You are on page 1of 18

3.

04 Advanced Drying Technologies of Relevance in the Food Industry


Henry Sabarez, CSIRO Agriculture & Food, Werribee, VIC, Australia
© 2021 Elsevier Inc. All rights reserved.

3.04.1 Introduction 64
3.04.2 The Food Drying Process 65
3.04.2.1 Transport Phenomena 65
3.04.2.2 Heat Transfer Mechanism 66
3.04.2.3 Mass Transfer Mechanism 66
3.04.3 Advanced Food Drying Technologies 66
3.04.3.1 Heat Pump Drying 67
3.04.3.1.1 Operating Principle and Conditions of Heat Pump Drying 67
3.04.3.1.2 Advantages and Disadvantages 68
3.04.3.2 Foam Mat Drying 68
3.04.3.2.1 Foam Mat Drying 68
3.04.3.2.2 Foaming 69
3.04.3.2.3 Drying of Foams 69
3.04.3.2.4 Microwave Foam Drying 70
3.04.3.3 Refractance Window Drying 70
3.04.3.3.1 Operating Principle of RW Drying Technology 71
3.04.3.3.2 Components of a Commercial RW Drying System 71
3.04.3.3.3 Case Studies 71
3.04.3.4 Ultrasound Drying Technology 73
3.04.3.4.1 Application of Ultrasound in Food Drying 73
3.04.3.4.2 Low Temperature Drying (Above Freezing) 74
3.04.3.4.3 Low Temperature Drying (Below Freezing) 75
3.04.3.4.4 Drying Performance of Various Ultrasonic Designs 75
3.04.3.5 Electrohydrodynamic Drying 76
3.04.3.5.1 Components of EHD Drying System 76
3.04.3.5.2 Principle of Operation in EHD Drying 76
3.04.3.5.3 Advantages and Disadvantages 77
3.04.3.5.4 Challenges and Opportunities for Industrial Implementation 77
3.04.4 Concluding Remarks 78
References 78

3.04.1 Introduction

The food and agribusiness sector (particularly in Australia) has traditionally built much of their businesses on trading raw commod-
ities (e.g., fresh fruits and vegetables, meat, grains, dairy, etc.). The biggest opportunity for future growth in this sector lies in value-
adding to these basic raw commodities, which has the potential to provide returns across the value chain. Adding value to these
short shelf-life and seasonal raw commodities through transformation into shelf-stable food products using drying processes is
a significant global industry with the possibility to provide an effective means of ensuring all year-round supply of various food
nutrients in a convenient way. Numerous food products are routinely preserved by drying, which include grains, marine products,
meat products, dairy products, as well as horticultural products (Jangam, 2011). Many dried foods are ready-to-eat while most man-
ufactured food products virtually contain ingredients that undergo some form of drying.
The conventional drying of food materials, however, is energy-intensive, time consuming and detrimental to product quality,
primarily due to the exposure of food materials to high temperatures or long drying times (Sabarez, 2015, 2016b). This is driving
interests for the continued research in the development of innovative new drying technologies that intensifies gentle processing (i.e.,
low temperatures) with improved energy efficiency while protecting products from quality degradation. In addition, there are
emerging challenges (e.g., food security, dwindling resources, changing climate, etc.) across the globe that are driving further
the food industry in the pursuit of technological innovations for value-adding to these commodities to remain sustainably compet-
itive and profitable. In particular, the increasing consumer demand for premium and healthy food products will continue to drive
the need for gentle (i.e., eco-friendly, efficient and sustainable) processing to maintain competitiveness with minimal impact on the
environment.
Many drying techniques have evolved due to the need to produce premium quality dried food products and ingredients that are
particularly heat-sensitive (Sabarez et al., 2018). Such drying systems include the utilization of low temperatures, but often require

64 Innovative Food Processing Technologies, Volume 3 https://doi.org/10.1016/B978-0-08-100596-5.23042-4


Advanced Drying Technologies of Relevance in the Food Industry 65

very long drying times, are highly energy consuming and detrimental to the product quality and susceptible to microbial contam-
ination. From a quality point of view, the process of drying at below freezing temperatures under vacuum operating pressure (i.e.,
vacuum freeze drying) is considered as one of the best methods of drying heat-sensitive food materials. However, vacuum freeze
drying (also known as lyophilization) is an expensive process when compared to other conventional drying techniques as it suffers
from high production costs, high energy consumptions, long drying times, and low throughputs (batch-wize process) (Ratti, 2001).
The process of drying at low temperatures under atmospheric pressure (without vacuum) has the advantage of low energy
consumption, and high production throughput (i.e., can be operated in a continuous process) with product quality comparable
to the vacuum freeze drying. However, working at low temperatures under atmospheric pressure usually leads to very long drying
times. In recent years, a number of nonthermal food processing technologies have been investigated and developed with the aim to
improve or replace the conventional food processing technologies. The limitations inherent in conventional drying processes may
be overcome by the combined application of these novel or emerging technologies (Sabarez et al., 2012; Sabarez, 2016a). These
technologies take advantage of other physical phenomena (e.g., sound waves, pressures, pulsed electric field and electromagnetic,
etc.) to intensify the drying process at low temperatures.
This chapter provides an evaluative review of advanced drying technologies relevant to the food industry specifically for drying of
food materials at low temperatures under atmospheric pressure. It starts with a brief overview of the drying process of food materials
and then focusses on the details of various advanced drying technologies particularly applied for low temperature drying of food
materials.

3.04.2 The Food Drying Process

The process of drying of food materials is complex, involving coupled transient mechanisms of heat, mass, and momentum transfer
processes accompanied by physical, chemical, and phase transformations (Sabarez, 2012, 2015). There are two distinct transport
mechanisms that simultaneously occur during drying, (i) heat transfer from the drying medium to the food material, which induces
phase changes of water from solid or liquid into a vapor phase (except in the case of osmotic dehydration where water is removed
without a change in phase), and (ii) water transport (either in liquid or vapor phase) from the interior of the food product to its
surface in which the moisture vapor is eventually transported away from the product by a carrier gas (or by the application of
vacuum for nonconvective drying systems). The heat and mass transfer phenomena (both occurring externally and internally to
the food matrix) are usually influenced by both temperature and water concentration differences, as well as the air velocity field,
together with the properties of the food material itself. The internal heat and mass transfer processes may be also affected by the
physical changes (e.g., shrinkage) that may occur in the product during drying. A conceptual representation of the transport
phenomena occurring during drying of a food material is illustrated in Fig. 1.

3.04.2.1 Transport Phenomena


In most cases, air is employed as the carrier gas to take away the moisture from the product during drying. A layer of air usually forms
at the air-food interface (i.e., a layer of air surrounding the food material) during the drying process in which the flow of air is

Heating Plate

Radiation

Temp Velocity vector


Hot Air RH
Velocity

Free stream airflow

Convection
Boundary layer
Boundary
Layer External Transfer
thickness Evaporation

Conduction Diffusion

Food Internal Transfer

Tray/Heating Plate

Figure 1 A conceptual representation of the transport phenomena occurring during drying of a food material. Adapted from Sabarez (2017).
66 Advanced Drying Technologies of Relevance in the Food Industry

inevitably slowed down (close to the food surface) due to friction (Sabarez, 2016a). This layer of air acts as a barrier to the external
transfer of both heat and water vapor during drying. Heat is mainly conducted through this (near) stagnant air primarily by molec-
ular collisions. Water vapor diffuses through this boundary layer caused by water vapor pressure gradients and is eventually carried
away by the moving stream of air. The thickness of this boundary layer is determined primarily by air velocity. In particular, the
external transfer rates for both heat and water vapor are greatly influenced by the air velocity field (fluid dynamics) and the air prop-
erties (i.e., temperature and relative humidity). If the air velocity is too low, the boundary layer becomes thicker and the resistance to
heat and water vapor transfer increases. In addition, the water vapor leaving the surface of the food material increases the humidity
of the air in the boundary layer due to the slow molecular motion of the moving stream of air. This causes a reduction in the water
vapor pressure gradient and slows down the externally controlled drying process.

3.04.2.2 Heat Transfer Mechanism


Energy is primarily required during drying to generate a phase change of water in food materials from liquid to vapor (vaporization)
or solid to vapor (sublimation) and to activate molecular movement. There are three modes of heat transfer (i.e., convection,
conduction, and radiation) that may be used alone or in combination to supply heat from the heat source to the food materials.
These different heat transfer mechanisms may be deployed simultaneously or sequentially depending on the individual application
to achieve an energy efficient process (Sabarez, 2017). All modes of heat transfer, except those using electromagnetic energy (micro-
wave and radio frequency), supply heat at the boundaries of the food material so that the heat must subsequently transfer within the
food material primarily by conduction. In convective heat transfer, air is used to supply heat to change the phase of water in the food
materials in addition to its role of carrying away the evaporated moisture from the product. The energy consumption in convective-
based drying is usually high due to the inherent inefficiency of using air as the drying medium. While in conductive heat transfer,
heat is supplied to the food materials through contact with heated metallic/nonmetallic solids (through molecular vibration) or
stationery fluids (primarily by molecular collision). The radiant energy, on the other hand, can be supplied in various forms of elec-
tromagnetic waves, categorized according to the region of the electromagnetic spectrum (e.g., radio frequency (RF), infrared (IR),
microwave (MW), etc.). In food drying applications, the main types of radiation applied are IR, MW, and RF, all of which employ
very different heating mechanisms as described elsewhere (Sabarez, 2015).

3.04.2.3 Mass Transfer Mechanism


The speed of moisture removal during drying may be governed by either the rate of moisture diffusion within the food matrix
(internal mass transfer) or by the rate of moisture evaporation from the food surface to the drying medium (external mass transfer).
The external moisture transfer from the food surface to the drying medium is mainly influenced by the properties of the drying air.
Once the rate of replenishment of moisture from the interior to the surface of the product is slower than the external moisture trans-
fer rate, then the internally controlled drying process, which is mainly influenced by temperature and the properties of food mate-
rials, becomes the dominant mass transfer mechanism. According to Heldman and Hartel (1997), the moisture within the food
product can migrate in several ways via a number of different mechanisms (liquid or vapor phase). The liquid transport mecha-
nisms include capillary flow, surface diffusion, and liquid diffusion, while the vapor transport mechanisms consist of Knudsen
diffusion, mutual diffusion, Stefan diffusion, Poiseuille flow, and condensation-evaporation (Sablani and Rahman, 2007). Vapor-
ization in some cases may occur within the product and thus water diffuses in the form of vapor through the food matrix, with the
difference in vapor pressure as the driving force for moisture transfer (Sabarez, 2016a). The differences in pressure between the
drying medium and the internal food structure (pressure flow) and the differences in temperature between the surface and the inte-
rior of the product (thermal flow) may also influence the internal mobility of moisture. The moisture (in liquid or vapor phase)
must travel to the boundary of the material before it is then transported away (in vapor phase) by the carrier gas (or by application
of vacuum for nonconvective dryers).

3.04.3 Advanced Food Drying Technologies

There are many different methods of drying food materials, each with their own advantages and disadvantages for particular appli-
cations. A vast number of drying methods have evolved due to the differences in the physical attributes of the product, modes of
heat input, operating temperatures and pressures, quality specifications on the dried product, and so on (Sabarez, 2015; Sabarez,
2018). The majority of dryers used in the food industry are of convective type, i.e., hot air is used both to supply heat for evaporation
of water and to carry away the evaporated moisture from the product. According to Mujumdar and Devahastin (2008), over 85% of
industrial dryers are of the convective type. This is by far the most common drying method used at industrial scale because it is
simple and easy to operate, in addition to relatively low capital costs (for the time being), although it is poor in energy efficiency.
The details of the various drying techniques for food materials and their classification can be found elsewhere (Sabarez, 2015;
Jangam, 2011; Bansal and Chung, 2007).
Several advanced drying technologies for low temperature drying of food materials have been investigated in recent years. These
studies have explored and developed innovative technologies which take advantage of the combined application of other physical
phenomena (e.g., ultrasound, electromagnetic field, pulsed electric field, etc.) with the convective drying method. In particular,
Advanced Drying Technologies of Relevance in the Food Industry 67

these nonthermal novel technologies were applied for the development of advanced drying concepts in improving the quality of
food products through efficient and gentle processing (i.e., low temperature drying). This section describes examples of advanced
drying technologies (i.e., heat pump drying, foam mat drying, refractance window drying, ultrasound drying and electrohydrody-
namic drying) applied for low temperature drying of food materials.

3.04.3.1 Heat Pump Drying


Heat pump drying (HPD) method is well recognized as one of the energy-efficient ways of drying heat-sensitive materials such as
bioactive food and pharmaceutical products at comparatively low temperatures. Heat pump dryers have been commercially used
particularly in Norway to dry uniformly high-quality dried cod (balacoa or clipfish) (Perera, 2016). Since the optimum temperature
for drying of agricultural produce by which a good product quality can be obtained is 60  C (Patel and Kar, 2011), HPD system is
well suited for such type of low temperature drying application. In HPD method, a shorter drying time at lower temperatures can be
achieved compared to the conventional convective air drying due to comparatively low levels of relative humidity achievable in the
system. Intensification of the drying process at low temperatures will increase throughput while minimizing microbial contamina-
tion, product quality degradation, and energy consumption. A basic HPD system consists of a drying chamber with trays or trolley of
trays to contain the products, a fan for air circulation, and a refrigeration unit (Sabarez, 2016b). The refrigeration unit is made up of
a condenser (hot heat exchanger), an expansion valve, a compressor, and an evaporator (cold heat exchanger), which functions
similar to a domestic refrigeration system. Fig. 2 shows a schematic diagram of a typical design of a basic HPD system.

3.04.3.1.1 Operating Principle and Conditions of Heat Pump Drying


The HPD system operates as the integration of a standard convective air dryer and a refrigeration unit. Warm dry air is drawn from
the condenser and blown across the drying chamber containing the wet material. The warm humid air exiting from the drying
chamber is blown over the evaporator (which acts as a heat sink to dehumidify the drying air), where it is rapidly cooled to a temper-
ature below its dew point, resulting in water condensing out, thereby removing the water from the drying air. This renders the air dry
and recovers the latent heat of evaporation through water vapor removal which subsequently permits air recirculation (Patel and
Kar, 2011). The latent heat given up by the moist air is taken up by the refrigerant, recirculated within the refrigeration circuit, and
the absorbed energy (latent heat) is released at the condenser. The cool dry air from the evaporator is then blown into the condenser
(which acts as a heat source) to elevate the temperature of the drying air. An auxiliary heating element (optional) can be added after
the condenser to achieve higher temperatures of the drying air (if needed). The reheated air is then returned to the drying chamber,
which completes the cycle.
It was reported that the HPD systems can operate over a wide range of drying conditions with the temperature ranging
from 16  C to 80  C, relative humidity (RH) ranging from 3% to 85%, air velocity ranging from 0.03 to 5.4 m/s, and drying
time ranging from 16 min to 80 h, depending on the dryer configuration, drying mode, and the dried material (Ong, 2016). In
a single-stage heat pump without auxiliary heating, the maximum achievable air temperature is approximately 57  C for heat
pump cycle charged with refrigerant R22, where the highest condensing temperature of the refrigerant is approximately 60  C
(Chou and Chua, 2007). Claussen et al. (2012) used HPD to study the drying characteristics of salted cod and found that an increase
in air velocity (from 0.8 to 3.0 m/s) and temperature (from 15  C to 22  C) and a decrease of relative humidity (from 60% to 30%)
increased the water removal rate in the first 12 h. These authors also showed that by lowering the relative humidity in the drying air
and utilization of the condenser, when the amount of wet salted fish is increased in the drying tunnel, the dryer capacity could be

motor blower

air distributor
evaporator

vent
condenser

heater

trolley

compressor

Figure 2 Schematic diagram of a heat pump drying system. Adapted from Sabarez (2007) and Sabarez (2016b).
68 Advanced Drying Technologies of Relevance in the Food Industry

increased by 50%. Shi et al. (2008) studied drying of horse mackerel using HPD and found that the specific moisture extraction rate
(SMER) was maximum when the bypass air ratio was 0.6–0.8, with optimum air velocity for drying of 2.0–3.0 m/s. The SMER for
a well-designed dehumidifier is in the range of 1–4 kg/kWh, with an average value of approximately 2.5 kg/kWh (Perera and
Rahman, 1997). It is useful to compare this figure with the latent heat of vaporization of water, which is 2255 kJ/kg at 100  C
or 1.596 kg/kWh.

3.04.3.1.2 Advantages and Disadvantages


One of the major advantages of HPD system over conventional hot air systems for drying of food materials is the higher energy
efficiency (Minea, 2013). The HPD system has evolved to enhance the energy efficiency of the drying process by recovering the latent
heat (Hawlader et al., 2001). It is usually a closed system, which was reported to consume less energy compared with other drying
methods. Perera and Rahman (1997) reported a drying efficiency of 95% for HPD compared to less than 70% for vacuum drying,
and between 35% and 40% for hot air drying. In this drying system, high energy efficiencies can be achieved because both the
sensible and the latent heat of vaporization are recovered. Because the water is condensed and removed in a liquid state rather
than in its vapor state, it allows the latent heat of vaporization to be captured, and only a small amount of sensible heat is lost
(Oliver, 1982). Also, the ability of HPD system to convert the latent heat of condensation into sensible heat makes the system
one of the unique heat-recovering devices for drying applications.
In addition to higher drying efficiency, Perera and Rahman (1997) stated that HPD also offers better product quality and it is
environmentally friendly. Another advantage of HPD system is its ability to operate independently outside ambient weather condi-
tions as water is removed from the product without outside ventilation, because the system is totally enclosed with air recirculated
fully. As drying using HPD system is done at low temperatures, low relative humidities and in a closed system, HPD is especially
suitable for drying of food materials that retain a high level of volatile aromatic constituents free of dust and microbial contami-
nation. However, HPD systems have higher initial capital costs and require a regular maintenance of the refrigeration system (Patel
and Kar, 2011). Also, the simplest designs of HPD system are only efficient in drawing out free moisture and loosely bound mois-
ture from the product. Efficiency drops dramatically once free moisture and loosely bound moistures are removed (i.e., at low mois-
ture contents). In most cases, the efficiency of HPD used for drying decreases with increase in temperature above 45  C
(Perera, 2016). However, these disadvantages could be overcome with proper design modifications, optimization of process condi-
tions and adequate control strategies.

3.04.3.2 Foam Mat Drying


The production of powders from liquid or semi-liquid food products is commonly carried out using various drying methods such as
spray drying, drum drying and freeze drying. However, these drying methods are either expensive (freeze drying) and/or not suitable
especially for heat-sensitive, viscous and sticky materials (spray drying and drum drying). Alternatively, convective air drying in
cabinet/tray dryers can be employed to dry a layer of highly viscous and sticky liquid or semi-liquid food products (e.g., fruit or
vegetable juices or purees) that cannot be dried using the other drying methods because of the state of the product. The conventional
air drying is inexpensive, simple and commonly used method in the food industry. For heat-sensitive food products, the conven-
tional air drying process can be carried out at low temperatures, however, this significantly slows down the drying process. Also, the
air drying process of a layer of highly viscous and sticky liquid or semi-liquid materials is usually very slow due to its dense structure.
In addition, the product would increasingly become more denser as drying progresses resulting from shrinkage. A dense structure
prolongs convective air drying as moisture diffusion from inside to the outside of the product is increasingly limited and this usually
leads to poor grindability of the dried product to produce powders.

3.04.3.2.1 Foam Mat Drying


Foam mat drying has evolved to effective dry a layer of viscous and sticky liquid or semi-liquid food materials using the conven-
tional air drying process. In this method, the dense liquid or semi-liquid food material is transformed into a stable porous structure
by a foaming step prior to drying. The foaming process is mainly carried out to incorporate substantial volume of air or other inert
gas by whipping or beating the raw material under controlled conditions using various means such a blender or specially designed
devices in the presence of a foaming agent (which works as a foam inducer and stabilizer). The stable foam is then spread out in
a relatively thin sheet or mat and dried by convection, using hot air flowing over or through the thin layer of the foamed material
until it is dried to the desired moisture. Foam mat drying is usually operated as a convective process at atmospheric pressure (Ratti
and Kudra, 2006). It results in a thin porous honeycomb sheet or mat which can easily be ground to a powder and reconstituted
(Hardy and Jideani, 2017). The dried sheet product is usually converted into fine powder by simple and easy grinding resulting in
free flowing powder. The whole process of foam mat drying is presented in Fig. 3.
The history of foam mat drying is dated back to 1917 when Campbell Food Company (Campbell, 1917) patented a method for
the drying of foamed evaporated milk as cited by Ratti and Kudra (2006). This is followed by patents for the drying of foamed egg
white (Mink, 1939 and 1940), further developed by Morgan and co-workers (1961). Foam mat drying is finding an increasing
demand and application in drying liquids/semi-liquids that render a high-quality concentrate such as milk, fruit juices, coffee
and tea, in a commercial scale (Bag et al., 2011). Commercial-scale foam mat dryers of belt type have been successfully operated
at Kikko Foods Corporation in Japan for fruit and vegetable powder production (Uno and Kobayashi, 1985). The foam mat drying
process can also be used to dry juice, milk, fruits, beverages, and jams (Widyastuti and Srianta, 2011). Foam mat drying is suitable
Advanced Drying Technologies of Relevance in the Food Industry 69

Figure 3 Schematic representation of the foam mat drying process. Adapted from Qadri et al. (2019).

for drying of viscous and heat sensitive liquid/semi-liquid foods with low glass transition temperature and high sugar content,
which are usually difficult to dry.
The drying process of foamed materials can systematically be performed at lower temperatures and shorter drying times
compared to non-foamed material in the same type of dryer and conditions (Widyastuti and Srianta, 2011), as the degree of drying
in the foam mat drying process is reasonably high because of the massive increase in the liquid-gas interface. The relatively high
drying rates (and thus fast drying) result from the larger surface area of the material exposed to the drying air, ensuring a much
higher moisture diffusion through the air present in the gas bubbles (Brygidyr et al., 1977). Capillary diffusion has also been re-
ported as the other main reason for the moisture movement within the product during foam mat drying (Sankat and
Castaigne, 2004). For convective foam drying, Sangamithra et al. (2015a,b) and Sankat and Castaigne (2004) hypothesized that
the foam bubbles open during drying, and new bubbles are created constantly by the moisture movement. The difference between
foam mat drying and the other drying methods is that in foam mat drying, the physical structure of raw food material is changed by
breakdown of the cell walls and gas (air) incorporation, while in other drying methods, the structure does not undergo any change
before drying except the case of pretreatments.

3.04.3.2.2 Foaming
Food foams can be made by three different techniques such as (i) whipping or beating, (ii) sparging or bubbling, and (iii) shaking.
The whipping or beating method of foam formation was widely used in the food processing industries. Commercially available
electric hand blender was used for about 3–5 min to obtain consistent foam (Valenzuela and Aguilera, 2013). However, many foods
do not possess foaming property and thus may not form a stable foam even if subjected to any of the above foaming techniques
(Qadri et al., 2019). Such foods can be converted into stable foams with the help of some additives. The additives that help the
formation of stable foams are divided into foaming agents and foam stabilizers.
Foaming agents reduce the surface tension between the interfaces and help in foaming of the liquid. The commonly used foam-
ing agents reported in different studies include egg albumin both as fresh and dry (Thuwapanichayanan et al., 2012;
Muthukumaran et al., 2008a,b; Qadri and Srivastava, 2014), whey protein as concentrate (Thuwapanichayanan et al., 2012) and
isolate (Darniadi et al., 2018) and soya protein isolate (Zheng et al., 2013; Thuwapanichayanan et al., 2008). On the other
hand, foam stabilizers are additives that enable a foam to retain its structure for longer duration and delay coalescence. The stabi-
lizing agents commonly used by researchers in foam mat drying studies include maltodextrin, methylcellulose, and carboxyl methyl
cellulose (Qadri et al., 2019). Some of the other foam stabilizers are thickening or gelling agents such as pectin, gelatin, starches,
xanthan gum, and Arabic gum (acacia). The effect of stabilizers in decreasing the instability of foam can be explained by increasing
the viscosity of the continuous phase or by forming a three-dimensional network that impedes the movement of components
within the foam (Walsh et al., 2008).
Foam stability is of primary importance for the success of foam mat drying. To assure a fast drying process of foamed materials, it
is of utmost importance that foams are able to retain their porous structure without collapsing throughout the process (Sankat and
Castaigne, 2004). An essential challenge, particularly in protein-based foams is their instability, which is triggered by gravitational
drainage. It was demonstrated that maintenance of foam structural stability, that is, high product porosity, throughout the whole
drying process is the key to success (Ambros et al., 2019). Hence, diffusion limitation, which normally occurs toward the end
of drying processes causing overheating and prolongation of drying time, can be effectively inhibited. Aside from drying methods
and conditions, the drying rates of foamed materials depend greatly on foam characteristics, namely on its layer thickness, stability,
density, and bubble size. Higher drying rates were obtained for foams of lower density. It appears, however, that the foams of the
same density most suitable for drying are those with the smallest bubbles and the greatest homogeneity, as they can be dried to
lower moisture contents under milder conditions.

3.04.3.2.3 Drying of Foams


Batch type hot air drying in cabinet dryers is the most commonly used method in drying foamed materials that has been reported in
literature (Falade et al., 2003; Sankat and Castaigne, 2004; Thuwapanichayanan et al., 2008; Kandasamy et al., 2014). Besides hot
70 Advanced Drying Technologies of Relevance in the Food Industry

air drying, other techniques of drying have also been studied by many researchers who reported encouraging results (Qadri and
Srivastava, 2014). Drying of foamed materials can also be carried out either by freeze drying and/or spray drying. The foam mat
drying process is reported to be considerably cheaper than vacuum, freeze and spray drying methods for the production of food
powders (Kadam et al., 2010). It is the simplest form of drying compared to other methods such as freeze drying, spray drying,
as it is less expensive, less complicated, and is less time-consuming (Febrianto et al., 2012). The hybridization of foam mat drying
technology has been shown to make this technology look even more promising than the conventional one (Hardy and
Jideani, 2017).
One of the shortcomings of foam mat drying is the poor heat transfer within the material because of the presence of gas in the
foamed materials (as thermal conductivity of a gas is much lower than that of a liquid). The poor heat conductivity of gas entrapped
in foam decreases the heat transfer rate within the material in conventional foam mat drying (Ratti and Kudra, 2006). Also, the
foamed materials are dried in a thin layer which results in limited throughput, although shorter drying time offsets the low
throughput, which indicates that optimum conditions must be established for each material to be dried. One of the approaches
that permit higher drying loads is by volumetric heating using microwaves, which can penetrate deep layers of the drying materials.
Microwave is a potential new way of assisting dehydration of food materials because of its ability to generate heat within the food
material rapidly with no thermal lag. The application of microwave energy could also overcome the poor heat transfer of the hot air
into the foam structure. Microwaves have the ability to penetrate a large volume of the product at once, independently of the low
conductivity (Ratti and Kudra, 2006).

3.04.3.2.4 Microwave Foam Drying


As a result, microwave foam drying (MFD) method was developed that combines the advantages of microwave heating and foam
mat drying to improve the drying efficiency and quality of the product. As a volumetric heating mode, the dissipation of microwave
energy induces the water evaporation inside the material to build interior pressure, which expands the volume of bubbles to enlarge
the surface of the moisture evaporation in MFD process. The drying temperature and microwave power are the two most critical
factors comprising the application of microwave. These two factors affect the drying parameters such as drying time, drying speed,
drying curve, drying efficiency, and final product quality. During a typical microwave assisted drying process, throughout the pro-
cessing system, a fixed microwave power level is applied and temperature control is non-existent (Kone et al., 2013; Hardy and
Jideani, 2017). However, Li et al. (2010) reported the possibility of temperature and power control of microwave drying application
in a food material. They found that employing different attributes of microwave power during drying of apples provides optimal
temperature control and product quality. Also, the initial cost of microwave system is relatively high and there are few food mate-
rials containing dielectric properties that will accommodate the use of microwave in drying system (Hardy and Jideani, 2017). These
require the application of microwave in drying processes to be optimized.
Nevertheless, the application of microwave to assist in foam drying was successfully carried out for blackcurrant pulp
(Zheng et al., 2011a), guava (Qadri and Srivastava, 2014), corn soaking water (Qiang et al., 2013; Chai et al., 2016), and berry puree
(Liu et al., 2017). Most of the studies on the use of microwave power as heating source in foam mat drying have concluded that there
is significant reduction in drying time making microwave foam mat drying faster than the conventional method. This is due to huge
reduction in processing time so that the destruction of such constituents is overcompensated even if the temperature of the product
reaches higher values (Qadri et al., 2019). Qadri and Srivastava (2014) studied the effect of microwave power on drying character-
istics and quality attributes of foamed tomato pulp-egg albumin mixture in a microwave-assisted foam mat drying system. They
observed that the drying time was reduced by about 15–16 times in case of microwave assisted foam mat drying as compared
to simple foam mat drying with no adverse effect on color, titrable acidity and pH of the product. These authors also found the
retention of ascorbic acid in samples dried in microwave assisted foam mat drying was also better as compared to that in samples
dried in air convection foam mat drying system. Comparing microwave energy and hot air as drying media for foam mat drying of
blackcurrant pulp, the quality attributes in terms of color and appearance of final product, as well as drying rate of pulp subjected to
microwave conditions are superior to that under hot air conditions (Zheng et al., 2009).
Although foam mat drying has appreciable advantages, the product dried by foam mat drying may not always be superior in
quality when compared with other techniques of drying especially spray drying (Qadri et al., 2019). In general, the foam mat dried
materials are reported to retain the volatiles better compared to the non-foam dried counterparts (Kadam et al., 2012). However, the
flavor retention has also been shown to be impaired by foam mat drying in some foods. For instance, the foam mat dried tomato
paste possessed good solubility but was poor in flavor retention (Lewicki, 1975). Acceleration in the oxidative reactions due to enor-
mous increase in liquid-gas contact surface area is one of the main reasons for the deterioration of product quality including flavor
loss. A study conducted on tamarind pulp reported that the acceptability of the product in terms of flavor was significantly affected
by the type of foaming agent used (Vernon-Carter et al., 2001). Optimization of the drying conditions of MFD for enhancing energy
efficiency and drying quality is therefore important to intensify adoption of this method in the food industry. Also, the need is to
focus on this promising area of drying technology for further upgrade of the process (Qadri et al., 2019).

3.04.3.3 Refractance Window Drying


Refractance windowÔ (RW) drying is a novel drying technique patented by Magoon (1986) and developed by MCD Technologies,
Inc. (Tacoma, Washington, USA). The technology is suitable for producing dried products from liquid and semi-liquid foods
(Bolland, 2000). RW drying is similar to drum drying in that the material is dried in a thin film layer on a heated surface, but in
Advanced Drying Technologies of Relevance in the Food Industry 71

RW drying process the heated surface is at much lower temperature (Abonyi et al., 2001). The equipment is simple to operate and
maintain and relatively inexpensive (Abonyi et al., 1999).

3.04.3.3.1 Operating Principle of RW Drying Technology


The RW drying technology utilizes the refractive principle of the surface of water, which is harnessed by creating a window for the
passage of infrared (IR) energy (Sabarez, 2019). Water is one of the most effective heat transfer media as the three modes of heat
transfer (i.e., convection, conduction, and radiation) function within it. If a heated water is placed in an insulated container, the heat
loss from the heated water to its surroundings occurs mainly through convection/evaporation at the water surface open to the atmo-
sphere while the infrared energy component at this water surface (i.e., water-air interface) is mainly reflected back due to the refrac-
tive index mismatch (Fig. 4A). If the surface of the heated water exposed to the atmosphere is covered by an infrared-transparent
material such as thin plastic film, then the heat loss from the heated water can only occur by means of conduction as the evaporation
and its associated heat loss is blocked.
However, if a moisture-laden material is placed on the plastic surface then a “window” for passage of infrared energy is created at
the point of contact between the material that bears moisture and the plastic surface. This greatly reduces the refractive index
mismatch between the water and food (i.e., water-plastic-food system), allowing the transfer of infrared energy from the body
of heated water through to the material to be dried (i.e., as if there is no plastic material present), as infrared-transparent material
such as plastic allows infrared energy to pass through without converting it into heat (Fig. 4B). Thus at this stage, all three modes of
heat transfer occur between the drying medium (water) and the material to be dried, inducing exceptionally effective heat transfer
and rapid moisture evaporation.
As drying progresses, however, the “window” for passage of infrared energy is slowly closing since the amount moisture of the
material in contact with the plastic surface is decreased as the material dries out (i.e., increases its refractive index creating
a mismatch). This results in a decrease in heat transfer as conduction becomes the dominant mode of heat transfer taking place
(Fig. 4C). As the thermal conductivity of the plastic material is relatively low, little energy is transmitted via conduction from
the heated water to the material on the plastic surface. This reduces the rate of heating protecting the product from overheating
and resulting color and flavor degradation (Bolland, 2000). In theory, the “window” provides the system a means by which it regu-
lates itself the coupled heat and mass transfer mechanisms. Sabarez (2019) also describes the details of the operating principle of
RW drying technology.

3.04.3.3.2 Components of a Commercial RW Drying System


A typical commercial scale RW drying system is made up of a number of heating modules arranged in series with sections for feed
application and product removal at both ends of the system (Fig. 5). Each heating module is approximately 2.4 m wide and 6 m
long. The number of modules in a commercial RW drying plant depends on the throughput requirement, typically around 4–5
modules. For each module, the hot water is continually circulated and maintained at a set temperature (maximum is usually 96  C
as higher temperatures create bubbles which reduces the efficiency of heat transfer). Raghavi et al. (2018) reported the temperature
of the circulating water during RW drying being maintained between 94 and 98  C. A thin layer of puree material to be dried with
a thickness ranging from 0.1 to 1.0 mm is evenly applied to the surface of a 0.2 mm thick transparent plastic (Mylar) conveyor belt
system that floats on the surface of the heated circulating water. The speed of the belt can be adjusted according to the drying time
required. During the drying operation, the moisture released from the material is extracted by ambient air circulated above the belt.
A cooling system (using ambient water) is introduced at the end of the belt to reduce the temperature of the dried product for its
ease of removal (i.e., stickiness issue) from the belt.

3.04.3.3.3 Case Studies


A number of studies were found to be relevant to the RW drying process (Ochoa-Martínez et al., 2012; Caparino et al., 2012; Nindo
et al., 2003a,b, 2004; Abonyi et al., 2001; Bolland, 2000; Clarke, 2004). Some studies have focused on the effects of RW drying
process on product quality attributes such as carotenoids in carrots and ascorbic acid and color of strawberries (Abonyi et al.,
1999); color and antioxidant activity of asparagus (Nindo et al., 2003b) and aloe (Nindo et al., 2006); flavor, color, carotenoids,

A air
B C
air air

convection wet food material dried food material


evaporation
thin plastic film thin plastic film
water-air interface

convection convection conduction convection conduction


radiation radiation radiation

HOT WATER BATH HOT WATER BATH HOT WATER BATH

container wall container wall container wall

Figure 4 Schematic representation of the operating principle of RW drying technology showing the suggested modes of heat transfer from the hot
water bath to the food material. Adapted from Sabarez (2019).
72 Advanced Drying Technologies of Relevance in the Food Industry

A B

Figure 5 Photos of (A) a commercial scale RW dryer facility (RWD5 Model, MCD Technologies, USA), (B) the wet-feed end of the dryer, and
(C) the dried-product exit end of the dryer. Adapted from Sabarez and Chessari (2006), Sabarez (2019).

and capsaicinoids of paprika (Topuz et al., 2009, 2011); encapsulated flavors of orange oil (Cadwallader et al., 2010); antioxidant
compounds of tomatoes (Abul-Fadl and Ghanem, 2011) and colored potatoes (Nayak et al., 2011); bioactive compounds of pota-
toes (Kaspar et al., 2012) and microbial reduction of pumpkins (Nindo et al., 2003a). These studies have reported major advantages
of the RW drying technology over other drying methods (i.e., drum drying and/or spray drying) in terms of excellent color, flavor
and nutrient retention as the food materials are exposed to milder temperatures.
RW drying has also been studied to compare with other drying processes in terms of physical aspects, including energy efficiency
(Abonyi et al., 1999; Nindo et al., 2003a, 2004; 2006; Abul-Fadl and Ghanem, 2011). According to Abonyi et al. (1999), products
can be dried in a few minutes with this technology, contrary to hot air or tunnel dryers which will take several hours. Nindo et al.
(2003a) reported that drying of pumpkin puree from 80% to 5% moisture content (wet basis) was achieved in less than 5 min in
both pilot and commercial scale RW dryers at a circulation water temperature of 95  C, with a 52%–70% energy efficiency of the RW
drying system. The full-scale commercial RW dryer model used by Nindo et al. (2003a,b) consisted of 4 heating modules covering
a length of 12.9 m and 1.8 m cooling section. Raghavi et al. (2018) presented a comprehensive review on the recent trends in refrac-
tance window drying of foods with emphasis on the underlying mechanism and effect on product quality.
Sabarez and Chessari (2006) undertook an industry sponsored study to determine the optimal design and operating conditions
in a commercial scale RW drying process (i.e., using a commercial RW dryer with 5 heating modules) of tomato puree through
computational modeling approach together with commercial scale drying trials to validate the model predictions conditions.
Fig. 6 shows a typical example of the drying curves predicted by the model together with the experimental drying kinetics during
commercial scale RW drying of tomato puree at two water temperatures (Tw). Analysis of the results shows that it took
about 3.1 min to dry tomato puree (feed thickness of 0.2 mm) down to 10% at 96  C water temperature, while it took 7.9 min
to dry at water temperature of 83  C (Sabarez, 2019). These demonstrate the importance of regulating water temperature and
belt speed in achieving the desired final moisture content of the product exiting the dryer. In addition, the work by Sabarez and
Chessari (2006) has also generated new perspectives on the heat and mass transfer mechanisms involved during RW drying process,
confirming that the radiation effects are indeed small and they do not contribute significantly in the commercial RW drying process,

100
90 Measured (Tw=83°C)
Predicted (Tw=83°C)
Moisture Content (%wb)

80
Measured (Tw=96°C)
70
Predicted (Tw=96°C)
60
50
40
30
20
10
0
051015202530
Dryer Length Posion (m)
Figure 6 Measured versus predicted drying kinetics for tomato puree during commercial scale RW drying (V ¼ 0.4–1.8 m/s; Belt speed ¼ 3.91 m/
min).
Advanced Drying Technologies of Relevance in the Food Industry 73

providing invaluable insights for improvements to the design and operation of the commercial drying system. The readers may refer
to the recent publication by Sabarez (2019) for further details of this technology.

3.04.3.4 Ultrasound Drying Technology


Ultrasound is a series of sound waves with frequencies above the threshold of human hearing, 18 kHz (Mason, 1998). It is trans-
mitted through any solid, liquid or gas that possesses elastic (in solids) or acoustic (in fluids) properties. The same basic system
components are needed to generate and transmit ultrasonic waves (no matter what industry or application involved). These include
an electrical power generator, transducer, and emitter (Povey and Mason, 1998). The transducer converts electrical energy (or
mechanical energy, in the case of the liquid whistle) into sound energy by vibrating mechanically at ultrasonic frequencies (Povey
and Mason, 1998). A transducer attached to an electrical generator, for example, will transform 20 kHz electrical energy from the
generator into ultrasound energy of the same frequency by vibrating at 20,000 mechanical cycles per second. The emitter (also called
the reactor) physically radiates the ultrasonic wave from the transducer into the medium. Emitters may also fulfill the role of ampli-
fying the ultrasonic vibrations while radiating them. Fig. 7 shows an example of the basic ultrasonic system components retrofitted
into an experimental drying setup. The ultrasonic unit retrofitted in this setup consisted of a piezoelectric transducer (working
at 20 kHz) with a rectangular flexural vibrating plate driven by an electronic generator. The details of the setup can be found else-
where (Sabarez et al., 2012; Sabarez, 2016a).
It has been known for many years that the energy generated by sound pressure waves could enhance a wide range of processes
due to a series of mechanisms activated by the ultrasonic energy such as heat, diffusion, mechanical rupture, chemical effects, and so
on (Gallego-Juarez et al., 2007). These sound pressure waves are often classified according to their frequency or power. Depending
on its application, ultrasound can be categorized into high frequency (low energy, low intensity) at frequencies >100 kHz (MHz
range), and low frequency (high energy, high intensity) at frequencies range of 20–100 kHz (Jambrak et al., 2008;
Onwude et al., 2017). High frequency ultrasound is mainly used as a non-destructive technique for the analysis and evaluation
of product quality, characterization of the properties of food materials, medical diagnosis, etc (Onwude et al., 2017; Ensminger
and Bond, 2011). While the low frequency ultrasound also known as power ultrasound is used for the intensification of the trans-
port phenomena and it is characterized by causing changes in the physical, chemical and/or biochemical properties of the media in
which it is applied. In recent years, power ultrasound has found practical applications in food processing, especially in drying.

3.04.3.4.1 Application of Ultrasound in Food Drying


A number of investigations have shown the potential of power ultrasound to improve the drying process of various food mate-
rials. In these studies, the ultrasonic energy was transmitted as either airborne to the surface of food material (Garcia-Perez et al.,
2009, 2007a, b, 2010; Khmelev et al., 2008, 2011; Ozuna et al., 2011; Soria and Villamiel, 2010; Kowalski and Mierzwa, 2015) or
in direct contact between the product and the vibrating element (Gallego-Juarez et al., 2007, 2010; Schossler et al., 2012). The

Exhaust
Balance

Vibrating Plate

Transducer Fan Motor


PLC
Airflow

Generator
Heater
Actuator

Steam
Generator

Sensors (x16) Sample Tray Cooling Coil


PLC Inlet
(movable)

Refrigeration Water Tank

Computer FRONT VIEW SIDE VIEW

Figure 7 Schematic diagram of a computerized ultrasound-assisted convective experimental drying system. Adapted from Sabarez et al. (2012),
Sabarez (2016a).
74 Advanced Drying Technologies of Relevance in the Food Industry

direct contact system can promote an accelerated drying process because this system permits good transfer of ultrasonic energy
from the vibrating element to the food material. Nevertheless, the main drawback of this technique may be its difficulty to adapt
to traditional air drying processes, difficulty in controlling the heating effect on the product and difficulty to operate in a contin-
uous mode. On the other hand, the airborne ultrasonic system works without direct contact between the vibrating element and
the food material, which seems to offer much better adaptability to conventional air drying processes. However, the main diffi-
culties in this system arise from the inefficient generation of ultrasonic energy in air and the transfer of such ultrasonic energy
from air into the product due to the acoustic impedance mismatch and the energy absorption by the air at ultrasonic frequencies
(Gallego-Juarez et al., 1999).
Despite several research papers and patents that were reported in this area across the globe, no commercial scale installation of
the application of ultrasound in food drying processes has been developed to date yet (Kowalski and Pawlowski, 2015; Soria and
Villamiel, 2010). This is due to the technological challenges in achieving an efficient transmission of acoustic energy and the prac-
tical difficulties in adapting the technology at an industrial scale drying operation. However, with the advances in ultrasonic designs,
recent progress has been made in bringing the application of ultrasound in drying processes closer to industrial scale operations. In
particular, a new ultrasonic design based on the indirect transmission of ultrasonic energy from the ultrasound emitter through to
the material to be dried was investigated to assist in low temperature drying of food materials (Sabarez et al., 2018, 2019). This new
ultrasonic design has been shown in the laboratory to be highly effective in intensifying low temperature air drying (temperatures
ranging from approximately 40  C to below freezing) of various food materials (e.g., fruits, coffee, and meat products) resulting in
shorter drying times with better product quality attributes.

3.04.3.4.2 Low Temperature Drying (Above Freezing)


Using the new ultrasonic transmission design, a number of drying experiments were carried out at low temperatures (from above
freezing up to 40  C). Fig. 8 shows a typical example of the drying kinetics during the convective air drying of 5 mm apple slices
at 40  C temperature (T) without and with ultrasound at 40 kHz frequency with power level of 466 W. These experiments were
conducted with drying air velocity (V) and relative humidity (RH) maintained at about 1.2 m/s and 25%, respectively. It can be
seen in the plot that the application of the new ultrasonic transmission design in combination with convective air drying signifi-
cantly reduced the overall drying time. To achieve the target moisture content of 25% in wet basis (i.e., typical moisture content
of ready-to-eat fruit snacks), analysis of the drying curves revealed that it took about 8.0 hours to dry the apple samples without
ultrasound and just 4.4 hours with ultrasound using the new ultrasonic transmission system. The results indicate a significant reduc-
tion in drying time of about 45% with the simultaneous application of ultrasound on the convective drying of apple slices (corre-
sponds to a 32% reduction of energy consumption).
Several studies were also reported by a number of authors investigating the effect of various ultrasound parameters and modes
of ultrasonic transmission (i.e., airborne and direct contact) particularly for low temperature drying application (above freezing).
Sabarez et al. (2012) found a significant reduction in drying time of up to 57% with the application of airborne ultrasound
during drying of 5 mm apple slices at 40  C. Under these drying conditions, an approximately up to 54% reduction in energy
consumption during drying could be achieved with the application of airbone ultrasonic energy. In addition, this study shows
that the effect of ultrasound in intensifying the drying process diminishes at high temperatures. Gallego-Juarez et al. (1999) also
showed that the effect of ultrasound on the drying kinetics of apple cubes in forced air drying assisted by airborne ultrasound is
temperature dependent and diminishes at high temperatures. Similarly, Ortuno et al. (2010) observed that the application of
ultrasound (power level of 90 W) provided an average reduction in drying time of over 45% for drying of orange peels. They
found that the energy saving was close to 30% with ultrasonic application. Garcia-Perez et al. (2009) found a significant reduc-
tion in drying time (up to 70% at acoustic power of 90 W) with the application of power ultrasound for convective drying of
eggplant cylinders.

90
without ultrasound
80 with ultrasound
Moisture Content (%wb)

70
60
50
40
30
20
10
0 1 2 3 4 5 6 7 8 9
Drying me (h)
Figure 8 Effect of ultrasound (40 kHz frequency; 466 W power) on the drying kinetics of apple slices (T ¼ 40  C; RH ¼ 25%; V ¼ 1.2 m/s; 5 mm
thickness). Adapted from Sabarez et al. (2018).
Advanced Drying Technologies of Relevance in the Food Industry 75

3.04.3.4.3 Low Temperature Drying (Below Freezing)


Drying of food materials below the freezing point temperature can also be performed under atmospheric pressure, known as atmo-
spheric freeze drying (AFD) (Meryman, 1959; Claussen et al., 2007). AFD has the advantage of low energy consumption because
there is no need of vacuum and the refrigeration requirement is reduced compared to the traditional vacuum freeze drying.
However, working at atmospheric pressure and below freezing temperatures can lead to very long drying times compared to the
vacuum freeze drying. The application of ultrasound has been shown to intensify the AFD process of various food materials. For
example, Fig. 9 shows the drying kinetics of 5 mm apple slices during drying at below freezing temperature (5  C) under atmo-
spheric pressure without and with ultrasound (using the new ultrasonic transmission design developed by Sabarez et al., 2019).
It is clear from plot that the application of ultrasound in combination with AFD significantly reduced the overall drying time.
Analysis of the drying curves revealed that it took about 72 hours to dry the apple slices without ultrasound and just 22 hours with
ultrasound to reach the target moisture content of 25% (wet basis). The results indicate a significant reduction in drying time of
about 69% with the simultaneous application of ultrasound on the atmospheric freeze drying of apple slices (corresponds to
a 41.9% reduction of energy consumption). In this study, the drying process was terminated when the target moisture content
of 25% was achieved. However, it was reported that ultrasound may produce some effects on the interfaces of intercellular spaces
or even cavitations, which may be beneficial in the removal of the most strongly attached moisture in the solid matrix during drying.
This phenomenon is demonstrated by Sabarez et al. (2012) who showed that drying of apple slices with ultrasound reduces the final
moisture content of the product to a very low level (about 2.5%) compared to those dried without ultrasound (about 8.5%) under
the same conditions of the drying air. Similarly, Gallego-Juarez et al. (2007) found a very low final moisture content (less than 1%)
in forced air drying of carrot slices with ultrasound directly coupled to the sample.
A number of studies have also shown the feasibility of employing power ultrasound to accelerate the AFD process (i.e., drying
below freezing temperatures under atmospheric pressure). Garcia-Perez et al. (2012) have reported a maximum drying time reduc-
tion of 77% by applying power ultrasound during atmospheric freeze drying of apple at 10  C. Santacatalina et al. (2015) have
demonstrated that the application of power ultrasound is the parameter with the greatest influence on reducing the atmospheric
freeze drying time of apple samples. Ultrasound-assisted atmospheric freeze drying was also investigated by Bantle and co-
workers (Bantle and Eikevik, 2011; Bantle et al., 2010). In these studies a heat-pump fluidized bed dryer was modified with an ultra-
sonic transducer as a source of airborne ultrasound for drying of apple and peas samples (Bantle and Eikevik, 2011; Bantle et al.,
2010). The results of these studies revealed that the ultrasound caused an increase in the rate of drying by 6%–17.5% and improved
the system’s energy efficiency by 3%–11% (Bantle et al., 2010). In terms of energy consumption, Wolff and Gilbert (1990) reported
that atmospheric freeze drying can provide an energy saving of 35% compared to vacuum freeze drying.

3.04.3.4.4 Drying Performance of Various Ultrasonic Designs


The differences in drying performance (i.e., time reduction and energy saving) of various ultrasonic drying systems reported in the
literature for drying of food materials could be due to the differences in the raw material properties, drying methods and conditions,
and ultrasonic parameters used (i.e., particularly ultrasonic frequency, power and mode of ultrasound transfer). It should be noted
that most of these studies were carried out at low frequencies (20–26 kHz) with drawbacks of low energy efficiency, difficulty in
controlling the product temperature and high noise levels. These limitations are also a result of the mode in which the ultrasonic
energy is transmitted, either in direct contact between the vibrating plate and the product (Gallego-Juarez et al., 2007; Schossler
et al., 2012) or airborne to the surface of the product (Beck et al., 2014; Sabarez et al., 2012).
The new design developed by Sabarez et al. (2019) and Sabarez et al. (2018) could offer a better applicability at industrial scale
as it allows easy adoption in a continuous operation, since no direct contact between the sample and the ultrasonic emitter is
needed and it operates at atmospheric pressure conditions. It also facilitates an efficient transmission of the ultrasonic energy as
the mismatch of acoustic impedance is minimized, and the use of high ultrasonic frequencies (40 kHz and above) enables for

90
without ultrasound
80
with ultrasound
Moisture Content (%wb)

70

60

50

40

30

20

10

0
0 10 20 30 40 50 60 70 80
Drying me (h)

Figure 9 Drying kinetics of apple slices during atmospheric freeze drying (AFD) without and with ultrasound at 40 kHz frequency & 466 W power
(T ¼ 5  C; RH ¼ 75%; V ¼ 1.0 m/s; 5 mm thickness). Adapted from Sabarez et al. (2019).
76 Advanced Drying Technologies of Relevance in the Food Industry

reduced noise levels. The innovation embeds a temperature controlling mechanism that provides a better control of the product
temperatures (i.e., removing/recovering heat generated by the ultrasonic system). Also, the design would allow the ultrasonic system
to provide the necessary energy required for sublimation (in case of atmospheric freeze drying) without the need for an additional
heating device. In general, this new approach has great potential as a low cost alternative to the expensive vacuum freeze drying and
would also allow intensification of lengthy low temperature (from 40  C to below freezing) drying under atmospheric pressure,
providing a promising nonthermal means for gentle (i.e., low temperature) drying of food materials to produce premium quality
food products.

3.04.3.5 Electrohydrodynamic Drying


Electrohydrodynamic (EHD) drying is a new nonthermal drying technology, based on ionic wind generation between an emitter
and a collector electrode (Defraeye and Martynenko, 2019). The technology is particularly suitable to improve drying of heat-
sensitive materials, predominantly food products. It is a promising method to dry high-value agricultural products convectively
without using additional heat (Bajgai et al., 2006; Kudra and Martynenko, 2015; Zhang et al., 2017). EHD drying has been studied
for over two decades on a laboratory scale (Barthakur, 1990; Chen et al., 1994; Hashinaga et al., 1999; Isobe et al., 1999). It has been
used to dry food products, including apple (Martynenko and Zheng, 2016), mushrooms (Taghian Dinani et al., 2014, 2015a,b;
Dinani and Havet, 2015), banana (Esehaghbeygi et al., 2014; Pirnazari et al., 2016), mango (Bardy et al., 2016), okara cake
(Li et al., 2006) or tomato (Esehaghbeygi and Basiry, 2011). Compared to traditional convective drying, EHD drying has been re-
ported to reduce drying time and product shrinkage, increase rehydration capacity, improve (soften) texture and preserve color and
flavor better (Bai et al., 2013; Esehaghbeygi and Basiry, 2011; Esehaghbeygi et al., 2014; Taghian Dinani et al., 2015a, 2014). It was
also found to preserve better nutritional content (vitamin C, proteins, carotene) (Taghian Dinani et al., 2015b; Yang and
Ding, 2016). Also, the nonthermal nature of EHD drying may present a wide range of applications in industrial drying, given its
capacity to produce high quality processed products (Bajgai and Hashinaga, 2001a,b).

3.04.3.5.1 Components of EHD Drying System


EHD drying system consists of a vertically moveable emitter electrode (a thin wire or a sharply pointed needle) projected to a hor-
izontally grounded collector electrode (such as a metallic plate) on which the material to be dried is placed (Fig. 10) (Bajgai and
Hashinaga, 2001a; Bajgai et al., 2006). A high voltage (in several kiloVolts) is applied at the emitter electrode (wire or needle). In
EHD drying, either AC or DC high voltage of ordinary frequency (60 Hz) can be used with DC current typically applied, but some
studies reported efficient AC applications as well (Isobe et al., 1999; Yang and Ding, 2016; Zheng et al., 2011b). The point electrode
is connected to an AC or DC high-voltage transformer which supplies either negative or positive high-voltage. In order to set
the desired high-voltage parameters for EHD drying, the transformer is connected to a voltage regulator, which is supplied from
a 60-Hz, 110-V, AC source (Bajgai et al., 2006). The electrode gap can be adjusted to the desired distance by moving the emitter
electrode up and down.

3.04.3.5.2 Principle of Operation in EHD Drying


The principle of operation behind EHD drying is that airflow is generated by means of high-voltage corona discharge, which induces
ionic wind (Defraeye and Martynenko, 2018). To generate corona discharge, the voltage should exceed the breakdown electric field
strength of air locally near the emitter. The required voltage for the initiation of corona discharge is strongly dependent on the

Figure 10 A schematic diagram of electrohydrodynamic (EHD) drying system. Adapted from Bajgai et al. (2006).
Advanced Drying Technologies of Relevance in the Food Industry 77

emitter and collector geometry, their curvature, and the distance between the electrodes (Defraeye and Martynenko, 2018). A too
high voltage should be avoided as this induces complete breakdown of the air, leading to spark-over or arcing. A high voltage differ-
ence is created between an emitter electrode, with very small radius thus large curvature (wire or needle), and a grounded collector
electrode with a much smaller curvature (Defraeye and Martynenko, 2018). As a result, for sufficiently high voltages (in kiloVolt
range), a small region in the air around the emitter is ionized. The ionized air is a cluster of molecules bound together by the
Coulomb force associated with an excess or deficiency of electrons (Bajgai et al., 2006). The electrostatic (Coulomb) force acting
on these ions makes them drift toward the collector electrode. In this process, the ions collide with the surrounding neutral air mole-
cules. The resulting exchange of momentum leads to a net air movement toward the collector electrode (i.e., impinging airflow onto
the food and collector electrode).
In EHD drying, the movement of the air ions in a strong electric field that generates an ionic wind disturbs the boundary layer
and enhances convective heat and mass transfer between the product that needs to be dried. During this process, a lowering of
entropy also occurs due to polarization of water molecules in the electric field, which in turn lowers the temperature of the material
being dried (Bajgai et al., 2006). However, in this system the moisture removal efficiency for impinging flow can dramatically reduce
if multiple products are placed on the collector electrode plate. According to Defraeye and Martynenko (2019), it is probably diffi-
cult to achieve a uniform drying rate for multiple food products, located at different distances from the emitter for a single wire-to-
plate configuration (Fig. 11A). As an EHD-driven air jet is directed toward the plate, it is then diverted to the sides over the products.
When air passes over successive products, partial saturation of the air with vapor will occur, which will likely reduce the drying rate
of products more downstream. If more emitters are placed (Fig. 11B), another problem arises due to the multiple air jets generated
that bounce back from the product and a part of the airflow is directed again upwards. This partially saturated moist air then recir-
culates back to the sample, which can also slow down the drying rate.

3.04.3.5.3 Advantages and Disadvantages


In EHD drying technology, there are no moving parts or induced vibrations as the airflow is produced locally by corona discharge
instead of mechanically by a fan. In particular, the EHD system can provide airflow much more locally than large fan systems, which
can yield more efficient drying (Defraeye and Martynenko, 2018). EHD drying has also a shorter response time and requires limited
power to generate corona discharge compared to mechanically-generated convective airflow generation (e.g., a fan) (Defraeye and
Martynenko, 2018). Compared to high-end drying technologies (e.g., freeze drying), the hardware requirements and operational
cost are less. The power required to produce a corona discharge is very low (typically 1–10 W), which is in the same range as
a computer fan. However, the total power consumption for EHD is typically much higher, caused by the other components of
the system. Martynenko and Zheng (2016) for example reported 55–75 W for their system. Kudra and Martynenko (2015) reported
that the high-voltage converter consumed 85%–99% of the total energy required for EHD drying. The energy consumed by the
auxiliary equipment (e.g., converter or fan) has a large impact on the total energy consumption for EHD airflow generation.
This fact, which is often ignored in the lab-scale research, forms a major limitation for industrial applications of EHD drying (Kudra
and Martynenko, 2015).

3.04.3.5.4 Challenges and Opportunities for Industrial Implementation


Although EHD drying technology has been known for several decades, it is still evolving toward industrial implementation, i.e.,
needs further development for industrial use (Defraeye and Martynenko, 2018). Working prototypes of the technology have
been reported (Lai, 2010), but no commercial EHD dryers are available yet (Defraeye and Martynenko, 2019). A key hurdle to
industrial implementation is the limited possibility to upscale the commonly studied configuration (i.e., needle-to-plate), where
airflow impinges onto the food and collector electrode plate (Fig. 11A). This configuration leads to drastically reduced moisture
removal rates when drying hundreds of products simultaneously as industrial processes require simultaneous drying of large
amounts of products. In this configuration, evaporation is also restricted from the bottom surface of the food product, thereby
inducing non-uniform drying conditions inside each food product. Moreover, the step toward industrial upscaling is further
hindered by the fact that the EHD system can cause electromagnetic interference and audible noise, and the electrodes attract
dust, which can affect the performance of the EHD over time (so a periodic cleaning of electrodes or a-priori filtering of the air

Figure 11 Types of configurations for EHD drying of multiple products (A) impinging flow for a single wire-to-plate, (B) impinging flow for
a periodic wire-to-plate, (C) flow around the products for a periodic wire-to-mesh (emitter ¼ red; collector ¼ black; periodic conditions indicate that
multiple products are placed sideways). Adapted from Defraeye and Martynenko (2019).
78 Advanced Drying Technologies of Relevance in the Food Industry

is required). As a byproduct, ozone is produced, but also nitrous oxide (Chen and Davidson, 2002; Hashinaga et al., 1999). It is
imperative that the safety thresholds of these chemicals are not exceeded.
According to Defraeye and Martynenko (2018), the key to upscaling EHD drying technology lies in novel electrode configura-
tions. Upscaling to an industrial level with the commonly studied configurations (i.e., traditional flow impingement to the plate)
could create problems (for example saturation of the air with moisture). To avoid water vapor accumulation in the drying zone,
a wire-to-mesh configuration is proposed by Defraeye and Martynenko (2018) (Fig. 11C). In this configuration, airflow is able
to pass alongside the food instead of impinging to it. The authors reported that the mesh collector minimizes interference of neigh-
boring airflows and avoids recirculation of moist air in the drying zone. As such, the wire-to-mesh configuration provides more
uniform drying between adjacent products, but also within a product, as it can dry from all its surfaces. The authors have also
demonstrated to achieve significantly faster drying rates (i.e., roughly up to twofold) when placing food on a mesh instead of a plate
and dried uniformly. This alternative wire-to-mesh configuration could provide a better potential for industrial upscaling with the
additional advantage that products can dry uniformly from all their surfaces.

3.04.4 Concluding Remarks

The advanced drying technologies described in this review that are already commercialized, including heat pump drying, foam mat
drying and refractance window drying, will continue to play a significant role in the industrial manufacturing of dried products and
ingredients as long as these technologies are still viable and have not reached their limit of performance. On the other hand, the
advanced drying technologies (i.e., ultrasound drying and electrohydrodynamic drying) yet to be commercialized will continue
to evolve toward commercial implementation as these technologies have shown great potentials as efficient and cost-effective alter-
native to current low temperature drying methods. In addition, efforts in the developments of innovative new drying technologies
will also continue to be at the forefront in the pursuit of innovations to meet the continually emerging challenges and new oppor-
tunities beyond the limits of the current drying technologies. These new innovations will be important for the sustainable growth of
the food industry into the future and to help achieve a reduction of the industries’ environmental footprint. This will also contribute
in improved health and well-being through increased availability of a new range of affordable and premium food products and
ingredients that can be conveniently prepared and demonstrated to retain high contents of flavors and nutrients.
There are still challenges to overcome in the developments of innovative new drying technologies. Scaling-up of these new inno-
vations at industrial scale remains to be a major challenge because of the highly nonlinear nature of the governing equations of the
transport processes during drying. This is coupled with the complex properties inherent in food materials together with the addi-
tional physics (e.g., ultrasound, pulsed electric field, electromagnetic, electrohydrodynamic, etc.) involved in the advanced drying
processes. A multidisciplinary approach is crucially important to pursue a better understanding of the underlying drying fundamen-
tals and the interplay between transport phenomena and the material properties. With advances in other relevant fields (i.e. sensing,
digital, computing, automation and visualization), the next level of sophistication in drying technologies should continue to evolve
to meet the emerging challenges into the future.

References

Abonyi, B.I., Fenh, H., Tang, J., Edwards, C.G., Chew, B.P., Mattison, D.S., Fellman, J.K., 2001. Quality retention in strawberry and carrot purees dried with refractance window
system. J. Food Sci. 67 (2), 1051–1056.
Abonyi, B.I., Tang, J., Edwards, C.G., 1999. Evaluation of Energy Efficiency and Quality Retention for the Refractance Window™ Drying System. Research Report. Washington State
University, Pullman.
Abul-Fadl, M., Ghanem, T., 2011. Effect of Refractance Window (RW) drying method on quality criteria of produced tomato powders as compared to the convection drying method.
World Appl. Sci. J. 15 (7), 953–965.
Ambros, S., Dombrowski, J., Boettger, D., Kulozik, U., 2019. The concept of microwave foam drying under vacuum: a gentle preservation method for sensitive biological material.
J. Food Sci. 84 (7), 1682–1691.
Bag, S.K., Srivastav, P.P., Mishra, H.N., 2011. Optimization of process parameters for foaming of bael (Aegle marmelos L.) fruit pulp. Food Bioprocess Technol. 4 (8), 1450–1458.
https://doi.org/10.1007/s11947-009-0243-6.
Bai, Y., Qu, M., Luan, Z., Li, X., Yang, Y., 2013. Electrohydrodynamic drying of sea cucumber (Stichopus japonicus). LWT - Food Sci. Technol. 54 (2), 570–576.
Bajgai, T.R., Raghavan, G.S.V., Hashinaga, F., Ngadi, M.O., 2006. Electrohydrodynamic drying - a concise overview. Dry. Technol. 24 (7), 905–910.
Bajgai, T.R., Hashinaga, F., 2001a. High electric field drying of Japanese radish. Dry. Technol. 19 (9), 2291–2301.
Bajgai, T.R., Hashinaga, F., 2001b. Drying of spinach with a high electric field. Dry. Technol. 19 (9), 2331–2341.
Bansal, P.R., Chung, K.Y., 2007. Food drying equipment and design. In: Hui, Y.H., Clary, C., Farid, M.M., Fasina, O.O., Noomhorm, A., Welti-Chenis, J. (Eds.), Food Drying Science
and Technology: Microbiology, Chemistry, Applications. DESTech Publications, Inc, Pennsylvania, USA, pp. 359–402 (Chapter 15).
Bantle, M., Eikevik, T.M., 2011. Parametric study of high-intensity ultrasound in the atmospheric freeze drying of peas. Dry. Technol. 29 (10), 1230–1239.
Bantle, M., Eikevik, T.M., Grüttner, A., 2010. Mass transfer in ultrasonic assisted atmospheric freeze drying. In: Proceeding of the 17th International Drying Symposium (IDS 2010),
Magdeburg, Germany, vol. B, pp. 763–768.
Bardy, E., Manai, S., Havet, M., Rouaud, O., 2016. Drying kinetics comparison of methylcellulose gel vs. mango fruit in forced convective drying with and without electro-
hydrodynamic enhancement. ASME J. Heat Transf. 138, 84504. https://doi.org/10.1115/1.4033390.
Barthakur, N., 1990. Electrohydrodynamic from NaCl solutions enhancement of evaporation. Desalination 78 (3), 455–465.
Beck, S.M., Sabarez, H.T., Gaukel, V., Knoerzer, K., 2014. Enhancement of convective drying by application of airborne ultrasound: a response surface approach. Ultrason.
Sonochem. 21, 2144–2150.
Bolland, K.M., 2000. A new low-temperature/short-time drying process. Cereal Foods World 45, 293–296.
Advanced Drying Technologies of Relevance in the Food Industry 79

Brygidyr, A.M., Rzepecka, M.A., McConnell, M.B., 1977. Characterization and drying of tomato paste foam by hot air and microwave energy. Can. Inst. Food Sci. Technol. J. 10 (4),
313–319.
Cadwallader, K., Moore, J., Zhang, Z., Schmidt, S., 2010. Comparison of spray drying and Refractance WindowTM drying technologies for the encapsulation of orange oil. In:
Ho, C.T., Mussinan, C.J., Shahidi, F., Contis, E.T. (Eds.), Recent Advances in Food and Flavor Chemistry: Food Flavors and Encapsulation. The Royal Society of Chemistry,
pp. 246–254.
Campbell, C.H., 1917. Drying Milk. US Patent 1250427.
Caparino, O., Tang, J., Nindo, C., Sablani, S., Powers, J., Fellman, J., 2012. Effect of drying methods on the physical properties and microstructures of mango (Philippine ‘Carabao’
var.) powder. J. Food Eng. 111, 135–148.
Chai, L., Qiang, L., Cheng-hai, L., Xian-zhe, Z., 2016. Process parameter study on microwave assisted foam-mat drying properties of corn soaking water. J. NE Agric. Univ. 23 (2),
65–77. https://doi.org/10.1016/S1006-8104(16)30049-6.
Chen, Y., Barthakur, N.N., Arnold, N.P., 1994. Electrohydrodynamic (EHD) drying of potato slabs. J. Food Eng. 23 (1), 107–119.
Chen, J., Davidson, J.H., 2002. Ozone production in the positive DC corona discharge: model and comparison to experiments. Plasma Chem. Plasma Process. 22 (4), 495–522.
Chou, S.K., Chua, K.J., 2007. Heat pump drying systems. In: Mujumdar, A.S. (Ed.), Handbook of Industrial Drying. Taylor & Francis Group, Boca Raton, FL, pp. 1104–1130.
Clarke, P.T., 2004. Refractance window – down under. In: Proceedings of the 14th International Drying Symposium (IDS 2004), Sao Paulo, Brazil, vol. B, pp. 813–820.
Claussen, I.C., Indergård, E., Magnussen, O.M., et al., 2012. Factors, Influencing the Drying Process of Salted Fish (Cod-Gadus Microcephalus and Saithe–Pollachius virens) Part B:
Water Removal Rate from Salted Fish Based on Weight and Drying Air Conditions.
Claussen, I.C., Ustad, T.S., Strommen, I., Walde, P.M., 2007. Atmospheric freeze drying – a review. Dry. Technol. 25, 957–967.
Darniadi, S., Ho, P., Murray, B.S., 2018. Comparison of blueberry powder produced via foam-mat freeze-drying versus spray-drying: evaluation of foam and powder properties.
J. Sci. Food Agric. 98 (5), 2002–2010. https://doi.org/10.1002/jsfa.8685.
Defraeye, T., Martynenko, A., 2019. Electrohydrodynamic drying of multiple food products: evaluating the potential of emitter-collector electrode configurations for upscaling. J. Food
Eng. 240, 38–42.
Defraeye, T., Martynenko, A., 2018. Electrohydrodynamic drying of food: new insights from conjugate modeling. J. Clean. Prod. 198, 269–284.
Dinani, S.T., Havet, M., 2015. Effect of voltage and air flow velocity of combined convective-electrohydrodynamic drying system on the physical properties of mushroom slices. Ind.
Crop. Prod. 70, 417–426.
Ensminger, D., Bond, L.J., 2011. Applications of high intensity ultrasonics basic mechanisms and effects. In: Ultrasonics: Fundamentals, Technologies, and Applications, third ed.
CRC Press, Taylor & Francis Group, Boca Raton, United States, pp. 459–494.
Esehaghbeygi, A., Pirnazari, K., Sadeghi, M., 2014. Quality assessment of electrohydrodynamic and microwave dehydrated banana slices. LWT - Food Sci. Technol. 55 (2),
565–571.
Esehaghbeygi, A., Basiry, M., 2011. Electrohydrodynamic (EHD) drying of tomato slices (Lycopersicon esculentum). J. Food Eng. 104 (4), 628–631.
Falade, K., Adeyanju, K., Uzo-Peters, P., 2003. Foam-mat drying of cowpea (Vigna unguiculata) using glyceryl monostearate and egg albumin as foaming agents. Eur. Food Res.
Technol. 217 (6), 486–491. https://doi.org/10.1007/s00217-003-0775-3.
Febrianto, A., Kumalaningsih, S., Aswari, A.W., 2012. Process engineering of drying milk powder with foam mat drying method: a study of the effect of the concentration and types
of filler. J. Basic Appl. Sci. Res. 2, 3588–3592.
Gallego-Juarez, J.A., Rodriguez-Corral, G., Galvez-Moraleda, J.C., Yang, T.S., 1999. A new high intensity ultrasonic technology for food dehydration. Dry. Technol. 17 (3), 597–608.
Gallego-Juarez, J.A., Riera, E., de la Fuente Blanco, S., Rodriguez-Corral, G., Acosta-Aparicio, V.M., Blanco, A., 2007. Application of high-power ultrasound for dehydration of
vegetables: processes and devices. Dry. Technol. 25 (11), 1893–1901.
Garcia-Perez, J.V., Carcel, J.A., Riera, E., Rosello, C., Mullet, A., 2012. Intensification of low temperature drying by using ultrasound. Dry. Technol. 30, 1199–1208.
Garcia-Perez, J.V., Carcel, J.A., Riera, E., Mulet, A., 2009. Influence of the applied acoustic energy on the drying of carrots and lemon. Dry. Technol. 27 (2), 281–287.
Garcia-Perez, J.V., Carcel, J.A., Benedito, J., Mulet, A., 2007a. Power ultrasound mass transfer enhancement in food drying. Food Bioprod. Process. 85, 247–254.
Garcia-Perez, J.V., Carcel, J.A., Benedito, J., Riera, E., Mulet, A., 2007b. Influence of process variables on hot air drying assisted by power ultrasound. In: Proceedings of the 19th
International Congress on Acoustics, Madrid, Spain.
Garcia-Perez, J.V., Puig, A., Perez-Munuera, I., Carcel, J.A., Riera, E., 2010. Kinetic and microstructural changes induced by power ultrasound application on convective drying of
eggplant. In: Proceedings of the 20th International Congress on Acoustics (ICA 2010), Sydney, Australia.
Hardy, Z., Jideani, V.A., 2017. Foam-mat drying technology: a review. Crit. Rev. Food Sci. Nutr. 57 (12), 2560–2572. https://doi.org/10.1080/10408398.2015.1020359.
Hashinaga, F., Bajgai, T.R., Isobe, S., Barthakur, N.N., 1999. Electrohydrodynamic (EHD) drying of apple slices. Dry. Technol. 17 (3), 479–495.
Hawlader, M.N.A., et al., 2001. On the steady-state modeling of a two-stage evaporator system. Int. J. Energy Res. 25, 859–880.
Heldman, D.R., Hartel, R.W., 1997. Principles of Food Processing. Chapman and Hall, NewYork.
Isobe, S., Barthakur, N.N., Yoshino, T., Okushima, L., Sase, S., 1999. Electrohydrodynamic characteristics of agar gel. Food Sci. Technol. Res. 5, 132–136.
Jambrak, A.R., Mason, T.J., Lelas, V., Herceg, Z., Herceg, I.L., 2008. Effect of ultrasound treatment on solubility and foaming properties of whey protein suspensions. J. Food Eng.
86 (2), 281–287. https://doi.org/10.1016/j.jfoodeng.
Jangam, S.V., 2011. An overview of recent developments and some R&D challenges related to drying of foods. Dry. Technol. Int. J. 29 (12), 1343–1357.
Kadam, D.M., Patil, R.T., Kaushik, P., 2010. Foam mat drying of fruit and vegetable products. In: Jangam, S.V., Law, C.L., Mujumdar, A.S. (Eds.), Drying of Foods, Vegetables and
Fruits, pp. 111–124 (Singapore).
Kadam, D.M., Wilson, R.A., Kaur, S., Manisha, 2012. Influence of foam mat drying on quality of tomato powder. Int. J. Food Prop. 15, 211–220.
Kandasamy, P., Varadharaju, N., Kalemullah, S., Maladhi, D., 2014. Optimization of process parameters for foam-mat drying of papaya pulp. J. Food Sci. Technol. 51 (10),
2526–2534. https://doi.org/10.1007/s13197-012-0812-y.
Kaspar, K., Park, J., Mathison, B.B.M.S., Boon, P., 2012. Processing of pigmented-flesh potatoes (Solanum tuberosum L.) on the retention of bioactive compounds. Int. J. Food Sci.
Technol. 47, 376–382.
Khmelev, V.N., Barsukov, R.V., Abramenko, B.S., Genne, D.V., 2008. Research and development of ultrasonic device prototype for intensification of drying process. In: 9th
International Workshop and Tutorials EDM’ 2008, Erlagol, Russia.
Khmelev, V.N., Shalunov, A.V., Barsukov, R.V., Abramenko, D.S., Lebedev, A.N., 2011. Studies of ultrasonic dehydration efficiency. J. Zhejiang Univ. Sci. A Appl. Phys. Eng. 12 (4),
247–254.
Kone, K.Y., Druon, C., Gnimpieba, E.Z., Delmotte, M., Duquenoy, A., Laguerre, J.C., 2013. Power density control in microwave assisted air drying to improve quality of food. J. Food
Eng. 119, 750–751.
Kowalski, S.J., Mierzwa, D., 2015. US-assisted convective drying of biological materials. Dry. Technol. 33, 1601–1613.
Kowalski, S.J., Pawlowski, A., 2015. Intensification of apple drying due to ultrasound enhancement. J. Food Eng. 156, 1–9.
Kudra, T., Martynenko, A., 2015. Energy aspects in electrohydrodynamic drying. Dry. Technol. 33 (13), 1534–1540.
Lai, F.C., 2010. A prototype of EHD-enhanced drying system. J. Electrost. 68 (1), 101–104.
Lewicki, P.P., 1975. Mechanisms Concerned in Foam-Mat Drying of Tomato Paste, vol. 55. Transactions of Agriculture and Academy, Warsaw, pp. 1–67.
Li, Z.F., Raghavan, G.S.V., Orsat, V., 2010. Temperature and power control in microwave drying. J. Food Eng. 97 (4), 478–483.
Li, F.D., Li, L.T., Sun, J.F., Tatsumi, E., 2006. Effect of electrohydrodynamic (EHD) technique on drying process and appearance of okara cake. J. Food Eng. 77 (2), 275–280.
Liu, C., Liu, C., Xue, H., Sun, Y., Lin, Z., Liu, H., Hou, J., Zheng, X., 2017. Effect of microwave energy dissipation on drying process of berry puree under microwave foam drying
conditions. Dry. Technol. 35 (11), 1388–1397.
80 Advanced Drying Technologies of Relevance in the Food Industry

Magoon, R., 1986. Method and Apparatus for Drying Fruit Pulp and the Like. US Patent 4, 631, 837.
Martynenko, A., Zheng, W., 2016. Electrohydrodynamic drying of apple slices: energy and quality aspects. J. Food Eng. 168, 215–222.
Mason, T.J., 1998. Power ultrasound in food processing - the wayforward. In: Povey, M.J.W., Mason, T.J. (Eds.), Ultrasound in Food Processing. Thomson Science, London, UK,
pp. 105–126.
Meryman, H.T., 1959. Sublimation. Freeze-drying without vacuum. Science 130, 628–629.
Minea, V., 2013. Heat pump-assisted drying: recent technological advances and R&D needs. Dry. Technol. 31, 1177–1189.
Mink, L.D., 1939. Egg Material Treatment. US Patent 2183516.
Mink, L.D., 1940. Treatment of Egg Whites. US Patent 2200963.
Morgan, A.I., Graham, R.P., Ginnette, L.F., Williams, G.S., 1961. Recent developments in foam-mat drying. Food Technol. 15, 37–39.
Mujumdar, A.S., Devahastin, S., 2008. Fundamental principles of drying. In: Mujumdar, A.S. (Ed.), Guide to Industrial Drying: Principles, Equipments and New Developments. Three
S Colors Publications, Mumbai, India, pp. 1–21.
Muthukumaran, A., Ratti, C., Raghavan, V.G.S., 2008a. Foam-mat freeze drying of egg white and mathematical modeling part I: optimization of egg white foam stability. Dry.
Technol. 26 (4), 508–512.
Muthukumaran, A., Ratti, C., Raghavan, V.G.S., 2008b. Foam-mat freeze drying of egg white mathematical modeling part II: freeze drying and modeling. Dry. Technol. 26 (4), 513–
518. https://doi.org/10.1080/07373930801929615.
Nayak, B., Berrios, J., Powers, J., Tang, J., Ji, Y., 2011. Colored potatoes (Solanum tuberosum L.) dried for antioxidant-rich value-added foods. J. Food Process. Preserv. 35,
571–580.
Nindo, C.I., Feng, H., Shen, H.Q., Tang, J., Kang, D.H., 2003a. Energy utilisation and microbial reduction in a new film drying system. J. Food Process. Preserv. 27, 117–136.
Nindo, C.I., Sun, T., Wang, S.W., Tang, J., Powers, J.R., 2003b. Evaluation of drying technologies for retention of physical quality and antioxidants in asparagus (Asparagus
officinalis L.). LWT- Food Sci. Technol. 36, 507–516.
Nindo, C.I., Tang, J., Powers, J.R., Bolland, K., 2004. Energy consumption during Refractance Window® evaporation of selected berry juices. Int. J. Energy Res. 28, 1089–1100.
Nindo, C., Tang, J., Cakir, E., Powers, J., 2006. Potential of Refractance Window technology for value added processing of fruits and vegetables in developing countries. In: ASABE
Annual International Meeting Sponsored, Portland, Oregon.
Ochoa-Martínez, C., Quintero, P., Ayala, A., Ortiz, M., 2012. Drying characteristics of mango slices using the Refractance WindowTM technique. J. Food Eng. 109, 69–75.
Oliver, T.N., 1982. Process drying with a dehumidifying heat pump. In: Paper C2 Presented at the Int Sym on the Industrial Application of Heat Pumps, Coventry, England, 24–26
March, pp. 73–88.
Ong, S.P., 2016. Advances in heat pump-assisted drying of fruits. In: Minea, V. (Ed.), Advances in Heat Pump-Assisted Drying Technology. Taylor & Francis Inc., pp. 149–176
(Chapter 4).
Uno, J., Kobayashi, N., 1985. Belt type foam mat dryers. In: Toei, R., Mujumdar, A.S. (Eds.), Drying ’85. Springer-Verlag, Berlin Heidelberg, pp. 345–348.
Onwude, D.I., Hashim, N., Janius, R., Abdan, K., Chen, G., Oladejo, A.O., 2017. Non-thermal hybrid drying of fruits and vegetables: a review of current technologies. Innovat. Food
Sci. Emerg. Technol. 43, 223–238.
Ortuno, C., Perez-Munuera, I., Puig, A., Riera, E., Garcia-Perez, J.V., 2010. Influence of power ultrasound application on mass transport and microstructure of orange peel during
hot air drying. Phys. Proc. 3, 153–159.
Ozuna, C., Carcel, J.A., Garcia-Perez, J.V., Mulet, A., 2011. Improvement of water transport mechanisms during potato drying by applying ultrasound. J. Sci. Food Agric. 91,
2511–2517.
Patel, K.K., Kar, A., 2011. Heat pump drying of agricultural produce – an overview. J. Food Sci. Technol. 49 (2), 142–160.
Perera, C.O., Rahman, M.S., 1997. Heat pump dehumidifier drying of food. Trends Food Sci. Technol. 8 (3), 75–79.
Perera, C.O., 2016. Advances in heat pump-assisted agro-food drying technologies. In: Minea, V. (Ed.), Advances in Heat Pump-Assisted Drying Technology. Taylor & Francis Inc.,
pp. 129–148 (Chapter 3).
Pirnazari, K., Esehaghbeygi, A., Sadeghi, M., 2016. Modeling the electrohydrodynamic (EHD) drying of banana slices. Int. J. Food Eng. 12 (1), 17–26.
Povey, M., Mason, T., 1998. Ultrasound in Food Processing. Blackie Academic and Professional, London.
Qadri, O.S., Srivastava, A.K., Yousuf, B., 2019. Trends in foam mat drying of foods: special emphasis on hybrid foam mat drying technology. Crit. Rev. Food Sci. Nutr. https://
doi.org/10.1080/10408398.2019.1588221.
Qadri, O.S., Srivastava, A.K., 2014. Effect of microwave power on foam-mat drying of tomato pulp. Agric. Eng. Int. CIGR J. 16, 238–244.
Qiang, L., Yu, S., Xiang-wen, X., Qin-quin, Z., Xian-zhe, Z., 2013. Drying characteristics of microwave assisted foam drying of corn soaking water. J. NE Agric. Univ. 20 (1), 53–59.
https://doi.org/10.1016/S1006-8104(13)60009-4.
Raghavi, L.M., Moses, J.A., Anandharamakrishnan, C., 2018. Refractance Window drying of foods: a review. J. Food Eng. 222, 267–275.
Ratti, C., 2001. Hot air and freeze-drying of high-value foods: a review. J. Food Eng. 49, 311–319.
Ratti, C., Kudra, T., 2006. Drying of foamed biological materials: opportunities and challenges. Dry. Technol. 24 (9), 1101–1108. https://doi.org/10.1080/07373930600778213.
Sabarez, H., 2019. Refractance Window™ drying: a mechanistic understanding of the drying process using modelling approach. In: Reference Module in Food Science. Elsevier,
pp. 1–14. https://doi.org/10.1016/B978-0-08-100596-5.21439-X.
Sabarez, H.T., 2018. Thermal drying of food materials. In: Rosenthal, A., Deliza, R., Welti-Chanes, J., Barbosa-Canovas, G.V. (Eds.), Fruit Preservation: Novel and Conventional
Technologies. Springer-Verlag, New York, USA, pp. 181–210 (Chapter 7).
Sabarez, H.T., 2017. Computational modeling of drying processes of food materials. In: Reference Module in Food Science. Elsevier, pp. 1–11.
Sabarez, H.T., 2016a. Airborne ultrasound for convective drying intensification. In: Knoerzer, K., Juliano, P., Smithers, G. (Eds.), Innovative Food Processing Technologies
– Extraction, Separation, Component Modification and Process Intensification. Woodhead Publishing, An Imprint of Elsevier, Elsevier B.V., pp. 361–386 (Chapter 14).
Sabarez, H.T., 2016b. Drying of food materials. In: Reference Module in Food Science. Elsevier, pp. 1–9.
Sabarez, H.T., 2015. Modelling of drying processes of food materials. In: Bakalis, S., Knoerzer, K., Fryer, P. (Eds.), Modelling Food Processing Operations Woodhead Publishing. An
Imprint of Elsevier, Elsevier B.V., pp. 95–123 (Chapter 4).
Sabarez, H.T., 2012. Computational modeling of the transport phenomena occurring during convective drying of prunes. J. Food Eng. 111 (2), 279–288.
Sabarez, H.T., 2007. Tomato dehydration. In: Hui, Y.H., Clary, C., Farid, M.M., Fasina, O.O., Noomhorm, A., Welti-Chenis, J. (Eds.), Food Drying Science and Technology:
Microbiology, Chemistry, Applications. DESTech Publications, Inc., Pennsylvania, USA, pp. 603–628 (Chapter 26).
Sabarez, H.T., Swiergon, P., Knoerzer, K., 2019. Ultrasound assisted atmospheric freeze drying of food materials. In: Proceedings of the 3rd Nordic Baltic Drying Conference (NBDC
2019), Saint Petersburg, Russia.
Sabarez, H.T., Keuhbauch, S., Knoerzer, K., 2018. Ultrasound assisted low temperature drying of food materials. In: Proceedings of the 21st International Drying Symposium (IDS
2018), Valencia, Spain, pp. 1245–1250. https://doi.org/10.4995/ids2018.2018.7329.
Sabarez, H.T., Gallego-Juarez, J.A., Riera, E., 2012. Ultrasonic-assisted convective drying of apple slices. Dry. Technol.: Int. J. 30 (9), 989–997.
Sabarez, H.T., Chessari, C., 2006. High Quality Fruit and Vegetable Ingredients for Functional Foods: Modeling and Optimisation of Refractance Window Drying Technology.
Unpublished report prepared by CSIRO for Nutradry Pty Ltd, Brisbane, Australia.
Sablani, S.S., Rahman, M.S., 2007. Fundamentals of food dehydration (Chapter 1). In: Hui, Y.H., Clary, C., Farid, M.M., Fasina, O.O., Noomhorm, A., Welti-Chenis, J. (Eds.), Food
Drying Science and Technology: Microbiology, Chemistry, Applications. DESTech Publications, Inc., Pennsylvania, USA, pp. 1–42.
Sangamithra, A., Sivakumar, V., Swamy, G.J., Kannan, K., 2015a. Foam mat drying of food materials: a review. J. Food Process. Preserv. 39, 3165–3174.
Sangamithra, A., Sivakumar, V., Kannan, K., Swamy, G.J., 2015b. Foam-mat drying of Muskmelon. Int. J. Food Eng. 11 (1), 127–137.
Advanced Drying Technologies of Relevance in the Food Industry 81

Sankat, C.K., Castaigne, F., 2004. Foaming and drying behaviour of ripe bananas. LWT – Food Sci. Technol. 37 (5), 517–525. https://doi.org/10.1016/S0023-6438(03)00132-4.
Santacatalina, J.V., Fissore, D., Carcel, J.A., Mullet, A., Garcia-Perez, J.V., 2015. Model-based investigation into atmospheric freeze drying assisted by power ultrasound. J. Food
Eng. 151, 7–15.
Shi, Q.L., Xue, C.H., Zhao, Y., Li, Z.J., Wang, X.Y., 2008. Drying characteristics of horse mackerel (Trachurus japonicus) dried in a heat pump dehumidifier. J. Food Eng. 84 (1),
12–20.
Soria, A.C., Villamiel, M., 2010. Effect of ultrasound of the technological properties and bioactivity of food: a review. Trends Food Sci. Technol. 21, 323–331.
Schossler, K., Jager, H., Knorr, D., 2012. Effect of continuous and intermittent ultrasound on drying time and effective diffusivity during convective drying of apple and red bell
pepper. J. Food Eng. 108, 103–110.
Taghian Dinani, S., Hamdami, N., Shahedi, M., Havet, M., 2015a. Quality assessment of mushroom slices dried by hot air combined with an electrohydrodynamic (EHD) drying
system. Food Bioprod. Process. 94, 572–580.
Taghian Dinani, S., Hamdami, N., Shahedi, M., Havet, M., Queveau, D., 2015b. Influence of the electrohydrodynamic process on the properties of dried button mushroom slices:
a differential scanning calorimetry (DSC) study. Food Bioprod. Process. 95, 83–95.
Taghian Dinani, S., Havet, M., Hamdami, N., Shahedi, M., 2014. Drying of mushroom slices using hot air combined with an electrohydrodynamic (EHD) drying system. Dry. Technol.
32 (5) https://doi.org/10.1080/07373937.2013.851086.
Thuwapanichayanan, R., Prachayawarakorn, S., Soponronnarit, S., 2012. Effects of foaming agents and foam density on drying characteristics and textural property of banana
foams. LWT - Food Sci. Technol. 47 (2), 348–357. https://doi.org/10.1016/j.lwt.2012.01.030.
Thuwapanichayanan, R., Prachayawarakorn, S., Soponronnarit, S., 2008. Drying characteristics and quality of banana foam mat. J. Food Eng. 86 (4), 573–583. https://doi.org/
10.1016/j.jfoodeng.2007.11.008.
Topuz, A., Dincer, C., Özdemir, K., Feng, H., Kushad, M., 2011. Influence of different drying methods on carotenoids and capsaicinoids of paprika (Cv., Jalapeno). Food Chem. 129,
860–865.
Topuz, A., Feng, H., Kushad, M., 2009. The effect of drying method and storage on color characteristics of paprika. LWT - Food Sci. Technol. 42, 1667–1673.
Valenzuela, C., Aguilera, J.M., 2013. Aerated apple leathers: effect of microstructure on drying and mechanical properties. Dry. Technol. 31 (16), 1951–1959. https://doi.org/
10.1080/07373937.2013.803979.
Vernon-Carter, E.J., Espinosa-Paredes, G., Beristain, C.I., Romero-Tehuitzil, H., 2001. Effect of foaming agents on the stability, rheological properties, drying kinetics and flavor
retention of tamarind foam-mats. Food Res. Int. 34 (7), 587–598. https://doi.org/10.1016/S0963-9969(01)00076-X.
Walsh, D.J., Russell, K., Fitzgerald, R.J., 2008. Stabilisation of sodium caseinate hydrolysate foams. Food Res. Int. 41 (1), 43–52. https://doi.org/10.1016/j.foodres.2007.09.003.
Widyastuti, T.E.W., Srianta, I., 2011. Development of functional drink based on foam-mat dried papaya (Carica papaya L.): optimization of foam-mat drying process and its
formulation. Int. J. Food Nutr. Public Health 4, 167–176.
Wolff, E., Gilbert, H., 1990. Atmospheric freeze drying Part 1: design, experimental investigation and energy-saving advantages. Dry. Technol. 8, 385–404.
Yang, M., Ding, C., 2016. Electrohydrodynamic (EHD) drying of the Chinese wolfberry fruits. Springerplus 5 (1), 909–929.
Zhang, M., Chen, H., Mujumdar, A.S., Tang, J., Miao, S., Wang, Y., 2017. Recent developments in high-quality drying of vegetables, fruits and aquatic products. Crit. Rev. Food Sci.
Nutr. 57 (6), 1239–1255.
Zheng, X., Wang, Y., Liu, C., Sun, J., Liu, B., Zhang, B., Lin, Z., Sun, Y., Liu, H., 2013. Microwave energy absorption behavior of foamed berry puree under microwave drying
conditions. Dry. Technol. 31 (7), 785–794. https://doi.org/10.1080/07373937.2012.761635.
Zheng, X.Z., Liu, C.H., Zhou, H., 2011a. Optimization of parameters for microwave assisted foam-mat drying of blackcurrant pulp. Dry. Technol. 29 (2), 230–238. https://doi.org/
10.1080/07373937.2010.484112.
Zheng, D., Zheng, D., Cheng, Y., Liu, H., Li, L., 2011b. Investigation of EHD-enhanced water evaporation and a novel empirical model. Int. J. Food Eng. 7 (2), 11.
Zheng, X.Z., Liu, C.H., Zhou, H., 2009. Drying characteristics of blackcurrant pulp by microwave-assisted foam mat drying. Trans. CSAE 25 (8), 288–293.

You might also like