt in the extruder barrel and the cooling die

0260-8774/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.jfoodeng.2013.11.023
68 R. Osen et al. / Journal of Food Engineering 127 (2014) 67–74

resulting in specific texture characteristics of the extrudates (Chen St. Joseph, MI, USA). The ash contents were determined in a ther-
et al., 2010). Consequently, SME values relate to the apparent mogravimetric system (TGA 601, Leco Corporation, St. Joseph, MI,
viscosity within the extruder barrel and can help to monitor and USA) at 950 °C until weight constancy (AOAC International,
compare changes in the proteinaceous matrix during extrusion 1990). The protein contents were calculated based on the nitrogen
texturization. content (N) according to the Dumas combustion method described
In order to improve the process stability and fiber formation in the German Food Act (2005) using a Protein/Nitrogen Analyzer
during HMEC, it is important to investigate the relationships be- FP 528 (Leco Corporation, St. Joseph, MI, USA) with a conversion
tween the protein ingredient properties, extruder responses, and factor of 6.25 (Shand et al., 2007; Sumner et al., 1981). The total
product texture. Earlier work on low moisture extrusion of soy lipid content was determined including fatty acids from phospho-
protein showed that both the processing conditions and texture lipids according to the method of Caviezel, DGF K-I 2c (00)
properties are affected by the protein properties, in particular the (DGF-Einheitsmethoden, 2004). Native and partially hydrolyzed
protein concentration and protein solubility (Riaz, 2004). Under starch was analyzed by determination of glucose units following
high moisture conditions, there are several published studies about complete hydrolysis with amyloglucosidase, using a Biopharm as-
the effects of process parameters on extruder responses and prod- say kit (R-Biopharm AG). All analyses were performed in duplicate.
uct properties such as texture and protein solubility (Chen et al.,
2010; Fang et al., 2014; Lin et al., 2002). Until now, it is not clear 2.1.2. Particle size distribution
which raw material properties might affect extruder responses The particle sizes of the protein powders were determined
and product texture. using a Malvern laser diffraction particle size analyzer (Mastersizer
Recent HMEC studies have focused on soy as a raw material S Long Bed Version 2.15, Malvern Instruments Ltd., Malvern, UK).
(Chen et al., 2011, 2010; Lin et al., 2000; Liu and Hsieh, 2008; Powders were dispensed in a wet dispersion unit using 1-butanol
MacDonald et al., 2009). However, a number of drawbacks are (VWR, Germany). Volume diameters D(v, 0.1), D(v, 0.5) and D(v,
associated with the use of soybean such as the presence of antinu- 0.9) were calculated from the particle volume distributions of the
tritional factors, its allergenic potential, and the introduction of respective isolates.
genetically modified organisms (Martínez-Villaluenga et al.,
2008). As an alternative to soy, pea protein is of special interest 2.1.3. Thermal properties
due to its nutritional characteristics and low potential for allergic Differential scanning calorimetry (DSC) was used to analyze the
responses (Nowak-Wegrzyn et al., 2003). Only a few studies have thermal properties of pea protein slurries (30% w/w) according to
been undertaken using pea seeds (Alonso et al., 2000), pea flour the method of Sousa et al. (1995). Approximately 10 mg of the
(Hood-Niefer and Tyler, 2010) and pea protein concentrate (Wang protein slurry was weighed into aluminum pans. The pans were
et al., 1999) for texturization under low moisture conditions. To hermetically sealed and heated from 40 °C to 120 °C at a rate of
the best of our knowledge no work has been published on HMEC 5 °C/min on a DSC instrument (Q2000, TA Instruments, USA). Each
with pea protein isolate (PPI) and, in detail, about the effect of sample was reheated one time to verify that there was no revers-
extrusion temperature on extruder responses and product texture. ibility of denaturation. The onset temperature (T0), peak transition
The objective of this study was therefore to investigate high temperature or denaturation temperature (Ts), and enthalpy of
moisture extrusion of three different PPIs with regard to the denaturation (DH) were computed from the thermograms. Tripli-
protein ingredient characteristics, extruder responses, extrudate cate measurements were carried out for each sample.
texture properties, and their interactions. This could lead to
improved understanding of the way proteins interact and form a 2.2. Functional properties
fibrous meat-like texture.
2.2.1. Protein solubility
Protein solubility was determined according to the procedure
2. Materials and methods
used by Morr et al. (1985) by mixing an aliquot of 1 g of protein
with 50 mL 0.1 M sodium chloride solution and incubating at
2.1. Characterization of protein ingredients
ambient temperature in a shaking water bath for 60 min. The
pH was adjusted using 0.1 M hydrogen chloride or sodium hydrox-
Three commercial pea protein isolates (Pisum sativum L.)
ide solution respectively. The non-dissolved fraction was separated
were used: PisaneÒM9 (Cosucra Groupe, Warcoing, Belgium),
by centrifugation at 20,000g for 15 min at ambient temperature.
EmvitalÒE7 (Emsland-Stärke GmbH, Emilchheim, Germany), and
The protein content in the supernatant was measured by a com-
NutralysÒF85M (Roquette Frères S.A., Lestrem, France), which were
bustion method based on an AOAC method according to Dumas
designated PPI 1, PPI 2, and PPI 3. These materials were all kindly
using a LECO analyzer. Duplicate measurements were undertaken
provided by the respective manufacturers.
for each sample.
The chemical compositions of PPI 1-3 are summarized in
Table 1.
2.2.2. Water binding capacity, oil binding capacity, emulsifying
capacity
2.1.1. Chemical composition The determination of the water binding capacity (WBC) was
The total dry matter was analyzed according to the German performed according to the AACC (1982) method and expressed
Food Act (2005). Samples were dried to weight constancy at as the weight of water bound by 1 g of sample.
105 °C in a thermogravimetric system (TGA 601, Leco Corporation, The oil binding capacity (OBC) was determined by the proce-
dure used by Lin et al. (1974) and expressed as grams of oil bound
by 1 g protein sample.
Table 1
Major chemical composition of PPIs. The emulsifying capacity (EC) was determined according to the
method of Wäsche et al. (2001) by continuous addition of oil to an
Material Dry matter (%) Ash (%) Protein (%) Lipid (%) Starch (%)
oil-in-water emulsion to the point of phase inversion of the
PPI 1 93.4 ± 0.1 5.1 ± 0.0 84.9 ± 0.4 7.7 ± 0.3 0.4 ± 0.0 emulsion. The volume of oil needed for phase inversion was used
PPI 2 94.2 ± 0.0 4.2 ± 0.0 87.3 ± 0.1 8.3 ± 0.1 0.0 ± 0.0
to calculate the emulsifying capacity (mL oil per g protein isolate).
PPI 3 94.3 ± 0.0 5.4 ± 0.0 83.2 ± 0.1 7.4 ± 0.1 0.4 ± 0.1
In order to compare the functional properties of the PPIs at the
R. Osen et al. / Journal of Food Engineering 127 (2014) 67–74 69

same pH value, the pH of PPI 2 was adjusted using 0.1 M sodium samples were thawed at ambient temperature. A square shaped
hydroxide solution. Duplicate measurements were undertaken sample (19  19 mm) was cut using a knife blade (A/LKB probe,
for each sample. Winopal Forschungsbedarf GmbH, Germany) to 75% of its original
thickness at a speed of 2 mm/s and the cutting strength was re-
2.2.3. RVA viscosity corded. Samples were evaluated vertical (longitudinal strength,
The viscosity of the proteins as a function of temperature was FL) and parallel (transverse strength, FT) to the direction of extru-
measured on a Rapid Visco Analyzer (Newport Scientific Pty. Ltd., date outflow from the extruder. Equal values indicate uniform tex-
Warriewood, Australia) using the AACC Method 76-21 STD 2 ture with low material anisotropy and without any fibrous texture.
(AACC, 1997). Protein slurries of the desired concentration were All determinations were performed with at least 12 replicates. Pre-
heated to 50 °C and stirred at 960 rpm for 10 s for thorough disper- liminary studies showed no significant effect of freezing and thaw-
sion. The slurry was held at 50 °C for 50 s, and then heated to 95 °C ing of the extrudates on FL and FT (p < 0.05).
at 6 °C/60 s, held at 95 °C for 300 s, and finally cooled to 50 °C.
Duplicate measurements were performed for each sample. 2.3.3. Statistical analysis
Data (extruder responses, product texture properties) were
2.3. High moisture extrusion cooking treated by ANOVA for statistical significance using OriginPro Soft-
ware, version 8.5 (OriginLab Corporation, Northampton, MA, USA).
Extrusion experiments were performed using a laboratory, co- Tukey’s HSD (honestly significant difference) test at a 5% level was
rotating, intermeshing twin screw extruder (Haake Rheocord, used to identify the significant difference of each treatment.
Thermo Fisher Scientific, Inc., UK) with a screw diameter of
16 mm, a smooth barrel, and a length–diameter ratio of 25:1. 3. Results and discussion
The screw profile is built with different screw elements that can
be assembled on hexagon-shaped shafts. The screw profile com- 3.1. Characterization of protein ingredients
prised (from feed to exit): 192 mm, twin lead feed screw; 8 mm,
45° forwarding paddles; 16 mm, twin lead feed screw; 16 mm, 3.1.1. Particle size distribution
45° forwarding paddles; 16 mm, twin lead feed screw; 8 mm, 45° Due to the short residence time in the laboratory-scale extru-
forwarding paddles; 32 mm, twin lead feed screw; 4 mm, 45° for- der, it is possible that large particles cannot be sufficiently melted
warding paddles; 8 mm, twin lead feed screw; 8 mm, 45° forward- in the cooking zone of the extruder. This could affect both extruder
ing paddles; 64 mm, twin lead feed screw; and 24 mm, single lead responses and fiber formation. Hence the average particle size was
screw. The barrel is segmented into 5 temperature-controlled measured. Table 2 shows the D(v, 0.1), D(v, 0.5), and D(v, 0.9) values
zones that are heated by an electric cartridge heating system and of the respective isolates.
cooled with water. To prevent expansion at the die exit, a long Particle size is reported to play a significant role in low moisture
die with dimensions of 19  2  210 mm (W  H  L) was attached extrusion with regard to the processing and texture properties of
to the end of the extruder, with water at 80 °C as a cooling medium the extrudates. Some raw materials are as fine as 38 lm or as
at a flow rate of 3.4 l/min. A twin screw gravimetric feeder type coarse as 180 lm (Riaz, 2006; Yada, 2004). The average particle
KCM (K-tron, Niederlenz, Switzerland) was used to feed the dry size of PPI 1 was the lowest, followed by PPI 2 and PPI 3 respec-
protein ingredients into the extruder at a feeding rate of 0.45 kg/ tively. The differences in the particle size distributions can be
h. While operating, water at ambient temperature was pumped attributed to the different manufacturing methods.
(Alpha 50 Plus, ECOM, Prague, Czech Rep.) into the top of the extru-
der barrel 130 mm downstream from the center of the feed port 3.1.2. Thermal properties
resulting in a moisture content of 55%. The screw speed was DSC is a well-established method for investigating the thermal
150 rpm. During the initial heating phase, barrel temperatures properties of proteins. The DSC thermograms of the protein slurries
were stepwise increased from 40 °C with a temperature profile of are shown in Fig. 1.
40, 60, 80 and 100 °C from the first (feeding zone) to the fourth Among the three PPIs, only PPI 2 showed a small endothermic
zone. The last zone (fifth) was set at the desired cooking tempera- peak at 88.5 ± 0.2 °C (dH 1.3 ± 0.0 J/g).
ture of 100, 120, 140 or 160 °C respectively. According to Arnfield and Murray (1981), a Td value of 86 °C
was observed for air-classified pea protein. Shand et al. (2007)
2.3.1. Extruder responses found two major peaks for laboratory prepared PPI, one at 67 °C
Once the extruder reached steady state conditions as indicated for the non-globulin fraction and one at 85 °C for the globulin frac-
by constant extrusion system parameters, the extruder responses tion. Considering our results, the endothermic peak obtained for
including torque (N m) and pressure (bar) in front of the cooling PPI 2 may represent the thermal transition of the partly native
die were recorded in-line using Polylab software (Thermo Fisher globulin fraction. In contrast, the absence of any peaks in the ther-
Scientific, Inc., UK). The SME was calculated from the screw speed mograms of PPI 1 and PPI 3 indicates the denaturation of the
n (rpm), motor torque T (N m), and mass flow rate MFR (g/min) by respective protein fractions.
the following equation (Chen et al., 2010; Godavarti and Karwe,
1997): 3.1.3. Functional properties
  According to Lampart-Szczapa et al. (2006), the functional prop-
kJ 2p  n  T
SME ¼ erties of proteins in food products result from interactions between
kg MFR

Table 2
2.3.2. Product texture properties Average particle size (volume diameter) of PPIs.
The texture properties of the extrudates were evaluated using a
Material D(v, 0.1) (lm) D(v, 0.5) (lm) D(v, 0.9) (lm)
TA.XT plus Texture Analyzer (Stable Micro Systems, UK) according
to the modified procedure of Thiebaud et al. (1996) and Chen et al. PPI 1 9.1 ± 0.2 39.5 ± 0.9 108.7 ± 4.6
PPI 2 29.0 ± 0.0 71.8 ± 0.4 137.2 ± 1.1
(2010). Samples were collected for each temperature and immedi-
PPI 3 40.8 ± 0.2 157.7 ± 1.7 344.5 ± 2.1
ately stored in airtight plastic bags at 20 °C. Prior to analysis, all
70 R. Osen et al. / Journal of Food Engineering 127 (2014) 67–74

(a) and PPI 3 had lower solubility values than PPI 2 which can be
attributed to their denatured state (Fig. 1).
PPI 1 According to Yada (2004), protein solubility reflects the heat
treatment history of proteins throughout the preparation process,
with a lower solubility following extensive heat treatment. The dif-
ferences in solubility of the PPIs may have been the result of the
processing conditions (mechanical loads from grinding, high tem-
PPI 2 peratures e.g. during spray-drying) to enrich the protein.

3.1.3.2. Water binding capacity, oil binding capacity, emulsifying


capacity. Selected functional properties of the PPIs were compared
PPI 3 and are summarized in Table 3.
The water binding capacity (WBC) depends on the number of
polar sites interacting with water. PPI 1 exhibited the highest
WBC of 5.4 ml/g, followed by PPI 3 and PPI 2 with 5.0 and 2.1 ml/
g respectively. As shown in Fig. 2, more protein was soluble at
pH 7.5 in PPI 2 (30%) than in the other proteins (15%). This af-
(b) fects the WBC determination, since the amount of soluble protein
in the discarded supernatant is not available for binding water.
Considering the thermal properties of PPI 2, its low WBC is in
accordance with the findings of Sumner et al. (1981). They
described a positive correlation between water absorption and
PPI 2 heat induced denaturation of PPI for differently dried protein
products.
The oil binding capacity (OBC) was between 0.7 and 1.7 ml/g
and in the range reported by Owusu-Ansah and McCurdy (1991).
Native proteins show lower OBC than their denatured counterparts
because of their structural folding (Zayas, 1997). Accordingly, the
low OBC of PPI 2 correlates with its thermal properties and can
be attributed to its structural folding with presumably a lower
amount of hydrophobic groups on the molecule surface compared
to the denatured PPI 1 and 3.
Emulsification properties are reported to be partly improved by
Fig. 1. Exemplary DSC thermograms of protein slurries (30% w/w) heated at 5 °C/ heat denaturation, correlating with an increase in surface hydro-
min. (a) PPI 1, PPI 2 and PPI 3 and (b) zoomed endothermic peak of PPI 2. phobicity (Zayas, 1997). This might explain the fact that PPI 2
had the lowest emulsifying capacity (EC) among the proteins. Sim-
ilar findings were reported by Sumner et al. (1981). They stated
the protein and other food components, depending on the physico- that drum-dried PPI had the highest EC-values compared to
chemical properties characterizing the protein structure. spray-dried and freeze-dried PPI.

3.1.3.3. RVA viscosity properties. The viscosity patterns of the PPIs


3.1.3.1. Protein solubility. Protein quality for texturization of soy are presented in Fig. 3.
based protein ingredients is usually measured by the level of pro- Both PPI 1 and 3 showed a high initial viscosity of around
tein solubility (Riaz, 2004). 2000 mPa s that decreased upon heating to 95 °C. The pasting pro-
The recorded protein solubility curves in the pH range of 3–8 files relate to the proteins’ functional properties, particularly a low
(Fig. 2) are in agreement with previous studies (Shand et al., protein solubility and high WBC. Upon hydration the powders
2007; Sosulski and Mccurdy, 1987; Tomoskozi et al., 2001). PPI 1 absorb water and swell resulting in a high starting viscosity that
subsequently decreases when the mechanical and thermal energy
input increases. In contrast, the viscosity profile of PPI 2 differs.
Following a low initial viscosity of <100 mPa s, the viscosity in-
creased sharply during heating to 95 °C. The viscosity curves of
PPI 2 (Fig. 3b) showed a sharp increase in viscosity at temperatures
around 80 °C, depending on the protein content of the slurry. The
low starting viscosity of PPI 2 can be attributed to its high solubil-
ity compared to the other proteins. Furthermore, the sudden in-
crease in viscosity at elevated temperatures can be explained by
the protein’s thermal properties (Section 3.1.2). When the temper-
ature increases above 75 °C, denaturation of the native proteins
causes the protein solubility to decrease which results in an in-
crease in viscosity.
In summary, comparison of the protein ingredients revealed
that the functional and rheological properties of the PPIs depended
on their thermal properties.
As indicated by DSC studies, only PPI 2 had partly native protein
fractions and exhibited higher protein solubility and lower WHC,
Fig. 2. Protein solubility (%) of PPIs as a function of pH. OBC, and EC. These properties in turn affected its viscosity profile
R. Osen et al. / Journal of Food Engineering 127 (2014) 67–74 71

Table 3
Functional properties of the PPIs.

Material pH Value Water binding capacity (ml/g) Oil binding capacity (ml/g) Emulsifying capacity (ml/g)
PPI 1 7.5 5.4 ± 0.3 1.6 ± 0.1 870.0 ± 21.2
PPI 2 6.7 2.1 ± 0.0 0.7 ± 0.0 580.0 ± 33.4
PPI 2a 7.5 2.5 ± 0.1 1.3 ± 0.0 395.0 ± 28.3
PPI 3 7.4 5.0 ± 0.1 1.7 ± 0.0 725.0 ± 0.0
a
Adjusted to pH 7.5 for comparison.

(a)

(b)

Fig. 4. The effect of raw material (PPI 1, PPI 2 and PPI 3) and cooking temperature
on SME at 55% moisture content (w/w).

viscosity at the die. At the same time, the SME decreased between
140 and 150 °C and increased at 160 °C. From previous research
studies it is known that the relationship between cooking temper-
ature and extruder responses is not only dependent on the melt
viscosity in the extruder barrel but also on the friction during tex-
turization within the cooling die.
Fig. 3. RVA viscosity curves of PPIs. (a) Paste profile of PPI 1 (15% w/w), PPI 2 (20% Fig. 4 compares SME values for different PPIs at various temper-
w/w) and PPI 3 (15% w/w). Viscosity profiles are displayed with concentrations
giving smooth curves without sharp peaks. (b) Paste profile of varied concentrations
atures. It shows that the extrusion of PPI 1 exhibited the lowest
of PPI 2. SME, followed by PPI 3 and PPI 2 respectively. No correlation was
found between the average particle size and extruder responses.
Although all powders were suitable for processing via HMEC, we
during heating to 95 °C. The current observations indicate that experienced that larger particles (PPI 3) were easier to process than
commercial protein products can vary in terms of their functional fine powders (PPI 1) because the latter tend to agglomerate in the
properties, which might affect their performance during feed section. Regarding factor interactions, the PPIs exhibited dif-
processing. ferent SME-temperature dependencies at cooking temperatures
above 100 °C, indicating complex protein interactions during the
3.2. High moisture extrusion cooking texturization process. With increasing temperature from 120 to
160 °C, the SME increased for PPI 1, decreased for PPI 2, and re-
3.2.1. Extruder responses mained constant for PPI 3.
The effects of the protein ingredients and cooking temperature Our results correlate with previous research studies, and show
on the SME are shown in Fig. 4. that extruder responses are affected by the cooking temperature
Extruder responses were affected by the protein ingredients and the protein ingredients. As the temperature decreases in the
and cooking temperature (p < 0.05), with a significant interaction cooling die, protein–protein interactions increase and crosslinking
between the factors. The SME was highest at 100 °C and decreased occurs (Liu and Hsieh, 2008). This causes the viscous melt to solid-
sharply upon further heating to 120 °C. Above 100 °C the PPIs ify in the die channel. The resulting friction increases back pressure
showed different temperature-SME dependencies. which affects the SME and pressure before the die inlet. According
Conventionally, it is expected that a higher temperature pro- to Riaz (2004), protein ingredients having lower protein solubility
duces a lower melt viscosity resulting in a lower SME. However, require more mechanical energy in the extruder to effectively tex-
Lin et al. (2002) observed, during HMEC of SPI and starch at a 9:1 turize the protein under low moisture conditions. However, we
ratio and a moisture content of 60–70%, that the extruder re- found that the most soluble protein (PPI 2) required the highest
sponses changed little at temperatures between 138 and 160 °C. SME during processing. From all these data it is not yet clear which
They assumed protein denaturation and texturization affected protein property affects the energy consumption during HMEC.
the temperature–viscosity relationship. Chen et al. (2010) found, Additionally, the SME can be used to monitor the rheological
during HMEC of SPI at a moisture content of 60%, that increasing behavior of the protein–water mixture during the initial phase of
the cooking temperature from 140 to 160 °C decreased the in-line HMEC, below the cooking temperature necessary for texturization.
72 R. Osen et al. / Journal of Food Engineering 127 (2014) 67–74

3.2.2. Product texture properties


The effect of cooking temperature on the cutting strength of
extrudates made from three commercial PPIs is shown in Fig. 6a–
c. At temperatures below 100 °C, the structure of the protein water
mixture was not yet generated and therefore cutting strength val-
ues are not shown within that temperature range.
The individual protein ingredients as well as the cooking tem-
perature significantly affected the extrudate texture (p < 0.05),
with a significant interaction among the factors. At low tempera-
ture (100 °C), both the longitudinal and transverse cutting strength
were similar and in the range of 6–8 N. Increasing cooking temper-
ature resulted in an increased tensile strength predominantly in
the longitudinal direction whereas the transverse strength re-
mained constant, except for PPI 2 which increased from 7 to
11 N. Similar results were obtained by Thiebaud et al. (1996). They
studied the influence of process variables on the resistance to
stretching of extrudates from a mix of fish surimi and soy protein
Fig. 5. SME values of PPIs during initial starting and heating phase until the cooking concentrate and found that an increase in cooking temperature in-
temperature reached 100 °C. SME profiles for each PPI were recorded during three
creased FL but did not significantly affect FT. Chen et al. (2010)
consecutive experiments at a constant feed rate of 1 kg/h and 65% moisture content
(w/w). The cooking temperature was read in the 5th barrel section. During the showed that cooking temperature had a significant effect on tensile
heating phase, the temperature profile was set at 40, 40, 40, 40, 40; 40, 60, 60, 60, strength and degree of texturization during HMEC of soy protein
60; 40, 60, 80, 80, 80 and 40, 60, 80, 100, 100 °C from the first to the fifth zone isolate.
respectively. Although fiber formation was affected by both the cooking
temperature and the raw materials used, the latter had only a
Fig. 5 shows the SME curves recorded during the initial stepwise minor effect. This can be attributed to the finding that the major
heating phase until a cooking temperature of 100 °C was reached. difference between the PPIs, namely their functional properties,
PPI 1 and PPI 3 exhibited similar patterns, showing a sharp in- becomes less relevant at temperatures above the proteins’ dena-
crease in SME during the first 4 min followed by a steady decrease turation temperature.
upon further heating. In contrast, PPI 2 exhibited low energy In summary, our observations are in accordance with the liter-
consumption until a sharp increase in SME at around 80 °C. The ature and can be described and interpreted as follows. Different
sudden rise in SME can be attributed to the denaturation of partly macrostructures were observed within the experimental domain.
native proteins leading to a sudden decrease in protein solubility. Below 120 °C, samples exhibited a dough-like soft texture without
This causes the viscosity to increase rapidly, as described in any fibrous structures that may be due to incomplete melting and
Section 3.1.3. partial unraveling of macromolecules. Upon further increase of the
In summary, we can conclude that at temperatures below the cooking temperature, extrudates started to display multilayered
denaturation temperature of proteins their functional properties structures with layers parallel to the die wall and fine fibers ap-
such as protein solubility and WHC have a great impact on extru- peared upon tearing. During this stage, energy input might have
der responses, especially during the initial phase. This should be caused macromolecules to unravel making bonding sites available
taken into account during processing to avoid blocking of the ex- for further crosslinking that were previously buried within the
truder barrel or the backing up of water. At temperatures above macromolecules. This would explain the positive correlation be-
the proteins’ denaturation temperature, their thermal properties tween the cooking temperature and the cutting strength of the
– which highly affect their functional properties as previously extrudates. At temperatures as high as 160 °C, the macrostructures
shown – become less significant in terms of processing. became more homogenous with a smooth surface. Only upon tear-
In order to determine the relationship between protein ing do predominantly lengthwise oriented fibers appear (Fig. 7).
properties and the SME, further studies focusing on other protein As described in Section 3.2.1, the melt viscosity decreases with
properties such as the molecular weight distribution, amino acid increasing temperature above 100 °C. This might affect the flow
composition, and protein–protein interactions are required. velocity profile during the moment of solidification in the cooling

Fig. 6. Cutting strength as a function of cooking temperature of PPI 1 (a), PPI 2 (b) and PPI 3 (c) at 55% moisture content (w/w) in longitudinal (FL) and transverse (FT)
direction.
R. Osen et al. / Journal of Food Engineering 127 (2014) 67–74 73

Fig. 7. Digital images of extruded PPI 1 (55% moisture content (w/w), 1 kg/h feed rate) displaying (a) predominant lengthwise fibrous structures (160 °C cooking temperature)
and (b) parabolic patterns (130 °C cooking temperature).

die. Assuming laminar flow of a viscous fluid in the die, the melt analyses, Mr. Michael Schott for the particle sizing, and Mr. Rainer
temperature and flow velocity are higher at the core of the flow Giggenbach for the texture analysis.
channel than close to the cooled die wall. During solidification this
might lead to a parabolic pattern caused by shear-oriented linear
unraveled macromolecules that crosslink when cooled below a References
characteristic temperature. That would explain why predomi-
nantly lengthwise oriented fibers appeared with increasing AACC, 1982. Hydration capacity of pregelatinized cereal products, 10 ed. American
Association of Cereal Chemists, St- Paul, MN, USA.
temperature. AACC, 1997. General Pasting Method for Wheat or Rye Flour or Starch Using the
In summary, it appears that functional properties play a minor Rapid Visco Analyser. American Association of Cereal Chemists International, St.
role during fiber formation due to the elevated cooking tempera- Paul, MN, USA.
Aiking, H., 2011. Future protein supply. Trends Food Sci. Technol. 22 (2–3), 112–
ture, which exceeds the proteins’ denaturation temperature. Also, 120.
particle size did not affect the product texture properties as the Alonso, R., Orue, E., Zabalza, M.J., Grant, G., Marzo, F., 2000. Effect of extrusion
particles were already hydrated, melted to a protein water mix- cooking on structure and functional properties of pea and kidney bean proteins.
J. Sci. Food Agric. 80 (3), 397–403.
ture, and solidified prior to texture analysis.
AOAC International, 1990. Official methods of analysis, Arlington, USA.
Arnfield, S.D., Murray, E.D., 1981. The influence of processing parameters on food
4. Summary and conclusion protein functionality. I. Differential scanning calorimetry as an indicator of
protein denaturation. J. – Canad. Inst. Food Sci. Technol. = J. l’Inst. Canad. Sci.
Technol. 14 (4), 289–294.
All the PPIs that were investigated were successfully texturized Caine, W.R., Aalhus, J.L., Best, D.R., Dugan, M.E.R., Jeremiah, L.E., 2003. Relationship
at 55% moisture content to highly fibrous extrudates. Their texture of texture profile analysis and Warner–Bratzler shear force with sensory
characteristics of beef rib steaks. Meat Sci. 64 (4), 333–339.
properties could be controlled by the cooking temperature. Consid- Chen, F.L., Wei, Y.M., Zhang, B., Ojokoh, A.O., 2010. System parameters and product
ering the aforementioned benefits of peas over soy (locally grown, properties response of soybean protein extruded at wide moisture range. J. Food
low allergy, high nutritional value), our findings showed that pea Eng. 96 (2), 208–213.
Chen, F.L., Wei, Y.M., Zhang, B., 2011. Chemical cross-linking and molecular
protein isolates are highly valuable protein ingredients for the aggregation of soybean protein during extrusion cooking at low and high
development of whole-muscle meat alternatives, opening up a moisture content. LWT – Food Sci. Technol. 44 (4), 957–962.
wide range of products for different consumer requirements. de Huidobro, F.R., Miguel, E., Blázquez, B., Onega, E., 2005. A comparison between
two methods (Warner–Bratzler and texture profile analysis) for testing either
Comparison of the 3 PPIs revealed that although their basic
raw meat or cooked meat. Meat Sci. 69 (3), 527–536.
chemical compositions were similar their functional properties DGF-Einheitsmethoden, 2004. Deutsche Gesellschaft für Fettwissenschaft e.V,
correlated with the proteins’ thermal properties. These in turn af- second ed., Münster, WVG, Stuttgart, Germany.
Fang, Y., Zhang, B., Wei, Y., 2013. Effects of the specific mechanical energy on the
fected the viscosity of the protein mass during the initial heating
physicochemical properties of texturized soy protein during high-moisture
phase of the extrusion process, but became less relevant at temper- extrusion cooking. J. Food Eng..
atures above the proteins’ denaturation temperature. In order to Fang, Y., Zhang, B., Wei, Y., 2014. Effects of the specific mechanical energy on the
screen unknown protein ingredients for their processing behavior physicochemical properties of texturized soy protein during high-moisture
extrusion cooking. J. Food Eng. 121, 32–38.
particularly during the critical heating phase, RVA viscosity tests German Food Act, 2005. Methods L. 16.01-2, L. 17.00-1, L. 17.00-3. In BVL
of protein slurries could be a simple and useful tool for simulation Bundesamt für Verbraucherschutz und Lebensmittelsicherheit (Ed.), Amtliche
of the extrusion process. Although all powders were suitable for Sammlung von Untersuchungsverfahren nach § 64 LFGB, § 35 Vorlaeufiges
Tabakgesetz, § 28b GenTG – I – Lebensmittel – Band I (L) Verfahren zur
HMEC, larger particles (PPI 3) were easier to process than fine pow- Probenahme und Untersuchung von Lebensmitteln. Beuth Verlag GmbH, Berlin,
ders (PPI 1). Fiber formation depended on cooking temperature Germany.
and was basically similar among the three PPIs, although consider- Godavarti, S., Karwe, M.V., 1997. Determination of specific mechanical energy
distribution on a twin-screw extruder. J. Agric. Eng. Res. 67 (4), 277–287.
ably different energy input was observed during their extrusion Guy, R., 2001. Extrusion Cooking: Technology and Applications, Guy Robin ed. CRC
texturization. In order to comprehensively explore the effect of Press, Cambridge, England.
the protein ingredients on HMEC, further investigations of the pro- Hood-Niefer, S.D., Tyler, R.T., 2010. Effect of protein, moisture content and barrel
temperature on the physicochemical characteristics of pea flour extrudates.
teins’ properties are necessary. Comparison of their molecular
Food Res. Int. 43 (2), 659–663.
weight distribution, amino acid composition, and protein–protein Lampart-Szczapa, E., Konieczny, P., Nogala-Kałucka, M., Walczak, S., Kossowska, I.,
interactions before and after extrusion could provide further in- Malinowska, M., 2006. Some functional properties of lupin proteins modified by
lactic fermentation and extrusion. Food Chem. 96 (2), 290–296.
sight into the structure-process-texture-relationship.
Lin, M.J.Y., Humbert, E.S., Sosulski, F.W., 1974. Certain functional properties of
sunflower meal products. J. Food Sci. 39 (2), 368–370.
Acknowledgement Lin, S., Huff, H.E., Hsieh, F., 2000. Texture and chemical characteristics of soy protein
meat analog extruded at high moisture. J. Food Sci. 65 (2), 264–269.
Lin, S., Huff, H.E., Hsieh, F., 2002. Extrusion process parameters, sensory
The authors would like to thank Ms. Sigrid Bergmann, Ms. Sigrid characteristics, and structural properties of a high moisture soy protein meat
Gruppe, Ms. Elfriede Bischof, and Ms. Evi Müller for the chemical analog. J. Food Sci. 67 (3), 1066–1072.
74 R. Osen et al. / Journal of Food Engineering 127 (2014) 67–74

Liu, K.S., Hsieh, F.H., 2007. Protein–protein interactions in high moisture-extruded Shand, P.J., Ya, H., Pietrasik, Z., Wanasundara, P.K.J.P.D., 2007. Physicochemical and
meat analogs and heat-induced soy protein gels. J. Am. Oil Chem. Soc. 84 (8), textural properties of heat-induced pea protein isolate gels. Food Chem. 102 (4),
741–748. 1119–1130.
Liu, K.S., Hsieh, F.H., 2008. Protein–protein interactions during high-moisture Smil, V., 2000. Feeding the World: A Challenge for the Twenty-first Century. MIT
extrusion for fibrous meat analogues and comparison of protein solubility Press, Cambridge, Mass.
methods using different solvent systems. J. Agric. Food Chem. 56 (8), 2681– Sosulski, F.W., Mccurdy, A.R., 1987. Functionality of flours, protein-fractions and
2687. isolates from field peas and faba bean. J. Food Sci. 52 (4), 1010–1014.
MacDonald, R.S., Pryzbyszewski, J., Hsieh, F.-H., 2009. Soy protein isolate extruded Sousa, I.M.N., Mitchell, J.R., Ledward, D.A., Hill, S.E., daCosta, M.L.B., 1995.
with high moisture retains high nutritional quality. J. Agric. Food Chem. 57 (9), Differential scanning calorimetry of lupin and soy proteins. Zeitschrift
3550–3555. Lebens.-Unter. Und-Forschung 201 (6), 566–569.
Martínez-Villaluenga, C., Gulewicz, P., Frias, J., Gulewicz, K., Vidal-Valverde, C., Steinfeld, H., Gerber, P., Wassenaar, T., Castel, V., Rosales, M., De Haan, C., 2006.
2008. Assessment of protein fractions of three cultivars of Pisum sativum L.: Livestock’s Long Shadow: Environmental Issues and Options. FAO, Rome, Italy.
Effect of germination. Eur. Food Res. Technol. 226 (6), 1465–1478. Sumner, A.K., Nielsen, M.A., Youngs, C.G., 1981. Production and evaluation of pea
Morr, C.V., German, B., Kinsella, J.E., Regenstein, J.M., Van Buren, J.P., Kilara, A., 1985. protein isolate. J. Food Sci. 46 (2), 364–366.
A collaborative study to develop a standardized Food protein solubility Thiebaud, M., Dumay, E., Cheftel, J.C., 1996. Influence of process variables on the
procedure. J. Food Sci. Technol. 50, 1715–1718. characteristics of a high moisture fish soy protein mix texturized by extrusion
Noguchi, A., 1990. Extrusion cooking of high moisture protein foods. In: Mercier, cooking. Food Sci. Technol. – Lebensm.-Wiss. Technol. 29 (5–6), 526–535.
C.L.P., Harper, J.M. (Eds.), Extrusion cooking. AACC, Minnesota, pp. 343–369. Tomoskozi, S., Lasztity, R., Haraszi, R., Baticz, O., 2001. Isolation and study of the
Nowak-Wegrzyn, A., Sampson, H.A., Wood, R.A., Sicherer, S.H., 2003. Food protein- functional properties of pea proteins. Nahrung-Food 45 (6), 399–401.
induced enterocolitis syndrome caused by solid food proteins. Pediatrics 111 Wang, N., Bhirud, P.R., Tyler, R.T., 1999. Extrusion texturization of air-classified pea
(4), 829–835. protein. J. Food Sci. 64 (3), 509–513.
Owusu-Ansah, Y.J., McCurdy, S.M., 1991. Pea proteins: a review of chemistry, Wäsche, A., Müller, K., Knauf, U., 2001. New processing of lupin protein isolates and
technology of production, and utilization. Food Rev. Int. 7 (1), 103–134. functional properties. Nahrung/Food 45, 393–395.
Riaz, M.N., 2004. Texturized soy protein as an ingredient. In: Yada, R.Y. (Ed.), Yada, R.Y., 2004. Proteins in food processing. CRC Press, Woodhead Pub., Cambridge,
Proteins in food processing. CRC Press, Woodhead Pub., Cambridge, Eng., pp. Eng.
517–558. Zayas, J.F., 1997. Functionality of proteins in food. Springer, New York.
Riaz, M.N., 2006. Soy applications in food. CRC, Boca Raton, Fla.

You might also like