You are on page 1of 13

Extrusion for the Production of Functional Foods and Ingredients

Mohammed Shafiq Alam and Raouf Aslam, Department of Processing and Food Engineering, Punjab Agricultural University,
Ludhiana, Punjab, India
© 2020 Elsevier Inc. All rights reserved.

Introduction 1
Extrusion Equipment 2
Single-Screw Extruders 2
Twin-Screw Extruders 3
Extrusion Process Variables and Their Effects on Quality Characteristics 3
Development of Functional Products Using Extrusion Techniques 4
Food Groups as a Source of Functional Ingredients 4
Cereals 5
Roots and Tubers 5
Legumes and Oilseeds 5
Fruits and Vegetables 6
Animal Products 6
Plant and Animal By-products 7
Functional Foods Developed With Extrusion Cooking 7
Snack Foods 7
Ready-to-Eat Cereals 7
Texturized Vegetable Protein 8
Pasta Products 8
Pet Food and Animal Feed Products 8
Impact of Extrusion Cooking on Nutritional Quality of Foods 9
Concluding Remarks 10
References 10

Introduction

With growing health awareness among consumers with diverse nutritional needs, foods are now being modified to provide
improved health benefits beyond basic nutrition. A wide variety of bioactive compounds, including phytochemicals (phenolics,
carotenoids, flavonoids, etc.), minerals (potassium, magnesium, calcium, etc.), and vitamins, are naturally found in several
plant-based foods, exhibiting numerous health benefits. By selectively incorporating valuable phytochemicals and other constit-
uents into food formulations, it has now become possible to reduce the risk of diseases and promote optimal well-being of indi-
viduals. These foods are collectively known as “functional foods” and are attracting widespread research interest across the world.
There are a number of food processing operations that conveniently allow modification or addition of a food component to serve
more than basic nutrition. Extrusion cooking is one such food processing technology by which functional foods are conveniently
and economically formulated with a better control of product specifications. The production of various foods by forcibly flowing
a material through a shaped hole or die under varied conditions of temperature and pressure at a convenient rate is called extru-
sion cooking. The process has led to the production of a wide range of cereal-based foods, protein supplements, and ready-to-use
products, like pasta, bread crumbs, breakfast cereals, snacks, macaroni, confectionary, pet foods, chewing gums, croutons etc.
(Chang and Ng, 2009). The apparatus used for extrusion cooking is the food extruder, which can be considered a high-
temperature–short-time bioreactor that applies thermal and shear energies to transform raw ingredients into extruded products.
The energy applied during extrusion cooking induces a diverse range of transformations in the structural and nutritional prop-
erties of the raw material, which include starch gelatinization and degradation, oxidation of lipids, protein denaturation, dietary
fiber solubility, flavor formation, and possible degradation of vitamins, phytochemicals, and antinutritional factors (Riaz et al.,
2009). Owing to its versatility, higher control of variables, high capacity, low energy costs, lack of effluent generation, continuity
of operations, and adaptability, extrusion cooking is becoming popular for the design and development of novel food products
having high quality, better nutrition, and unique shapes and characteristics (Faraj et al., 2004). Extruded products with a modified
structure are also reported to exhibit a better solubility, swelling power, water-holding capacity, water hydration, and viscosity. As
a result the process could be used to increase the soluble fiber content of plant cell wall–rich by-products like brans and hulls of
legumes and cereals to improve their usability (Ralet et al., 1990; Gourgue et al., 1994; Hwang et al., 1997; Gaosong and Vasan-
than, 2000; Rouilly et al., 2006).

Reference Module in Food Sciences https://doi.org/10.1016/B978-0-08-100596-5.23041-2 1


2 Extrusion for the Production of Functional Foods and Ingredients

Extrusion Equipment

The basic instrument required in the process of extrusion cooking is the extruder, which can be classified on the basis of its convey-
ance mechanism as a screw, piston, or roller extruder (Brennan and Grandison, 2012). Among all of them, screw extruders are the
most popular in the food processing industry and may operate with single, twin, or multiple screws rotating inside a fixed barrel.
The basic design of extruders consists of a fixed barrel tightly enclosing a rotating Archimedes screw with an opening, called a die, at
the discharge (Park et al., 2014). Properly mixed and pre-conditioned raw materials are fed through a hopper into the fixed barrel to
be conveyed along to the discharge end by the rotating screws. The design and shape of the extrudate is imparted by the die as the
product leaves the extruder (Bordoloi and Ganguly, 2014). Small-scale extrusion involving less complex raw materials and low cost
is usually carried out with single-screw designs, whereas in food processing applications handling a higher constant throughput and
dealing with more complex ingredients, twin screw extruders are advised (Clextral, 2010; Karwe, 2009; Ramachandra and Thejas-
wini, 2015; Steel et al., 2012).

Single-Screw Extruders
Single-screw extruders employ only one rotating screw inside a fixed barrel to discharge the extrudate from a size-specific die. These
extruders rely on flow under drag that moves the material toward the discharge die and hence mounts pressure. To understand the
process of extrusion, we need to look at the forces required to facilitate the flow of material toward the discharge. To achieve the
forward movement of material, the destructured ingredients should not rotate with the rotating screw. This can be explained using
the example of a nut and bolt, where if the nut turns too, it will not be tightened. To comply with such a motion, the material is held
to the barrel by the frictional forces developing due to drag that make it advance along to the discharge end. As such, the forward
movement of the material would depend on the frictional force and the speed of rotation. The forward travel would be higher if the
travel speed were less and the frictional force were high. The recirculating flow of the material between the screw flights, known as
cross-channel flow, enhances the mixing of the ingredients inside the barrel and does not contribute to the net forward movement.
As such, raw materials having a high coefficient of friction, such as whole grains, maize grits, and rice cones, exhibit a good drag flow
in single-screw extruders. Owing to a positive extrusion at excessive die pressures due to a high coefficient of friction, these materials
are preferred and widely used to make breakfast cereals and other snacks. Due to high die pressures, one more flow, known as the
pressure flow, in the backward direction, is usually encountered in single-screw extruders. This also helps in proper mixing of the
ingredients.
The net effect of all these flows is the forward extrusion of the material, which has undergone a certain amount of mixing. There
are a number of modeling equations predicting the nature of the flow of the extrudate as effected by these forces in the literature.
Rossen and Miller (1973) have described a simplified model as:
Qnet ¼ Qd þ Qp (1)

where Qd is the drag flow, Qnet is the net flow, and Qp is the pressure flow.
The drag flow is a function of screw geometry, a, and the screw speed, N, and can be given as:
Qd ¼ aN (2)
and
1 2 2  ne
a¼ p D h 1 SinFCosF (3)
12 t
where N is the screw speed, D is the inside diameter of the barrel, h is the channel depth of screw, n is the number of parallel
screw channels, e is the axial flight land width, t is the pitch length, and V is the screw helix angle.
The rate of flow in the case of pressure flow is dependent on the die pressure, screw geometry, and material viscosity. It can be
expressed as:
bP
Qp ¼  (4)
mL
where b is the screw constant given by:
1  ne 2
b¼ pD3 1  Sin F (5)
12 t
and P is the pressure at the die, L is the length of screw required to generate pressure at the die (also referred to as the degree of fill),
and m is the Newtonian viscosity.
A single-screw extruder can be divided into three zones, conveying, swelling, and melting/degradation, according to the cooking
transition of rice starch. On the other hand, based on temperature differences, the conveying and swelling zones fall into the cooling
zone of the extruder, whereas the melting zone falls into the heating region. As the flow pattern changes from plug flow reactor to
a continuous stirred tank reactor, the degree of mixing of raw ingredients is enhanced due to prolonged residence time. For example,
Davidson et al. (1984) and Diosady et al. (1985) stated that shear degradation of starch was witnessed only in fully cooked wheat
Extrusion for the Production of Functional Foods and Ingredients 3

amylopectin during single-screw extrusion. Similarly, Rodis et al. (1993) postulated that the fragmentation of corn starch was
affected by both shear and thermal properties at temperatures above 100 C and a moisture content less than 30%. In addition,
the length and length-to-diameter ratio of extruder barrels also affect the properties of the final product in the case of a single-
screw extruder. Van Zuilichem et al. (1990) favored a length:diameter ratio of greater than 30 for reasonably high dextrose equiv-
alent values. Some researchers have also indicated that longer barrels, such as 1222 mm (Hakulin et al., 1983), also yield higher
dextrose equivalent values. Single-screw extruders are useful in low-cost applications where the required throughput of extruded
material is not quite as high.

Twin-Screw Extruders
A twin-screw extruder handles more capacity and more complex raw ingredients and has thus found widespread applications in the
food industry. A typical twin-screw extruder comprises two co- or counterrotating heavy duty screws rotating tightly inside a barrel.
Although there are some examples of counterrotating twin-screw extruders, they are not normally used in food applications owing
to their poor mixing characteristics and being less economical because of the high pressure involved. Corotating extruders can,
however, be operated at higher speeds and offer enhanced mixing of raw materials. In such a system, the material is transferred
from one screw to the other and flow occurs because of a combination of drag flow and positive displacement flow, as in
a single-screw extruder. A simplified flow modeling equation as used by Booy (1980), Denson and Hwang (1980), and Yacu
(1985) can be presented as:
b DP
Qnet ¼ aN  (6)
h L
where N is the screw speed, h is the viscosity, DPDP is the pressure rise, L is the fill length, and a and b are constants based on screw
geometry. To provide constant pumping and to produce high extrusion pressure, it is common to have a section of screw elements
behind the die head zone to be spirally flighted. However, some researchers like Roberts and Guy (1987) have suggested that such
a configuration may be more susceptible to catastrophe such as sudden change and metastability at the equilibrium operating state.
The product quality and the stability of the system depend on the optimization of process variables and feed material. The primary
process variables in a twin-screw extruder consist of screw design and the speed, feed rate, design of the die, barrel length, and barrel
temperature, which are under the direct control of the operator.

Extrusion Process Variables and Their Effects on Quality Characteristics

A number of variables, including screw speed, feed rate, die geometry, barrel length, temperature, moisture content, etc., affect
product quality and selection of raw materials in extrusion cooking (Alam et al., 2016). The control of feed rate, screw speed, barrel
temperature, and barrel pressure will determine the crispness, hardness, and various other quality parameters that are going to
decide the success of the product (Harper, 1981). The high temperatures and strong shear forces developed during extrusion cause
the disintegration of the quaternary and tertiary structure of macromolecules associated with various food components, releasing
moisture. This enhances the moisture level in the feed, which further influences the rheological activity of the food product in the
extruder by increasing the shear load (Harper, 1988). The process of extrusion cooking is to a large extent dependent on pressure
buildup in the barrel, degree of filling, and slippage inside the barrel wall, which determines the residence time and degree of cook
of products. The rate of shear development, and hence the mean residence of ingredients inside the barrel, and heat dissipation into
the dough, which in turn affects dough viscosity, are also affected by the screw speed. Barrel temperature determines the degree of
starch gelatinization and product expansion, possibly due to higher evaporation rate while exiting from the die (Muthukumarappan
and Karunanithy, 2012). However, elevated temperatures can unfavorably affect the product color. The feed rate of an extruder
depends primarily on the screw type, speed, and feed moisture and has a profound effect on residence time, barrel pressure, torque
requirement, and dough temperature. The expansion of extrudates changes inversely with moisture content (Davidson et al., 1984;
Miller, 1985). At higher moisture content, the dough experiences more slippage, hence reducing the forward movement of materials
and subsequent increase of pressure at the die orifice. The cooking process induces macromolecular degradation and structural
changes (Colonna and Mercier, 1983), expressed as changes in melt rheology (Vergnes and Villemaire, 1987) and functional prop-
erties of the product, like water solubility (Fitton, 1986), dispersion viscosity, and water absorption (Doublier et al., 1986). A
number of studies have shown the extent of effect on the functional and structural properties of extruded products as affected
by the aforementioned process variables. In one study on the impact of processing variables like barrel temperature, moisture
content, and screw speed on cassava starch’s physical characteristics, researchers found that some of the starch performance param-
eters (shape, specific volume, expansion index, water absorption index, initial viscosity, and peak and final viscosity) were affected
by barrel temperature, while others (specific volume, color, final viscosity, and retrogradation) were affected by the feed moisture
content (Leonel et al., 2009).
Expansion of extrudates is an important property that governs the quality of end products and is dependent on product moisture
content and degree of cook. Some researchers have proposed empirical models to describe the extrudate expansion process (Tayeb
et al., 1992; Bruin and Jongen, 2003; Ding et al., 2005; Altan et al., 2008). In a study by Perez et al. (2008) on extrusion of a maize
and soybean mixture, it was observed that the extrusion conditions, with respect to expansion and flour dispersion viscosity, of
4 Extrusion for the Production of Functional Foods and Ingredients

moisture content around 16% and intermediate temperatures of 170 C yielded the best-quality products. It has also been observed
that with the addition of 5%–20% soy protein, the mechanical strength of extrudates showed a significant improvement (Launay
and Lisch, 1983). At intermediate temperatures, high moisture, and low screw speeds, starch experiences less breakdown in struc-
ture. In particular, pasting properties, color, and expansion of the dough have been observed to be profoundly influenced by barrel
temperature. The effects of barrel temperature (150 C–210 C), screw speed (120–180 rpm), feed moisture (16%–20%), and bran
composition (10%–50%) on cassava extrudates using a single-screw extruder were studied by Hashimoto and Grossmann (2003). It
was observed that water solubility indices and water absorption were significantly influenced by screw speed, barrel temperature,
and the levels of cassava bran. In another study, screw speed and feed moisture were observed to have a greater effect than barrel
temperature on expansion ratios of starch extrudates from water yam (Oke et al., 2013). At low moisture and high temperature,
Serbio and Chang (2000) observed the highest expansion in yam flour that was extruded using a single-screw extruder.
One of the other essential parameters that decide the consistency of the extruded components is the temperature of the extruder.
Sacchetti et al. (2004) explored the impact of extrusion temperature and feed composition on the physical, functional, and sensory
properties of snack-like products based on chestnut and rice flour. It was noticed that chestnut flour is appropriate for the method of
extrusion cooking when correctly combined with rice flour. Mahungu et al. (1999) found that the profile of isoflavones in a maize/
soy mixture is greatly influenced by the temperature of the extrusion barrel followed by the amount of moisture with small changes
in the content of isoflavone. Chestnut flour (30%) was processed at 120 C together with rice flour to generate a snack-like product
with reduced browning and density. In an experiment conducted by Smithey et al. (1995), the storage stability of the extrudate was
not influenced by either the extrusion temperature or the form of cereal flour. Higher temperature improved the activation energy in
a single-screw extruder for starch conversion (Wang et al., 1992; Zheng and Wang, 1994). Zhao et al. (2011) studied food extrusion
reaction kinetics and observed that, during concurrent heating and shearing, raw food materials underwent a nonisothermal cycle
and, due to variation in starch gelatinization and viscosity of different components present in the food material, a distribution of
residence time in the extruder was observed. In addition, extruder flow has a narrower residence time distribution than pipe flow
because the extruder has cross-channel flows and thus improved mixing. Starch gelatinization with excess water approaches pseudo-
first-order Arrhenius kinetics (Okechukwu and Rao, 1996; Turhan and Gunasekaran, 2002).
While temperature and moisture content have a profound effect on the product quality, numerous researchers have also studied
the simultaneous effects of process parameters on a number of product-specific quality parameters. For yam flour extruded using
a single-screw extruder, a high water solubility index was recorded at high barrel temperature and high feed humidity, but the great-
est expansion was observed at low humidity and high temperature (Serbio and Chang, 2000). In comparison, the properties of
double-extruded potato starch, obtained by grinding and reextruding potato starch, varied from those of single-extruded starch
and were determined by manufacturing temperatures (Tomaszewska-Ciosk et al., 2012). Bisharat et al. (2015) studied the effects
of feed moisture (14%–19%, wb), screw speed (150–250 rpm), and extrusion temperature (140 C–180 C) on antioxidant func-
tion, phenolic quality, and sensory properties of vegetable-enriched corn-based extruded snacks using broccoli flour (4%–10%)
and olive paste (4%–8%) blends in broccoli/corn and olive paste/corn formulations. A rise in the temperature of extrusion and
the amount of broccoli flour in broccoli/corn products resulted in increased antioxidant activity and phenolic material. Antioxidant
production improved with higher levels of olive paste in olive paste/corn products. Crispness and cohesiveness values were found to
be low at a higher feed moisture. Similarly, increased density, water absorption index, and hardness and decreased expansion and
water solubility index were recorded for cocoyam extrusion processing at high levels of feed moisture, while high barrel temperature
increased the expansion ratio and water solubility index (Peluola-Adeyemi and Idowu, 2014).
A reciprocal association between expansion ratio and both moisture content and fish flour was recorded in a study on the effects
of extrusion parameters and Labeo rohita fish flour (10%, 15%, 20%, 25%, and 30%) on the properties of extruded products (Singh
et al., 2014). In extruded snacks derived from thin ribbonfish and rice flour using a twin-screw extruder at various barrel temper-
atures, the most appropriate substance was formulated from 10% thin fish at 90 C barrel temperature, with higher lysine, glutamic
acid, and leucine content of products containing thin fish (Dileep et al., 2010). Barrel temperature had a noticeable effect on the
expansion ratio of products made with 10% fish mince, and product breaking strength increased with the addition of 20% fish
mince. Another study tested the physical properties and oxidative stability of 0.5%–2.5% fortified extruded snacks with fish oil
(Pankyamma et al., 2014). Feed moisture mostly influenced the porosity, expansion ratio, and crispness followed by the amount
of fish oil. The amount of incorporation of fish oil was oxidatively stable up to 0.8%.

Development of Functional Products Using Extrusion Techniques

A wide variety of food materials can be processed in extruders depending on the raw materials used and processing conditions
employed. To develop functional food such as snack foods, ready-to-eat (RTE) cereals, texturized vegetable proteins (TVPs), pasta
products, pet foods, animal feed products, etc., using extrusion cooking, the preferable raw materials, other than cereals, are pulses,
legumes, fruits, vegetables, oilseeds, roots, tubers, animal products, nuts, seeds, and their by-products.

Food Groups as a Source of Functional Ingredients


The most popular raw ingredient for extrusion cooking has been cereal flour with minor additions of diverse food materials that
comply with the material characteristics of the equipment. The most commonly used raw materials are flours and grits of starchy
Extrusion for the Production of Functional Foods and Ingredients 5

foods, like maize, wheat, rice, and potatoes, with starch-containing foods such as sorghum, rye, buckwheat, cassava, and oats
finding importance lately. The raw materials are added in varied combinations and proportions, yielding products of different
colors, textures, functionalities, and nutritional properties depending upon the type of raw ingredient and cooking conditions
(Riaz, 2000). The details are discussed next.

Cereals
Cereals form the raw material of choice for extrusion cooking processes because of their starch content and abrasive nature that suits
the material characteristics of most extruders. Most cereals have a similar overall composition, which is often low in proteins and
high in carbohydrates, with some exceptions of oats and maize, which have higher lipid content (7%–9%) than other cereals (1%–
2%) (Haard, 1999; Riaz, 2010). The most popular extruded products from cereals are breakfast cereals, snacks, and other products
like breads, pasta, modified starches, confectionary, and croutons (Ilo et al., 2000). The primary functional ingredient in RTE cereal
formulations is grain or secondary-derived components like flour, whole grain, or fractions (Whalen et al., 2000). There are two
types of breakfast cereals: RTE cold and hot cereals (traditional cereals) (Offiah et al., 2019). Extrusion techniques have evolved
over time to reduce the in-home preparation time of breakfast cereals. As such, highly convenient, RTE products are being manu-
factured using more sophisticated extrusion methods (Neulicht and Shular, 1995). In the hot cereal category, the majority of the
products are made from wheat and oats, while corn and rice make up a smaller percentage. Based on the product form, RTE extruded
cereal products can be classified into extruded shredded cereals, extruded flaked cereals, extruded puffed cereals, granola, and
directly expanded cereals (Neulicht and Shular, 1995; Whalen et al., 2000). In the production of extruded flaked cereals, the
cereal-based ingredients are kneaded through a discharge orifice and the dough is cut into pellets of specific size (Neulicht and Shu-
lar, 1995). To reduce the in-home cooking time of extruded cereals, the starch is cooked or gelatinized in the extrusion process,
which depends on temperature, time, water availability, and shear (Ilo et al., 2000), hence affecting the product quality. Starch gela-
tinization is often accompanied by expansion of the products, and the best-suited cereals for expansion are degermed corn/grits and
rice (Riaz, 2006). A higher lipid content in cereals restricts the forward motion of the dough because of slippage inside the barrel (Ilo
et al., 2000; Riaz, 2010). In addition, lipids are more susceptible to rancidity and can thus reduce product quality with time (Guy,
2012). On the other hand, starches containing 5%–20% amylose content would undergo a favorable expansion, thus improving the
textural attributes and consumer appeal of breakfast cereals and snacks (Riaz, 2010). Since degermed corn expands to a higher
degree than whole corn, it is often preferred as a raw material for extruded products that require some degree of expansion
(Riaz, 2010). The granulation of starch determines the texture of the extruded product and is decided by the type of extruder
used and the nature of the breakfast cereal or snack needed. A fine granulation will impart a fine texture and softness in bite, whereas
crunchy texture in the product is due to the coarse granulation of starch used. Wheat flour is often used in a wide range of extruded
products, including pasta, snacks, breads, and flavored crackers. In more recent times, cereal processing firms have been developing
a particular interest in using underutilized and pseudo-cereals, like sorghum, millets, rye, and primitive wheat species, in extrusion
applications (Offiah et al., 2019).

Roots and Tubers


Major root and tuber crops common to the tropics include potato (Solanum spp.), cassava (Manihot esculenta Crantz), sweet potato
(Ipomoea batatas L.), yam (Dioscorea spp.), and cocoyams (Colocasia spp. and Xanthosoma sagittifolium) (Chandrasekara and Kumar,
2016). Among these, potato and cassava are the most commonly used in extruded snacks, as flour, granules, starch, or flakes (Ilo
et al., 2000). These crops are rich in carbohydrates and have high potential in the production of extruded products. Potato and cas-
sava, in the form of flour, starch, flakes, or granules, are most commonly used to produce ready-to-use extruded snacks and other
products (Ilo et al., 2000). Potato starch is constituted of 20%–25% amylose and has a low lipid content (0.1%–0.2%) (Guy, 2012;
Riaz, 2006), and can thus impart additional expansion and better functional properties to extruded products. Dough made with
potato starch has been reported to show better swelling and binding properties and flow in a highly viscous state, too. In contrast,
tapioca starch contains around 17% amylose and is mostly used in third-generation snack preparations (Riaz, 2006). On the other
hand, granules from these crops are produced from diced and tempered tubers, which are cooked and dried in a standard process
(Riaz, 2006). The effectiveness of these granules changes with the differences in raw materials and processing conditions. Guy
(2012) observed that dough made from potato granules had lower stiffness and stickiness compared with dough made from potato
flour. Extrusion cooking has added value and functionality to root and tuber crops. The use of flours in extrusion has not only diver-
sified the range of snacks and other products, but also added economic importance to these crops. Modified starches exhibit a higher
viscosity, suppressed retrogradation tendency, and improved gel strength compared with their native counterparts (Santana et al.,
2014).

Legumes and Oilseeds


Pulses form an important source of protein in the diets of people especially from developing countries. Some of the important
pulses include pigeon pea (Cajanus cajan), dry beans (Phaseolus spp.), dry peas (Pisum spp.), dry cowpea (Vigna unguiculata), dry
broad beans (Vicia faba), chickpea/garbanzo (Cicer arietinum), Bambara groundnut (Vigna subterranea), and lentils (Lens culinaris)
(Berrios, 2016). Pulses like dry beans, lentils, and peas have been classified as high protein and vitamin-rich vegetables by the
USDA food pyramid dietary guidelines (Berrios, 2016). These proteins are mainly in the form of albumins and globulins. In addi-
tion, pulses also contain a number of antinutritional components, such as saponins, lectins, cyanogenic compounds, and trypsin
and chymotrypsin inhibitors, which can be destroyed by extrusion cooking due to their heat sensitivity (Soetan and Oyewole,
6 Extrusion for the Production of Functional Foods and Ingredients

2009). In the formulation of RTE snacks and other products, flours from pulses are often used in addition to other starchy compo-
nents (Hulse, 2012). The incorporation of leguminous products can be instrumental in the enhancement of flavors, nutrition, and
some physical attributes. Moreover, protein-rich fractions such as protein concentrates from legumes are being used to process meat
analogs and TVPs (Hulse, 2012). Meat analogs are extruded products having a high protein content and similar in texture and
mouthfeel to meats (Kearns et al., 2013), whereas texturized proteins can be formulated from a range of raw materials. General
specifications that favor the extrusion of texturized proteins usually include a protein dispersibility index (PDI) of 20–80, fat
content of 0.5%–6.5%, up to 8 mesh particle size, and a fiber content of around 7% (Kearns et al., 2013). Despite the diversity
in legumes, soybean protein in the form of defatted soy flour has been the raw material of choice in the processing of texturized
proteins owing to its properties of having a minimum protein content of 50% and PDI of 60–70. The addition of legumes to extru-
sion formulations has been observed to improve protein digestibility and overall protein content, making the formulated foods
more nutritious. For example, Patil et al. (2016a,b) observed that extruded products made from wheat flour supplemented with
green pea, yellow pea, chickpea, and lentils had 1%–1.5% higher protein content and exhibited an improved protein digestibility
of up to 62%. The added nutritious value in extruded snacks is attracting health-conscious consumers and is hence gaining popu-
larity (Riaz, 2006). Several formulations, including adding full-fat soybean and maize (Rweyemamu et al., 2015) or partially defat-
ted soybean with maize (Obatolu et al., 2006), have been successfully used in the processing of maize-based snacks. Similarly,
protein-supplemented pasta was observed to have the best quality when the ingredients included 15% mung bean flour, 10%
whey protein concentrate, and 6% egg albumen (Savita et al., 2013).

Fruits and Vegetables


Fruits and vegetables are rich in dietary fibers and other functional compounds like anthocyanins, flavonoids, and carotenoids (Dji-
las et al., 2009) and are also used as additives in raw material formulations in extrusion processes in the form of ground pomace.
The use of natural additives in RTE products and breakfast cereals enhances their health appeal and nutritional quality. The presence
of a high moisture content in fresh produce is a limiting factor for their use in extrusion cooking; however, concentrates or spray-
dried products have been used successfully in product formulations. Camire et al. (2007) used a blend of degermed white cornmeal
and dehydrated powders from Concord grape, cranberry, blueberry, and red raspberry fruits to produce functional extruded prod-
ucts. It was observed that products having cranberry fruit powder showed the highest antioxidant activity compared with those con-
taining other fruit powders. In another study, directly expanded snacks were developed from whole cornmeal with 20% or 50%
additions of dehydrated powders of Indian gooseberry, Kulfa leaves, pumpkin, or curry leaves in a twin-screw extruder (Karkle
et al., 2009). It was observed that snacks made with 50% dehydrated powder showed lower radial expansion compared with those
made with only 25%, except for gooseberry. Although snacks made with gooseberry, pumpkin, or curry leaves had a lower breaking
force than 100% cornmeal, used as control, they obtained acceptable scores from a sensory panel for all five kinds of dehydrated
powders used. Furthermore, rice/maize-based products supplemented with grapes, guava, bananas, beans, tomatoes, and pumpkin
pulps have been developed to enhance nutritional value and taste. While the introduction of 10% fruit or vegetable pulp to the
products provided additional nutritional and structural advantages, maximum expansion was recorded for both rice and maize
products mixed with banana pulps (Jain et al., 2013). In another study, increased levels of apple pomace in extruded corn and
hydrodynamic cavitated sorghum flour combined with low die temperature and screw rate resulted in increased overall phenolic
content and antioxidant activity (Lohani and Muthukumarappan, 2016). Improved products with higher phenolic content and
increased antioxidant activity, texture, and functional properties were obtained at higher levels of apple pomace. In addition,
product formulations of rice flour and cactus pear fruit pulp were extruded using a twin-screw extruder at 15 kg/h feed frequency,
13% (w/w) feed humidity, 400 rpm screw speed, and 40:1 L/D ratio (Sarkar et al., 2011). Changes in the amount of cactus pear fruit
improved breaking force and apparent product density, but the ratio of porosity and radial expansion decreased with increases in
fruit solids level.

Animal Products
Animal products are a rich source of fats and proteins that can supplement the nutritional requirements of extruded foods. Products
like minced fish, egg white powder, meat, milk powder, fish flour/powder, and cheese have found wide applications in formulations
of novel extruded products (Dileep et al., 2010; Dubey, 2011; Muralidharan, 1999). The reasoning behind these ingredients being
incorporated into extruded products includes new product development, improved nutritional value, market creation, flavor gener-
ation, and diversification of ingredients. The integration of underused seafood has drawn recent interest in science, and the intro-
duction of these high-protein fractions into extruded goods creates value for low-cost and underused seafood, thus increasing their
use (Surasani, 2016). Typically, with proper packaging, these extruded products have a shelf life of about 4–6 months. Compared
with fish powders, fish pastes and minced fish are more economical. Fish- and meat-based products are manufactured using extru-
sion cooking, with high barrel temperature, feed humidity, protein and starch content, and screw speeds being the main factors
during the cycle that influence product quality (Surasani, 2016). Fish (catla catla) mince and cheese flavor were used in a mix
with refined wheat flour, green gram, semolina, and black gram for the extrusion of pasta and it was observed that the product,
which contained refined wheat flour, semolina, black gram dal, cheese flavor, and fish mince in a ratio of 32.5:32.5:10:5:20, had
the most preferred quality (Devi et al., 2013). In another study, pasta formulated with fish (Oreochromis niloticus) flour contained
higher amounts of protein, lipid, ash, total essential amino acids, and polyunsaturated fatty acids. Furthermore the incorporation of
fish powder decreased lightness and water activity while showing negligible effects on the chemical stability of the developed pasta
stored at 25 C for 21 days (Monteiro et al., 2016). The most desirable nutrient-rich RTE snacks in terms of quality and acceptability
Extrusion for the Production of Functional Foods and Ingredients 7

were developed from croaker fish flour (Johnius dussumieri) and blends of rice flour, corn flour, and soybean flour, using a twin-screw
extruder, when used in the proportions of 18% fish flour, 45% rice flour, 30% corn flour, and 5% soybean flour, whereas the
product containing less fish flour (12%) had better expansion (Mulye and Zofair, 2015). Extruded snacks and products containing
fish meal from four fishes (O. niloticus, Salmo salar, Thunnus spp., and Sardinella brasiliensis) were observed to naturally contain higher
protein content (Goes et al., 2015). In other examples, snacks have also been developed using chicken and other meats as additives
with better functionality and nutritional values. In one study, chicken meat noodles were prepared from chicken meat mixed with
whole wheat flour with meat inclusion levels of 0%, 30%, 40%, and 50%, and it was observed that with increased meat levels, crude
fiber, yield, water solubility index, volume, and cooking loss were reduced, whereas an increase in water absorption index, ash, fat,
and protein was observed with increased meat levels (Verma et al., 2014). In this study, the most desired noodles were those that
contained 30% meat.

Plant and Animal By-products


Plant and animal by-products are a rich source of nutritional and functional compounds that are discarded as food loss from various
processing operations (Shilev et al., 2006). The food losses per capita in South/Southeast Asia and sub-Saharan Africa have been
estimated to be approximately 120–170 kg/year (FAO, 2011). A significant amount of food products are rendered nonusable by
food processing operations globally. The Food and Agriculture Organization has estimated that around 33% of food products
intended for human consumption are wasted during the food processing chain (FAO, 2011). Extrusion cooking has offered an
invaluable technological platform to recover these food wastes by using them as supplementary ingredients in the production of
extrudates. There have been numerous efforts to integrate food wastes into extrusion cooking, and as a result a number of products
have been formulated that exhibit varied sensorial, functional, nutritional, textural, and physical characteristics (Altan and Maskan,
2016). The residues used in extrusion can be from diverse food groups, including fruits and vegetables, meat and dairy products,
sugar and confectionary products, fat and oil processing, grain mill and bakery products, and other sundry food preparation resi-
dues (Kasapidou et al., 2015). Of all these, by-products from fruits and vegetables are estimated to amount to 30% of the processed
material (Gowe, 2015; Kasapidou et al., 2015). These by-products are valuable sources of vitamins, antioxidants, dietary fiber,
minerals, essential fatty acids, and phytochemicals, including phytosterols, polyphenols, carotenoids, and hesperidin (Ezejiofor
et al., 2014; Kasapidou et al., 2015; Varzakas et al., 2016), having numerous health benefits (Altan and Maskan, 2016). Some fruit
and vegetable residues and by-products that have been used to develop extruded foods include pineapple waste pulp (Kothakota
et al., 2013), tomato peel and seed (Devi et al., 2016), cauliflower trimmings (Stojceska et al., 2008), milled orange peel, carrot
pomace (Alam and Kumar 2014; Kumar et al., 2010), tomato and grape pomace (Altan et al., 2009), avocado seeds (Olaeta
et al., 2007), cranberry pomace (White et al., 2010), and dehydrated naranjita fruit bagasse (Ruiz-Armenta et al., 2018).

Functional Foods Developed With Extrusion Cooking


Snack Foods
The adoption of extruded foods by consumers is primarily due to the ease, taste, desirable appearance, and texture considered to be
unique to these foods, especially when it comes to snack items (Harper, 1981). Snack food extrusion involves chosen grains being
exposed to a variety of complex physical processes to create snacks of varying shapes and textures (White, 1994). In the extrusion of
snack foods, grain and other materials are combined and cooked under pressure, with shearing in a barrel, which is also called
a pipe, at high temperatures. The resultant mass is compressed through a die and sliced into individual pieces and takes different
shapes that customers expect in the retail snack food aisles (Harper, 1981). Current materials, state-of-the-art extrusion technolo-
gies, and creative processing methods are merged to deliver new snack foods with ever-increasing appeal to health-conscious
customers searching for different textures and mouth sensations in convenient foods (Pamies et al., 2000). The processing variables,
such as feed rate, screw speed, barrel pressure, and temperature, determine the quality characteristics of extruded snacks that deter-
mine the acceptability of that product (Harper, 1981). A modern extruded snack food product’s success or failure is directly related
to sensory characteristics, where texture plays an important role. Texture is of great importance in such foods, in which expansion is
required and crispness is one of the most important attributes (Pamies et al., 2000). Snacks were prepared from normal maize and
high-quality protein maize in a single-screw extruder (Martinez et al., 1996), chickpea flour (Carrillo et al., 2000), and sweet whey
solids in combination with cornmeal and potato flour (Onwulata et al., 2001). To improve expansion, sometimes high-temperature
short-time air puffing is incorporated to give better extruded products. For example, in the production of potato–soy RTE snack
foods, air puffing was used to produce lightly textured and highly porous snacks, ideally at puffing temperature in the range of
185 C–255 C and puffing time of 20–60 s (Nath and Chattopadhyay, 2008).

Ready-to-Eat Cereals
RTE cereals have become popular among people across all age groups, saving cooking time and addressing most of the nutritional
needs. Extrusion cooking offers advantages, including low processing costs, faster processing times, less area required for processing
equipment, and greater flexibility in design and structure of end products. It has become very common to supplement the cereal
ingredients with by-products and other additives to target a specific class of consumers. RTE products made with cassava flour
and semi-defatted Brazil nut cake, rich in vegetable proteins, was made by Souza and Menezes (2008). In another study, blends
of rice and barley flour were used to prepare RTE extruded products (Gupta et al., 2008). It was observed that a feed moisture
content of 30%, barley flour content of 20%, and die temperature of 175 C yielded the most acceptable product. Whey protein
8 Extrusion for the Production of Functional Foods and Ingredients

isolate and cornmeal used to produce nutritious RTE crunchy snacks (Onwulata et al., 2010), Mesona blumes gum and rice flour to
make rice extrudates (Zhuang et al., 2010), and instant cereal–legume products made from a blend of cereal products, legumes, and
fats (Lorenz and Jansen, 1980) are some further examples of RTE formulations.

Texturized Vegetable Protein


TVPs are becoming an important source of high-density plant-based proteins, especially for vegetarian people. TVPs are made by
restructuring protein molecules, usually soy protein, into a cross-linked, layered mass whose structure resists deformation upon
application of heat or further processing. They can be of two types: extrusion cooked meat analogs and meat extenders (Alam
et al., 2016). Barrel temperature is one of the most critical factors for texturization of plant proteins. Areas (1992) reported that
a proportional decrease in disulfide linkages in soy protein isolates was observed when temperature was increased from 140 C
to 180 C, while temperatures less than 90 C are not favorable for expansion and layer formation (Cheftel et al., 1992). TVPs
have also been used as animal feed due to their nutritional quality and ease of processing. Perilla et al. (1997) extruded poultry
diets from full-fat soybeans at 122 C and 126 C, which led to maximum growth in broiler chickens, whereas diets extruded at
118 C and 120 C resulted in significantly lower body weights. Cheftel et al. (1992) also derived nutrient-rich fat analogs from
whey protein isolates and cheese analogs from caseinate and butter oil. In another study, protein-rich puffed snacks were made
from cornmeal using a twin-screw extruder (Onwulata et al., 2010). Whey protein isolate was texturized to minimize the water-
binding property of dairy proteins and nonfat dried milk. Moreover, the protein quality of texturized whey protein isolate remain
unaffected with the change in moisture content.

Pasta Products
Pastas are wheat-based products that are formed from dough, with or without leavening. The ingredients generally comprise flour
and water, although eggs are sometimes added. The most suitable material for making flour for pasta is durum semolina. It is much
harder than common wheat and gives pasta its yellow color. Wang et al. (1999) processed a pasta-like pea product that exhibited
a compact, dense structure, with relatively less swollen starch granules. The developed products appeared to be completely coated
with a gelatinized starch and protein matrix. Wojtowicz and Moscicki (2011) produced precooked pasta-like products enriched with
wheat bran. The physicochemical properties and water status of fresh pasta (both laminated and extruded) produced with recently
designed mixers were evaluated by Carini et al. (2010). The mixers prompted the formation of a dough in 1–2 s with uniform
hydration of the solids. Manthey et al. (2008) investigated the effects of semolina, flaxseed flour concentration, and hydration level
on the physical and cooking characteristics of freshly extruded pasta. Another study used a single-screw extrusion cooker to process
precooked pasta-like products. It was observed that, with increased bran addition (20% and 25% of bran in the recipe), the firmness
of hydrated products decreased, while the product quality deteriorated, with excess stickiness and adhesiveness when the processing
was carried out at low revolutions per minute. At low screw speed, the bran fractions were also found to be unmodified in micro-
structure, but higher speed unsettled the cell walls of wheat bran. By mixing an edible acid, a farinaceous material, and water, Barnes
et al. (1997) produced an acidified pasta product. The processes included forming the acidified dough into a shape followed by
blending water with the surface-gelatinized product to form a wet, acidified pasta product. After proper steaming, the acidified
wet product was packed.

Pet Food and Animal Feed Products


An extensive application of extrusion technology can be observed in the animal feed product industries. Plavnik and Wan (1995)
carried out a study on feeds for broilers and examined the impact of short-time extrusion and expansion on the nutritional values.
Broiler chicks (18–21 days of age) were used to study the impact of these factors on digestibility and energy of feeds. Corn-based
feed or barley or wheat, which were treated by extrusion or expansion, were compared with the same untreated materials and were
subjected to milling for size equalization. During the process of extrusion of whole feeds, increases in the apparent metabolizable
energy by 3.5% and gross energy digestion by 1.5% were recorded. It was concluded that the energy of common feeds for broilers
was enhanced by the processes of high-temperature short-time extrusion and expansion. The same concept was used by Mathew
et al. (1999) for extruded high-protein pet food to obtain better quality parameters, by the injection of water into the precondi-
tioner. Marsman et al. (1995) studied the extrusion of soybean meal and examined how the shear forces and addition of
a protease–hemicellulase mixture affected the physical, physiological, and chemical parameters during the process. After extrusion,
the chemical analysis indicated 25%–41% reduction in trypsin inhibitor activity and a decrease in lectin content to below detectable
levels. The in vitro protein digestibility in toasted soybean meal showed a negative correlation with trypsin inhibitor activity and was
found to be increased from 60.7% to 81.1%, during extrusion with two twin lead slotted screws. The indices of nitrogen solubility
and protein dispersibility showed a decline due to extrusion, but both failed to segregate between the different shear levels. Nitrogen
solubility index, being used as a quality parameter of soybean meal extrusion, at the levels of 50–558, enhanced feed conversion
and in vitro protein digestibility of broiler chickens. The lipid oxidation rate was significantly higher in pet foods extruded at
300 rpm, compared with those produced at 200 and 400 rpm. As explained by Hsieh et al. (1998), both the fat content and the
feed moisture content had similar effects; the extrudates with larger moisture content caused a lower lipid oxidative rate. A brief
overview of experiments involving the formulation of various extruded products from different raw materials and processing condi-
tions is presented in Table 1.
Extrusion for the Production of Functional Foods and Ingredients 9

Table 1 Various extruded products from different raw materials and processing conditions

Process parameters
Screw speed Feed
Raw Formulation Temperature ( C) (rpm) moisture (%) Final product References

Rice, pea protein isolate 130 400–600 19–23 Pea-fortified rice snack Philipp et al. (2017)
Rice, beetroot, soy, milk powder, potato, 80–150 200–350 d Gluten-free snack Stojceska et al. (2008)
carrot, corn starch, teff flour, cranberry, apple
Rice, taro 79–100 150 40 Floating fish pellets Kamarudin et al. (2018)
Rice grit, durum flour, partially defatted 150–175 200–280 12–18 Rice grit snack Yagcı and Gögüş (2008)
hazelnut flour, and fruit waste
Chestnut flour, rice flour 90–120 80 35 Rice–chestnut flour–based Sacchetti et al. (2004)
snack
Rice, orange, yellow, and red cactus 100–160 250 16 Cactus pear rice–based El-Samahy et al. (2007)
pear fruit extrudate
Rice, soybean, finger millet 184 18 285 Soy-fortified millet–based Seth and Rajamanickam
snack (2012)
Whey protein concentrate, rice 106.36–173.64 140 16–23 Extruded flour da Silva Teba et al.
(2017)
Corn grit, defatted soybean meal, germinated 120–180 30 2–3 Fortified extruded snacks Korkerd et al. (2016)
brown rice meal, and mango peel fiber
Wheat flour and ground yellow 80 240–400 18 Extruded cereals Azzollini et al. (2018)
mealworm larvae
Corn grits, Jerusalem artichoke tubers, 120–160 200–300 – Functional corn snacks Pe˛ ksa et al. (2016)
amaranth seeds, and pumpkin flesh
Wheat bran and flour 50–140 250 – Extruded snacks Fleischman et al. (2016)
Lentil, chickpea, green pea, and yellow pea flour 180 210 12 Wheat-based extrudates Patil et al. (2016a,b)
Oat fiber, corn flour, and curcuminoids 80–110 200–300 21–35 Extruded snack products Sayanjali et al. (2019)

Impact of Extrusion Cooking on Nutritional Quality of Foods

Extrusion cooking involves mechanical and heat effects that permanently or semipermanently change the food structure. Hence
effects on several nutritional factors are expected. These effects can be beneficial or otherwise. Among the beneficial effects, starch
gelatinization, destruction of antinutritional factors, reduction of lipid oxidation, and increase in dietary fiber content are the most
notable ones. However, the process may also reduce the nutritional value of proteins by inducing Maillard reactions or causing
a loss in heat-labile vitamins and other compounds. Effects on phytochemicals, some nonnutrient healthful components, and
minerals may or may not be beneficial. Among the beneficial effects, it has been observed that the protein digestibility of extruded
products is higher than nonextruded counterparts. This may possibly be because of a combined effect of protein denaturation and
inactivation of antinutritional factors that restrict digestibility (Singh et al., 2007). Research has shown an increase in digestibility of
vegetable proteins after mild extrusion cooking (Hakansson et al., 1987; Colonna et al., 1989; Areas, 1992). Protein denaturation
and enzyme inactivation caused by extrusion may expose new sites for enzymatic activity (Colonna et al., 1989). The feed ratio,
among all the process variables, was found to have a maximum effect on the digestibility of proteins, followed by process temper-
ature in the case of a fish–wheat flour blend (Bhattacharya et al., 1988; Camire et al., 1990). With an increase in the process temper-
ature (100 C–140 C), protease inhibitors are inactivated more effectively, which also contributes to an increase in the protein
digestibility. For proper residence times inside the extruder, extrusion temperatures as high as 140 C do not have any unwanted
effect on the protein digestibility. One of the major advantages of extrusion cooking is the destruction of antinutritional factors,
including hemagglutinins, phytates, tannins, and trypsin inhibitors, all of which restrict the digestibility of proteins (Bookwalter
et al., 1971; Lorenz and Jansen, 1980; Armour et al., 1998; Alonso et al., 1998). The inactivation of trypsin inhibitors has been
observed to increase with extrusion temperature, residence time, and moisture content (Bjorck and Asp, 1983). Among the
unwanted changes, the Maillard reaction, which involves amino and carbonyl groups, leading to browning and unwanted flavor
production, can reduce consumer appeal and acceptability of the products. The importance of this effect is more significant for
animal feeds, which are formulated for special nutritional needs like weaning or used as a whole item in the diet (Fukui et al.,
1993). The cooking conditions of high barrel temperature and low feed moisture usually favor the Maillard reaction. Beaufrand
et al. (1978) reported a lysine loss of around 80% at 170 C and 10%–14% feed moisture and 60 rpm screw speed. One more factor
due to high temperatures that may unfavorably affect the product is the formation of acrylamides. They may be formed due to extru-
sion, baking, and roasting processes in cereal-based products (Studer et al., 2004). The use of twin-screw extruders with high
mechanical and thermal inputs can promote acrylamide formation. Kretschmer (2004) has, however, reported that their concen-
tration can be reduced by reducing the temperatures and subsequently increasing the moisture content. It has also been observed
that free amino acids are more sensitive to damage due to extrusion cooking than those in proteins. In a study by Maga and Sizer
(1978), tyrosine, serine, isoleucine, phenylalanine, and lysine content decreased considerably during the extrusion of potato flakes
10 Extrusion for the Production of Functional Foods and Ingredients

at 70 C–160 C temperature, 100 rpm screw speed, and 48% feed moisture. In addition, a number of studies have confirmed sugar
losses due to extrusion (Noguchi et al., 1982; Camire et al., 1990). It may be because of the conversion of sucrose into glucose and
fructose and their subsequent reaction with proteins due to the Maillard reaction. In the case of starch, extrusion cooking is very
unique for the gelatinization process as it occurs at very low moisture levels (12%–22%) compared with other conventional
methods (Qu and Wang, 1994). The shear-induced molecular weight reductions are based on the molecular size of amylopectin
molecules; the higher the initial molecular size, the higher the reduction. Screw configurations in an extruder can be designed specif-
ically to minimize or maximize the breakdown of starch in raw ingredients (Gautam and Choudhury, 1999).
The extent of the impact of extrusion cooking on vitamins varies widely according to their chemical structure and composition.
Environmental variables, like moisture, time, oxygen, temperature, light, and pH, also determine the extent of vitamin degradation.
Among the water-insoluble vitamins, vitamins D and K are quite stable, while vitamins A and E and related compounds, like carot-
enoids and tocopherols, degrade significantly under heat and oxygen (Killeit, 1994). Furthermore, high-quality ingredients and only
legally permitted additives should be used (Guy, 2004). The structural changes encountered by macromolecules like proteins and
starches define the underlying texture and design of the final cooked products. In protein-based extruded foods, proteins are pref-
erably chosen from oilseeds like soybean or from fractionated cereal proteins (Asgar et al., 2010).

Concluding Remarks

Extrusion cooking provides an excellent opportunity to process tailored foods with targeted functional properties. Foods with
improved functionality can be processed to manage lifestyle diseases or for the general well-being of specific consumers. Extrusion
is one of the fastest growing food processing techniques for the production of new, nutritive, and creative products using inexpen-
sive raw ingredients and with short cooking time. Although conventional food extruders have advanced tools to handle mostly
cereal- or starch-based raw materials, recent advances and innovations are being explored to accommodate new ideas and customer
demands for healthy foods. Moreover, extrusion cooking allows restructuring and conversion of plant-based proteins into meat
analogs. These products not only taste, feel, and smell better, but are also attracting a huge consumer demand. This only confirms
the fact that extrusion cooking can become an instrumental tool in the generation of novel foods or even creating new markets for
these products. A processor, however, needs to adhere to the optimum operating conditions, as extreme extrusion and/or improper
raw material formulations can hinder the intended functionality in the extruded products. To ensure a higher retention of nutrients
and phytonutrients, improved starch digestibility, and high protein content, mild extrusion processing conditions should be used.
Future research should be aimed at exploring the possibilities of developing specific functional foods using advanced extrusion tech-
nologies exhibiting versatile health benefits. Furthermore, the inclusion of underutilized crops having the desired functionality and
utilization of by-products should be encouraged to produce healthy extruded products. Thus, extrusion cooking will continue to
remain a mainstay technology for producing nutritionally rich formulated foods with targeted functionality in the near future.

References

Alam, S., Kumar, S., 2014. Optimization of extrusion process parameters for red lentil-carrot pomace incorporated ready-to-eat expanded product using response surface. Food Sci.
Technol. 2, 106–119.
Alam, M.S., Kaur, J., Khaira, H., Gupta, K., 2016. Extrusion and extruded products: changes in quality attributes as affected by extrusion process parameters: a review. Crit. Rev.
Food Sci. Nutr. 56 (3), 445–473.
Alonso, R., Orue, E., Marzo, F., 1998. Effects of extrusion and conventional processing methods on protein and antinutritional factor contents in pea seeds. Food Chem. 63,
505–512.
Altan, A., Maskan, M., 2016. Development of extruded foods by utilizing food industry by-products. In: Maskan, M., Altan, A. (Eds.), Advances in Food Extrusion Technology. CRC
Press, Boca Raton, pp. 121–160.
Altan, A., McCarthy, K.L., Maskan, M., 2008. Extrusion cooking of barley flour and process parameter optimization by using response surface methodology. J. Sci. Food Agric. 88,
1648–1659.
Altan, A., McCarthy, K.L., Maskan, M., 2009. Effect of extrusion process on antioxidant activity, total phenolics and B-glucan content of extrudates developed from barley-fruit and
vegetable by-products. Int. J. Food Sci. Technol. 44, 1263–1271.
Areas, J.A.G., 1992. Extrusion of food proteins. Crit. Rev. Food Sci. Nutr. 32, 365–392.
Armour, J.C., Perera, R.L.C., Buchan, W.C., Grant, G., 1998. Protease inhibitors and lectins in soya beans and effect of aqueous heat-treatment. J. Sci. Food Agric. 78, 225–231.
Asgar, M.A., Fazilah, A., Huda, N., Bhat, R., Karim, A.A., 2010. Non meat protein alternatives as meat extenders and meat analogs. Compr. Rev. Food Sci. Food Saf. 9, 513–529.
Azzollini, D., Derossi, A., Fogliano, V., Lakemond, C.M.M., Severini, C., 2018. Effects of formulation and process conditions on microstructure, texture and digestibility of extruded
insect-riched snacks. Innovat. Food Sci. Emerg. Technol. 45, 344–353.
Barnes, G.J., Collins-Thompson, D., Hsu, J.Y., 1997. Acidified pastas. Trends Food Sci. Technol. 8 (10), 348.
Beaufrand, M.J., De la Gueriviere, J.F., Monnier, C., Poullain, B., 1978. Influence of the extrusion process on the availability of proteins. Ann. Nutr. Food 32, 353–364.
Berrios, J.D.J., 2016. Extrusion processing of main commercial legume pulses. In: Maskan, M., Altan, A. (Eds.), Advances in Food Extrusion Technology. CRC Press, Boca Raton,
pp. 209–236.
Bisharat, G.I., Lazou, A.E., Panagiotou, N.M., Krokida, M.K., Maroulis, Z.B., 2015. Antioxidant potential and quality characteristics of vegetable-enriched corn-based extruded
snacks. J. Food Sci. Technol. 52 (7), 3986–4000.
Bjorck, I., Asp, N.G., 1983. The effects of extrusion cooking on nutritional value. J. Food Eng. 2, 281–308.
Bhattacharya, S., Das, H., Bose, A.N., 1988. Effect of extrusion process variables on in-vitro protein digestibility of fish-wheat flour blends. Food Chemistry 28 (3), 225–231.
Bookwalter, G.N., Mustakas, G.C., Kwolek, W.F., Mcghee, J.E., Albrecht, W.J., 1971. Full-fat soy flour extrusion cooked-properties and food uses. J. Food Sci. 36, 5–9.
Booy, M. L., 1980. Polym. Eng. Sci. 20, 1220.
Extrusion for the Production of Functional Foods and Ingredients 11

Bordoloi, R., Ganguly, S., 2014. Extrusion technique in food processing and a review on its various technological parameters. Ind. J. Sci. Res. and Tech. 2, 1–3.
Brennan, J.G., Grandison, A.S., 2012. Food Processing Handbook. John Wiley and Sons, United Kingdom.
Bruin, S., Jongen, T.R.G., 2003. Food process engineering: the last 25 years and challenges ahead. Compr. Rev. Food Sci. Food Saf. 2, 42–812.
Camire, M.E., Camire, A.L., Krumhar, K., 1990. Chemical and nutritional changes. Crit. Rev. Food Sci. Nutr. 29, 35–57.
Camire, M.E., Dougherty, M.P., Briggs, J.L., 2007. Functionality of fruit powders in extruded corn breakfast cereals. Food Chem. 101, 765–770.
Carini, E., Vittadini, E., Curti, E., Antoniazzi, F., Viazzani, P., 2010. Effect of different mixers on physicochemical properties and water status of extruded and laminated fresh pasta.
Food Chem. 122 (2), 462–469.
Carrillo, J.M., Moreno, C.R., Rodelo, E.A., Trejo, A.C., Escobedo, R.M., 2000. Physicochemical and nutritional characteristics of extruded flours from fresh and hardened chickpeas
(Cicer arietinum L). LWT - Food Sci. Technol. 33, 117–123.
Chang, Y.H., Ng, P.K., 2009. Effects of extrusion process variables on extractable ginsenosides in wheat-ginseng extrudates. J. Agric. Food Chem. 57 (6), 2356–2362.
Chandrasekara, A., Kumar, T., 2016. Roots and tuber crops as functional foods: a review on phytochemical constituents and their potential health benefits. Int. J. Food Sci. 1–15.
Cheftel, J.C., Kitagawa, M., Queguiner, C., 1992. New protein texturization processes by extrusion cooking at high moisture levels. Food Rev. Int. 8, 235–275.
Clextral, 2010. Twin Screw versus Single Screw in Feed Extrusion Processing. Clextral Technologies and Lines, Firminy, France.
Colonna, P., Mercier, C., 1983. Macromolecular modifications of manioc starch components by extrusion-cooking with and without lipids. Carbohydrate Polymers 3 (2), 87–108.
Colonna, P., Tayeb, J., Mercier, C., 1989. Extrusion cooking of starch and starchy products. In: Mercier, C., Linko, P., Harper, J.M. (Eds.), Extrusion Cooking. American Association
of Cereal Chemists, Inc., St. Paul, MN, pp. 247–319.
da Silva Teba, C., da Silva, E.M.M., Chávez, D.W.H., de Carvalho, C.W.P., Ascheri, J.L.R., 2017. Effects of whey protein concentrate, feed moisture and temperature on the
physicochemical characteristics of a rice-based extruded flour. Food Chemistry 228, 287–296.
Davidson, V.J., Paton, D., Diosady, L.L., Rubin, L.J., 1984. A model for mechanical degradation of wheat starch in a single-screw extruder. J. Food Sci. 49, 1154–1157.
Denson, C.D., Hwang, B.K., 1980. The influence of the axial pressure gradient on flow rate for Newtonian liquids in a self wiping, co-rotating twin screw extruder. Polym. Eng. Sci.
20, 965–971.
Devi, B.K., Kuriakose, S.P., Krishnan, A.V.C., Choudhary, P., Rawson, A., 2016. Utilization of by-product from tomato processing industry for the development of new product.
J. Food Process. Technol. 7, 608.
Devi, L.N., Aparna, K., Kalpana, K., 2013. Utilization of fish mince in formulation and development of pasta products. Int. Food Res. J. 20 (1), 219.
Dileep, A.O., Shamasundar, B.A., Binsi, P.K., Howell, N.K., 2010. Composition and quality of rice flour-fish mince based extruded products with emphasis on thermal properties of
rice flour. J. Texture Stud. 41, 190–207.
Ding, Q., Ainsworth, P., Tucker, G., Marson, H., 2005. The effect of extrusion conditions on the physicochemical properties and sensory characteristics of rice-based expanded
snacks. J. Food Eng. 66, 283–289.
Diosady, L.L., Paton, D., Rosen, N., Rubin, L.J., Athanassoulias, C., 1985. Degradation of wheat starch in a single-screw extruder: mechano-kinetic breakdown of cooked starch.
J. Food Sci. 50, 1697–1699.
Djilas, S., Canadanovic-Brunet, J., Cetkovic, G., 2009. By-products of fruits processing as a source of phytochemicals. Chem. Ind. Chem. Eng. Q. 15, 191–202.
Doublier, J.L., Colonna, P., Mercier, C., 1986. Extrusion cooking and drum drying of wheat starch. II Rheological characterization of starch pastes. Cereal Chem. 63 (3), 240–246.
Dubey, A., 2011. Use of extrusion technology and fat replacers to produce high protein, low fat cheese. Utah State University, Digital Commons.
El-Samahy, S.K., El-Hady, E.A., Habiba, R.A., Moussa-Ayoub, T.E., 2007. Some functional, chemical, and sensory characteristics of cactus pear rice-based extrudates. J. Prof.
Assoc. Cactus Dev. 9, 136–147.
Ezejiofor, T.I.N., Enebaku, U.E., Ogueke, C., 2014. Waste to wealthvalue recovery from agro-food processing wastes using biotechnology: a review. Br Biotechnol J. 4, 418–481.
FAO, 2011. Global Food Losses and Food Waste-Extent, Causes and Prevention. SAVE FOOD: An Initiative on Food Loss and Waste Reduction. Rome.
Faraj, A., Vasanthan, T., Hoover, R., 2004. The effect of extrusion cooking on resistant starch formation in waxy and regular barley flours. Food Res. Int. 37, 517–525.
Fitton, M.G., 1986. Gums and stabilisers for the food industry. In: Phillips, G.O., Wedlock, D.J., Williams, P.A. (Eds.), Gums and Stabilisers for the Food Industry, vol. 3. Elsevier
Applied Science Publishers, London, pp. 213–220.
Fleischman, E.F., Kowalski, R.J., Morris, C.F., Nguyen, T., Li, C., Ganjyal, G., Ross, C.F., 2016. Physical, textural, and antioxidant properties of extruded waxy wheat flour snack
supplemented with several varieties of bran. J. Food Sci. 81 (11), E2726–E2733.
Fukui, K., Aoyama, T., Hashimoto, Y., Yamamoto, T., 1993. Effect of extrusion of soy protein isolate on plasma cholesterol level and nutritive value of protein in growing male rats.
J. Jpn. Soc. Nutr. Food Sci. 46, 211–216.
Gaosong, J., Vasanthan, T., 2000. Effect of extrusion cooking on the primary structure and water solubility of b-glucans from regular and waxy barley. Cereal Chem. 77, 396–400.
Gautam, A., Choudhury, G.S., 1999. Screw configuration effects on starch breakdown during twin screw extrusion of rice flour. J. Food Process. Preserv. 23, 355–375.
Goes, E.S., de Souza, M.L.R., Campelo, D.A.V., Yoshida, G.M., Xavier, T.O., Moura, L.B., Monteiro, A.R.G., 2015. Extruded snacks with the addition of different fish meals. Food Sci.
Technol. 35, 683–689.
Gourgue, C., Champ, M., Guillon, F., Delort-Laval, J., 1994. Effect of extrusion cooking on the hypoglycemic properties of citrus fibre: an invitro study. J. Sci. Food Agric. 64,
493–499.
Gowe, C., 2015. Review on potential use of fruit and vegetables by-products as a valuable source of natural food additives. Food Science and Quality Management 45, 47–61.
Gupta, M., Bawa, A.S., Semwal, A.D., 2008. Effect of barley flour on development of rice-based extruded snacks. Cereal Chem. 85 (2), 115–122.
Guy, R.C., 2004. Extrusion technologies. In: Wrigley, C., Corke, H., Walker, C. (Eds.), Encyclopedia of Grain Science. Academic press, Amsterdam, pp. 366–371.
Guy, R.C.E., 2012. Raw materials for extrusion cooking process. In: Frame, N.D. (Ed.), The Technology of Extrusion Cooking. Springer Science and Business Media, New York,
pp. 52–73.
Haard, N.F., 1999. Fermented cereals: a global perspective. In: Food and Agriculture Organization: FAO Agricultural Services Bulletin, vol. 138. Food and Agriculture Organization of
the United Nations, Rome.
Hakansson, B., Jagerstad, M., Oste, R., Akesson, B., Jonsson, L., 1987. The effects of various thermal processes on protein quality, vitamins and selenium content in whole-grain
wheat and white flour. J. Cereal. Sci. 6, 269–282.
Hakulin, S., Linko, Y.Y., Linko, P., Seiler, K., Seibel, W., 1983. Enzymatic conversion of starch in twin screw HTST extruder. Starch 35, 411–414.
Harper, J.M., 1981. Extrusion of Foods, vol. 1. CRC Press Inc, Boca Raton, FL.
Harper, J.M., 1988. Effects of extrusion processing on nutrients. In: Karmas, E., Harris, R.S. (Eds.), Nutritional Evaluation of Food Processing, third ed. Van Nostrand Reinhold
Company, New York, NY, pp. 360–365.
Hashimoto, J.M., Grossmann, M.V.E., 2003. Effects of extrusion conditions on quality of cassava bran/cassava starch extrudates. Int. J. Food Sci. Technol. 38, 511–517.
Hsieh, S., Lin, F., Huff, H.E., 1998. Effects of lipids and processing conditions on lipid oxidation of extruded dry pet food during storage. Anim. Feed Sci. Technol. 71, 283–294.
Hulse, J.H., 2012. Nature, composition and utilization of food legumes. In: Muelbauer, F.J., Kaiser, W.J. (Eds.), Expanding the Production and Use of Cool Season Food Legumes: A
Global Perspective of Persistent Constraints and of Opportunities and Strategies for Further Increasing the Productivity and Use of Pea, Lentil, Faba Bean, Chickpea and Grass Pea
in Different Farming Systems. Springer Science and Business Media, Dordrecht, pp. 77–97.
Hwang, C.F., Riha, W.E., Jin, B., Karwe, M.V., Hartman, T.G., Daun, H., Ho, C.H., 1997. Effect of cysteine addition on the volatile compounds released at the die during twin-screw
extrusion of wheat flour. LWT - Food Sci. Technol. 30 (4), 411–416.
Ilo, S., Schoenlechner, R., Berghofe, E., 2000. Role of lipids in the extrusion cooking processes. Grasas Aceites 51 (1), 97–110.
Jain, D., Devi, M., Thakur, N., 2013. Study on the effect of machine operative parameters on physical characteristics of rice/maize based fruit/vegetable pulp fortified extrudates.
Agric. Eng. Int. CIGR J. 15, 231–242.
12 Extrusion for the Production of Functional Foods and Ingredients

Kamarudin, M.S., De Cruz, C.R., Saad, C.R., Romano, N., Ramezani-Fard, E., 2018. Effects of extruder die head temperature and pre-gelatinized taro and broken rice flour level on
physical properties of floating fish pellets. Anim. Feed Sci. Technol. 236, 122–130.
Karkle, E.L., Alavi, S., Dogan, H., Jain, S., Waghray, K., 2009. Development and Evaluation of Fruit and Vegetable-Based Extruded Snacks. AACC International Cereal Science
Knowledge Database.
Karwe, M.V., 2009. Food extrusion. In: Barbosa-Canovas, G.V. (Ed.), Food Engineering, Encyclopedia of Life Support Systems, vol. III. EOLSS Publishers., Oxford, pp. 227–236.
Kasapidou, E., Sossidou, E., Mitlianga, P., 2015. Fruit and vegetable coproducts as functional feed ingredients in farm animal nutrition for improved product quality. Agriculture 5,
1020–1034.
Kearns, J.P., Rokey, G.J., Huber, G.R., 2013. Extrusion of Texturized Proteins. American Soybean Association, Missouri, United States.
Killeit, U., 1994. Vitamin retention in extrusion cooking. Food Chem. 49, 149–155.
Korkerd, S., Wanlapa, S., Puttanlek, C., Uttapap, D., Rungsardthong, V., 2016. Expansion and functional properties of extruded snacks enriched with nutrition sources from food
processing by-products. J. Food Sci. Technol. 53 (1), 561–570.
Kothakota, A., Jindal, N., Thimmaiah, B., 2013. A study on evaluation and characterization of extruded product by using various by-products. Afr. J. Food Sci. 7, 485–497.
Kretschmer, P., 2004. Acrylamide in Cereal-Based Foodstuffs. URL: http://www.ilu-ev.de/acrylamide_eng/Presse/040422_5.htm. (Accessed 5 December 2019).
Kumar, K., Sarkar, B.C., Sharma, H.K., 2010. Development and characterization of extruded product of carrot pomace, rice flour and pulse powder. Afr. J. Food Sci. 4, 703–717.
Launay, B., Lisch, J.M., 1983. Twin-screw extrusion cooking of starches: flow behaviour of starch pastes, expansion and mechanical properties of extrudates. J. Food Eng. 2,
259–280.
Leonel, M., Freitas, T. S. de, Mischan, M.M., 2009. Physical characteristics of extruded cassava starch. Sci. Agric. 66 (4), 486–493.
Lohani, U.C., Muthukumarappan, K., 2016. Effect of extrusion processing parameters on antioxidant, textural and functional properties of hydrodynamic cavitated corn flour,
sorghum flour and apple pomace-based extrudates. J. Food Process. Eng. 40 (3), 1–15.
Lorenz, K., Jansen, G.R., 1980. Nutrient stability of full-fat soy flour and corn-soy blends produced by low cost extrusion. Cereal Foods World 25 (4), 161–162, 171–172.
Maga, J.A., Sizer, C.E., 1978. Ascorbic acid and thiamine retention during extrusion of potato flakes. LWT Lebensmitt Wissensch. Technol. 11, 192–194.
Mahungu, S.M., Diaz-Mercado, S., Li, J., Schwenk, M., Singletary, K., Faller, J., 1999. Stability of isoflavones during extrusion processing of corn/soy mixture. J. Agric. Food Chem.
47, 279–284.
Manthey, F.A., Sinha, S., Wolf-Hall, C.E., Hall, C.A., 2008. Effect of flaxseed flour and packaging on shelf life of refrigerated pasta. J. Food Process. Preserv. 32 (1), 75–87.
Marsman, G.J.P., Gruppen, H., Van Zuilichem, D.J., Resink, J.W., Voragen, A.G.J., 1995. The influence of screw configuration on the in vitro digestibility and protein solubility of
soybean- and rapeseed meals. J. Food Eng. 26, 13–28.
Martinez, B.F., Figueroa, J.D.C., Larios, S.A., 1996. High lysine extruded products of quality protein maize. J. Sci. Food Agric. 71 (2), 151–155.
Mathew, J.M., Hoseney, R.C., Faubion, J.M., 1999. Effect of corn moisture on the properties of pet food extrudates. Cereal Chem. 76 (6), 953–956.
Miller, R.C., 1985. Low moisture extrusion: effects of cooking moisture on product characteristics. J. Food Sci. 50, 249–253.
Monteiro, M.L.G., Marsico, E.T., Soares Junior, M.S., Magalhaes, A.O., Canto, A.C.C.S., Costa-Lima, B.R.C., Alvares, T.S., Conte Junior, C.A., 2016. Nutritional profile and
chemical stability of pasta fortified with tilapia (Oreochromis Niloticus) flour. PLoS One 11, 1–17.
Mulye, V.B., Zofair, S.M., 2015. Utilization of croaker (Johniusdussumieri) to develop ready to eat puff snack product using extrusion technology. Food Sci. Res. J. 6, 184–191.
Muralidharan, S., 1999. Studies on the Utilization of Deboned Trout (Oncorhynchus mykiss) Frames in Fish Snack. MS thesis. Digital Commons, Utah State University, Utah, United
States.
Muthukumarappan, K., Karunanithy, C., 2012. Extrusion cooking process. In: Ahemed, J., Rahman, M.S. (Eds.), Handbook of Food Process Design, first ed., vol. 1. Blackwell
Publishing Ltd., pp. 710–742
Nath, A., Chattopadhyay, P.K., 2008. Effect of process parameters and soy flour concentration on quality attributes and microstructural changes in ready-to-eat potato–soy snack
using high-temperature short time air puffing. LWT - Food Sci. Technol. 41, 707–715.
Neulicht, R., Shular, J., 1995. Cereal Breakfast Food. Emission Factor Documentation, Kansas City, Missouri.
Noguchi, A., Mosso, K., Aymanrd, C., Jevnink, J., Cheftel, J.C., 1982. Millard reactions during extrusion cooking of proteinenriched biscuits. Lebensm. Wiss. Technol. 15,
105–110.
Obatolu, A.V., Olusola, O.O., Adebowale, E.A., 2006. Qualities of extruded puffed snacks from maize/soybean mixture. J. Food Process. Eng. 29, 149–161.
Offiah, V., Kontogiorgos, V., Falade, K.O., 2019. Extrusion processing of raw food materials and by-products: a review. Crit. Rev. Food Sci. Nutr. 59 (18), 2979–2998.
Oke, M.O., Awonorin, S.O., Sanni, L.O., Asiedu, R., Aiyedun, P.O., 2013. Effect of extrusion variables on extrudates properties of water yam flour-a response Surface analysis.
J. Food Process. Preserv. 37 (5), 456–473.
Okechukwu, P.E., Rao, M.A., 1996. Kinetics of cowpea starch gelatinization based on granule swelling. Starch 48, 43–47.
Olaeta, J.A., Schwartz, M., Undurraga, P., Contreras, S., 2007. Use of Hass avocado (Perseaamericana mill.) seed as a processed product. In: Proceedings VI World Avocado
Congress, Chile, pp. 1–8.
Onwulata, C.I., Smith, P.W., Konstance, V.H., 2001. Incorporation of whey products in extruded corn, potato of rice snacks. Food Res. Int. 34, 679–687.
Onwulata, C.I., Phillips, J.G., Tunick, M.H., Qi, P.X., Cooke, P.H., 2010. Texturized dairy proteins. J. Food Sci. 75 (2), E100–E109.
Pamies, B.V., Roudaut, G., Dacremont, C., Meste, M.L., Mitchell, J.R., 2000. Understanding the texture of low moisture cereal products: mechanical and sensory measurements of
crispness. J. Sci. Food Agric. 80, 1679–1685.
Pankyamma, V., Basu, S., Bhadran, S.S., Chouksey, M.K., Gudipati, V., 2014. Fish oil-fortified extruded snacks: evaluation of physical properties and oxidative stability by response
surface methodology. J. Food Process. Eng. 37, 349–361.
Peluola-Adeyemi, O.A., Idowu, M.A., 2014. Effect of extrusion parameters on the physical and functional properties of cocoyam (Colocasia Esculenta) flour. IOSR J. Environ. Sci.
Toxicol. Food Technol. 8, 29–34.
Park, S.H., Lamsal, B.P., Balasubramaniam, V.M., 2014. Principles of food processing. In: Clark, S., Jung, S., Lamsal, B. (Eds.), Food Processing: Principles and Applications. John
Wiley and Sons, West Sussex, pp. 1–15.
Patil, S.S., Brennan, M.A., Mason, S.L., Brennan, C.S., 2016a. The effects of fortification of legumes and extrusion on the protein digestibility of wheat based snack. Foods 5 (2), 26.
Patil, S.S., Rudra, Varghese, S.G.E., Kaur, C., 2016b. Effect of extruded finger millet (Eleusine coracan L.) on textural properties and sensory acceptability of composite bread. Food
Biosci. 14, 62–69.
Pe˛ ksa, A., Kita, A., Carbonell-Barrachina, A.A., Miedziana, J., Kolniak-Ostek, J., Tajner-Czopek, A., Droz_ dz_ , W., 2016. Sensory attributes and physicochemical features of corn
snacks as affected by different flour types and extrusion conditions. LWT-Food Sci. Technol. 72, 26–36.
Perez, A.A., Drago, S.R., Carrara, C.R., De Greef, D.M., Torres, R.L., Gonzalez, R.J., 2008. Extrusion cooking of a maize/soybean mixture: factors affecting expanded product
characteristics and flour dispersion viscosity. J. Food Eng. 87, 333–340.
Perilla, N.S., Cruz, M.P., De Belalcazar, F., Diaz, G.D., 1997. Effect of temperature of wet extrusion on the nutritional value of full-fat soyabeans for broiler chickens. British Poultry
Science 38 (4), 412–416.
Philipp, C., Buckow, R., Silcock, P., Oey, I., 2017. Instrumental and sensory properties of pea protein-fortified extruded rice snacks. Food Res. Int. 102, 658–665.
Plavnik, I., Wan, D., 1995. Nutritional effects of expansion and short time extrusion on feeds for broilers. Anim. Feed Sci. Technol. 55, 247–251.
Qu, D., Wang, S.S., 1994. Kinetics of the formation of gelatinized and melted starch at extrusion cooking conditions. Starch 46, 225–229.
Ralet, M.C., Thibault, J.F., Della Valle, G., 1990. Influence of extrusion-cooking on the physico-chemical properties of wheat bran. J. Cereal. Sci. 11 (3), 249–259.
Ramachandra, H.G., Thejaswini, M.L., 2015. Extrusion technology: a novel method of food processing. Int. J. Innovat. Res. Sci. Eng. Technol. 2, 358–369.
Riaz, M.N., 2000. Introduction to extruders and their principles. In: Riaz, M.N. (Ed.), Extruders in Food Applications. CRC Press, Boca Raton, pp. 1–24.
Extrusion for the Production of Functional Foods and Ingredients 13

Riaz, M.N., 2006. Extruded snacks. In: Hui, Y.H., Castell-Perez, E., Cunha, L.M., Guerrero-Legarreta, I., Liang, H.H., Lo, Y.M., Marshall, D.L., Nip, W.K., Shahidi, F., Sherkat, F.,
Winger, R.J., Yam, K.L. (Eds.), Handbook of Food Science, Technology and Engineering. CRC Press, Boca Raton, p. 168.
Riaz, M.N., 2010. Extrusion of Cereals. Kent, UK: New Food Magazine, 4.
Riaz, M., Asif, M., Ali, R., 2009. Stability of vitamins during extrusion. Crit. Rev. Food Sci. Nutr. 49, 361–368.
Roberts, S.A., Guy, R.C.E., 1987. Metastable states in a food extrusion cooker. J. Food Eng. 6, 103–112.
Rodis, P., Wen, L.F., Wasserman, B.P., 1993. Assessment of extrusion induced starch fragmentation by gel permeation chromatography and methylation analysis. Cereal Chem. 70
(2), 152–157.
Rossen, J.L., Miller, R.C., 1973. Food extrusion. Food Technol. 27 (8), 46–53.
Rouilly, A., Jorda, J., Rigal, L., 2006. Thermo-mechanical processing of sugar beet pulp. I. Twin-screw extrusion process. Carbohydr. Polym. 66, 81–87.
Ruiz-Armenta, X.A., Zazueta-Morales, J.J., Aguilar-Palazuelos, E., Delgado-Nieblas, C.I., Lopez-Diaz, A., Camacho-Hernandez, I.L., Gutierrez-Dorado, R., Martınez-Bustos, F., 2018.
Effect of extrusion on the carotenoid content, physical and sensory properties of snacks added with bagasse of naranjita fruit: optimization process. CyTA - J. Food 16, 172–180.
Rweyemamu, L.M.P., Yusuph, A., Mrema, G.D., 2015. Physical properties of extruded snacks enriched with soybean and moringa leaf powder. Afr. J. Food Sci. Technol. 6, 28–34.
Sacchetti, G., Pinnavaia, G.G., Guidolin, E., Dalla Rosa, M., 2004. Effects of extrusion temperature and feed composition on the functional, physical and sensory properties of
chestnut and rice flour-based snack-like products. Food Res. Int. 37 (5), 527–534.
Santana, A.L., Angela, M., Meireles, A., 2014. New starches are the trend for industry applications: a review. Food Publ. Health 4, 229–241.
Sarkar, P., Setia, N., Choudhury, G.S., 2011. Extrusion processing of cactus pear. Adv. J. Food Sci. Technol. 3, 102–110.
Savita, S., Arshwinder, K., Gurkirat, K., Vikas, N., 2013. Influence of different protein sources on cooking and sensory quality of pasta. Int. J. Eng. Res. Afr. 3 (2), 1757–1763.
Sayanjali, S., Sanguansri, L., Ying, D., Buckow, R., Gras, S., Augustin, M.A., 2019. Extrusion of a curcuminoid-enriched oat fiber-corn-based snack product. J. Food Sci. 84 (2),
284–291.
Serbio, L., Chang, Y.K., 2000. Effects of selected process parameters in extrusion of yam flour (Dioscorea rotundata) on physicochemical properties of the extrudates. Mol. Nutr.
Food Res. 44, 96–101.
Seth, D., Rajamanickam, G., 2012. Development of extruded snacks using soy, sorghum, millet and rice blend-A response surface methodology approach. Int. J. Food Sci. Technol.
47 (7), 1526–1531.
Shilev, S., Naydenov, M., Vancheva, V., Aladjadjiyan, A., 2006. Composting of food and agricultural wastes. In: Oreopoutou, V., Russ, W. (Eds.), Utilization of Byproducts and
Treatment of Wastes in the Food Industry. Springer Science and Business Media, New York, pp. 283–302.
Singh, S., Gamlath, S., Wakeling, L., 2007. Nutritional aspects of food extrusion: a review. Int. J. Food Sci. Technol. 42 (8), 916–929.
Singh, R.K., Majumdar, R.K., Venkateshwarlu, G., 2014. Optimum extrusion-cooking conditions for improving physical properties of fish-cereal based snacks by response surface
methodology. J. Food Sci. Technol. 51, 1827–1836.
Smithey, S.L., Badding, Huff H.E., Hsieh, E., 1995. Processing parameters and product properties of extruded beef with nonmeat cereal binders. LWT Food Sci. Technol. 28,
386–394.
Soetan, K.O., Oyewole, O.E., 2009. The need for adequate processing to reduce the anti- nutritional factors in plants used as human foods and animal feeds: a review. Afr. J. Food
Sci. 3 (9), 223–232.
Souza, M.L.D., Menezes, H.C.D., 2008. Extrusion of Brazil nut and cassava flour mixtures. Ciência Tecnol. Aliment. 28 (2), 451–462.
Steel, C., Oro, M., Schmeler, M., Ferrera, R., Chang, Y., 2012. Thermoplastic extrusion in food processing. In: Adel, E.-S. (Ed.), Thermoplastic Elastomers. In Tech, Croatia.
Stojceska, V., Ainsworth, P., Plunkett, A., Ibanoglu, S., 2008. The recycling of brewer’s processing by-product into ready-to-eat snacks using extrusion technology. J. Cereal. Sci. 47
(3), 469–479.
Studer, A., Blank, I., Stadler, R.H., 2004. Thermal processing contaminants in foodstuffs and potential strategies of control. Czech J. Food Sci. 22, 1–10.
Surasani, V.K.R., 2016. Application of food extrusion process to develop fish meat-based extruded products. Food Eng. Rev. 8, 448–456.
Tayeb, J., Valle, G.D., Barres, C., Vergnes, B., 1992. Simulation of transport phenomena in twin-screw extruders. In: Kokini, J.L., Ho, C.T., Karwe, M.V. (Eds.), Food Extrusion
Science and Technology. Marcel Dekker, Inc, New York, p. 41.
Tomaszewska-Ciosk, E., Golachowski, A., Drozdz, W., Boruczkowski, T., Boruczkowska, H., Zdybel, E., 2012. Selected properties of single- and double-extruded potato starch. Pol.
J. Food Nutr. Sci. 62 (3), 171–177.
Turhan, M., Gunasekaran, S., 2002. Kinetics of in situ and in vitro gelatinization of hard and soft wheat starches during cooking in water. J. Food Eng. 52, 1–7.
Van Zuilichem, D.J., Van Roekel, G.J., Stolp, W., Vant Riet, K., 1990. Modeling of the enzymatic conversion of cracked corn by twin-screw extrusion cooking. J. Food Eng. 12 (1),
13–28.
Varzakas, T., 2016. Quality and safety aspects of cereals (wheat) and their products. Critical Reviews in Food Science and Nutrition 56 (15), 2495–2510.
Vergnes, B., Villemaire, J.P., 1987. Rheological behaviour of low moisture molten maize starch. Rheol. Acta 26 (6), 570–576.
Verma, A.K., Pathak, V., Singh, V.P., 2014. Quality characteristics of value added chicken meat noodles. J. Nutr. Food Sci. 4, 255.
Wang, N., Bhirud, P., Sosulski, F., Tyler, R., 1999. Pasta-like product from pea flour by twin-screw extrusion. J. Food Sci. 64, 671–678.
Whalen, P.J., Des Rochers, J.L., Walker, C.E., 2000. Ready-to-eat breakfast cereals. In: Kulp, K. (Ed.), Handbook of Cereal Science and Technology. Marcel Dekker Inc., New York,
pp. 615–705.
White, G., 1994. Defining the true meaning of snacks. Food Technol. Int. Eur. 2, 115–117.
White, B.L., Howard, L.R., Prior, R.L., 2010. Polyphenolic composition and antioxidant capacity of extruded cranberry pomace. J. Agric. Food Chem. 58 (7), 4037–4042.
Wojtowicz, A., Moscicki, L., 2011. Effect of wheat bran addition and screw speed on microstructure and textural characteristics of common wheat precooked pasta-like products.
Pol. J. Food Nutr. Sci. 61 (2), 101–107.
Yacu, W.A., 1985. Modelling a twin-screw co-rotating extruder. J. Food Eng. 8, 1–21.
Yagcı, S., Gögüş, F., 2008. Response surface methodology for evaluation of physical and functional properties of extruded snack foods developed from food-by-products. J. Food
Eng. 86 (1), 122–132.
Zhao, X., Yiminwei, Wang, Z., Chen, F., Ojokoh, A.O., 2011. Reaction kinetics in food extrusion: methods and results. Crit. Rev. Food Sci. Nutr. 51, 835–854.
Zheng, X., Wang, S.S., 1994. Shear induced starch conversion during extrusion. J. Food Sci. Techno. 59 (5), 1137–1143.
Zhuang, H., Feng, T., Xie, Z., Toure, A., Xu, X., Jin, Z., Su, Q., 2010. Effect of mesona blumes gum on physicochemical and sensory characteristics of rice extrudates. Int. J. Food
Sci. Technol. 45 (11), 2415–2424.

You might also like