Defraeye 2018

Download as pdf or txt
Download as pdf or txt
You are on page 1of 47

Accepted Manuscript

Electrohydrodynamic drying of food: New insights from conjugate modeling

Thijs Defraeye, A. Martynenko

PII: S0959-6526(18)31909-7
DOI: 10.1016/j.jclepro.2018.06.250
Reference: JCLP 13396

To appear in: Journal of Cleaner Production

Received Date: 10 November 2017


Revised Date: 13 June 2018
Accepted Date: 24 June 2018

Please cite this article as: Defraeye T, Martynenko A, Electrohydrodynamic drying of food: New insights
from conjugate modeling, Journal of Cleaner Production (2018), doi: 10.1016/j.jclepro.2018.06.250.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
Electrohydrodynamic drying of food: new insights from conjugate modeling
ACCEPTED MANUSCRIPT

Thijs Defraeye a,b, A. Martynenko3


a
Empa, Swiss Federal Laboratories for Materials Science and Technology, Laboratory for Biomimetic Membranes and
Textiles, Lerchenfeldstrasse 5, 9014 St. Gallen, Switzerland

b
Empa, Swiss Federal Laboratories for Materials Science and Technology, Multiscale Studies in Building Physics,
Überlandstrasse 129, 8600 Dübendorf, Switzerland

PT
3
Department of Engineering, Dalhousie University, Agricultural Campus PO Box 550 Truro, Canada

RI
Abstract

Electrohydrodynamic (EHD) drying is a non-thermal technology with promising perspectives to dehydrate heat-

SC
sensitive materials such as foods. With EHD drying, corona discharge is used to generate airflow, which enhances

U
convective drying of the food product. Further development and upscaling of this technology are hindered by a lack
AN
of insight both the airflow and the dehydration process inside the material during drying. This study is the first to

develop a conjugate continuum model which couples EHD-generated airflow directly to the convective heat transfer
M

and moisture removal from the food. For the wire-to-plate configuration with impinging flow, the impact of different

geometrical and operational parameters (wire radius, distance to collector electrode, emitter voltage) on the EHD-
D

driven airflow and resulting drying kinetics is quantified. The fruit drying time is found to increase linearly with
TE

increasing distance between electrodes or increasing emitter electrode radius, but decreases in a non-linear way

with increasing voltage between the electrodes. For other emitter-collector configurations, where airflow passes
EP

around the fruit, such as wire-to-mesh, wire-to-plate, wire-to-wires and wire-to-parallel plates, significant
C

differences in drying kinetics were quantified. Such configurations provide better perspectives towards upscaling to
AC

dry large amounts of products uniformly. Of all tested configurations, the wire-to-mesh configuration provided the

highest drying rate. Placing the fruit on a mesh also showed to be advantageous since the fruit can be dried more

uniformly. The developed conjugate approach has the distinct advantage that the impact of EHD process parameters

and geometrical configurations on the drying rate can be quantified very swiftly. Such modeling approach is thereby

a valuable tool for further optimization of EHD drying technology towards industrial implementation.

1
Keywords: Electrohydrodynamic, computational fluid dynamics, dehydration, airflow, corona discharge, COMSOL
ACCEPTED MANUSCRIPT
Corresponding author

• E-mail [email protected]

• Tel. +41 (0)58 765 4790

• Fax. +41 44 633 10 41

PT
RI
U SC
AN
M
D
TE
C EP
AC

2
1. Introduction
ACCEPTED MANUSCRIPT
Electrohydrodynamic drying is a promising technology to dry foods convectively without using additional heat (Bajgai

et al., 2006; Kudra and Martynenko, 2015; Martynenko and Kudra, 2016; Singh et al., 2012; Zhang et al., 2017).

Electrohydrodynamic (EHD) airflow is generated by producing a high voltage difference between an emitter

electrode, for example a thin wire or needle, and a grounded collector electrode, such as a plate (Figure 1). Due to

the high voltage and the large curvature of the emitter, the air around it is ionized locally and corona discharge is

PT
produced. The ions are accelerated by the Coulomb force and move towards the grounded collector electrode. The

RI
drift of the ions to the collector and their subsequent collisions with neutral air molecules induce corona (ionic) wind

(≈ 10-1-101 m s-1). This airflow enhances moisture removal from the wet product by increasing convective mass

SC
transfer rates. However, also other mechanisms than the additional convection have been argued to increase the

drying rate (Bajgai et al., 2006; Singh et al., 2012; Zhang et al., 2017). Since for drying processes, fluid flow is

U
produced in air – not water – one could argue to call the process electro-aerodynamic (EAD) drying (Defraeye and
AN
Martynenko, 2018).
M

EHD drying is a non-thermal technology. As such, it is particularly suitable to dehydrate heat-sensitive biomaterials

(Singh et al., 2012), which are often challenging to dry appropriately (Bando et al., 2017; Bantle and Eikevik, 2014;
D

Horttanainen et al., 2017; Prosapio et al., 2017; Sahoo et al., 2017). Compared to mechanically-generated airflow
TE

(e.g. a fan), EHD drying does not induce vibrations, has a shorter response time and requires limited power to

generate corona discharge. There is also more flexibility in the size and geometry of the system, as airflow is
EP

produced locally, i.e. below each needle, instead of centrally, as with an axial fan. These advantages make EHD a

promising alternative drying technology. It has been used to dry food products, amongst others apple (Martynenko
C

and Zheng, 2016), mushrooms (Taghian Dinani and Havet, 2015a, 2015b), banana (Pirnazari et al., 2016), mango
AC

(Bardy et al., 2016), carrot (Ding et al., 2015) or wheat (Singh et al., 2015). Working prototypes of EHD dryers have

been built (Lai, 2010), but no commercial EHD dryers are available yet, although the technology has been known for

several decades. Electrohydrodynamics have also been used in other fields, for example for ionic pumps (solid-state

fans, plasma actuators), electrostatic precipitators, propulsion engines, cooling devices, active flow control,

electrospinning, printing, solar chimneys and building technology (Adamiak, 2013; Fylladitakis et al., 2014;

Ghalamchi et al., 2017).

3
ACCEPTED MANUSCRIPT

PT
RI
U SC
Figure 1. Schematic illustration of EHD airflow generation process for positive corona discharge (not to scale).
AN
The majority of the work on EHD drying has been experimental (Defraeye and Martynenko, 2018). Most of the

studies report changes of the bulk moisture content change as a function of time for different applied voltages or
M

corona currents. Product temperatures or (bulk) airflow fields are rarely measured. There are advanced non-
D

destructive techniques available to monitor in real-time the internal moisture content distribution, such as neutron
TE

imaging (Defraeye et al., 2016). However, these have not been used for EHD drying. Advanced laser diagnostics, such

as particle image velocimetry, have been used to study EHD-generated airflow fields (Kocik et al., 2009; Podlinski et
EP

al., 2013). This technique has not been applied for EHD drying studies yet and also does not provide information

about moisture or temperature fields in the air. Numerical modeling of the airflow together with the convective
C

drying process using a continuum approach is an alternative to get such information (Caccavale et al., 2016; Curcio et
AC

al., 2016; Defraeye, 2014; Defraeye et al., 2012; Defraeye and Radu, 2017; Halder and Datta, 2012). Such a conjugate

modeling strategy provides high spatial and temporal resolution, which enables not only to track the internal

moisture distribution, temperature and water activity throughout the entire drying process, but also the heat and

moisture transport in the airflow (Curcio et al., 2016; Halder and Datta, 2012; Kurnia et al., 2013; Lamnatou et al.,

2010, 2009; Marra et al., 2010)). Both surface-averaged and volume-averaged quantities can be extracted, such as

the overall moisture loss. Furthermore, extensive parametric evaluations can be performed in relatively short

4
timeframes, without being inherently hindered by experimental or biological uncertainty. Despite these advantages,
ACCEPTED MANUSCRIPT
no simulation-based method was developed which explicitly combines EHD-generated flow with drying of fresh

foods, such as fruits or vegetables. Several studies however focused on modeling EHD flow (Ayuttaya et al., 2012;

Fylladitakis et al., 2014; Ghazanchaei et al., 2015; Ould Ahmedou and Havet, 2009; Oussalah and Zebboudj, 2006).

Most studies in EHD drying are focused on evaluation of rather simple wire/needle-to-plate configurations with

airflow that impinges onto the food placed on the collector electrode plate (Ding et al., 2015; Esehaghbeygi and

PT
Basiry, 2011; Singh et al., 2015; Taghian Dinani and Havet, 2015a; Yang and Ding, 2016) . This might not be the

optimal configuration to maximize moisture removal when drying multiple products (Defraeye and Martynenko,

RI
2018), as saturation of the air with vapor can occur, as air passes over successive products. This leads to a reduced

SC
drying rate of products more downstream. In addition, when multiple wires/needles are considered, multiple

streams of air bounce back from the food and interfere with each other. Also here, saturation of the recirculating air

U
with vapor can occur, which is detrimental for the drying rate. To enable industrial upscaling, or at least verify its
AN
feasibility, configurations need to be designed that are able to dry large amounts of products uniformly. Moreover, a

more detailed knowledge of the EHD airflow generation and especially the coupling to the resulting convective
M

drying process in the product are essential.


D

In this paper, a 2D numerical continuum model is developed that couples EHD-driven airflow to a dehydration model
TE

for fruit. The base case is a wire-to-plate configuration (Figure 2b), where a single fruit is placed on a long plate. The

impact of different geometrical parameters (wire radius, distance to collector electrode) and operational parameters
EP

(emitter voltage) on the EHD-driven airflow and resulting drying kinetics is evaluated. In addition, the impact of other

emitter-collector configurations on the drying kinetics is evaluated. In these configurations, including wire-to-mesh,
C

wire-to-wires and wire-to-parallel plates (Figure 2c), the air flows around the fruit. The aim is to identify which
AC

configuration maximizes the fruit drying rate and the dehydration uniformity of the products. To our knowledge, this

is the first model that links ionic wind directly to the resulting convective drying process of a food product.

5
2. Materials and methods
ACCEPTED MANUSCRIPT

2.1 Computational model

A two-dimensional model is made to calculate drying of a single food product, representing a long, rectangular fruit

slice. EHD airflow is generated for a wire-to-plate configuration. The computational model includes a cylindrical wire

(with radius rw), i.e. the emitter electrode, onto which a positive high voltage (Vw) is applied. A grounded collector

electrode is placed at a distance dec from the emitter electrode. Ions are assumed to be generated at the emitter

PT
electrode. They are accelerated by the Coulomb force in the drift region and are thereby transported to the collector

RI
electrode. Under this electrostatic action, airflow is generated, which draws dry air at a temperature Tref of 20 °C and

a relative humidity RHref of 30% from the inlet towards the fruit to be dried. These conditions are typical for

SC
convective drying of fruit in the ambient environment. Such EHD-driven convective drying, without additional

application of heating, is considered a non-thermal technology. The fruit that is dried convectively is a rectangular

U
piece of apple fruit (L x H = 10 x 5 mm). Apple fruit is chosen as it is a frequently applied model material to study fruit
AN
drying and since it has a high commercial value.
M

Types of configurations
D

Three types of configurations are simulated, which are depicted in Figure 2. The specifications of the simulations are
TE

given in Table 1. Note that the range of EHD process parameters (Vw, dec, rw) is chosen based on typically used values

in previous EHD drying studies (Defraeye and Martynenko, 2018).


C EP
AC

6
(a) Type 1: Impinging flow configuration without food product
ACCEPTED MANUSCRIPT
wire-to-plate
Emitter electrode (Vw, rw)

dec Collector electrode

(b) Type 2: Impinging flow configuration with food product

PT
wire-to-plate

RI
Emitter electrode (Vw, rw)
Product to be dried

SC
dec
Collector electrode

U
(c) Type 3: Configurations with airflow around the food product
wire-to-plate wire-to-mesh wire-to-wire wire-to-parallel
AN
Emitter electrode (Vw, rw) plates
M
Symmetry conditions

dec
D
TE

Collector
electrode
Product to be dried
EP

Figure 2. Types of configurations: (a) EHD impinging flow for wire-to-plate configuration (without food product),

(b) EHD drying of a food product for impinging flow for wire-to-plate configuration and (c) EHD drying of a food
C

product for airflow around the food product for various configurations (emitter = red, collector = black; not to
AC

scale; legend is the same in each graph; symmetry conditions indicate that multiple products are placed sideways).

As a first type, a wire-to-plate configuration is evaluated for a long solid plate without any fruit being placed on the

plate. As such, flow impinges on the plate and is directed sideward to the outlet. This reference case used to get

insight in the impact of different process parameters (radius of the wire rw, distance between emitter and collector

dec and voltage applied at the wire Vw) and of modeling assumptions on the EHD-driven airflow. For this case, the

results are also verified against analytical solutions and experimental data.
7
As a second type, a wire-to-plate configuration is evaluated for a long plate, but now a fruit is placed on the plate
ACCEPTED MANUSCRIPT
below the wire. The impact of different process parameters (rw, dec and Vw) on the EHD-driven airflow but also on the

fruit drying kinetics is quantified. This impinging flow configuration is very relevant as most studies in EHD drying

looked at similar wire-to-plate or wire-to-needle configurations.

As a third type, configurations with other emitter-to-collector geometries are evaluated. In these cases however,

airflow can pass around the fruit. As the moist air is not directed upwards or sideways towards other products, such

PT
configurations are considered more optimal for drying multiple products in a uniform way compared to placing them

on a large plate. The evaluated configurations have been sourced from EHD flow research and include: wire-to-short

RI
plate, wire-to-mesh, wire-to-parallel plates and wire-to-wires. These configurations have not been extensively tested

SC
for EHD drying, but could provide enhanced drying kinetics, compared to the impinging flow configuration. For the

wire-to-short plate configuration, the fruit is supported by the collector electrode, which is a short metal plate. For

U
all other configurations, the fruit is assumed to lie on a very open-porous wire mesh (with a negligible impact on
AN
airflow). In the wire-to-mesh configuration, a metallic (grounded) mesh is assumed, but in wire-to-wires and wire-to-

parallel plates configurations the mesh is made out of a dielectric material (e.g. plastic). As such, for most Type 3
M

configurations, the fruit can dry from the bottom surface as well through the mesh.
D

Table 1. Simulation conditions for different types of configurations. The base case values are indicated in bold. If
TE

no value is specified (indicated by a dash), it is taken the same as the base case. The parameter ‘x’ in the

nomenclature is variable.
EP

Description Name Wire radius Distance emitter to Voltage at wire

rw [µm] collector electrode (emitter)


C

dec [mm] Vw [kV]


AC

Type 1: Wire-to-plate (without fruit) for impinging flow (W2P-IF-

NoFruit)

Base case W2P-IF-NoFruit-Base 250 20 20

Impact of wire radius W2P-IF-NoFruit-rw= x 50, 100, 200, 250, - -

µm 300, 400, 500

Impact of distance emitter-collector W2P-IF-NoFruit-dec= x - 5, 10, 20, 40, 80 -

mm

Impact of voltage W2P-IF-NoFruit-Vw= x - - 12, 15, 20, 25,

8
kV 30, 40
ACCEPTED MANUSCRIPT
Type 2: Wire-to-plate (with fruit) for impinging flow (W2P-IF)

Base case W2P-IF-Base 250 20 20

Impact of wire radius W2P-IF-rw= x µm 50, 100, 200, 250, - -

300, 400, 500

Impact of distance emitter-collector W2P-IF-dec= x mm - 10, 20, 40, 80 -

Impact of voltage W2P-IF-Vw= x kV - - 12, 15, 20, 25,

30, 40

PT
Type 3: Various configurations for air flow around the fruit (FAF)

Wire-to-short plate W2SP-FAF - - -

Wire-to-mesh W2Mesh-FAF - - -

RI
Wire-to-parallel plates W2Plates-FAF - - -

Wire-to-wires W2Wires-FAF - - -

U SC
AN
M
D
TE
C EP
AC

9
Computational model
ACCEPTED MANUSCRIPT
The computational models for these configurations are specified in Figure 3. The dimensions of the computational

domains are taken sufficiently large to avoid an influence from the boundaries on the potential, space charge and

airflow fields in the vicinity of the wire (emitter) as well as on the drying process of the fruit. The ionization layer at

the vicinity of the wire is not explicitly included as a separate region in the domain, as detailed in section 2.2.1. To be

able to compare the efficacy of the different geometrical configurations (Figure 2), similar electrostatic conditions

PT
should be applied. This is not that straightforward as all configurations (Type 3) have different collector geometry. To

this end, the applied voltage at the emitter was taken the same for all configurations (Type 3, Table 1) and the

RI
distance between emitter and collector dec was chosen in such a way to achieve the same bulk electric field strength

SC
Ebulk [V m-1] in each configuration. This parameter is used in several studies and is defined as:

Vw
Ebulk =

U
(1)
d ec
AN
As the voltage is taken the same for all Type 3 configurations, this means that also the distance between the
M

electrodes needs to be the same in order to achieve the same Ebulk. Appropriate grids were built for the air and fruit

domains, based on a grid sensitivity analysis. The grid of the base case for type 2 consists of 191 298 tetrahedral and
D

quadrilateral finite elements for example (25 574 elements for the fruit domain and 165 724 elements for the air
TE

domain). A grid refinement towards the electrode boundaries and the air-tissue interface was applied to enhance

numerical accuracy and stability, as the largest gradients occur there.


EP

(a) Impinging flow configuration 30L (b) Configuration for airflow around the fruit product 30L

Air domain Inlet Air domain Inlet


C

- Electrical potential zero static pressure - Electrical potential zero static pressure
- Space charge density (SCD) Tref = 20°C - Space charge density (SCD) Tref = 20°C
- Turbulent airflow RHref = 30% - Turbulent airflow RHref = 30%
- Heat transfer (to extract convective heat - Heat transfer (to extract convective heat
AC

transfer coefficient) transfer coefficient)


Symmetry
zero static pressure

Fruit domain Fruit domain


30L

- Electrical potential - Electrical potential


- Heat and mass transfer (drying process) - Heat and mass transfer (drying process)
Outlet

Emitter electrode Emitter electrode


V = Vw V = Vw
ρe = ρe,w ρe = ρe,w
Collector electrode
θ Grounded (Vc = 0) θ
y dec y dec
No-slip wall L Airflow ρe,w = 0 No-slip wall L Airflow
H H
60L

x x0 x

Collector electrode
Outlet Grounded (Vc = 0)
zero static pressure ρe,w = 0

10
Figure 3. Computational models for EHD drying for impinging (a) and air flow around the fruit (b) configurations, as illustrated
ACCEPTED MANUSCRIPT
for the wire-to-plate configuration (not to scale).

PT
RI
U SC
AN
M
D
TE
C EP
AC

11
2.2 Numerical modeling
modeling
ACCEPTED MANUSCRIPT
Continuum modeling of EHD drying requires that following phenomena are captured: (1) the electrostatic potential

field, (2) the associated charge transport due to ion drift, caused by the corona discharge under the high-voltage

field, (3) the resulting airflow generation due to ion movement and collision with neutral air molecules, (4) heat and

mass transport in the fruit tissue due to convective, EHD-generated, airflow, leading to dehydration. These individual

physics, the corresponding transport equations and the coupling between air and fruit domains are described below.

PT
In summary, the numerical procedure first solves for steady-state airflow, generated by the electrohydrodynamic

RI
effect. From this airflow field, the convective heat and mass transfer coefficients on the fruit surface are determined.

These are used as an input to solve the transient dehydration process of the fruit.

SC
All the used submodels for EHD, the associated airflow and drying process were validated previously by the authors

and others (Defraeye et al., 2013; Defraeye and Verboven, 2017; Oussalah and Zebboudj, 2006; Tirumala and Go,

U
2014). A verification of the EHD model with experimental and analytical data is presented below. Note that existing
AN
models are implemented, but the coupling between fruit and airflow was not yet performed for EHD drying. Also
M

note that the boundary and initial conditions are detailed in the Supplementary Material.
D
TE
C EP
AC

12
2.2
2.2.1 EHD airflow modeling
ACCEPTED MANUSCRIPT
Electrostatics and space charge transport modeling

The governing equations for EHD-driven flow are well established in several works (Ayuttaya et al., 2012; Fylladitakis

et al., 2014; Ghazanchaei et al., 2015; Martynenko and Kudra, 2016; Ould Ahmedou and Havet, 2009; Oussalah and

Zebboudj, 2006). The electrical potential V (in [V]) in air and the fruit is described by Poisson’s equation:

∇⋅ ( −ε0εr∇V ) = ρe

PT
(2)

where ρe is the space charge density [C m-3], ε0 is the dielectric permittivity of vacuum (8.854 x 10-12 C V-1 m-1), and εr

RI
is the relative permittivity of the material (1 for air and 54 for apple fruit, (Marra et al., 2010)), which is also called

SC
the dielectric constant. The electrical potential is linked to the electric field (intensity) E [V m-1] by:

E = −∇V (3)

U
AN
The resulting electric current in the drift zone (air) is caused by three phenomena: (1) ions that move from emitter to

collector electrode due to the presence of the electric field, where this drift is also called conduction; (2) transport of
M

charged particles due to airflow (advection), (3) diffusion of ions. As such, the electric current density J [C m-2s-1]

becomes:
D

J = −µe ρe∇V + ρeu − Di∇ρe (4)


TE

where µe is the ion mobility in air (1.8 x 10-4 m2 V-1s-1), Di is the diffusivity of the ions [m2 s-1] and u is the velocity
EP

vector for air [m s-1]. The conductive term is usually dominant and therefore other two terms are often neglected,

resulting in:
C

J = −µe ρe∇V = µe ρeE


AC

(5)

In this study, ion transport is assumed to only occur in the air so not in the fruit. In steady-state conditions, the

continuity equation for current density becomes:

∇⋅J = 0 (6)

13
Airflow modeling
ACCEPTED MANUSCRIPT
Conservation of mass, momentum and heat in the air, in stationary state, are modeled by solving the Navier-Stokes

equations. The impact of the electric field on the fluid flow is represented by including a volumetric source term Fe

[kg m-2 s-2] in the momentum equation:

ρ au ⋅ ∇u = −∇p + µa∇2u + Fe (7)

PT
where ρa is the air density (1.20 kg m-3 at 20°C), µa is the dynamic viscosity of air (1.81 x 10-5 kg m-1s-1 at 20°C). The

volumetric source term Fe, namely the Coulomb (electrophoretic) force, is:

RI
Fe = −ρe∇V (8)

SC
The Coulomb force quantifies the net momentum gain due to the charged particles (ions) that undergo momentum-

U
transfer collisions with neutral air molecules. Some studies also included the dielectrophoretic force, caused by
AN
spatial variations in dielectric properties of the air, and the electro-restrictive force due to inhomogeneity in the

electric field (Ayuttaya et al. 2012; Singh, Orsat, and Raghavan 2012). For single phase, isothermal fluids, these two
M

terms can be neglected (Dalvi-Isfahan et al., 2016; Heidarinejad and Babaei, 2015).
D

In order to calculate the convective scalar exchange with the fruit, the convective heat transfer in the air is solved

through the heat transfer equation:


TE

ρacp,au ⋅∇T = −∇⋅ ( −ka∇T ) (9)


EP

where cp,a is the specific heat capacity of air (1005 J kg-1K-1 at 20°C), ka is the thermal conductivity of the air (0.0258
C

W m-1K-1 at 20°C) and T is the air temperature [K]. Vapor transport in the air is not directly modeled but its influence
AC

on the drying process is accounted for by using the heat and mass transfer analogy, as detailed further in

Supplementary Material.

These equations are solved to obtain EHD-generated airflow and appropriate boundary conditions are specified on

the fruit surface (Supplementary Material). Note that, the potential field (Eq. (2)) is also solved inside the fruit. The

Navier-Stokes equations for turbulent flow are solved by applying the Reynolds-averaged Navier-Stokes (RANS)

approach in combination with the standard k-ε turbulence model. This turbulence model is still the most commonly

14
used model in computational fluid dynamics (CFD) engineering (Casey and Wintergerste, 2000). Wall functions are
ACCEPTED MANUSCRIPT
used to model transport in the boundary layer. The grid resolution in the boundary layer region was made

sufficiently dense to have very low values of the dimensionless wall distance (y+ value). The y+ values were typically

below 5 on the fruit surface. At these low y+ values, the boundary layer is actually fully resolved down to the viscous

sublayer, by which it corresponds to low-Reynolds number modeling (Defraeye et al., 2013, 2010). Buoyancy effects

are not taken into account in the simulations. Long-wave radiation exchange between fruit and the surrounding

PT
surfaces is also not included. To solve EHD generated airflow, additional relations are required to specify appropriate

boundary conditions for the corona discharge at the emitter electrode. These relations are given in the

RI
Supplementary Material.

U SC
AN
M
D
TE
C EP
AC

15
2.2
2.2.2 Dehydration modeling of fruit tissue
ACCEPTED MANUSCRIPT
To calculate heat and moisture transfer in the fruit tissue during drying, a previously developed model is used

(Defraeye and Verboven, 2017), so only the main characteristics are highlighted here. The main model assumptions

are that evaporation is assumed to occur only at the tissue surface and that shrinking and swelling of the tissue are

neglected as often assumed in multiphysics modeling of fruit drying (Defraeye, 2014).

Conservation equations

PT
The following conservation equations for moisture and energy are solved to the dependent variables temperature T

RI
[K] and water potential ψ [Pa]:

∂wm ∂ψ
+ ∇ ⋅ ( − K m∇ψ ) = 0

SC
(10)
∂ψ ∂t

U
∂wm ∂ψ ∂T
hl + ( c p ,s ws + c p ,l wm ) + ∇ ⋅ ( −hl K m∇ψ ) + ∇ ⋅ ( −λPM ∇T ) = 0 (11)
∂ψ ∂t ∂t
AN
where ws is the dry matter density (solid, 130 kg m-3) and wm is the moisture content of the tissue [kg m-3]. Km is the
M

moisture permeability of the tissue (8 x 10-16 s), hl is the enthalpy of liquid water [J kg-1], λPM is the thermal

conductivity of the tissue (porous medium, 0.418 W m-1 K-1), cp,s and cp,l are the specific heat capacities of dry matter
D

(1634 J kg-1 K-1) and liquid water (4182 J kg-1 K-1), respectively. For Km and the sorption isotherm, the values for the
TE

apple cultivar Braeburn are taken (Aregawi et al., 2013).


EP

Constitutive equations
C
AC

The enthalpies of liquid water and water vapor, hl and hv [J kg-1], are:

hl = c p ,l (T − Tref ,0 ) (12)

hv = c p ,v ( T − Tref ,0 ) + Lv (13)

where Lv is the heat of vaporization (2.5 x 106 J kg-1), also called latent heat, which is the energy needed for the phase

change from liquid to vapor. Tref,0 is a reference temperature, taken equal to 273.15 K (0°C) and cp,v is the specific

16
heat capacity of water vapor (1880 J kg-1K-1). To determine the moisture capacity Cm = ∂wm , the sorption isotherm
ACCEPTED MANUSCRIPT ∂ψ

(wm vs. water activity aw) is required, as well as the relation of the water activity aw to the water potential ψ. The

latter is given by:

ψ = ρl RvT ln ( aw ) (14)

where ρl is the density liquid water (1000 kg m-3) and Rv is the specific gas constant for water vapor (461.52 J kg-1K-1).

PT
The sorption isotherm equals:

RI
1
  0.97014
 
0.15926 

SC
wm ( aw ) = ws  (15)
  1.0177  
 ln  
  aw  

U
AN
M
D
TE
C EP
AC

17
ACCEPTED MANUSCRIPT
2.3
2.3 Numerical simulations

This model was implemented in COMSOL Multiphysics (version 5.2a), which is a finite-element based commercial

software. For numerical convergence and stability, different physics were solved sequentially. First, the electrical

potential and space charge density were solved together as a stationary problem. A parametric sweep for different

values of the space charge density at the emitter surface was performed. This was required to determine the correct

PT
SCD at which the breakdown electric field strength at the corona surface is reached, which leads to corona formation

(section 2.2.1). Linear shape functions were used together with a fully-coupled direct solver, relying on the MUMPS

RI
(MUltifrontal Massively Parallel sparse direct Solver) solver scheme. Afterwards, turbulent airflow was solved as a

SC
stationary problem, in order to extract the convective transfer coefficient distribution on the fruit surface. The

computational fluid dynamics (CFD) module of COMSOL was used, using a segregated solver, relying on the PARDISO

U
(PARallel DIrect sparse SOlver Interface) solver scheme. Finally, the drying process was solved as a transient problem.
AN
The partial differential equation interface (coefficient form) was used to calculate the hygrothermal transport in the

fruit. Quadratic shape functions were used together with a fully coupled direct solver, relying on the MUMPS solver
M

scheme.
D

The tolerances for convergence and other solver settings were determined based on sensitivity analysis in such a
TE

way that increasing the tolerance further did not alter the solution results anymore. After the steady airflow

calculation, the CTCs on the fruit surface were extracted and a drying process for 40 hours was simulated, starting
EP

from the specified initial conditions. These simulations applied adaptive time stepping, with a maximal time step of

60 s, as determined from a temporal sensitivity analysis.


C
AC

2.4
2.4 Drying performance
performance evaluation

The critical drying time (tcrit) is used to compare drying efficiency in a more straightforward quantitative way than

just based on drying curves(Defraeye and Verboven, 2017). It is the time needed for the sample to reach the critical

moisture content (wcrit). The latter is defined as the (volume-)averaged moisture content in the sample that

corresponds, via the sorption isotherm, to an equilibrium water activity below which no spoilage occurs (aw,crit). For

18
dried fruit, this aw,crit is about 0.6 (Bonazzi and Dumoulin, 2011). In principle, if the drying process is stopped at wcrit
ACCEPTED MANUSCRIPT
and the sample is equilibrated at the same or lower water activity, the average final water activity in the tissue will

be below aw,crit. For the present study, wcrit was 37.8 kg m-3, leading to a Xcrit of 0.29 kg kgdm-1.

With tcrit, only one characteristic value is obtained per drying curve, which facilitates quantifying differences between

drying kinetics of different fruits. It is a useful control parameter to decide when to stop a drying process. By

stopping the process at the right point in time, over-drying is avoided and energy can be saved.

PT
RI
U SC
AN
M
D
TE
C EP
AC

19
3. Results and discussion
ACCEPTED MANUSCRIPT

3.1 EHD airflow impinging on a plate (without


(without fruit)
fruit)

This section focusses on EHD-generated airflow impinging on a large plate without the presence of a fruit. The aim is

to (1) verify the obtained model for electrostatics and space charge transport, (2) evaluate the impact of several

modeling assumptions, and (3) provide insight in the impact of multiple EHD process parameters (Vw, dec, rw) on the

resulting airflow for the well-known wire-to-plate configuration without fruit.

PT
3.1.1
3.1.1 Verification of electrostatics

RI
A common verification of the EHD model is presented, namely a comparison of the current density J (Eq.(5)) right

SC
above the collector plate with experimental data (Oussalah and Zebboudj, 2006; Yuen, 2006) and analytical

predictions (Figure 4). The electric current density can be described analytically with Warburg’s law:

J = J0 cosm (θ )
U (16)
AN
where J0 is the electric current density below the emitter, θ is the incident angle (Figure 3) with tan(θ) = x0/dec, where
M

x0 is the distance along the plate and equals 0 in the center of the plate (below the emitter). The exponent m varies

usually between 4.5 and 5 (Dalvi-Isfahan et al., 2016), which are both depicted in Figure 4. The agreement of the
D

simulation data with the analytical equation and experimental results is satisfactory. The electrical current density (J)
TE

of the simulations lies within 0.05J0 of the experimental values of J.


C EP
AC

20
1
ACCEPTED
Analytical (Warburg's law) - m=4.5 MANUSCRIPT
0.9 Analytical (Warburg's law) - m=5
Yuen (2006)
0.8
Electrical current density J/J0 [-]

Oussalah and Zebboudj (2006)


0.7 Simulations

0.6
θ
0.5
dec
0.4

0.3

PT
x0
0.2

0.1

RI
0
0 0.5 1 1.5 2 2.5 3
Distance along plate x0/dec [-]

SC
Figure 4. Electrical current density on the plate surface (scaled to the value below the wire J0) as a function of the

U
distance along the collector electrode (x0, scaled to the distance of the emitter to the collector dec) for the
AN
simulations, analytical predictions using Warburg’s law and experimental data (Oussalah and Zebboudj, 2006;

Yuen, 2006).
M
D

3.1.2 Impact of modeling assumptions


TE

The impact of a few common modeling assumptions on EHD airflow is verified to give modelers insight in which

approach to choose when developing the model. First, a case is evaluated where the size of the ionization region is
EP

not neglected, but instead the ionization region is explicitly modelled in the following way: instead of taking the wire
C

as the physical boundary in the computational domain (so at rw) to prescribe the voltage and SCD on, the ionization
AC

region (at rc) is specified as the physical boundary. The wire is thereby not included in the model. The voltage and

size of the ionization region can be calculated from Eqs.(11-13), based on the voltage at the emitter. The resulting

radius of the ionization region for the base case rc is 737 µm and the applied voltage Vc (at rc) is 17530 V. As a second

case, breakdown at the wire is assumed to occur if the surface-average value of the electric field strength over the

emitter electrode reaches the breakdown electric field strength, instead of taking the maximal value of the electric

field strength at the emitter electrode as the breakdown criterion. Using the surface-average electric field strength

value will result in a different SCD at the wire in order to have breakdown, which also needs to be determined
21
iteratively. In Table 2, the resulting average air speed of the total incoming airflow in the domain (Uref) and the
ACCEPTED MANUSCRIPT
average convective heat transfer coefficient (CHTC) on the plate are shown for the different cases. The average

inflow air speed over all domain boundaries (Uref) is calculated based on the total mass flow entering the domain

over all domain boundaries, as some air can also enter the domain via the lateral boundaries (outlet in Figure 3).

Table 2. Impact of modeling assumptions on average air speed of the total incoming airflow and surface-averaged

convective heat transfer coefficient (CHTC) on collector surface.

PT
Air speed CHTC

RI
m s-1 Difference with base W m-2K-1 Difference with

case base case

SC
Base case (W2P-IF-NoFruit) 0.0906 - 19.6 -

Ionization region modelled 0.0950 4.9% 20.0 2.3%

Average electric field strength 0.0853 5.9%


U 19.8 0.9%
AN
on emitter used
M

It is clear that modeling the wire (and neglecting the ionization zone) versus modeling the ionization zone leads to
D

very similar results, given that the correct voltage Vc and size of the ionization zone rc are taken. A previous study
TE

(Tirumala and Go, 2014) mentions that neglecting the ionization zone is justified if its size is small compared to the

electrode gap. In the present study, this is the case (0.737 mm << 20 mm). The value of the electric field strength
EP

used to determine the SCD iteratively at the emitter electrode (either maximal or surface-averaged), only affects the
C

resulting airflow and heat transfer in a limited way, so both options can be considered.
AC

3.1.3
3.1.3 Impact
Impact of geometrical parameters and operating conditions on airflow and heat transfer

The impact of EHD process parameters, such as the applied voltage at the emitter (Vw), the distance between emitter

and collector (dec) and the wire radius (rw), on the airflow is quantified for the wire-to-plate configuration. In Figure

5a, the surface-averaged space charge density over the emitter electrode is shown as a function of Vw, dec and rw. In

22
Figure 5b and 5c, similar graphs are given for the average air speed of the total incoming airflow in the domain (Uref)
ACCEPTED MANUSCRIPT
and the surface-averaged convective heat transfer coefficient (CHTC) over the collector plate. Derived correlations

based on the data points are also shown. Note that the results with the fruit are also included for reasons of

comparison, but they are only discussed later. The Reynolds number, defined based on the average air speed at the

domain boundaries Uref (Figure 5) and the fruit length (Lf), ranges from 8-166 (for Vw), 23-136 (for dec) and 38-78 (for

rw). Although these values suggest laminar convective flow, the local air speeds are much higher, as will be discussed

PT
below, and airflow can become actually turbulent.

RI
U SC
AN
M
D
TE
C EP
AC

23
(a) Space charge density
W2P-IF-NoFruit ACCEPTED MANUSCRIPT
W2P-IF
7.0E-03 4.0E-03
Poly. (W2P-IF-NoFruit)
Poly. (W2P-IF) 1.0E-02 W2P-IF-NoFruit W2P-IF-NoFruit

6.0E-03 y= 6.14E-12x2- 9.96E-08x + 3.26E-04 W2P-IF 3.5E-03 W2P-IF


R² = 1.00E+00 Power (W2P-IF-NoFruit) Expon. (W2P-IF-NoFruit)
y = 4.06E-12x2 - 6.59E-08x + 2.06E-04 Power (W2P-IF) 3.0E-03 Expon. (W2P-IF)
5.0E-03 R² = 1.00E+00 1.0E-03

y = 1.26E-09x-3.50E+00 2.5E-03

SCD [C/m3]
SCD [C/m3]

SCD [C/m3]
4.0E-03 R² = 9.92E-01
y = 3.66E-03e-5.66E-03x
2.0E-03
1.0E-04 R² = 9.75E-01
3.0E-03
1.5E-03
y = 1.20E-08x-2.71E+00 y = 2.61E-03e-6.09E-03x
2.0E-03 R² = 9.99E-01 R² = 9.82E-01
1.0E-05 1.0E-03

1.0E-03 5.0E-04

PT
0.0E+00 1.0E-06 0.0E+00
0 10000 20000 30000 40000 50000 0 0.02 0.04 0.06 0.08 0.1 0 200 400 600
Voltage Vw [V] Distance dec [m] Radius rw [µm]

RI
(b) Air speed
y = -5.83E-11x2 + 1.21E-05x - 1.16E-01
0.3 R² = 1.00E+00 0.25 0.16
y = -3.66E-11x2 + 1.06E-05x - 1.09E-01 W2P-IF-NoFruit W2P-IF-NoFruit

SC
R² = 1.00E+00 W2P-IF 0.14 W2P-IF
0.25 Power (W2P-IF-NoFruit)
0.2 Poly. (W2P-IF-NoFruit)
Power (W2P-IF) 0.12 Poly. (W2P-IF)
0.2
Air speed [m/s]

Air speed [m/s]

Air speed [m/s]


0.15 y = 7.54E-03x-6.28E-01 0.1

U
R² = 9.96E-01
0.15 0.08
y = 4.85E-03x-7.89E-01
0.1 R² = 9.99E-01 0.06
AN
0.1
y = 7.55E-08x2 - 1.67E-04x + 1.40E-01
0.04 R² = 9.95E-01
W2P-IF-NoFruit 0.05
0.05 W2P-IF
0.02 y = 5.47E-08x2 - 1.68E-04x + 1.30E-01
Poly. (W2P-IF-NoFruit) R² = 9.95E-01
Poly. (W2P-IF)
M

0 0 0
0 10000 20000 30000 40000 50000 0 0.02 0.04 0.06 0.08 0.1 0 100 200 300 400 500 600
Voltage Vw [V] Distance dec [m] Radius rw [µm]

(c) Convective heat transfer coefficient


D

60 70 30
y = -2.23E-08x2 + 2.97E-03x - 3.07E+01
W2P-IF-NoFruit W2P-IF-NoFruit
TE

R² = 9.97E-01
60 Power (W2P-IF-NoFruit)
50 W2P-IF-NoFruit 25 Poly. (W2P-IF-NoFruit)

Poly. (W2P-IF-NoFruit)
50
CHTC Plate [W/m2K]

20
CHTC Plate [W/m2K]
CHTC Plate [W/m2K]

40

40
EP

30 15
30
y= 6.53E-01x-8.51E-01
y = 3.10E-05x2 - 5.17E-02x + 3.05E+01
R² = 9.94E-01
20 10 R² = 9.93E-01
20
C

10 5
10
AC

0 0 0
0 10000 20000 30000 40000 50000 0 0.02 0.04 0.06 0.08 0.1 0 100 200 300 400 500 600
Voltage Vw [V] Distance dec [m] Radius rw [µm]

Figure 5. (a) Surface-averaged space charge density (SCD) over the emitter (wire) surface; (b) average inflow air

speed over all domain boundaries (Uref); (c) Surface-averaged CHTC on the collector electrode plate. These

quantities are all represented as a function of the voltage at the emitter (Vw), the distance between emitter and

collector electrode (dec) and the wire radius (rw) for the wire-to-plate impinging-flow (W2P-IF) configuration.

24
Results without (empty dots) and with (filled dots) fruit in the computational model are shown. The base case
ACCEPTED MANUSCRIPT
parameters are Vw = 20 kV, dec = 20 mm, rw = 250 µm.

Figure 5a shows that the space charge density (SCD) clearly has a nonlinear dependency on Vw, dec and rw. This SCD at

the emitter that is required to initiate corona discharge increases quadratically with increasing Vw. A power-law

correlation approximates the relation of SCD with dec and rw quite well. The SCD is found to vary several orders of

magnitude over the range of dec that is evaluated. For rw, the variation of the SCD is relatively small.

PT
Figure 5b shows that the air speed increases with increasing voltage at the emitter. A quadratic relationship fits

RI
slightly better than a linear one. Linear and quadratic relations between air speed and emitter voltage were also

found in other studies (Kocik et al., 2009; Zhang et al., 2015). A drastic increase in air speed occurs when the

SC
electrode gap is reduced, for which a power-law correlation was found as a good approximation. Changing the wire

diameter, for example from 500 µm to 50 µm, only increases the air speed with a factor of about 2. In general,

U
higher air speeds are achieved at higher voltages, smaller emitter-collector distances and thinner wires. Such
AN
numerical quantification of the relation between corona voltage and the resulting air speed serves as a promising

alternative for a fast identification of optimal emitter-collector geometrical configurations.


M

Figure 5c shows that the CHTC on the collector plate surface has a very similar relation with Vw, dec and rw as the air
D

speed. This is due to the fact that the CHTC for a flat plate is typically related to the air speed with a power-law
TE

relation: CHTC = a.Ub. Despite the rather low average air speeds (Uref) and Reynolds numbers, which suggest laminar

flow, the CHTCs are quite high which indicate turbulent flow. The reason is that the local air speeds close to the
EP

collector are much higher than the average values over the domain boundaries, namely ~ 2 m s-1 versus ~ 0.2 m s-1,
C

respectively for the base case. If Reynolds numbers would be based on the maximal air speed, they would be at least
AC

10 times higher, which can induce turbulent flow. As such, the impinging flow configuration is very successful for

enhancing local scalar (heat or mass) transfer. In summary, to achieve the highest air speed or CHTC, the voltage

should be maximized, the electrode gap should be reduced and a thin wire (thus large curvature) should be used.

The validity of using the electric field strength (Ebulk = Vw/dec) as a simple parameter to characterize the EHD-driven

airflow and convective transfer is evaluated. The reason is that Ebulk is often used to characterize and compare

different EHD airflow studies (Defraeye and Martynenko, 2018; Esehaghbeygi and Basiry, 2011). To this end, the

25
corresponding air speed and CHTC data of Figure 5b-c are presented as a function of the bulk electric field strength
ACCEPTED MANUSCRIPT
in Figure 6. Despite Ebulk being used often in the past, Figure 6 demonstrates that there is no correlation between the

Vw, dec and rw datasets. As such, for a particular Ebulk, completely different air speeds and CHTCs are obtained,

depending on the specific values of Vw, dec and rw. This lack of correlation indicates that Ebulk is not a suitable variable

to evaluate and compare different configurations.

(a) Air speed

PT
0.3

RI
0.25

SC
0.2
Air speed [m/s]

0.15

U
AN
0.1

Distance
0.05
Voltage
M

Radius
0
0 10 20 30 40 50
D

Bulk electric field strength Ebulk [kV/cm]

(b) CHTC
TE

60
EP

50
CHTC Plate [W/m2K]

40
C
AC

30

20

Distance
10
Voltage
Radius
0
0 10 20 30 40 50
Bulk electric field strength Ebulk [kV/cm]

26
Figure 6. Average inflow air speed over all domain boundaries (Uref) and surface-averaged CHTC on the collector
ACCEPTED MANUSCRIPT
electrode plate as a function of the bulk electric field strength (Ebulk) for the wire-to-plate impinging-flow (W2P-IF)

configuration (without fruit). Results for different emitter voltages Vw (empty dots), distances between the

electrodes dec (filled dots) and wire radii (empty squares) are shown. The base case parameters are Vw = 20 kV, dec

= 20 mm, rw = 250 µm.

Finally, the ion current between the electrodes Ic [A] was estimated as follows (Jewell-Larsen et al., 2008):

PT
Ic = JwAw (17)

RI
where Jw is the surface-averaged ion current density at the emitter [C m-2 s-1] (Eq.(5)) and Aw [m2] is the outer surface

SC
area of the wire, which equals 2πrwLw, where Lw is the length of the wire. The average ion current at the emitter is

shown in Figure 7 as a function of the voltage at the emitter, assuming a wire length of one meter (Lw = 1 m). The ion

U
current is small, in the order of a few milliamps per meter of wire. Previous experimental studies (Martynenko and
AN
Zheng, 2016; Singh et al., 2017) reported slightly lower currents (below 1 mA), which are most likely related to

shorter wires or needle electrodes that were used. A quadratic relation between ion current and emitter voltage is
M

found, in correspondence with previous studies (Dalvi-Isfahan et al., 2016; Martynenko and Kudra, 2016).
D
TE

14
y = 1.27E-08x2 - 1.76E-04x + 3.02E-01
EP

R² = 1.00E+00
12

10 y = 8.85E-09x2 - 1.27E-04x + 2.49E-01


C

R² = 1.00E+00
Ion current Ic [mA]

8 W2P-IF-NoFruit
AC

W2P-IF
6 Poly. (W2P-IF-NoFruit)
Poly. (W2P-IF)

0
0 10000 20000 30000 40000
Voltage Vw [V]

27
Figure 7. Ion current from emitter to collector (Ic, for a 1 meter long wire) as a function of the applied voltage at
ACCEPTED MANUSCRIPT
the emitter for the wire-to-plate impinging-flow (W2P-IF) configuration. Results without (empty circles) and with

fruit (filled circles) are shown.

PT
RI
U SC
AN
M
D
TE
C EP
AC

28
3.2 EHD drying of fruit
ACCEPTED MANUSCRIPT

3.2.1 EHD drying for impinging flow on a plate

For the base case of impinging flow on a plate on which a fruit is placed, the voltage, electric field strength, space

charge density, air speed and temperature distribution are illustrated in Figure 8. The complex spatial distribution of

these parameters around the emitter electrode and the fruit indicates that they strongly depend on the geometrical

configuration that is considered. Note that the heat transport calculation and the resulting temperature distribution

PT
are only used to determine the CHTC in a steady-state a-priori calculation (Supplementary Material), so the reported

RI
temperature field is not representative for the actual air temperatures during drying.

U SC
AN
M
D
TE

Figure 8. Voltage, electric field strength, space charge density, air speed and temperature distribution for the base

case (wire-to-plate configuration for impinging flow with fruit of 5x10 mm). The range of the color scale is always
EP

depicted for each parameter. Note that these ranges do not always depict the highest calculated values.

The impact of EHD process parameters, such as Vw, dec and rw, on the surface-averaged space charge density over the
C

emitter electrode and on the average air speed of the total incoming airflow in the domain (Uref) during drying of a
AC

fruit for the wire-to-plate impinging-flow configuration was presented earlier in Figure 5a-b. The dependencies of

SCD and air speed on Vw, dec and rw with fruit versus without fruit demonstrate very similar trends. In the presence of

a fruit, however, slightly higher SCD and air speeds are found, which is attributed to the higher electric permittivity

of the fruit. As such, the fruit surface behaves almost as a grounded surface (Vfs << Vw). Thereby, the actual distance

between the emitter and the grounded surface is smaller, namely dec-H instead of dec. This results in a stronger

electric field and higher air speed. This effect is also the reason that a higher ion current is found in the presence of a

29
fruit in Figure 7. In summary, also in the presence of a fruit, higher air speeds are achieved at higher voltages, smaller
ACCEPTED MANUSCRIPT
emitter-collector distances and thinner wires. In Figure 9, the air speed distribution is shown for different voltages,

distances between the electrodes and wire radii. It is clear that the jet gets concentrated much more around the

emitter when the voltage increases or the wire radius decreases. Over the current range of parameters considered,

the air speed distribution changes most drastically when altering the distance between the electrodes.

PT
RI
U SC
AN
M
D
TE
EP

Figure 9. Air speed distribution represented as a function of the voltage at the emitter (Vw), the distance between
C

emitter and collector electrode (dec) and the wire radius (rw) for the wire-to-plate impinging-flow configuration
AC

(W2P-IF, fruit of 5x10 mm, the base case is indicated). The range of the color scale is always depicted for each

voltage.

In Figure 10, the corresponding surface-averaged convective heat transfer coefficient (CHTC) over the fruit surface,

along its three exposed surfaces, is given. The dependencies of the CHTC of the fruit surface on Vw, dec and rw show

similar trends to the CHTC of the plate surface (without a fruit, Figure 5c). The convective transfer coefficients on the
30
fruit surface are relatively high and in the same order of magnitude as for forced convective drying of fruits
ACCEPTED MANUSCRIPT
(Defraeye and Radu, 2017). These surface-averaged CHTC give an overall indication of the CHTC magnitude, but

actually the CHTC is not constant over the fruit surface. The CHTC distribution over the fruit surface (half of top

surface and entire side surface) is shown in Figure 11 at different voltages. The local CHTC is scaled here with respect

to the surface-averaged CHTC over the fruit surface. The CHTC distribution over the fruit surface is clearly dependent

on the applied voltage, which causes differences in the air speeds and airflow fields around the fruit.

PT
RI
SC
90 100 60
y = -2.35E-08x2 + 3.90E-03x - 3.83E+01 W2P-IF W2P-IF
80 90
R² = 9.98E-01
Power (W2P-IF) 50 Poly. (W2P-IF)
70 80

70

U
CHTC Fruit [W/m2K]
CHTC Fruit [W/m2K]

CHTC Fruit [W/m2K]


60 40
60
50
AN
50 30
40
40 y = 1.94E-01x-1.31E+00
30 R² = 9.96E-01 20
30
20
20
W2P-IF 10
M

y = 1.00E-04x2 - 1.15E-01x + 5.31E+01


10 10
Poly. (W2P-IF) R² = 9.94E-01
0 0 0
0 10000 20000 30000 40000 50000 0 0.02 0.04 0.06 0.08 0.1 0 200 400 600
Voltage Vw [V] Distance dec [m] Radius rw [µm]
D

Figure 10. Average CHTC over the fruit surface as a function of the voltage at the emitter (Vw), the distance
TE

between emitter and collector electrode (dec) and the radius of the wire (rw) for the wire-to-plate impinging-flow

configuration. The base case parameters are Vw = 20 kV, dec = 20 mm, rw = 250 µm.
C EP
AC

31
2 W2P-IF-Vw = 12kV

1.8
ACCEPTED MANUSCRIPT
W2P-IF-Vw = 15kV
W2P-IF-Vw = 20kV (base)
W2P-IF-Vw = 25kV
1.6 W2P-IF-Vw = 30kV
W2P-IF-Vw = 40kV
1.4
CHTC/CHTCavg [-]

1.2

0.8

0.6

PT
0.4 xa
0.2

RI
0
0 0.2 0.4 0.6 0.8 1

SC
xa/L [-]

Figure 11. CHTC distribution over the fruit surface (along red dotted line xa), scaled with the surface-averaged

U
CHTC over the fruit surface, for different voltages for the wire-to-plate impinging-flow (W2P-IF) configuration.
AN
In Figure 12, the dry-base moisture content change (X [kg kg-1]) over time is shown for various values of Vw, dec and

rw. From these drying curves, the critical drying time is also extracted for all cases (as defined in section 2.4), and is
M

shown in Figure 13. As shown in Figure 10, increasingly higher CHTCs are obtained by increasing the voltage Vw,
D

decreasing dec or decreasing rw. However, these changes in CTCs do not affect the drying process to the same extent
TE

(Figure 12), as the drying time does not seem to reduce proportionally to the increase in CTCs. The reason is that for

high CTCs, the resistance to moisture transfer of the boundary layer actually is much smaller than the moisture
EP

transport resistance of the fruit tissue. Thereby the internal moisture transfer dominates the drying kinetics at high

CTCs, by which a further increase in CTCs clearly has a limited impact on the drying rate. This behavior is especially
C

pronounced with varying voltage. This trend highlights the specific advantage of the present (semi-)conjugate
AC

coupling of fruit drying to the airflow, namely that the impact of EHD process parameters on the drying rate can be

directly included and quantified. As with the CTCs, the impact of rw on the drying rate is the smallest for the range

considered. Figure 13 shows that the critical drying time increases almost linearly with increasing distance between

electrodes or increasing wire radius, as indicated by the excellent linear regression. The critical drying time decreases

in a very non-linear way with increasing voltage.

32
(a) Voltage (b) Distance emitter-collector (c) Radius emitter (wire)
6 W2P-IF-Vw = 12kV
ACCEPTED
6
MANUSCRIPT 6
W2P-IF-rw = 50um
W2P-IF-dec = 10mm
W2P-IF-Vw = 15kV W2P-IF-rw = 100um
W2P-IF-dec = 20mm (base)
5 W2P-IF-Vw = 20kV (base) 5 W2P-IF-rw = 200um
5
W2P-IF-dec = 40mm W2P-IF-rw = 250um (base)

Dry-base moisture content X [-]


Dry-base moisture content X [-]

Dry-base moisture content X [-]


W2P-IF-Vw = 25kV
W2P-IF-rw = 300um
4 W2P-IF-Vw = 30kV 4 W2P-IF-dec = 80mm 4
W2P-IF-rw = 400um
W2P-IF-Vw = 40kV Xcrit W2P-IF-rw = 500um
3 Xcrit 3 3 Xcrit

2 2 2

1 1 1

PT
0 0 0
0 10 20 30 40 0 10 20 30 40 0 5 10 15 20
Time [h] Time [h] Time [h]

Figure 12. Dry-base moisture content of the fruit (X) as a function of time for different voltages at the emitter (Vw),

RI
distances between electrodes (dec) and wire radius (rw) for the wire-to-plate impinging-flow configuration. The

SC
base case parameters are Vw = 20 kV, dec = 20 mm, rw = 250 µm.

2.5

U
Voltage
Radius
AN
Distance
2 Linear (Radius)
Linear (Distance)
M
tcrit/tcrit,base [h]

1.5
y = 0.4186x + 0.5697
R² = 0.9991
D

1
TE

0.5
y = 0.1544x + 0.8500
R² = 0.9934
EP

0
0 1 2 3 4
C

Vw/Vw,base, rw/rw,base, dec/dec,base [-]


AC

Figure 13. Critical drying time as a function of the voltage at the emitter (Vw), distance between electrodes (dec)

and wire radius (rw) for the wire-to-plate impinging-flow configuration. All quantities are scaled with respect to

the base case (Vw = 20 kV, dec = 20 mm, rw = 250 µm).The linear regressions are also given for the radius and

distance cases.

33
3.2.2 EHD drying for flow going around the fruit
ACCEPTED MANUSCRIPT
The impact of different geometrical configurations (Figure 2), where air is enabled to flow around the fruit, on the

airflow and fruit drying kinetics is quantified. The aim is to identify which configuration is most efficient to dry fruit

using EHD airflow within the perspective of industrial upscaling. In Figure 14, the airflow field around the fruit and

the water activity inside the fruit (after 5 hours of drying) are shown for all configurations. The magnitude of the

surface-averaged CHTC on the fruit surface is also reported. In Figure 15a, the mass loss (dry base moisture content

PT
X versus time) is shown for the five different configurations. Note that for the wire-to-plate configurations, the fruit

cannot dry from the bottom surface in contact with the collector, which is clear from the water activity profiles as

RI
well. For the wire-to-parallel plates and wire-to-wires configurations, the fruit is assumed to lie on a plastic mesh, so

SC
air can flow through, by which the fruit can dry from its four sides. In Figure 15b, the air velocity component

perpendicular to the outlet is shown for the five different configurations, which is u for the impinging flow

U
configuration and v for the configurations where flow goes around the fruit. This streamwise component is chosen as
AN
it is representative for the mass flow going out (or in) of the domain along that boundary.

When comparing both wire-to-plate configurations, the small plate produces a slower drying rate than the infinite
M

plate. This is directly related to the differences between flow fields for impinging flow and flow around the fruit, and
D

the resulting CTCs. The impinging flow configuration has much higher CHTCs, particularly on the windward surface,
TE

due to the air jet impinging onto the fruit. This air jet is then directed sideways towards the outlets. The impinging

flow configuration is however expected to pose problems when multiple products are dried together on the plate.
EP

Since the flow that is produced below the wire is directed sideways, it will gradually pick-up more moisture. This

increasingly humid air will reduce the drying rate of fruit located more downstream, which will result in non-uniform
C

drying for different fruits (Defraeye and Martynenko, 2018). Therefore, flow passing around the fruit for the wire-to-
AC

plate configuration seems a better option as the air is directed over and away from each individual fruit.

The highest drying rate is obtained for the wire-to-mesh configuration. It was better even compared to the

impinging-flow wire-to-plate configuration, despite the former having a lower surface-averaged CHTC. The reason is

that the wire-to-mesh configuration enables the fruit to dry from the bottom surface as well, in addition to the three

other surfaces. In addition to a faster overall drying rate, the product dries also more uniformly (Figure 14). This will

also result in less heterogeneity in quality within the fruit. Note however that for a 3D case, for example a fruit cube,

34
also drying from the lateral sides (out-of-plane direction) is enabled. As such, these differences in heterogeneity
ACCEPTED MANUSCRIPT
between wire-to-mesh and wire-to-plate will be reduced for a 3D case. The wire-to-mesh configuration remains in

any case more suitable to dry many fruits, as they can be placed next to one another on the mesh.

The wire-to-parallel plates and the wire-to-wires configurations dry slower than the wire-to-mesh configuration,

which is related to their lower CHTCs. This could be improved by changing dec, which was now taken the same for all

cases in order to achieve the same bulk electric field strength (Eq.(1)). However, as shown in Figure 6, Ebulk is not a

PT
suitable parameter to compare different geometrical configurations, but it is currently the only one at hand. It is

difficult and maybe not possible to define a representative parameter due to the complex dependency of the electric

RI
field and the resulting flow field on the geometry. Future studies are required to identify the optimal combination of

SC
dec and Vw for each geometry.

When comparing the streamwise air speed (Figure 15b), the maximal air speed is observed downstream of the

U
geometrical center of the fruit for the configurations where air flows around the fruit. For the impinging flow
AN
configuration, the maximal air speed is found close to the collector plate. This difference in airflow patterns

obviously affects the difference in drying patterns and also creates the difference in volumetric moisture distribution
M

inside of the fruit during drying. The consequences of this uneven airflow distribution on the fruit shrinkage and
D

quality should be the subject of further studies.


TE
C EP
AC

35
Figure 14. Airflow field around a fruit of 5x10 mm and the water activity inside the fruit (range 0-1, after 5 hours of
ACCEPTED MANUSCRIPT
low-temperature drying) for all configurations. The maximal air speed and surface-averaged CHTC on the fruit

surface are also reported.

(a) Moisture content (b) Streamwise air velocity


wire-to-plate (impinging flow) wire-to-plate (impinging flow)
6 12
wire-to-plate
wire-to-plate
wire-to-mesh

PT
wire-to-parallel plates 10 wire-to-mesh
5
wire-to-wires wire-to-parallel plates
Dry-matter moisture content X [-]

Xcrit wire-to-wires
8

RI
4

u/Uref, v/Uref [-]


6
3

SC
4

2
2

U 0
AN
0 -2
0 5 10 15 20 0 10 20 30
Time [h] do/L [-]
M

Figure 15. (a) Dry-base moisture content (X) versus time for 5 different configurations; (b) Streamwise air velocity
D

component at the outlet domain boundary (scaled with the average air speed Uref) versus distance along outlet do
TE

(scaled with the length of the fruit L). The parameter d0 shows the distance along the outlet (at y = 0 for impinging

flow and x = 0 for flow around the fruit in Figure 2).


C EP
AC

36
4. Conclusions
ACCEPTED MANUSCRIPT
The electrohydrodynamic drying (EHD) process of a food product was investigated numerically. The aim was to

increase our understanding of the airflow around a fruit and of the resulting EHD-driven dehydration process, and to

identify the most optimal configurations for industrial upscaling. To this end, the airflow field and resulting

convective heat transfer and moisture removal from the fruit were coupled to fruit dehydration in a conjugate way.

This study is the first to develop such a conjugate continuum model for EHD drying of fruit, to link ionic wind to the

PT
resulting convective drying process. Following conclusions are drawn:

RI
• The bulk electric field strength (Ebulk), which is the voltage difference over the electrodes (Vw) divided by the

distance between them (dec), is not a suitable parameter to evaluate and compare the effect of different

SC
geometrical configurations of electrodes.

• The wire electrode does not need to be explicitly included in the numerical model, since also only the

U
AN
ionization region around the wire can be used as a physical boundary in the model.

• The space charge density at the emitter electrode, the approach flow air speed and the convective transfer
M

coefficient at the fruit surface have a non-linear dependency on following EHD process parameters, namely

the applied voltage at the emitter (Vw), the distance between emitter and collector (dec) and the wire radius
D

(rw). These dependencies were quantified for the wire-to-plate configuration for impinging flow. For the
TE

typical range of EHD process parameters considered, the impact of the wire radius on these quantities was

the smallest.
EP

• Regarding the drying kinetics, the critical drying time increases linearly with increasing distance between
C

electrodes or increasing wire radius. The critical drying time decreases in a very non-linear way with
AC

increasing voltage between the electrodes.

• Clear differences in the drying kinetics were quantified between impinging flow (wire-to-plate configuration)

and configurations enabling flow around the fruit, such as wire-to-mesh, wire-to-short plate, wire-to-wires

and wire-to-parallel plates. Impinging flow towards a plate was shown to be not the best solution to achieve

fast and uniform drying of the fruit. Of all tested configurations, the wire-to-mesh configuration highest

EHD-driven drying rate. Placing the fruit on a mesh showed to be advantageous for drying uniformity, as

then fruit can dry from all its surfaces, and also to dry large amounts of fruits together.

37
The developed conjugate approach has the clear advantage that the impact of EHD process parameters and
ACCEPTED MANUSCRIPT
geometrical configurations on the drying rate can be directly quantified. Such modeling approach could be a valuable

tool to further optimize EHD drying technology towards industrial implementation.

Acknowledgements

PT
We acknowledge the support of the World Food System Center (WFSC) of ETH Zürich

(http://www.worldfoodsystem.ethz.ch).

RI
U SC
AN
M
D
TE
C EP
AC

38
References
ACCEPTED MANUSCRIPT
Adamiak, K., 2013. Numerical models in simulating wire-plate electrostatic precipitators: A review. J. Electrostat. 71,

673–680. doi:10.1016/j.elstat.2013.03.001

Aregawi, W., Defraeye, T., Saneinejad, S., Vontobel, P., Lehmann, E., Carmeliet, J., Derome, D., Verboven, P., Nicolai,

B., 2013. Dehydration of apple tissue: Intercomparison of neutron tomography with numerical modelling. Int. J.

Heat Mass Transf. 67, 173–182. doi:10.1016/j.ijheatmasstransfer.2013.08.017

PT
Ayuttaya, S.S.N., Chaktranond, C., Rattanadecho, P., Kreewatcharin, T., 2012. Effect of ground arrangements on

RI
swirling flow in a rectangular duct subjected to electrohydrodynamic effects. J. Fluids Eng. 134, 51211.

SC
Bajgai, T.R., Raghavan, G.S.V., Hashinaga, F., Ngadi, M.O., 2006. Electrohydrodynamic drying - a concise overview.

Dry. Technol. 24, 905–910. doi:10.1080/07373930600734091

U
Bando, K., Kansha, Y., Ishizuka, M., Tsutsumi, A., 2017. Innovative freeze-drying process based on self-heat
AN
recuperation technology. J. Clean. Prod. 168, 1244–1250. doi:10.1016/j.jclepro.2017.09.088
M

Bantle, M., Eikevik, T.M., 2014. A study of the energy efficiency of convective drying systems assisted by ultrasound

in the production of clipfish. J. Clean. Prod. 65, 217–223. doi:10.1016/j.jclepro.2013.07.016


D

Bardy, E., Manai, S., Havet, M., Rouaud, O., 2016. Drying kinetics comparison of methylcellulose gel vs. mango fruit in
TE

forced convective drying with and without electrohydrodynamic enhancement. ASME J. Heat Transf. 138,
EP

84504. doi:10.1115/1.4033390

Bonazzi, C., Dumoulin, E., 2011. Quality changes in food materials as influenced by drying processes, in: Tsotsas, E.,
C

Mujumdar, A.S. (Eds.), Modern Drying Technology - Product Quality and Formulation. Wiley-VHC Verlag GmbH,
AC

Weinheim, Germany, pp. 1–20.

Caccavale, P., De Bonis, M.V., Ruocco, G., 2016. Conjugate heat and mass transfer in drying: A modeling review. J.

Food Eng. 176, 28–35. doi:10.1016/j.jfoodeng.2015.08.031

Casey, M., Wintergerste, T., 2000. Special Interst Group on “Quality and Trust in Industrial CFD” Best Practice

Guidelines, First edit. ed. ERCOFTAC.

39
Curcio, S., Aversa, M., Chakraborty, S., Calabr??, V., Iorio, G., 2016. Formulation of a 3D conjugated multiphase
ACCEPTED MANUSCRIPT
transport model to predict drying process behavior of irregular-shaped vegetables. J. Food Eng. 176, 36–55.

doi:10.1016/j.jfoodeng.2015.11.020

Dalvi-Isfahan, M., Hamdami, N., Le-Bail, A., Xanthakis, E., 2016. The principles of high voltage electric field and its

application in food processing: A review. Food Res. Int. 89, 48–62. doi:10.1016/j.foodres.2016.09.002

Defraeye, T., 2014. Advanced computational modelling for drying processes - A review. Appl. Energy 131, 323–344.

PT
doi:10.1016/j.apenergy.2014.06.027

RI
Defraeye, T., Blocken, B., Carmeliet, J., 2012. Analysis of convective heat and mass transfer coefficients for

convective drying of a porous flat plate by conjugate modelling. Int. J. Heat Mass Transf. 55, 112–124.

SC
doi:10.1016/j.ijheatmasstransfer.2011.08.047

U
Defraeye, T., Blocken, B., Carmeliet, J., 2010. CFD analysis of convective heat transfer at the surfaces of a cube
AN
immersed in a turbulent boundary layer. Int. J. Heat Mass Transf. 53, 297–308.

doi:10.1016/j.ijheatmasstransfer.2009.09.029
M

Defraeye, T., Martynenko, A., 2018. Future perspectives for electro-hydrodynamic drying of biomaterials. Dry.
D

Technol. 36, 1–10.


TE

Defraeye, T., Nicolaï, B., Mannes, D., Aregawi, W., Verboven, P., Derome, D., 2016. Probing inside fruit slices during

convective drying by quantitative neutron imaging. J. Food Eng. 178, 198–202.


EP

doi:10.1016/j.jfoodeng.2016.01.023
C

Defraeye, T., Radu, A., 2017. Convective drying of fruit: a deeper look at the air-material interface by conjugate
AC

modelling. Int. J. Heat Mass Transf. 108, 1610–1622. doi:10.1016/j.ijheatmasstransfer.2017.01.002

Defraeye, T., Verboven, P., 2017. Convective drying of fruit: Role and impact of moisture transport properties in

modelling. J. Food Eng. 193, 95–107. doi:10.1016/j.jfoodeng.2016.08.013

Defraeye, T., Verboven, P., Nicolai, B., 2013. CFD modelling of flow and scalar exchange of spherical food products:

Turbulence and boundary-layer modelling. J. Food Eng. 114, 495–504. doi:10.1016/j.jfoodeng.2012.09.003

Ding, C., Lu, J., Song, Z., 2015. Electrohydrodynamic drying of carrot slices. PLoS One 10, 1–12.
40
doi:10.1371/journal.pone.0124077
ACCEPTED MANUSCRIPT
Esehaghbeygi, A., Basiry, M., 2011. Electrohydrodynamic (EHD) drying of tomato slices (Lycopersicon esculentum). J.

Food Eng. 104, 628–631. doi:10.1016/j.jfoodeng.2011.01.032

Fylladitakis, E.D., Theodoridis, M.P., Moronis, A.X., 2014. Review on the history, research, and applications of

electrohydrodynamics. IEEE Trans. Plasma Sci. 42, 358–375. doi:10.1109/TPS.2013.2297173

PT
Ghalamchi, M., Kasaeian, A., Ghalamchi, M., Fadaei, N., Daneshazarian, R., 2017. Optimizing of solar chimney

performance using electrohydrodynamic system based on array geometry. Energy Convers. Manag. 135, 261–

RI
269. doi:10.1016/j.enconman.2016.12.074

SC
Ghazanchaei, M., Adamiak, K., Castle, G.S.P., 2015. Predicted flow characteristics of a wire-nonparallel plate type

electrohydrodynamic gas pump using the Finite Element Method. J. Electrostat. 73, 103–111.

U
doi:10.1016/j.elstat.2014.11.003
AN
Halder, A., Datta, A.K., 2012. Surface heat and mass transfer coefficients for multiphase porous media transport

models with rapid evaporation. Food Bioprod. Process. 90, 475–490. doi:10.1016/j.fbp.2011.10.005
M

Heidarinejad, G., Babaei, R., 2015. Numerical investigation of electro hydrodynamics (EHD) enhanced water
D

evaporation using Large Eddy Simulation turbulent model. J. Electrostat. 77, 76–87.
TE

doi:10.1016/j.elstat.2015.07.007.
EP

Horttanainen, M., Deviatkin, I., Havukainen, J., 2017. Nitrogen release from mechanically dewatered sewage sludge

during thermal drying and potential for recovery. J. Clean. Prod. 142, 1819–1826.
C

doi:10.1016/j.jclepro.2016.11.102
AC

Jewell-Larsen, N.E., Karpov, S. V, Krichtafovitch, I.A., Jayanty, V., Hsu, C.-P., Mamishev, A. V, 2008. Modeling of

corona-induced electrohydrodynamic flow with COMSOL multiphysics. Proc. ESA Annu. Meet. Electrost. 1–13.

Kocik, M., Podlinski, J., Mizeraczyk, J., Chang, J.S., 2009. Particle image velocimetry measurements of wire-

nonparallel plates type electrohydrodynamic gas pump. IEEE Trans. Dielectr. Electr. Insul. 16, 312–319.

doi:10.1109/TDEI.2009.4815158

Kudra, T., Martynenko, A., 2015. Energy aspects in electrohydrodynamic drying. Dry. Technol. 33, 1534–1540.
41
doi:10.1080/07373937.2015.1009540
ACCEPTED MANUSCRIPT
Kurnia, J.C., Sasmito, A.P., Tong, W., Mujumdar, A.S., 2013. Energy-efficient thermal drying using impinging-jets with

time-varying heat input - A computational study. J. Food Eng. 114, 269–277.

doi:10.1016/j.jfoodeng.2012.08.029

Lai, F.C., 2010. A prototype of EHD-enhanced drying system. J. Electrostat. 68, 101–104.

doi:10.1016/j.elstat.2009.08.002

PT
Lamnatou, C., Papanicolaou, E., Belessiotis, V., Kyriakis, N., 2010. Finite-volume modelling of heat and mass transfer

RI
during convective drying of porous bodies - Non-conjugate and conjugate formulations involving the

aerodynamic effects. Renew. Energy 35, 1391–1402. doi:10.1016/j.renene.2009.11.008

SC
Lamnatou, C., Papanicolaou, E., Belessiotis, V.G., Kyriakis, N., 2009. Conjugate heat and mass transfer from a drying

U
rectangular cylinder in confined air flow. Numer. Heat Transf. Part A 56, 379–405.
AN
doi:10.1080/10407780903244353

Marra, F., De Bonis, M.V., Ruocco, G., 2010. Combined microwaves and convection heating: A conjugate approach. J.
M

Food Eng. 97, 31–39. doi:10.1016/j.jfoodeng.2009.09.012


D

Martynenko, A., Kudra, T., 2016. Electrically-induced transport phenomena in EHD drying - A review. Trends Food
TE

Sci. Technol. 54, 63–73. doi:10.1016/j.tifs.2016.05.019


EP

Martynenko, A., Zheng, W., 2016. Electrohydrodynamic drying of apple slices: Energy and quality aspects. J. Food

Eng. 168, 215–222. doi:10.1016/j.jfoodeng.2015.07.043


C

Ould Ahmedou, S., Havet, M., 2009. Effect of process parameters on the EHD airflow. J. Electrostat. 67, 222–227.
AC

doi:10.1016/j.elstat.2009.01.055

Oussalah, N., Zebboudj, Y., 2006. Finite-element analysis of positive and negative corona discharge in wire-to-plane

system. Eur. Phys. J. Appl. Phys. 34, 215–223. doi:10.1051/epjap:2006063

Pirnazari, K., Esehaghbeygi, A., Sadeghi, M., 2016. Modeling the electrohydrodynamic (EHD) drying of banana slices.

Int. J. Food Eng. 12, 17–26. doi:10.1515/ijfe-2015-0005

42
Podlinski, J., Niewulis, A., Berendt, A., Mizeraczyk, J., 2013. Pumping effect measured by PIV method in a multilayer
ACCEPTED MANUSCRIPT
spike electrode EHD device for air cleaning. IEEE Trans. Ind. Appl. 49, 2402–2408.

doi:10.1109/TIA.2013.2265214

Prosapio, V., Norton, I., De Marco, I., 2017. Optimization of freeze-drying using a Life Cycle Assessment approach:

Strawberries’ case study. J. Clean. Prod. 168, 1171–1179. doi:10.1016/j.jclepro.2017.09.125

Sahoo, N.K., Gupta, S.K., Rawat, I., Ansari, F.A., Singh, P., Naik, S.N., Bux, F., 2017. Sustainable dewatering and drying

PT
of self-flocculating microalgae and study of cake properties. J. Clean. Prod. 159, 248–256.

RI
doi:10.1016/j.jclepro.2017.05.015

Singh, A., Orsat, V., Raghavan, V., 2012. A comprehensive review on electrohydrodynamic drying and high-voltage

SC
electric field in the context of food and bioprocessing. Dry. Technol. 30, 1812–1820.

U
doi:10.1080/07373937.2012.708912 AN
Singh, A., Vanga, S.K., Nair, G.R., Gariepy, Y., Orsat, V., Raghavan, V., 2015. Electrohydrodynamic drying (EHD) of

wheat and its effect on wheat protein conformation. LWT - Food Sci. Technol. 64, 750–758.
M

doi:10.1016/j.lwt.2015.06.051
D

Singh, A., Vanga, S.K.K., Raveendran Nair, G., Gariepy, Y., Orsat, V., Raghavan, V., 2017. Electrohydrodynamic drying
TE

of sand. Dry. Technol. 35, 312–322. doi:10.1080/07373937.2016.1170028

Taghian Dinani, S., Havet, M., 2015a. Effect of voltage and air flow velocity of combined convective-
EP

electrohydrodynamic drying system on the physical properties of mushroom slices. Ind. Crops Prod. 70, 417–

426. doi:10.1016/j.indcrop.2015.03.047
C
AC

Taghian Dinani, S., Havet, M., 2015b. The influence of voltage and air flow velocity of combined convective-

electrohydrodynamic drying system on the kinetics and energy consumption of mushroom slices. J. Clean. Prod.

95, 203–211. doi:10.1016/j.jclepro.2015.02.033

Tirumala, R., Go, D.B., 2014. Comparative study of corona discharge simulation techniques for electrode

configurations inducing non-uniform electric fields. J. Electrostat. 72, 99–106. doi:10.1016/j.elstat.2013.12.003

Yang, M., Ding, C., 2016. Electrohydrodynamic (EHD) drying of the Chinese wolfberry fruits. Springer Plus 5, 909–929.

43
doi:10.1186/s40064-016-2546-1
ACCEPTED MANUSCRIPT
Yuen, A.W., 2006. Collector current density and dust collection in wire-plate electrostatic precipitators. University of

New South Wales.

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, 1239–1255.

doi:10.1080/10408398.2014.979280

PT
Zhang, Y., Liu, L., Chen, Y., Ouyang, J., 2015. Characteristics of ionic wind in needle-to-ring corona discharge. J.

RI
Electrostat. 74, 15–20. doi:10.1016/j.elstat.2014.12.008

U SC
AN
M
D
TE
C EP
AC

44
ACCEPTED MANUSCRIPT

Highlights

• This conjugate model is the first to couple ionic wind to food dehydration.
• Dependency of drying time on voltage, electrode distance and radius is quantified
• The conventional wire-to-plate configuration is not the best to achieve fast and uniform drying
• The wire-to-mesh configuration shows the highest EAD-driven food drying rate.
• Placing food on a mesh is preferred to dry large amounts of products uniformly.

PT
RI
U SC
AN
M
D
TE
C EP
AC

You might also like