Kudra 2015

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

This article was downloaded by: [University of Montana]

On: 07 April 2015, At: 17:47


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House,
37-41 Mortimer Street, London W1T 3JH, UK

Drying Technology: An International Journal


Publication details, including instructions for authors and subscription information:
http://www.tandfonline.com/loi/ldrt20

Energy Aspects in Electrohydrodynamic Drying


a a
T. Kudra & A. Martynenko
a
Department of Engineering, Dalhousie University, Truro, Canada
Accepted author version posted online: 05 Feb 2015.

Click for updates

To cite this article: T. Kudra & A. Martynenko (2015): Energy Aspects in Electrohydrodynamic Drying, Drying Technology: An
International Journal, DOI: 10.1080/07373937.2015.1009540

To link to this article: http://dx.doi.org/10.1080/07373937.2015.1009540

Disclaimer: This is a version of an unedited manuscript that has been accepted for publication. As a service
to authors and researchers we are providing this version of the accepted manuscript (AM). Copyediting,
typesetting, and review of the resulting proof will be undertaken on this manuscript before final publication of
the Version of Record (VoR). During production and pre-press, errors may be discovered which could affect the
content, and all legal disclaimers that apply to the journal relate to this version also.

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) contained
in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no
representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the
Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and
are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and
should be independently verified with primary sources of information. Taylor and Francis shall not be liable for
any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever
or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of
the Content.

This article may be used for research, teaching, and private study purposes. Any substantial or systematic
reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any
form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http://
www.tandfonline.com/page/terms-and-conditions
ENERGY ASPECTS IN ELECTROHYDRODYNAMIC DRYING

T. Kudra1, A. Martynenko1
1
Department of Engineering, Dalhousie University, Truro, Canada

Corresponding author E-mail: tadeusz.kudra@gmail.com

Abstract

Acknowledging favorable quality of products obtained through electrohydrodynamic

(EHD) drying, there is a lack of knowledge about energy aspects of this promising, yet
Downloaded by [University of Montana] at 17:47 07 April 2015

not commercialized, technology. This paper is a critical review of EHD research which

may be crucial for future studies and industrial applications. In particular, effects of

electrodes configuration and operating parameters on drying kinetics, energy

consumption and energy efficiency of EHD as compared to conventional drying are

examined. Some engineering considerations for using EHD in industrial dryers are also

discussed.

KEYWORDS: Ionic wind; Biomaterials; Drying rate; Energy; Quality

INTRODUCTION

Electrohydrodynamic (EHD) drying refers to the removal of water from a drying material

placed in the strong electric field due to the so-called "corona wind". This wind originates

from a sharp end of the electrically conducting pin as a result of ions leaving this pin and

impinging the surface of the material being dried on their way towards the plate

electrode.[1,2]

1
To avoid ambiguities and misinterpretation encountered in published papers, related to

treatment of biomaterials in a high-voltage electric field, a distinction for EHD should be

made at the outset from:

(i) electroporation (electropermeabilization), where the high-voltage electric field applied

usually in a pulsed mode (typically, exposure time is in the order of microseconds),

results in electric discharge. Pulsed electric field (PEF) perforates the hardly permeable

skin and disrupts the cellular structure of fruits, berries or vegetables, facilitating mass

transfer for osmotic dehydration[3] or extraction of valuable compounds.[4,5] A


Downloaded by [University of Montana] at 17:47 07 April 2015

comprehensive yet concise overview of PEF applications, including modification of

material properties, inactivation of enzymes and microorganisms is given in the technical

literature.[6]

(ii) altering the properties of fruits, vegetables and other biomaterials by exposure to a

high-voltage electric field between two parallel plate electrodes to extend the shelf life of

perishable foods.[7]

(iii) pretreatment in electric field prior to osmotic dehydration[8,9], air drying,[10,11] and

vacuum freeze-drying.[12]

(iv) any other technologies, exemplified by roller pressing assisted by electroosmosis,[13]

which exploits the characteristics of either high- or low- voltage electric field.[14,15]

Following the aforementioned exceptions, the EHD drying/dewatering is further referred

only to these applications, where high-voltage electric field generates the corona wind

which imposes a specific hydrodynamic effect when impinging the drying material.

2
Although hydrodynamic effects in fluids are well documented, the EHD drying has

recently gained extensive interest because it is regarded as a non-thermal technology,

particularly suitable for heat-sensitive biomaterials. With respect to product quality, the

following can be quoted as examples: lower shrinkage [16,17], higher rehydration ratio[16]

and preserved vitamin content.[18] Moreover, no discernible degradation is generally

noted in terms of color[16-19] though Li et al. [20] reported distinctive browning of okara

cake just under the needle electrode. These quality attributes are well enhanced when

EHD is combined with vacuum freeze drying.[21]


Downloaded by [University of Montana] at 17:47 07 April 2015

Regarding energy consumption it is claimed to be much lower in both EHD and

combined EHD-hot air drying than in simply hot air drying.[1,2,19,22,23] One of the reasons

is the EHD enhancement of drying rate. Numerous researchers indicate significant

increase of drying rate in the range from 1.7 to 4.52, depending on the type of

biomaterial, initial/final moisture content and operating conditions.[20,24-26] It results in

appreciably shorter drying time by 15 to 40%.[20,22] Interestingly, the energy consumption

of EHD combined with vacuum freeze drying was lower than freeze drying or EHD

drying alone.[21]

However, even though a prototype has been designed and tested,[27] yet no industrial

dryer exploiting the EHD principle was commercially available at the time of writing this

paper. Apparently, the widely quoted low electric energy consumption based on the only

power delivered with corona wind needs careful re-examination. Thus, this study aims to

3
analyze certain energy-related issues in EHD drying, which could be useful for industrial

scaling of this innovative technology towards process and equipment design.

ENERGY RELATED ISSUES IN EHD DRYING

With due respect to authors of fundamental research, the EHD drying of commodity

materials such as wheat, apples and carrots is rather an academic curiosity than a

technically and economically justifiable large-scale processing. EHD application could be

justified only for low temperature drying of valuable medicinal plants, probiotics and
Downloaded by [University of Montana] at 17:47 07 April 2015

nutraceuticals. The key criterion in industrial viability of EHD drying appears to be the

energy efficiency leading to process economics. Therefore, this section presents some

key issues related to energy consumption and energy efficiency of EHD drying.

Principle Of EHD Drying

The principle of EHD drying is based on the use of the "corona" wind, also called the

"ionic" or "electric" wind, where electric discharge from a high-voltage electrode creates

the jet of high-energy ions and gas molecules targeting the dried material. For this reason,

the electrode generating this specific gas jet is considered as the ionic wind (EHD)

generator.

The corona discharge appears when a high voltage is applied between two electrodes

with substantially different radii of curvature, such as a flat surface and needle (sharp

pin), or a flat surface and thin wire, giving point-to-plate and wire-to-plate

configurations, respectively.

4
For example, the corona wind emitted from the needle forms the gas jet on its way to the

plate electrode. Observation of the jet in dark environment indicated its conical form with

the cone angle of 2α (Fig. 1). According to the widely quoted Warburg empirical law[28]

(also confirmed by analytical solutions[29]), the maximum semi-vertical cone angle of the

corona wind in point-to-plate discharge in air is equal to 60o. This rule has been proven

for direct current (DC) of positive polarity in point-to-plate and wire-to-plate

configurations. It is reasonable to assume validity of the same jet geometry for other
Downloaded by [University of Montana] at 17:47 07 April 2015

modes of electric current supply; however this assumption should carefully be tested for

various alternating current (AC) wave forms. So far, no consensus exists on the

performance of alternating current (AC), which was found more efficient than DC for

water evaporation[30] or less efficient for drying of carrots[26] and paper.[31]

Typically, the point-to-plate electrode produces a highly non-uniform electric field, so the

strongest evaporation takes place just below the point electrode and diminishes towards

the periphery. In the case of evaporation from liquid solutions and suspensions, the

vortex-like turbulent motion was observed in water[32] and aqueous suspension of whey

protein,[33] which disappeared when the suspension started to solidify. Thus, such vortex

motion is unlikely to exist in drying of solid materials, like carrots and potatoes.

It should be noted that the jet of corona wind impinging the dried material rebounds from

the material surface, which affects the neighboring jets emitted from a multi-needle or

5
multi-wire electrode. However, this boundary phenomenon was not considered in

previous configurations of multi-needle electrodes.

Electrode Configuration

Although the plate electrode may be either solid or perforated, the emitting electrode is

usually a vertical single- and multi-needle type [16-20,25,26,31-37,39,40] or the wire-type placed

in parallel to the plate electrode.[21-24,27,29,30,38,41-43] The single- and multiple-wire

electrodes were studied because they better fit the foreseen industrial units such as band
Downloaded by [University of Montana] at 17:47 07 April 2015

dryers.[27] As indicated by Lai and Wong [36] the performance of the wire electrode is

better than the needle type when the applied voltage is lower than 15 kV. Over this

critical voltage, the needle electrode performs better. This phenomenon can be attributed

to the corona wind geometry,[36] which for the wire electrode resembles a slot-type jet

(Fig. 2a), whereas the needle electrode produces a conical-type jet with circular area on

the material surface (Fig. 2b). Fig. 2b presents the theoretically optimal staggered

arrangement of the multiple needle electrode, which would provide close to uniform

exposure of a material under drying.

Rationally, the best configuration of the multi-needle electrode is with the needles spaced

to produce impinging corona jets with circular areas almost touching each other. The

distance between the needles and their configuration should be determined

experimentally because of possible interference of neighboring jets and the effect of jets

rebounding from the surface of dried material. The gap between the electrode and

material is another important design parameter, which has to be properly chosen to

6
optimize the both electric field strength and material surface area covered by the jet of

electric wind. Not only electric field strength, but also electric charge density has to be

controlled in the gap.

Air Temperature

Despite the fact that heat generally enhances drying, it was found that the efficiency of

EHD drying drops with increasing air temperature.[39] The experiments on combined

EHD-hot air drying showed EHD enhancement of drying rate as 1.9, 1.6, and 1.5 for
Downloaded by [University of Montana] at 17:47 07 April 2015

temperatures 20, 35, and 50°C, respectively.[39] Decrease in EHD efficiency of drying

with temperature might be related to the non-Fickian mechanism of drying with

negligible effect of thermodiffusion. These experiments clearly demonstrated advantages

of EHD application for low-temperature drying. In experiments with partially wetted

glass beads[43] it was found that EHD drying is mostly a surface phenomenon. This is an

obvious obstacle for EHD in the final stages of drying to remove moisture trapped deep

inside the material. To overcome this obstacle, it has been proposed to create the vertical

thermal gradient by auxiliary heating from below.[35] A preliminary study has shown that

low-power heating at 118.9 W/m2 can effectively accelerate the EHD drying due to

enhanced thermodiffusion from the sample core to its surface.[35] Hence, with the backup

of auxiliary heating, the EHD phenomenon could be used for industrial drying at low air

temperature. However, possible accumulation of moisture in the gap between electrodes

might require additional air convective flow.

Convective Cross-Flow

7
Cross-flow convection was initially applied in EHD drying with the intention of

enhancing drying rates. However, numerous experiments in a wide range of air velocities

from 0.1 to 5 m/s revealed a generally negative impact of air flow on the efficiency of

EHD drying because of suppression of ionic wind.[23,24,36,38,43] The exception was the

drying of kiwi fruits with a 17-needle electrode where the effect of cross-flow air velocity

was found insignificant.[37] The effect of cross- flow on EHD performance was

thoroughly examined by Lai and Sharma[34] and Dalvand et al.[37]


Downloaded by [University of Montana] at 17:47 07 April 2015

To quantify interaction between convective airflow and ionic wind, the dimensionless

EHD number as the ratio between ionic wind velocity ue and cross-flow air velocity u has

been introduced:[42]

ue
N EHD (1)
u

The EHD number reflects interaction of two orthogonal forces: electric force of ionic

wind Fe and inertial force of airflow F. It was found that ionic wind velocity is directly

proportional to the electric field strength and can be calculated from the following

relationship [32]:

0
ue E (2)

The silent assumption incorporated in this formula is that air density is constant,

independent of water vapor density and electric charge density but this requires further

experimental verification.

8
The effect of EHD was significant only at low air velocity (NEHD 1), when ionic wind

was not suppressed by cross-flow convection.[23,26,36,37,42] The velocity of ionic wind is

usually in the order of several meters per second, which definitely entails the

aerodynamic effect on a drying material, disturbing the boundary layer at the material’s

surface. Under experimental conditions in the study by Pogorzelski et al.,[26] the ionic

wind generated by the needle electrode at 5.27 kV/cm was 1.45 m/s, which is much

higher than the cross-flow air velocity of 0.1 m/s. These authors confirmed the findings
Downloaded by [University of Montana] at 17:47 07 April 2015

by others that ionic wind is the major driving force in the first drying period.

The considerable reduction of the EHD effect by convective airflow (NEHD 1) could be

explained by domination of the inertial force over electric force, thus resulting in classical

convective air drying. In these experiments the drying rate was independent of electric

force, increasing with air velocity due to the direct effect on the boundary layer mass-

transfer resistance (boundary condition of the third kind). Multifactorial experimentation

proved negligible effect of low air velocity in the range 0 to 0.4 m/s on EHD efficiency

and significant effect of air velocity on energy consumption.[37] It could be explained

through substantial energy consumption by convective blower, which is 10 to 20 times

higher than energy consumption of the equivalent EHD wind generator.[44] Moreover, the

experimental data on EHD/hot air drying indicate that the overall energy consumption

could be by 1000 times higher than energy consumption of the EHD generator.[22]

It follows that the direct effect of convective cross-flow is not beneficial for industrial

EHD drying for at least two reasons: (i) reducing the effect of the vertical ionic wind due

9
to perpendicular air flow, and (ii) increasing the overall energy consumption. Indirect

effect of convective cross-flow on the EHD efficiency due to reduced humidity should be

further investigated. However, this effect is significant only for the first period of drying.

An alternative solution to control humidity is an air dehumidifier, sufficient to maintain

equilibrium vapor pressure.

Effect Of Humidity And Pressure

In spite of the fact that humidity and air pressure play significant role in the process of
Downloaded by [University of Montana] at 17:47 07 April 2015

drying, the effect of these conditions on the efficiency of EHD drying has never been

studied. Humidity measurements were used in experiments of Lai and

coworkers[34,36,42,43] to calculate the Sherwood number, representing conditions of mass

transfer. Bai et al.[21] presented results of vacuum freeze drying, which revealed better

performance of EHD drying at ambient temperature 18oC and relative humidity of 45%

versus vacuum freeze drying (conditions were not specified, however). Unfortunately, the

presented experimental design did not allow separate evaluation of temperature and low

pressure effects. This gap in the knowledge requires careful research of these factors,

which are critical for the industrial scaling of the EHD technology.

Drying Kinetics

The kinetics of EHD drying was found similar to the kinetics of convective drying, which

means that depending on the processed material and moisture content both constant and

falling rate periods are recognized. In some cases, the short increasing rate period was

10
also noted.[20,22] Figure 3 presents drying curves for two kinds of food products with

definitely constant drying rate at the beginning of drying.[26]

One of the measures to quantify the efficiency of EHD drying is the ratio of the mass

flow rate of EHD-dried sample over the control one dried under ambient conditions

without an electric field.[45] Based on this definition, the EHD enhancement has been

expressed through the drying rate ratio for the materials dried in the first drying period.[26]

Hence, for experiments presented in Fig. 3, the enhancement ratio was calculated as 2.01
Downloaded by [University of Montana] at 17:47 07 April 2015

for miscanthus and 4.52 for carrot.[26] Alternatively, drying enhancement could be

expressed through the ionic wind velocity (Eq. 2), which is proportional to easily

measurable applied voltage.[45] For solid biomaterials with complex diffusivity, such as

biscuits, with the assumption of coupled heat/mass transfer and negligible boundary

effects, it was proposed to characterize drying enhancement with a ratio of heat transfer

coefficients.[45]

It is important to notice that EHD enhancement of drying rate is not constant over the

entire drying process. EHD enhancement in the falling rate period is reported to be in the

range from 0.97 to 3.2 for a variety of biomaterials. [20,23-25,36] However, much higher

values of 4.52 for carrot[26] and 6.8 for paper towel[45] have been obtained for the constant

rate period, which point out to the prevailing hydrodynamic effect of the corona wind on

the evaporation of surface water. In the case of open surface water, the evaporation was

enhanced by 7.3 at cross-air velocity of 0.125 m/s.[38] Our own attempts to calculate the

11
EHD enhancement for the materials dried in both the constant and falling rate periods are

underway.

Energy Consumption

Electric energy consumption is widely acclaimed to be much lower in EHD drying than

in convective drying. For example, Yang et al.[46] reported energy savings due to EHD

drying at 50°C in the range from 50 to 85% as compared to oven drying and fluid bed

drying at 150°C. This large-scale EHD-based dryer with 2 m2 surface area and heating
Downloaded by [University of Montana] at 17:47 07 April 2015

power of 1.5 kW provided 0.4 kW of corona power at 40 kV of operating voltage.

Unfortunately, no specific information was given on EHD design details and experiments

with the oven and fluid bed dryers.

In general, energy consumption depends on many parameters, such as applied voltage,

positive or negative polarity, electrode configuration, and the electrode-to-material

spacing. The complex dependence of energy consumption on numerous parameters can

be exemplified through non-linear decrease of energy consumption with decreasing

voltage and electrodes gap, thus increasing the electric field strength up to the threshold

value of 5.2 kV/cm.[37] Also, the total moisture removed in a given time with three wires

was found higher than with one wire, which translates into higher energy efficiency.[47]

In the review by Singh et al.[2] it was pointed out that energy consumption depends on the

electrode configuration and water availability. The lowest values of energy consumption

in the range from 14.7 to 2727 kJ/kg referred to evaporation from the open water surface

12
or free water saturating the bed of 3- and 6-mm glass beads whereas the higher values

were pertinent for food products.[2,36,42]

Although numerous publications refer to energy advantages of EHD drying, the scale-up

route towards the industrial apparatus requires considering all aspects of energy

consumption. Namely, it should be noted that the energy consumption in published

papers is based on the "net" energy calculated from the applied voltage and current

passing from the corona electrode to plate electrode[19,24,26,31,39,40,48] and rarely determined
Downloaded by [University of Montana] at 17:47 07 April 2015

from the consumed electric power.[41,44]

However, typical EHD apparatus is composed of various electricity-based units (Fig. 4)

each of them being characterized by certain energy efficiency. Thus, in any comparison,

the factual electric energy delivered to the drying system should be considered. It does

not mean that the energy consumed by the EHD drying will exceed the one for air drying

but the savings effect might not be so appealing. For instance, Zander [49] indicated that

the major fraction of energy was consumed by peripheral equipment, so the ratio of the

net power to the supplied power varied from 0.01 to 0.17, depending on the apparatus and

experimental conditions. Our own studies were performed with a multiple-needle

electrode formed from 60 steel nails with 60o conical tip arranged in 6 rows and 10

columns with 10-mm square spacing, connected to the DC voltage source (SPELLMAN,

Model RHR20P10/FG/RC, USA).[44] The experiments indicated from 85 to 99% energy

lost in the low-to-high voltage convertor. Clearly, determination of the factual energy use

13
for EHD drying is needed to get the real picture of energy consumption for the

comparison of various competitive drying technologies.

Energy Efficiency

The energy performance of a dryer and drying process is characterized by various indices

such as volumetric evaporation rate, surface heat losses, steam consumption, unit heat

consumption, and energy (thermal) efficiency.[50] The ratio of energy used for moisture

evaporation to the total energy supplied to the dryer, is the most frequently quoted as an
Downloaded by [University of Montana] at 17:47 07 April 2015

efficiency in technical literature. However, the superior index to compare efficiency of

drying operation is specific energy consumption, defined as the amount of energy

required to evaporate unit mass of water and therefore given in kJ/kg.[50] This index,

further termed as energy efficiency, can be determined from two measurable variables,

namely the electric (thermal) power (kW) and the drying rate (kg/s). The energy

efficiency over the entire period of EHD drying varies broadly in the range from 90 to

720 kJ/kg,[26] which corresponds well with 100-800 kJ/kg for single- and multiple-needle

electrode and 200-5000 kJ/kg for wire electrode.[2] As a reference, the energy efficiency

of convective belt dryer, which so far is the only configuration of the foreseen industrial

EHD dryer,[27] attains 3800-3950 kJ/kg.[51,52] It should be noted that the energy

performance in kg/kJ, quoted in some papers,[34,35,42,43] is a reciprocal to energy

efficiency, but mutual conversion is quite straightforward.

From the analysis of literature data it follows that the energy efficiency of EHD drying

varies in time. Thus, the concept of instantaneous drying indices[50] could be used to

14
determine the temporal characteristic of the energy performance to be further exploited

for optimization of EHD drying. In such a case, the real-time quantification of energy

performance requires simultaneous measurements of two variables, namely the power

consumption and the evaporation rate. Instrumentation for real-time measurements on the

lab scale is available,[53] but the procedure of experimental verification is needed for

further scaling-up to the industrial prototype.

CONCLUSIONS
Downloaded by [University of Montana] at 17:47 07 April 2015

1. EHD proved to be a promising tool for acceleration of low-temperature drying of high

value materials like medicinal plants, probiotics, nutraceuticals and other heat-sensitive

biomaterials. Yet, detailed analysis of quality/energy issues should be performed.

2. The kinetics of EHD drying is similar to conventional drying, which means that

depending on the processed material, both constant and falling rate periods are

recognized. In general, drying rate in EHD processing is higher, typically from 1.3 to

4.52 and thus, drying time is shortened by 15 to 40%, depending on the material,

moisture content and operating conditions. When combined with convective drying, the

accelerating effect of EHD diminishes with air temperature and cross-flow air velocity.

3. Energy consumption in EHD drying is discernibly lower than that in convective drying

yet not as small as calculated from applied voltage and current of EHD generator.

4. Because of various hypotheses regarding the mechanism of water removal due to direct

hydrodynamic impact of electric wind and possibly turbulent vortex phenomenon, as well

as other non-thermal effects, such as lowering of entropy, this technology requires further

15
studies towards optimization and careful engineering design for the industrial

applications.

NOMENCLATURE

E electric field strength, V/m

u superficial velocity, m/s

NEHD dimensionless number, -

db dry basis
Downloaded by [University of Montana] at 17:47 07 April 2015

Indices

e electric

Greek letters

α semi-vertical cone angle, o

εo permittivity of free space, F/m

ρ density, kg/m3

Acronyms

AC alternating current

DC direct current

EHD electrohydrodynamic

REFERENCES

1. Bajgai, T.R.; Raghavan, G.S.V.; Hashinaga, F.; Ngadi, M.O.

Electrohydrodynamic drying – A concise overview. Drying Technology 2006, 7(7), 905 -

910.

16
2. Singh, A.; Orsat, V.; Raghavan, G.S.V. A comprehensive review on

electrohydrodynamic drying and high-voltage electric field in the context of food and

bioprocessing. Drying Technology 2012, 30(16), 1812-1820.

3. Amami, E., Vorobiev, E.; Kechaou, N. Effect of pulsed electric field on

the osmotic dehydration and mass transfer kinetics of apple tissue. Drying Technology

2005, 23(3), 581-595.

4. Sack, M.; Sigler, J.; Frenzel, S.; Eing, Chr.; Arnold, J.; Michelberger, Th.;

Frey, W.; Attmann, F.; Stukenbrock, L.; Müller, G. Research on industrial-scale


Downloaded by [University of Montana] at 17:47 07 April 2015

electroporation devices fostering the extraction of substances from biological tissue.

Food Engineering Review 2010, 2, 147-156.

5. Loginova, K.V.; Vorobiev, E.; Bals, O.; Lebovka, N.I. Pilot study of

countercurrent cold and mild heat extraction of sugar from sugar beets, assisted by pulsed

electric fields. Journal of Food Engineering 2011, 102, 340-347.

6. Barbosa-Canovas, G.V.; Zhang, Q.H.; Tabilo-Munizaga, G. Pulsed

Electric Fields in Food Processing; Technomic Publishing Co., Inc. Lancaster,

Pennsylvania, USA, 2001.

7. Bajgai, T.R.; Hashinaga, F.; Isobe, S.; Raghavan, G.S.V.; Ngadi, M.O.

Application of high electric field (HEF) on the shelf-life extension of emblic fruit

(Phyllanthus emblica L.). Journal of Food Engineering 2006, 74, 308-313.

8. Rastogi, N.K.; Eshtiaghi, M.N.; Knorr, D. Accelerated mass transfer

during osmotic dehydration of high intensity electrical filed pulse pretreated carrots.

Journal of Food Science 1999, 64(6), 1020-1023.

17
9. Ade-Omowaye, B.I.O.; Rastogi, N.K.; Angerbach, A.; Knorr, D.

Combined effects of pulsed electric field pre-treatment and partial osmotic dehydration

on air drying behaviour of red bell pepper. Journal of Food Engineering 2003, 60, 89-98.

10. Gachovska, T.K.; Simpson, M.V.; Ngadi, M.O.; Raghavan, G.S.V. Pulsed electric

field treatment of carrots before drying and rehydration. Journal of the Science of Food

and Agriculture 2009, 89(14), 2372-2376.

11. Gachovska, T.K.; Adedeji, A.A.; Ngadi, M.O.; Raghavan, G.S.V. Drying

characteristics of pulsed electric field-treated carrot. Drying Technology 2008, 26(10),


Downloaded by [University of Montana] at 17:47 07 April 2015

1244-1250.

12. Wu, Y.; Guo, Y.; Zhang, D. Study of the effect of high-pulsed electrical field

treatment on vacuum freeze-drying of apples. Drying Technology 2011, 29(14), 1714-

1720.

13. Lightfoot, D.; Raghavan, G.S.V. Combined field dewatering of seaweed with a

roller press. Applied Engineering in Agriculture 1995, 11(2), 291-295.

14. Orsat, V.; Raghavan, G.S.V.; Norris, E.R. Food processing waste dewatering by

electro-osmosis. Canadian Agricultural Engineering 1996, 38(1), 63-67.

15. Ng, S.K.; Plunkett, A.; Stojceska, V.; Ainsworth, P.; Lamont-Black, J.; Hall, J.;

White, C.; Glendenning, S.; Russell, D. Electro-kinetic technology as a low-cost method

for dewatering food by-product. Drying Technology 2011, 29(14), 1721-1728.

16. Bajgai, T.R.; Hashinaga, F. High electric field drying of Japanese radish. Drying

Technology 2001, 19(9), 2291-2302.

18
17. Alemrajabi, A.A.; Rezaee, F.; Mirhosseini, M.; Esehaghbeygi, A. Comparative

evaluation of the effects of electrohydrodynamic, oven, and ambient air on carrot

cylindrical slices during drying process. Drying Technology 2012, 30(1), 88-96.

18. Bajgai, T.R.; Hashinaga, F. Drying of spinach with a high electric field. Drying

Technology 2001, 19(9), 2331-2341.

19. Esehaghbeygi A.; Basiry M. Electrohydrodynamic (EHD) drying of tomato slices

(Lycopersicon esculentum). Journal of Food Engineering 2011, 104(4), 628-631.

20. Li, F.D.; Li, L.T.; Sun, J.F.; Tatsumi, E. Effect of electrohydrodynamic (EHD)
Downloaded by [University of Montana] at 17:47 07 April 2015

technique on drying process and appearance of okara cake. Journal of Food Engineering

2006, 77(2), 275-280.

21. Bai, Y.; Yang, Y.; Huang, Q. Combined electrohydrodynamic (EHD) and vacuum

freeze drying of sea cucumber. Drying Technology 2012, 30(10), 1051-1055.

22. Dinani, S.T.; Havet, M.; Hamdami, M.; Shahedi, M. Drying of mushroom slices

using hot air combined with an electrohydrodynamic (EHD) drying system. Drying

Technology 2014, 32(5), 597-605.

23. Ahmedou, S.A.O.; Rouaud, O.; Havet, M. Assessment of the hydrodynamic

drying process. Food Bioprocess Technology 2009, (2), 240-247.

24. Dinani, S.T.; Havet, M. Drying kinetics and energy consumption of combined

convective-electrohydrodynamic EHD) drying of mushroom slices. In Proceedings of

International Drying Symposium IDS 2014. Lyon, France. Aug. 24-27, 2014. CD ROM.

25. Isobe, S.; Barthakur, N.; Yoshino, T.; Okushima, L.; Sase, S.

Electrohydrodynamic drying characteristic of agar gel. Food Science Technology

Research 1999, 5(2), 132-136.

19
26. Pogorzelski, M.; Zander, Z.; Zander, L.; Wrotniak, M. Drying kinetics of plant

material using the electrohydrodynamic (EHD) method. Chemical Engineering and

Equipment 2013, 44(6), 552-553 (in Polish).

27. Lai, F.C. A prototype of EHD-enhanced drying system. Journal of Electrostatics

2010, 68(1), 101-104.

28. Warburg, E. Über die stille entladung in gases (About silent discharge in gases).

In Handbuch der Physik; Geiger, H., School, K., Westphal, W. Eds.; Springer: Berlin-

Heidelberg, 1927; 149-170 (in German).


Downloaded by [University of Montana] at 17:47 07 April 2015

29. Yanallah, K.; Pontiga, F.; Chen, J.H. A semi-analytical study of positive corona

discharge in wire–plane electrode configuration. Journal of Physics D: Applied Physics

2013, 46, 345202 (12 p).

30. Hashinaga, F.; Kharel, G.P.; Shintani, R. Effect of ordinary frequency high

electric fields on evaporation and drying. Food Science Technology International 1995,

1(2), 77-81.

31. Kirschvink-Kobayashi, A.; Kirschvink, J.L. Electrostatic enhancement of

industrial drying process. Industrial and Engineering Chemistry Process Design and

Development 1988, 25, 1030-1033.

32. Chen, Y.; Barthakur, N.N.; Arnold, N. Electrohydrodynamic (EHD) drying of

potato slabs. Journal of Food Engineering 1994, 23(1), 107-119.

33. Xue, X.; Barthakur, N.N.; Alli, I. Electrohydrodynamically-dried whey protein:

an electrophoretic and differential calorimetric analysis. Drying Technology 1999, 17(3),

467-478.

20
34. Lai, F.C.; Sharma, R.K. EHD-enhanced drying with the multiple needle electrode.

Journal of Electrostatics 2005, 63(3-4), 223-237.

35. Wong, D.S.; Lai, F.C. EHD-enhanced drying with auxiliary heating from below.

Journal of Energy Resources Technology 2004, 126, 133-139.

36. Lai, F.C.; Wong, D.S. EHD-enhanced drying with needle electrode. Drying

Technology 2003, 21(7), 1291-1306.

37. Dalvand, M.J.; Mohtasebi, S.S.; Rafiee, S. Modeling of electrohydrodynamic

drying process using response surface methodology. Food Science&Nutrition 2014, 2(3),
Downloaded by [University of Montana] at 17:47 07 April 2015

200-209.

38. Karami, R.; Ayazi, M.; Samiee, L.; Mobarake, M.D.; Goodarzvand-Chegini, F.

An experimental study of the effect of high electric field on mass transfer enhancement.

Journal of Petroleum Science and Technology 2012, 2(2), 40-49.

39. Cao, W.; Nishiyama, Y.; Koide, S. Electrohydrodynamic drying characteristics of

wheat using high voltage electrostatic field. Journal of Food Engineering 2004, 62(3),

209-213.

40. Li, L.; Li, F.; Tatsumi, E. Effects of high voltage electrostatic field on evaporation

of distilled water and okara drying. Biosystem Studies 2000, 3, 43-52.

41. Bai, Y.; Hu, Y.; Li, X. Influence of operating parameters on energy consumption

of electrohydrodynamic drying. International Journal of Applied Electromagnetics and

Mechanics 2011, 35, 57-65.

42. Lai, F.C.; Lai, K.-W. EHD-Enhanced drying with wire electrode. Drying

Technology 2002, 20(7), 1393-1405.

21
43. Alem-Rajabi, A., Lai, F.C. EHD-enhanced drying of partially wetted glass beads.

Drying Technology 2005, 23(3), 597-609.

44. Martynenko, A. EHD drying of apple slices. Technical Report, October 2014

(unpublished).

45. Goodenough, T.I.J.; Goodenough, P.W.; Goodenough, S.M. The efficiency of

corona wind drying and its application to the food industry. Journal of Food Engineering

2007, 80(4), 1233–1238.

46. Yang, J.; Ding, C.J.; Bai, A.Z.; Liang, Y.Z. An application of drying technology with
Downloaded by [University of Montana] at 17:47 07 April 2015

high-voltage electric fields on bioproducts. Acta Scientiarum Naturalium Universitatis

NeMongol 2004, 35(5), 510-511.

47. Balcer, B.E.; Lai, F.C. EHD-enhanced drying with multiple-wire electrode.

Drying Technology 2004, 22(4), 821-836.

48. Basiry, M.; Esehaghbeygi, A. Electrohydrodynamic (EHD) drying of rapeseed

(Brassica napus L.). Journal of Electrostatics 2010, 68, 360-363.

49. Zander, L. Personal information. 2014. (lidia.zander@uwm.edu.pl)

50. Kudra, T. Energy aspects in drying. Drying Technology 2004, 22(5), 917-932.

51. van't Land, C.M. Industrial Drying Equipment: Selection and Application. Marcel

Dekker: New York, 1991.

52. Couper, J.R.; Penny, W.R.; Fair, J.R.; Walas, S.M. Dryers and cooling towers. In

Chemical Process Equipment: Selection and Design. Elsevier: Amsterdam 2005; 225-

276.

53. Martynenko, A. Porosity evaluation of ginseng roots from real-time imaging and

mass measurements. Food and Bioprocess Technology 2011, 4(3), 417-428.

22
Figure 1. Schematics of the EHD system in needle-to-plate configuration: 1- high voltage

power supply, 2- needle electrode, 3- ions, 4- corona wind cone, 5- drying material, 6-

plate electrode.
Downloaded by [University of Montana] at 17:47 07 April 2015

23
Figure 2. Projected area of corona wind impinging the surface of drying material: a)

single wire electrode, b) multi-needle electrode in staggered arrangement.


Downloaded by [University of Montana] at 17:47 07 April 2015

24
Figure 3. Drying curves of carrot and miscanthus at DC positive polarity; 7 needles, air

velocity 0.1 m/s (excerpt from [26])


Downloaded by [University of Montana] at 17:47 07 April 2015

25
Figure 4. Typical components of EHD setup in DC configuration: 1-voltage

controller/stabilizer, 2 -step-up transformer, 3 - rectifier, 4 - polarity switch, 5 - safety

resistor, 6 - capacitor, 7 - applicator. In AC configuration the applicator (7) is directly

connected to the step-up transformer (2); measuring instruments and air supply system

are not shown.


Downloaded by [University of Montana] at 17:47 07 April 2015

26

You might also like