Download as pdf or txt
Download as pdf or txt
You are on page 1of 55

Accepted Manuscript

Title: Quantitative Understanding of Refractance WindowTM


Drying

Author: Monica Jimena Ortiz Jerez Tushar Gulati Ashim K.


Datta Claudia Isabel Ochoa Martı́nez

PII: S0960-3085(15)00066-8
DOI: http://dx.doi.org/doi:10.1016/j.fbp.2015.05.010
Reference: FBP 611

To appear in: Food and Bioproducts Processing

Received date: 19-7-2014


Revised date: 8-5-2015
Accepted date: 23-5-2015

Please cite this article as: Jerez, M.J.O., Gulati, T., Datta, A.K., Martínez,
C.I.O.,Quantitative Understanding of Refractance WindowrmTM Drying, Food and
Bioproducts Processing (2015), http://dx.doi.org/10.1016/j.fbp.2015.05.010

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.
1
2 Research Highlights:

3  A mathematical model for understanding Refractance WindowTM Drying is


4 presented

t
5  “Window” aspect of Refractance WindowTM drying is not observed as claimed

ip
6 previously

cr
7  Conduction heat transfer is the dominant heating mode for this drying process

us
8  Radiation heat transfer contributes only a small fraction of the total heat flux

an
10

11
M
d
p te
ce
Ac

1 Page 1 of 54
11 Quantitative Understanding of Refractance Window™ Drying

12 Monica Jimena Ortiz Jerez


13 Graduate student, Universidad del Valle
14 School of Food Engineering

t
15 Calle 13 100-00, Cali, Colombia

ip
16 Tel.: +57 2 3212482; Fax: +57 2 3212392.
17 Email: monica.ortiz@correounivalle.edu.co

cr
18
19 Tushar Gulati

us
20 Graduate Student, Cornell University
21 Dept. of Biological and Environmental Engineering
22 Riley-Robb Hall, Ithaca, NY 14853, USA

an
23 Email: tg237@cornell.edu
24
25 Ashim K. Datta*
M
26 Professor, Cornell University
27 Dept. of Biological and Environmental Engineering
28 Riley-Robb Hall, Ithaca, NY 14853, USA
d

29 Tel.: +1 607 255 2482; Fax: +1 607 255 4080


30 Email: akd1@cornell.edu
te

31
32 Claudia Isabel Ochoa Martínez
p

33 Professor, Universidad del Valle


ce

34 School of Food Engineering,


35 A.A. 25360 Cali, Colombia
36 Tel.: +57 2 3212482; Fax: +57 2 3212392.
Ac

37 Email: claudia.ochoa@correounivalle.edu.co
38
39

40

41

42

43 * Address all correspondence to this author

2 Page 2 of 54
44 ABSTRACT

45 Refractance WindowTM (RWTM) drying is a novel drying technique in which the material to be dried is
46 placed on a thin plastic sheet and, the plastic sheet is heated from below via circulating hot water. Current
47 understanding suggests that the use of a thin plastic sheet that is transparent to infrared radiation (IR)
48 creates a "window" for thermal radiation from hot water to the wet material. This “window” gradually

t
ip
49 closes as the material dries out cutting off thermal radiation and prevents the sample to reach water bath
50 temperatures. It is suggested that radiation heat transfer is one of the dominant modes of heat transfer that

cr
51 promotes faster drying. However, no mathematical model-based mechanistic understanding of this
52 process is available to support the above hypotheses, which is the objective of the present work. A

us
53 conjugate heat and mass transfer model is developed to simulate the drying of pumpkin slices and
54 investigate in particular the effect of the optical properties of a commonly used plastic sheet (Mylar®) on
55 the radiative component of heat transfer. Computed results indicate that there is only a 5% increase in

an
56 transmission of IR radiation through Mylar between a dry and wet product. A major portion of thermal
57 energy is transferred via conduction (99%) through the plastic sheet. The relatively low sample
M
58 temperature observed for RWTM drying of thin samples is attributed to the development of a dried,
59 thermally resistive layer at the base that prevents heat transfer from the plastic sheet during the later
60 portions of the drying process. However, for thick-sized samples, the low sample temperature is a result
d

61 of development of air spaces between the product and the plastic sheet, which reduces the heat flux from
te

62 the hot water. As a consequence, quality of the final product is preserved when compared with other
63 drying techniques.
p

64
ce

65 Keywords: Refractance WindowTM drying, porous media, Mylar®, pumpkin

66
Ac

3 Page 3 of 54
66

67 1 Introduction

68 Refractance WindowTM (RWTM) drying is a more recent method for drying heat-sensitive products such as
69 fruit and vegetable purees, slices, juices into powders, flakes or sheets. Retention of color, flavor, aroma,

t
ip
70 vitamin and antioxidants can be improved in this drying process and product quality can generally be
71 comparable to or, in some instances, better than what can be achieved in freeze drying. The technique is

cr
72 also used to dry pharmaceuticals, nutraceuticals, cosmetic and pigments.

73 Literature studies of RWTM drying have focused on quality characteristics such as carotenoids in carrots

us
74 and ascorbic acid and color of strawberries (Abonyi et al., 1999); color, antioxidant activity of asparagus
75 (Nindo et al., 2003b) and aloe (Nindo et al., 2006); flavor, color, carotenoids, capsaicinoids of paprika

an
76 (Topuz et al., 2009 and 2011); encapsulated flavors of orange oil (Cadwallader et al., 2010); antioxidants
77 compound of tomatoes (Abul Fadl & Ghanem, 2011) and colored potatoes (Nayak et al., 2011); bioactive
78 compounds of potatoes (Kaspar et al., 2012) and microbial reduction of pumpkins (Nindo et al., 2003a);
M
79 Physical aspects studied of this drying technology include energy efficiency (Abonyi et al., 1999; Nindo
80 et al., 2003a, 2004; Nindo et al., 2006; Abul Fadl and Ghanem, 2011), moisture sorption properties such
d

81 as rehydration ratio of dried tomato juice (Abul Fadl & Ghanem, 2011) and mango juice (Capariño et al.,
te

82 2012), water activity of aloe (Nindo et al., 2006); effective diffusivity and shrinkage of mango slices
83 (Ochoa et al., 2012); isotherms of açai juice (Pavan et al., 2012), and microstructure analysis of tomato
p

84 (Capariño et al., 2012).


ce

85 Refractance WindowTM drying has primarily been studied to compare with other drying processes in
86 terms of various quality attributes such as nutrition, flavor, color, aroma and antioxidant compounds.
87 Studies focussing on engineering analysis of this drying process, which is the overall goal of this study,
Ac

88 are not available. More detailed discussions of this follow in the next section.

89 1.1 Current qualitative understanding of Refractance WindowTM drying

90 Here we present the best qualitative understanding of RWTM drying available today. RWTM drying was
91 patented by Magoon (1986) and developed by MCD Technologies, Inc., Tacoma, Washington, USA. As
92 shown in Figure 1, heat is transferred from hot water to a plastic conveyor belt with which the water is in
93 contact. The material to be dried is spread thinly on the plastic conveyor belt and thus moves continuously
94 with the belt at a desired speed (Capariño et al., 2012).

4 Page 4 of 54
95 Current qualitative understanding is as follows: when there is no product, the refractive index mismatch
96 between water and air is so high that radiative energy incident at the plastic-water interface is mostly
97 reflected back and very little of it is transmitted into (or lost to) the air. When the product to be dried,
98 typically a puree containing a significant (~90%) amount of water, is placed on the plastic, the refractive
99 index mismatch between the water and the food at the water-plastic-food system is greatly reduced. The

t
100 refractive indices become closer, reducing the reflection at this interface and increasing the transmission

ip
101 of the radiative energy into the food. The plastic material is considered a good transmitter of thermal
102 radiation (it is considered nearly transparent to thermal radiation at the relevant wavelengths). The

cr
103 refractive “window” refers to this allowance of thermal radiation to pass through the plastic material
104 when a wet material is in contact with it, i.e., as if the plastic material is not present, corresponding to the

us
105 window being open. This refractive “window” mentioned does not stay open for the entire duration of the
106 drying process. As the wet material dries, its refractive index increases creating a mismatch and
107 thereafter, most of the thermal radiation is reflected into water that stops drying.

108
109 an
Mylar® (Dupont Teijin Films) is a polyethylene terephthalate (PET) and is the plastic material typically
used for the RWTM drying. The amount of energy conducted and radiated varies with type and thickness
M
110 of the plastic sheet being used. Likewise, the absorbance of the puree depends on its thickness and
111 moisture content (Ratti and Mujumdar, 1995).
d

112 1.2 Difficulties with current qualitative understanding of RWTM drying and
te

113 lack of quantitative models


p

114 Qualitative understanding of the drying process is rather vague. For example, all the modes of conduction
ce

115 (through the plastic sheet into the food), convection (between water and the plastic, and between the food
116 and air) and radiation (through the conveyor material into the food) were considered utilized here (Nindo
117 et al., 2006; Nindo and Tang, 2007) but the contributions from each mode or their relative importance
Ac

118 were neither measured experimentally nor analyzed theoretically. It has been claimed that the use of a
119 thin plastic that is transparent to relevant infrared radiation allows efficient transfer of radiation into the
120 product compared to a thick plastic, but no information exists on the appropriate thickness. Radiative heat
121 transfer has been thought to decrease, making conduction heat transfer dominate as the product dries, this
122 reduced rate of heating protecting the product from overheating and resulting color and flavor degradation
123 (Bolland, 2000; Vega-Mercado et al., 2001). But again, no detailed measurement or analysis is available.

124 1.3 Lack of quantitative models

5 Page 5 of 54
125 Quantitative understanding of either the contribution of individual modes of energy or the moisture
126 transport details of the drying process is not available. Only experimentally measured moisture and
127 temperature history of thin samples have been reported and compared with other drying techniques
128 (Abonyi et al. 1999, 2001; Nindo et al., 2003a, 2007). Relative contributions of different heating modes
129 have not been reported to confirm that radiation is also a dominant mode for heat transfer to the product

t
130 being dried. Thermal efficiency of RWTM drying has been carried out taking into account the different

ip
131 heat transfer modes and comparing with other drying techniques (Nindo et al., 2003a). In this study,
132 authors mention that radiant heat transfer from the circulating hot water and conduction heat transfer from

cr
133 the plastic film are critical factors responsible for quick drying of samples. The only mathematical
134 analysis of RWTM drying that is available is a dimensional analysis-based model relating drying

us
135 coefficient and moisture ratio as function of time, for drying strawberry juices (Ghanem, 2010). In this
136 work, all three modes of heat transfer were considered during drying of strawberry juice on a glass plate
137 heated via hot water from beneath. Again, as before, there is no mention of the relative contributions from
138
139
an
different heating modes and does not confirm that radiation heat transfer makes a significant contribution
towards quick drying of samples.
M
140 1.4 Objectives and the organization of the manuscript

141 The objectives of this study are to: 1) develop a quantitative understanding of the contribution of thermal
d

142 radiation in this process; 2) develop a heat and mass transfer model of the RWTM drying process; 3)
te

143 validate the model for RWTM drying of a thin layer of wet material on the plastic conveyor by comparing
144 temperature and moisture histories with experimental studies by Nindo et al. (2003a); 4) perform RWTM
p

145 drying experiments on thick slabs of pumpkin to study the applicability of the drying process to thick
ce

146 sized samples and 5) compare experimental moisture and temperature during drying of thick samples with
147 the model predictions developed in (2) above.
Ac

148 We first estimate the magnitude of thermal radiative heat flux entering from the bottom surface of the
149 food material based on optical properties of Mylar®. Using this understanding as a boundary condition,
150 we develop a heat and mass transfer model for drying of a thin layer of pumpkin puree. We then extend
151 the drying model to thick sized samples of pumpkin dried using the RWTM drying technique and comment
152 on the applicability of the process to thick sized samples. Finally, a parametric sensitivity analysis of
153 some key parameters is carried out to study the effects of their variability as affecting moisture loss from
154 the material being dried.

155 2 Estimation of the Magnitude of Thermal Radiation

6 Page 6 of 54
156 The thermal radiative transport through the plastic film was estimated based on the published optical
157 properties of polyester films (i.e., Krimm, 1960; Fonseca et al., 2013). In a comparative study (Tsilingiris,
158 2003) of various polymer films, Mylar® had a total transmittance as shown in Table 1. Most of these
159 polymer films have a low transmission window between 7-12 μm , followed by significantly higher
160 transmission for wavelengths longer than 12 μm .

t
ip
161 Figure 2 shows the spectral transmissivity of the Mylar® film. As seen in Figure 2, there are bands of
162 high transmissivity (~90%) of thermal radiation near wavelengths of 3, 4.7 and 6  m. At these

cr
163 wavelengths, water has high absorption for infrared radiation (Sandu, 1986). The values reported in
164 Figure 2, however, show almost an opposite trend to those reported in a study by Tsilingiris (2003), for

us
165 example. Therefore, there are discrepancies in optical properties of Mylar® film. Possible causes of such
166 discrepancies can be attributed to: 1) thickness of the Mylar® sheet, 2) manufacturing processes and 3)
167 large variations in transmissivity within a small wavelength range. As noted by Tislingiris (2003), even
168
169
an
slight differences in thickness, chemical composition and measurement procedures can cause significant
changes in optical properties of the Mylar® sheet and thus there is discrepancy in data from various
M
170 sources as can be seen in Table 2. For more discussion, the reader is referred to the work of Tislingris
171 (2003).
d

172 Water at 90 °C can be approximated as a blackbody and the wavelength at which maximum emission
173 from water occurs is (Siegel and Howell, 2002):
te

2898.7 2898.7
174 max    7.98 μm (1)
p

T 90  273.15
ce

175 Spectral transmissivity values of Mylar® around this wavelength are shown in Table 3 calculated as
176 follows. Spectral reflectivity values of Mylar® were calculated from Eqn. (2) (Siegel and Howell, 2002):
Ac

2
   
177    0 1  (2)
  2  1 

178 where 1 and 2 are the refractive index of Mylar® and water, respectively, and 0 is the refractive
179 index of either air, water or food based on three different scenarios: Case I: Mylar® sheet is placed on the
180 hot water bath with air on the other side; Case II: Mylar® sheet is placed on the hot water bath with water
181 ( 0 =1.33) replacing air on the other side and; Case III: similar to Case I but, with air replaced by a model

182 food (a 75% concentrated sucrose solution) with 0 =1.4774 on the other side (see figure 3). This simple

7 Page 7 of 54
183 analysis would allow calculating the transmissivity of Mylar® sheet to transmit IR radiation and also to
184 comment on the “refractance window” aspect of the drying process.

185 Spectral absorptivity,  of the Mylar® was obtained from Eqn. 3 using absorption coefficient of
186 Mylar® at different wavelengths and thickness of the Mylar® sheet, d (Almeida et al., 2006). Data was
187 obtained directly from DuPont which is the manufacturer of Mylar® films (Tech Service Hopewell,

t
ip
188 Tech-Service.Hopewell@usa.dupont.com). They provided with the absorption coefficient, K, of Mylar®
189 sheet at different wavelengths. Reflectivity was assumed to be constant over the wavelength range

cr
190 considered (Almeida et al., 2006, Adamov & Savinich, 1979). Using the reflectivity and absorption
191 coefficient, spectral transmissivity was calculated from the following equations (Siegel and Howell,

us
192 2002):


193 1  exp   Kd  (3)
1 

194
an
       1 (4)
M
195 Table 3 shows the percentage transmissivity of Mylar® sheet at the four wavelengths considered. These
196 numbers were found to be consistent with those reported in literature (e.g. Tsilingiris et al., 2003) and
197 were pursued for further calculations. Although Eqn. 1 shows maximum intensity of thermal energy
d

198 emission from water at close to 8  m, Mylar® does not transmit any of the energy into food (zero
te

199 transmissivity at this wavelength, as shown in Table 3) while at other wavelengths, the transmissivity is
200 about 20% even when water is present on both sides of the Mylar® sheet (Case II). However, the most
p

201 critical observation from Table 3 is that the transmission seems to change so little (less than 5%) between
ce

202 water (CASE I) and a relatively dry product (CASE III) or with air (CASE I), that it makes it hard to
203 support the “refractance window” concept explained earlier in Section 1.1. The Mylar® sheet does not
204 transmit as much radiant heat from the hot water and thus questions the “window” aspect of RWTM drying
Ac

205 as mentioned in previous studies. Nevertheless, the results of this exercise would be used to provide
206 radiative heat flux boundary conditions when developing the heat and mass transport model for the RWTM
207 drying process. The magnitude of radiative heat flux would be compared with conductive heat flux from
208 the plastic sheet to comment on the relative contributions of both fluxes for which no quantitative data
209 exists in literature. Details on obtaining the amount of radiative heat flux from the hot water to the food
210 material through the Mylar® (based on results presented in Table 3) are discussed in Section 4.4.1.

211 3 Experimental Measurements

8 Page 8 of 54
212 For the present study, two different sets of experimental data for temperature and moisture were used to
213 validate the heat and mass transport model developed (Section 4) to describe the RWTM drying process.
214 For the first set, experimental studies conducted by Nindo et al. (2003a) on pilot scale RWTM drying
215 equipment, were used. Experiments were conducted on thin layers of pumpkin puree (0.4 mm - 0.6 mm)
216 that were dried at three different water bath temperatures (55 °C, 75 °C and 95 °C). For a detailed

t
217 description of their experimental methodology, the reader is referred to the work of Nindo et al. (2003a).

ip
218 The next set of experiments was carried out on thick samples of pumpkin (1.25 cm) to study the

cr
219 applicability of the RWTM drying for thick-sized samples since most studies have dealt with thin samples
220 (up to 5 mm). Fresh pumpkins procured from a local supermarket and weighing approximately 3.5 kg

us
221 were used in a state of maturation with a total soluble solids content of 5.27  1.7 °Brix, and initial
222 moisture content of 92.23  1.85% (wet basis). The fruit was washed, peeled and the seeds were removed
223 and the flesh was cut into large pieces parallel to the axis of the fruit. Thereafter, the slices were passed

an
224 through a manual cutter (Home Collection brand) with transverse cuts across the length to produce
225 samples having an approximate size of 6 x 1.1 x 1.25 cm3 (length x width x thickness).
M
226 RWTM drying experiments were conducted in a stainless steel-thermostatic bath with a capacity of 5.5
227 liters (Precision General Purpose Water Baths brand) and filled with tap water. The water surface was
228 covered with a plastic film (Mylar®) having an area of 17 x 30 cm2 and thickness (d) of 0.26 mm.
d

229 Samples (40 homogeneous pieces) were carefully and neatly arranged together on the film for each test.
te

230 All measurements were performed at ambient laboratory conditions (24 °C and 73% RH) and were made
231 in duplicate.
p

232 3.1 Temperature history measurements


ce

233 From among 40 pieces of pumpkin, 3 samples were randomly selected and 2 thermocouples were inserted
234 into each sample. Temperatures were recorded using a thermo-electronic recorder (data logger) with type
Ac

235 K thermocouples having an accuracy of  0.01°C and response time of 30 sec. The thermocouples were
236 inserted into the pumpkin to record temperatures at the bottom and top surfaces of the sample during
237 drying. The samples were placed on the Mylar® sheet with a perfectly controlled water bath temperature
238 of 902 °C. The bath temperature was recorded every 5 min with a mercury bulb thermometer (0-100
239 °C). A household fan was located 1.5 m in front of the system to blow air at a speed of 1 m/s parallel to
240 the samples, which was turned on just before the temperature recordings began. Air velocity
241 measurements were taken every 30 min. During experiments, it was ensured that the thermocouples did

9 Page 9 of 54
242 not leave the samples to prevent erroneous temperature recordings. The drying time of samples was
243 approximately 5 h.

244 3.2 Moisture history measurements

245 The initial moisture content of the samples was determined by oven method at 60 °C and constant

t
246 humidity for 24 h (n.20.013 AOAC, 1980). Moisture loss from samples during drying were carried out by

ip
247 weighing the samples at 10 minute intervals until five hours of drying using a digital weighing scale
248 (Ohaus ® Adventurer ™ brand) having a precision of 0.01 g.

cr
249 4 Heat and mass transfer modeling of the drying process

us
250 In this section, a multiphase porous media-based heat and mass transfer model of the drying process is

an
251 developed. Mass and energy conservation equations are developed that include diffusion, capillary and
252 convective modes of transport. The rate of radiative heat transfer to the food product through the plastic
253 film is captured as a heat flux boundary condition while conductive heat flux from the Mylar® (plastic)
M
254 sheet is obtained by solving a conjugate heat transfer problem between the food and Mylar®. Evaporation
255 is considered distributed throughout the domain. Key assumptions, governing equations, initial and
256 boundary conditions are discussed followed by input parameters and model implementation using
d

257 commercial finite element software. The variables of interest for predicting quality (color, nutrient loss
te

258 etc.) of materials dried using this technique are temperature and moisture histories.
p

259 4.1 Problem description


ce

260 A schematic of the problem description is shown in Figure 4a. Pumpkin (puree or slab) is assumed to be a
261 rigid porous media comprising of two transportable phases: liquid water and gas (mixture of air and
262 water–vapor). A 2D model is simulated with conjugate heat transfer between the plastic sheet and the
Ac

263 food material in order to directly obtain the conductive heat flux through the plastic; radiative heat flux,
264 however, is supplied as a boundary flux term at the food-plastic interface.

265 4.2 Assumptions

266 A number of assumptions need to be made in formulating the drying problem. There are three phases in
267 continuum: solid, liquid water and gas. Gas pressure is shared by all phases and the effects of capillary
268 pressure have been included as a diffusion term. Thermal equilibrium exists between all phases. There is
269 non-equilibrium between water in solid and water–vapor in the gas phase and, therefore, their

10 Page 10 of 54
270 concentrations are not given by the moisture isotherms. Bound water is ignored so all the liquid water is
271 available for transport. Radiative heat flux is used only as a boundary term in order to raise the
272 temperature of the bottom surface, i.e., all the radiation is assumed lost at the surface and not transmitted
273 within the food. Maximum value of transmissivity at  =4.7  m (from Table 3) is used to obtain the
274 transmitted radiative flux through the Mylar® sheet. Shrinkage during drying is not considered and the

t
275 effects of gravity have been ignored.

ip
276 4.3 Governing equations

cr
277 Mass Conservation equations are written for the different phases that exist in the system, they being liquid

us
278 water phase and gas phase (water vapor and air) in the food material:
279 Liquid water phase:
cw 
280    nw   I (5)

an
t

281 Water vapor in gas phase:


M
cv 
282    nv  I (6)
t
d

283 The volumetric concentration of vapor in pumpkin, cv, is expressed in terms of mass fraction of vapor and
te

284 concentration of gas phase as v cg , where v is the mass fractions of water vapor in the material. It is

285 assumed that the gas phase follows ideal gas law. In Eqns. 5 & 6 above, I denotes the evaporation rate of
p

286 liquid water to vapor (later). The different concentrations ( cw , cg ) are related to their respective saturation
ce

287 or volume fraction as ci  i Si ( i  w, g ). Here, i is the density (liquid water or gas),  is the total

288 porosity (gas + liquid water) inside the medium and Si is the saturation of the transportable phase inside
Ac

289 the pore:

Vi
290 Si  (7)
V

291 4.3.1 Fluxes


292 The liquid flux, nw , is due to the net pressure on liquid, which is the difference between the total gas
293 pressure (this can be significant if there is sufficient evaporation), P and the capillary attraction of the
294 water to the food matrix (a negative pressure), pc, on the liquid:

11 Page 11 of 54
 k k 
295 nw    w  in , w r , w  ( P  pc )  (8)
 w 
296 Capillary pressure is a function of temperature and moisture content of the porous material. As the
297 material dries out, it becomes substantially unsaturated resulting in very high capillary pressures inside.
298 Writing pc in terms of temperature and liquid water concentration, i.e., pc  pc (cw ,T ) , the flux of liquid

t
299 water can be written as:

ip
 k k k k  p p 
300 nw  - w in , w r , w P   w in , w r , w  c cw  c T  (9)
w  w  cw T 

cr
301 The first term on the right hand side denotes flow due to gas pressure gradients within the material and

us
302 the second and third terms denote flow due to capillarity. The effect of temperature on capillarity is
303 assumed small and neglected. Writing concentration of water in terms of its relative saturation
304 ( cw   w Sw ), the second term in Eqn. 9 above is written in terms of capillary diffusivity (Dw,cap) defined

an
305 as:
kin , w kr ,w pc
306 Dw,cap   (10)
 w S w
M
307 The gas phase comprises a binary mixture of air and water vapor. Concentrations of air and water vapor
308 during drying are obtained from their respective mass conservation equations by solving them in terms of
d

309 their mass fractions, a and v , respectively as:


te

310

  g S g v  
   (u g  g v )     S g
C g2 
M a M v Dbin xv   I (11)
t  g 
 
p

311 v  a  1 (12)
ce

312 Darcy’s law is used to solve for momentum conservation of different phases in a porous media and
313 replaces the standard Navier-Stokes equation. The velocity of different fluid phases is due to gradients in
Ac

314 gas pressure given by:


kin ,i kr ,i
315 ui   P (13)
i
316 where, i = w, g. The components of a phase (e.g., water vapor and air) share the same velocity.
317 The total pressure, P, is calculated by solving the overall mass balance equation for the gas phase:

 ( S g  g )  kin, g kr , g 
    g P   I
t  g 
318   (14)

319 4.3.2 Energy Balance

12 Page 12 of 54
320 Thermal equilibrium is assumed to exist between the different phases inside the food material and one
321 energy balance equation is solved for the mixture (Eqn. 15) to calculate the temperature, T:

T
322 eff c p ,eff    (( c p u ) fluid T )    (keff T )   I (15)
t
323 The properties of the mixture e.g. eff , c p,eff , keff are averages of phase properties weighted by either their

t
ip
324 mass or volume fractions:
325 eff  (1   )s   (Sw w  Sg g ) (16)

cr
326 c p,eff  xg (v c p,v  a c p,a )  xwc p,w  xs c p,s (17)

keff  (1   )ks   (Swkw  Sg (v kv a ka ))

us
327 (18)

328 (  c p u) fluid  ( wuw  Dw,cap cw )c p,w   g ug (v c p,v  a c p,a ) (19)

an
329 In order to obtain the conductive heat flux from the plastic sheet, conjugate heat transfer between the food
330 and plastic sheet is implemented. This requires solving for the energy balance equation for the plastic
331 sheet also. As will be explained in Section 4.4.1, conjugate modeling requires continuity of temperature
M
332 and continuity of heat flux at the common boundary between the food and plastic sheet.
333 The plastic sheet is heated via conduction due to the presence of hot water below. Heat transfer
334 within the plastic is obtained by solving:
d

Tmylar
335  mylar c p ,mylar    (kmylar Tmylar ) (20)
te

t
336 4.3.3 Evaporation rate
p

337 Evaporation of liquid water to vapor is assumed to be spatially distributed inside the food as opposed to
ce

338 being only a surface phenomenon. An explicit expression for the evaporation rate ( I ) is used due to its
339 ease of implementation in a commercial software (Halder et al., 2007). A non-equilibrium formulation in
Ac

340 which the evaporation rate is proportional to the difference between actual vapor density and density of
341 vapor at equilibrium (also known as saturation vapor density) is given as:

342 I  K evap (  v ,eq   v ) S g  (21)

343 where, v  gv is the vapor density at any location inside the material (obtained as a model output) and

344 v,eq is the saturation vapor density (obtained from moisture sorption isotherm of the material).

345 4.4 Boundary and initial conditions

13 Page 13 of 54
346 4.4.1 Boundary Conditions
347 Governing equations (Eqns. 5, 11, 14, 15 and 20) were solved for different temperatures of water bath
348 below the plastic sheet. A schematic of the process is shown in Figure 4a. To closely represent the
349 experimental conditions, the bottom surface of the food is considered to be heated via conduction from
350 the Mylar® sheet resting on the temperature controlled water bath and radiation from the hot water. The

t
351 top and side surfaces are exposed to the surrounding atmosphere and moisture exchange takes place via

ip
352 these surfaces. At the bottom surface, for the pumpkin puree, there is no moisture lost from the bottom
353 due to perfect contact with the plastic sheet. Therefore, the bottom surface for the puree sample is

cr
354 considered insulated for moisture exchange. However, for the pumpkin slab, a perfect contact was not
355 experimentally observed during experiments between the solid and the plastic sheet (see Figure 10).

us
356 Hence, moisture lost via evaporation is considered at the bottom surface as well; however, convective
357 moisture exchange is small compared to top and side surfaces due to reduced air flow at the base. This
358 effect is taken into account by reducing the mass transfer coefficient at the bottom surface by a factor of
359
360
an
10 with respect to the top surface. A sensitivity analysis on the reduction factor is carried out to comment
on the role the bottom surface plays towards moisture loss from the material.
M
361 The boundary conditions are then given as:
362 Top & side surfaces
d

363 B.C for Eqn. (5): nw,surf  hm Sw ( g ,surf v,surf  v,amb ) (22)
te

364 B.C. for Eqn. (11): nv,surf  hm Sg ( g ,surf v,surf  v,amb ) (23)
p

365 B.C. for Eqn. (14): Psurf  Pamb (24)


ce

366 B.C for Eqn. (15): qsurf  ht (Tamb  Tsurf )  (  c p,wT )nw,surf  c p,vTnv,surf (25)
Ac

367 Bottom surface

368 B.C for Eqn. (5): nw,surf  hm,bottom Sw ( g ,surf v,surf  v,amb ) (26)

369 B.C for Eqn. (11): nv,surf  hm,bottom S g ( g ,surf v,surf  v,amb ) (27)

370 B.C. for Eqn. (14): Psurf  Pamb (28)


371 B.C for Eqn. (15): qsurf  n  (kmylar Tmylar )  qrad  (  c p,wT )nw,surf  c p,vTnv,surf (29)
   
Conductive flux Radiative
flux

14 Page 14 of 54
372 Eqn. 22 & 23 (and 26 & 27) explain how the amount of liquid water and water vapor are convected away
373 from the different surfaces of the food material. Convective mass transfer coefficient is different at the top
374 (and side) and bottom surfaces, respectively, based on discussion above. At the top and side surfaces
375 (Eqn. 25), heat is convected away by the surrounding air, heat loss due to evaporation of liquid water, and
376 loss due to bulk flow of liquid water and water vapor.

t
377 Equation 29 is implemented at the conjugate boundary between food material and the bottom

ip
378 surface and its implementation require detailed explanation. Note that the bottom boundary is an interior
379 boundary that typically would not need a boundary condition specified. However, there is evaporative

cr
380 heat flux (third term) at this interior boundary, making heat flux discontinuous. The condition at this
381 boundary is implemented as follows: The two domains of plastic sheet and food are solved in sequence,

us
382 not simultaneously. 1) First, Eqn. 20 is solved only in the Mylar® domain, using water temperature, Tw,
383 as the bottom surface boundary condition and the food temperature (Eqn. 30) available from the previous

an
384 time step.
385 Tmylar |top  Tfood |surf (30)

386 2) Once Mylar® temperature is available, Eqns. 5-15 are solved only in the food domain. On the bottom
M
387 boundary of the food domain, Eqn. 29 is implemented. The first term in Eqn. 29, conductive heat flux, is
388 computed from the Mylar® temperature calculated in the previous step. The radiative flux from the water
d

389 beneath the Mylar® sheet is given by qrad . Based on discussion provided in Section 2 above, the amount
390 of radiative heat flux from the water bath to the food material is obtained as:
te

391 qrad   mylar  (Tw4  Tbottom


4
) (31)
p

392 where  mylar is the transmissivity of the Mylar® sheet at wavelength   4.7 μm for which the
ce

393 transmissivity is the highest. The emissivity of water (  ) is assumed as 0.95. The complete system is
394 assumed to be under atmospheric conditions (Eqns. 24 & 28).
Ac

395 4.4.2 Initial Conditions

396 Initial moisture content of pumpkin puree was specified as 6 (dry basis) (Nindo et al., 2003a) and a value
397 of 13.4 (dry basis) was obtained experimentally for thick pumpkin slabs. Initial liquid water
398 saturation S w 0 , was assumed 0.99 for the puree to avoid singularity during computations. However, for

399 the raw pumpkin slabs, initial liquid water saturation, S w 0 , and total porosity,  , were calculated based on

400 initial gas porosity ( S g 0  (1  S w 0 ) ) of 14.7% (Mayor et al., 2011) and initial moisture content of 13.4

401 d.b. using Eqn. 36. The initial amount of water vapor fraction ( v 0 ) is obtained from equilibrium vapor

15 Page 15 of 54
402 pressure inside the food material ( pv ,eq ) at initial dry basis moisture content (M0), vapor pressure of pure

403 water ( Psat ) at initial temperature of the food material ( T0  298.15 K ) given by the Clausius-Clayperon
404 equation and Dalton’s law (Eqns. 33-35). The equilibrium vapor pressure is obtained from moisture
405 sorption isotherm or water activity curve of the food material. Water activity of pumpkins as function of
406 moisture content has been measured experimentally (Molina Filho et al., 2011):

t
ip
0.5
  1  1 
2

407 aw  0.395426  0.018052    21.9051    547.392  
 (32)
 M  M 

cr
  

408 pv,eq  aw (M ) psat (T ) (33)

us
pv , eq 0
409 xv 0  (34)

an
Pamb

xv 0 M v
410 v 0  (35)
xv0 M v  (1  xv 0 ) M a
M
411 In Eqn. 32, M is the dry basis moisture content defined in terms of concentration of liquid water and
412 concentration of solids
d

 S 
te

cw
413 M  w w (36)
cs (1   )  s
p

414 4.5 Input parameters:


ce

415 Input parameters used for modeling the RWTM drying of pumpkin are summarized in Table 4. Transport
416 of liquid water due to capillarity has been modeled as diffusion. Capillary diffusivity of water inside the
Ac

417 pumpkin is written in terms of dry basis moisture content (Eqn. 37). Effective diffusivity of water in
418 pumpkins reported in published literature is in the range of 10-8 and 10-11 m2/s (Guiné et al., 2011).
419 Equation 37 is an empirical equation that covers the entire range of diffusivity from a low to high
420 moisture content and was arrived at from data available in literature. A similar methodology was used for
421 estimating the diffusivity of liquid water in potatoes by Ni (1997). Binary diffusion of water vapor (Eqn.
422 38) in air is a function of temperature and pressure inside the food material (Vargaftik, 1975):

423 Dw,cap  1 x 10 8 e ( 2.4 1.2 M ) (37)

16 Page 16 of 54
1.8
2.13  T 
424 Dbin    (38)
P  273 

425 Specific heat of liquid water, water vapor and air are functions of temperature, as is the thermal
426 conductivity of water (Halder et al., 2007)

t
427 c p , w  4176.2  0.0909(T  273)  5.4731x10 3 (T  273)2 (39)

ip
428 c p , a  1004.828  0.01185(T  273)  4.300 x 10 4 (T  273) 2 (40)

cr
429 c p ,v  1790  0.107(T  273)  5.856x10 4 (T  273)2 (41)

us
430 kw  0.57109  1.762x103 (T  273)  6.7036x106 (T  273)2 (42)

an
431 Among other parameters, permeability of the porous matrix is an important property that quantifies the
432 ability of a material to allow pressure driven (or bulk) flow of fluids through it. The nature of pores
433 inside the food material and transport properties of the fluid phase are key factors that govern the
M
434 permeability of a material. Unlike most other properties, very little data exists for permeability and, as
435 such, no predictive model is available (Gulati and Datta, 2013). Permeability of the material with respect
d

436 to a particular fluid phase (ki) is written in terms of intrinsic permeability (kin,i) and relative permeability
437 (kr,i) as:
te

438 ki  kin,i kr ,i (43)


p

439 Intrinsic permeability is a property of the material and depends on structural parameters such as porosity,
440 for example. Intrinsic permeability is difficult to measure for food systems since it requires extremely
ce

441 large pressures to pump the fluid across the food material. For pumpkins, no data on intrinsic
442 permeability exists. However, intrinsic liquid permeability of raw potato has been measured
Ac

443 experimentally as 1 x 1017 m2 (Datta, 2006) and is used as a starting value. This permeability value may
444 be an approximation but a sensitivity analysis around this value would either result in extremely large
445 pressures inside the food material at lower permeability values or very high rates of moisture loss from
446 the material at higher permeability values. Intrinsic permeability for the gas phase is obtained from liquid
447 water permeability via the Klinkenberg correction factor through the following (Warning et al., 2012):
 0.15( kin , w ) 0.37 
448 ki , g  kin , w  1   (44)
 P 
 

449 Relative permeabilities of gas and liquid phase are written in terms of their relative saturations, S g

17 Page 17 of 54
450 and Sw , defined as (Bear, 1972):

 S w  0.08 3
 S w  0.08
kr , w   0.92 
451 0
(45)
 S w  0.08

t
1  1.1S w S w  1 / 1.1
452 kr , g   (46)

ip
0 S w  1 / 1 .1

cr
453 Density of water is modeled as a function of temperature and is given by McCabe et al. (2005):

454 w  1001.4  0.1276(T  273)  0.0029(T  273)2 (47)

us
455 Both air and water vapor are considered as ideal gases and the density of the gas phase is obtained using
456 the ideal gas law:

an
PM g
457 g  (48)
RT

458 M g  a M a  v M v (49)
M
459 Here, xa and xv are the mass fractions of air and water vapor, respectively. The density of the solid
phase  s , was assumed a constant value (Table 4).
d

460
te

461 The heat transfer coefficient ht was calculated using the following relationship (Earle, 1983):
p

ht d
462 Nu   0.23Re0.8 Pr 0.4 (50)
ka
ce

463 The Chilton-Colburn analogy (Incropera and DeWitt, 1990) was used to calculate the mass transfer
464 coefficient at the top surface ( hm ). At the bottom surface of the pumpkin slab, the mass transfer
Ac

465 coefficient was reduced by a factor of 10 based on discussion above (Section 4.4.1):

ht Dbin Le 0.33
466 hm  (51)
kv

kv
467 Le  (52)
 g c p , g Dbin

hm
468 hm,bottom  (53)
10

18 Page 18 of 54
469 Other parameters, such as viscosity of liquid water and gas, were assumed constant. Latent heat of
470 vaporization was also assumed as a constant value. The entire set of input parameters are summarized in
471 Table 4.

472 4.6 Model implementation

t
473 Commercially available finite element software COMSOL Multiphysics 4.3a (COMSOL Inc., Burlington,

ip
474 MA) was used to solve the governing equations described in Section 4.3. The “Transport in Dilute
475 Species” module was used to solve for liquid water conservation in the pumpkin (Eqn. 5). For gas

cr
476 conservation, “Transport in Concentrated Species” module with “Maxwell-Stefan Diffusion and
477 Convection” was used to solve for the mass fraction of vapor (Eqn. 11). “Darcy’s law” (Eqn. 14) module

us
478 was used to solve for the pressure inside the system and “Heat Transfer” module was used to solve for
479 temperature inside the food material and the Mylar® sheet.

an
480 The computational domain comprised a 2D Mylar® sheet-food system created according the dimensions
481 mentioned in Table 4. The mesh consisted of mapped quadrilaterals with linear distribution. A mesh-
482 independent study was conducted to ensure that any dependent variable did not change by more than 1%
M
483 of the total change when the element size was reduced by half. Simulating for 10 minutes of drying thin
484 puree samples and 5 hour of drying thick pumpkin slabs using an adaptive time stepping required 15 min.
485 and 2.5h of CPU time on a desktop with Intel Xeon processor having 24 GB random access memory
d

486 (RAM). The results are presented and discussed below.


te

487 5 Results and Discussion


p

488 In this section, moisture loss and temperature history at the top surface of the food material are compared
ce

489 with experimental results available for drying a thin sheet of pumpkin puree at three different water bath
490 temperatures. Simulation results are compared with experimental data on moisture loss and temperature
Ac

491 histories obtained using thick samples of pumpkin. The relative contributions of conductive and radiative
492 fluxes from the hot water to the food material are discussed next. Finally, a sensitivity analysis of the
493 drying process to various parameters is presented.

494 5.1 Thin samples

495 5.1.1 Moisture profiles

496 Figure 5 shows a comparison between predicted and experimentally observed changes in dry basis
497 moisture content for the pumpkin puree sample for three different temperatures of the water 55°C, 75°C

19 Page 19 of 54
498 and 95°C. Predicted moisture content matches the experimentally observed values well, with maximum
499 difference between predicted and observed moisture content being less than 10%. At the end of the drying
500 process, the (computed) dry basis moisture contents of the sample are reduced to 1.36, 0.28 and 0.24 for
501 55°C, 75°C and 95°C, respectively. When the water bath temperature is 95°C, a drastic reduction in
502 moisture content of the puree is observed with the sample drying out in less than three minutes. For a

t
503 water bath temperature of 55 °C, the sample completely dries out in about five minutes, whereas for 55

ip
504 °C temperature of the water bath, the moisture content of the sample remains higher than its other
505 counter- parts.

cr
506

us
507 5.1.2 Temperature profiles

508 Predicted temperature histories at a point on the top surface of the sample are shown in Figure 6.

an
509 Experimental data from literature (Nindo et al., 2003a) was available only for one water bath temperature
510 (90°C) that is superimposed in Figure 6. The temperature profile as seen in Figure 6 is typical of a drying
511 process. A rapid rise in temperature of the puree top surface is reached during the initial portion of the
M
512 drying process. Thereafter, the temperature of the surface remains fairly constant at T = 70°C showing a
513 slight dip at around 300 seconds due to evaporative cooling. A gradual increase in temperature towards
d

514 the end of the process is observed due to decreased drying rates resulting in much less amounts of
515 evaporation.
te

516 Results from the model can be used to develop a mechanistic understanding of the heat transfer
517 process. Initially, the conductive flux from the water beneath the Mylar® sheet is very high due to large
p

518 differences in temperatures between the food material and water. This causes the temperature in the food
ce

519 material to shoot up very rapidly. Thereafter, evaporation of water causes the temperature to plateau until
520 much of the moisture is lost from the material. Once the evaporation rate decreases and with low moisture
521 diffusivity at lower moisture content values, moisture migration decreases resulting in a gradual increase
Ac

522 in temperature towards the end of the drying process.

523 5.1.3 Different water bath temperatures

524 The effect of final moisture content of thin samples with respect to different water bath temperatures is
525 shown in Figure 5. As the water bath temperature increases, the rate of moisture loss increases resulting in
526 quick drying of the sample. Figure 7 compares the temperature histories of the product surface
527 temperature heated for the three water bath temperatures. With increasing temperatures, the sample
528 temperature increases, however, in all cases, it stays well below the water bath temperature. The reason is

20 Page 20 of 54
529 both internal and external resistances to heat transfer dominate accompanied with evaporative cooling.
530 Figure 8 shows the variation of thermal conductivity of the sample with drying time at a water bath
531 temperature of 75°C. As is seen in Figure 8, there is a sharp decrease in the effective thermal conductivity
532 value ( keff ) of the material as it dries out. Figure 8 also shows the evolution of gas phase within the

533 material that is comprised mostly of air. Since air has a very low thermal conductivity, it therefore

t
534 decreases the overall thermal conductivity of the material to a low value. Forced convective heat transfer

ip
535 coefficient over the sample is estimated to be about 10 W/m2K and, for a thin sheet of material (puree in
536 this case), the volume to surface ratio is of the order of 1 mm. This gives a heat transfer Biot number

cr
537 ( BiT ) of about 0.25. A Biot number greater than 0.1 indicates that both internal and external resistances

us
538 contribute towards temperature changes within the material. Even with a relatively higher water bath
539 temperature (e.g., 95°C), the temperature of the material is comparatively low when compared with, for
540 example, hot-air drying. Such low product temperatures and shorter drying times can explain the better

an
541 color and nutrition retention as noted by Nindo and Tang (2007) since quality degradation reaction
542 kinetics are temperature and time dependent. Therefore, high water bath temperatures would result in
543 quick dehydration along with relatively low product temperatures, retaining better quality product.
M
544 5.1.4 Relative contribution of conductive and radiative fluxes
d

545 Current theories suggest that during RWTM drying, thermal energy from the hot water is transmitted
546 through the Mylar® sheet (or any other plastic membrane) via conduction and radiation (e.g., Nindo and
te

547 Tang, 2007). However, no data as such exists on the relative contributions of conductive and radiative
548 fluxes. Figure 9 compares the relative magnitudes of computed conductive flux (first term in Eqn. 29) and
p

549 radiative flux (Eqn. 31) experienced by the bottom surface of the sample (section A in Figure 1) for a
ce

550 water bath temperature of 90 °C. As is seen, the conductive flux is very high during the initial stages of
551 the drying process and is the major contributor for the rapid rise in the temperature of the sample. The
radiation flux, however, is roughly 1/100th of the conductive flux initially and stays lower during the
Ac

552
553 entire process. As drying proceeds, both conductive and radiative flux decrease due to increase in the
554 product temperature. Thus, conductive flux is primarily responsible for temperature rise in the sample.
555 Figure 6 further confirms this result, showing surface temperature of the puree with and without radiation
556 effects are about the same. Thus, radiation effects are indeed small and they do not contribute
557 significantly in the RWTM drying process, as claimed in previous publications.

558 To summarize, the proposed model is able to predict the transient temperature and moisture content for
559 drying thin samples. Conduction is a major role player during the RWTM drying process while the
560 contribution from radiation is almost negligible. The next section explores the applicability of the process

21 Page 21 of 54
561 to thick samples and sensitivity of the process to various important parameters in order to identify the rate
562 determining ones.

563 5.2 Thick samples

564 Most of the experiments conducted using RWTM drying have used thin samples (few mm thickness) since

t
565 superior product quality has been obtained using thin samples. However, no experimental data on

ip
566 temperature or moisture exist for thick samples. Therefore, drying experiments were conducted for
567 pumpkin slabs having average thickness of about 1.25 cm (Figure 10), a relatively large thickness

cr
568 compared with sample thin samples. Details of the experiment are in Section 3.5. The geometry was
569 redrawn in COMSOL according to the dimensions of the pumpkin and the model developed earlier for

us
570 thin puree was run for the thick sample for a total drying time of five hours.

571 5.2.1 Moisture profile

572
573 an
Figure 11 compares predicted (no air gap) and experimentally observed changes in moisture content for a
thick slab of pumpkin during five hours of RWTM drying. The model slightly over predicts the moisture
M
574 loss at all instants of time. This is possibly due to the high conductive heat flux from the Mylar® sheet as
575 predicted by the model causing rapid moisture removal from the pumpkin. In experiments, it was
d

576 observed that the bottom surface of the pumpkin slabs was not always in contact with the plastic sheet.
577 Figure 10 shows the pumpkin pieces at different times of the drying process. One can clearly see the
te

578 development of air spaces in between the bottom surface of the food and the plastic sheet. The material
579 underwent large deformations, shrinking and bending in this process. This resulted in the development of
p

580 air spaces between the bottom surface and the plastic sheet creating an additional resistance for heat
ce

581 transfer from the Mylar® sheet that prevents rapid moisture removal. From visual observations, the sides
582 of the material could be seen bending upwards up to 25 mm from the base during the drying process. This
583 air gap reduced the net heat flux from the Mylar® sheet, reducing rate of moisture loss from the material.
Ac

584 5.2.2 Temperature profiles

585 Figure 12 shows predicted point temperature histories at the top and bottom surfaces of the pumpkin slab
586 for various assumed air gaps. Experimental results are superimposed for comparison. For the bottom
587 surface, for all three air gap values, there is an initial rise in predicted temperature similar to what is seen
588 during experiments. This rise is followed by a plateau region due to intense evaporation of liquid water
589 that causes local cooling. Once much of the moisture is lost, the temperature starts to increase
590 approaching the water bath temperature. However, the predicted rise in temperature during later portions

22 Page 22 of 54
591 of drying was not observed in experiments. This is likely to be due to the progressive increase in the air
592 gap between the food surface and the plastic sheet as shown in Figure 10 that results in a progressive
593 decrease in heat flux as drying proceeds. For the top surface, a plateau region is seen for predicted
594 temperature for the first few hours into the drying process following an initial rapid rise. The dip in the
595 predicted temperature after a few hours is attributed to 1) very low thermal conductivity of the material at

t
596 the base that reduces the heat transfer to the top portion of the material and, 2) evaporative cooling. Figure

ip
597 13 shows the spatial profile of effective thermal conductivity, keff , at the instant when the predicted

598 temperature starts to decrease at the top surface in Figure 12. As is seen, the bottom surface dries out

cr
599 much faster resulting in a very thin thermally resistive layer that decreases the rate of heat transfer to the
600 top. At this stage moisture loss starts to occur from the top surface as well, contributing towards

us
601 decreasing the temperature of the top surface via evaporation. Once the material has lost all the moisture,
602 the predicted temperature at the top surface starts to increase following a trend similar to that of the

an
603 bottom surface.

604 5.2.3 Effect of air gap: Reformulation of the problem


M
605 In order to study the effect of air gap, an air domain between the food and the Mylar® had to be
606 incorporated, requiring reformulation of the computational problem, as shown in the new schematic of
d

607 Figure 4b. Figure 10 shows that as the material dries, it shrinks and curls, creating increasing gap
608 thickness that is also non-uniform across the length of the product. Effect of the thermal resistance of this
te

609 complex and evolving air gap had to be included in a simple way through an average gap thickness across
610 the entire length. It was estimated from visual observations that with time, the average thickness varied
p

611 linearly from 0.1 mm to 15 mm during drying. An air domain is therefore incorporated between the
ce

612 plastic sheet and the food material and conjugate boundary conditions (continuity of heat fluxes and
613 continuity of temperature) are provided at the boundaries between the plastic sheet and food with air
614 domain. Continuity of heat flux between plastic sheet-air and food-air interfaces is given as:
Ac

615
616 Air-sheet interface:


617 qair  sheet  n  (kair Tair ) (55)

618 Air-food interface:



619 qair  food  n  (kair Tair )  qrad  (  c p,wT )nw,surf  c p,vTnv,surf (56)

620 Note that Eqn. 56 is similar with Eqn. 29 except for the first term which now represents the conductive
621 heat flux from the air-domain in between the pumpkin slab and plastic sheet.

23 Page 23 of 54
622 Continuity of temperature is implemented at the respective interfaces as:

623 Air-sheet interface:

624 T |air  sheet  Tmylar (57)

625 Air-food interface:

t
ip
626 T |air  food  T food (58)

cr
627 Insulation boundary conditions are assumed at the side surfaces of the air domain. Temperature (and
628 moisture) histories in pumpkins were simulated for various air gap thicknesses (Figures 11 and 12). As is

us
629 expected, for a smaller air gap, the temperature rise in the material would be higher along with rapid
630 moisture removal. Further, an attempt was made to study the effect of variable air gap in order to better
631 predict temperature and moisture profiles. The increasing thermal resistance contributed by the increase in

an
632 air gap is incorporated in the model in a computationally efficient way as follows: the heat flux from the
633 Mylar® sheet to the bottom surface of the food via the air gap is written as:
M
Tair
634 qmylar  food  kair (59)
y
d

635 Assuming an imperfect contact between the food and Mylar® surface during the start of the drying
te

636 process, an initial small air gap thickness (corresponding to contact resistance) of y0 , would exist

637 at t  0 . Multiplying and dividing Eqn. 59 with y0 and rearranging the heat flux from the Mylar® to the
p

638 food surface at any instant can be written in terms of heat flux at the start of the drying process (i.e. at
ce

639 t  0 ) and variable air gap thickness ( y ) as follows:

y T
640 qmylar  food   kair 0 (60)
Ac

y y0
 
kvar

T
641   kvar (61)
y0

642 where kvar is the “variable” thermal conductivity of air that changes with air gap thickness. Figures 14
643 and 15 show that for the variable air gap thickness, the predicted temperature and moisture profiles are
644 much closer to the experimentally measured values than in Figures 11 and 12 for fixed air gap
645 thicknesses, i.e., model predictions improve significantly when air gap thickness is allowed to vary with

24 Page 24 of 54
646 time. These comparisons indicate that air gap indeed plays a major role during contact drying of thick
647 sized materials especially during the later portions of the drying process. The drastically reduced heat flux
648 prevents rapid rise in temperature and resulting rapid moisture removal. This is in contrast with drying
649 thin materials with either perfect contact or with very little air gap (Figures 6 and 12). The conductive
650 heat flux from the Mylar® sheet is much lower compared to that with perfect contact and decreases with

t
651 air gap thickness, decreasing the temperature profile at the bottom surface in Figure 14. For better

ip
652 predictions, the deformation aspect of the material during drying will have to be incorporated into the
653 modeling framework which is beyond the scope of the present study.

cr
654 For thick samples, drying takes place primarily by conduction heat transfer from the hot plastic sheet as

us
655 was with thin samples. As drying proceeds, developing air gap decreases the net rate of heat transfer from
656 the plastic sheet. This could explain how this drying prevents the sample to reach high temperatures and
657 results in improved quality (color and nutrients) of the dried product.

658 5.3 Sensitivity Analysis


an
M
659 In this section the effect of important process parameters on the amount of moisture loss is discussed. In
660 particular, effect of liquid water permeability of the pumpkin and mass transfer coefficient value used at
661 the bottom surface of the pumpkin slab are described. Results discussed in this section are based on the
d

662 model developed that includes the effects of varying air gap thickness between the Mylar® sheet and the
te

663 food material.

664 5.3.1 Effects of liquid water permeability


p

665 There is no data reported in literature for liquid water permeability in pumpkins and a value of
ce

666 1017 m 2 was used as discussed in Section 4.5. Permeability measurements of biological materials are still
667 not a standard technique owing to several difficulties in allowing the fluid phases to pass through the
Ac

668 material (Halder et al., 2011). Also, a large variation in permeability values in performing sensitivity
669 analysis is not useful without simultaneously varying other properties. Water permeability values were
670 varied by  50% of the original value and the corresponding gas permeabilities were updated based on
671 Eqn. 44. As shown in Figure 16, change in permeability values did not affect the total moisture loss as
672 much since capillary diffusion dominates over pressure driven flow.

673 5.3.2 Mass transfer coefficient at the bottom surface

674 As was noted in Section 4.4.1, the bottom surface has a reduced mass transfer coefficient compared to the
675 top and side surfaces. Computed results in Figure 17 compares the moisture history for various values of

25 Page 25 of 54
676 reduced mass transfer coefficient at the bottom surface, showing sensitivity at later times. During the first
677 1.5 h of drying, capillary diffusivity is high and capillary diffusion controls moisture loss from the
678 material. However, during later portions, a high mass transfer coefficient value decreases the resistance to
679 moisture removal from the material and thus leads to higher rates of moisture loss.

680 6 Conclusions

t
ip
681 A first-principles based understanding of RWTM drying of foodstuffs is developed. A conjugate heat and

cr
682 mass transfer model is developed to simulate the drying of pumpkin slices and investigate in particular the
683 effect of the optical properties of a commonly used plastic sheet (Mylar®) on the radiative component of

us
684 heat transfer. Only 5% increase in transmission of IR radiation occurs when a dry material is replaced by
685 a wet material thereby questioning the “window” aspect of RWTM drying process as reported in literature.
686 For thin sample sizes, the moisture loss is sensitive to water bath temperatures. Relatively low sample

an
687 temperatures are attributed to a sharp decrease in the effective thermal conductivity of the food material
688 preventing the product temperature to approach water bath temperatures. Conduction is found to be the
689 prime mode of heat transfer to the material while thermal radiation contributes to only 1% of the total heat
M
690 transfer. For thick sized samples, continuous deformation of the food material leads to a progressive
691 development of air spaces between the food and the plastic sheet. This results in a decrease in the net heat
d

692 flux to the food material keeping the temperature of the material well below the water bath temperature. A
693 sensitivity study on mass transport parameters indicates that moisture loss does not depend as much on
te

694 intrinsic permeability of liquid water while it is reasonably sensitive to the mass transfer coefficient
695 values at the bottom surface of the material.
p

696
ce

697 Acknowledgements
Ac

698 The first author is thankful to COLCIENCIAS (Colombian Institute for the Development of Science and
699 Technology, Francisco José de Caldas) for the financial assistance received during the course of her
700 Doctoral studies at Universidad del Valle.

701

702

703

26 Page 26 of 54
704

705

706

t
707

ip
708

cr
709

us
710

711

712 Nomenclature
an
M
713 Symbol Description Units

714 aw water activity -


715 Bi Biot Number -
d

716 c concentration kg/m3


te

717 Cg molar density kmol/m3


718 cp specific heat J/kg  K
p

719 D diffusivity m2 /s
ce

720 d thickness of Mylar® sheet m


721 hm mass transfer coefficient m/s
722 ht heat transfer coefficient -
Ac

723 I rate of evaporation kg/m3  s


724 K absorption coefficient -
725 k thermal conductivity W/m  K
726 kin intrinsic permeability m2
727 Kevap evaporation rate constant 1/s
728 Le Lewis Number -
729 MC moisture content (dry basis) kgH2O /kgdry solid
730 n mass flux kg/m2  s

731 n unit normal -

27 Page 27 of 54
732 Nu Nusselt Number -
733 P gas pressure Pa
734 p partial pressure Pa
735 pc capillary pressure Pa
736 Pr Prandtl Number -
737 q energy flux W/m2

t
738 Re Reynold’s Number -

ip
739 S saturation of fluid phase, i -
740 t time s

cr
741 T temperature K
742 u velocity of fluid phase, i m/s
743 volume m3

us
V
744 x mole fraction -
745 y value on Y-axis m
746

an
747
748
749 Greek symbols
M
750  wavelength m
751  spectral reflectivity -
752  refractive index -
d

753  spectral absorptivity -


754  spectral transmissivity -
te

755  mass fraction -


756  porosity -
p

757  viscosity Pa-s


758  density kg/m3
ce

759  emissivity -
760  Stefan-Boltzmann constant (5.67x 108 ) W/m2  K 4
761  gradient 1/m
Ac

762
763 Subscripts
764
765 w,v ,a , g , s water, vapor, air, gas, solid
766 r relative
767 eff effective
768 surf surface
769 eq equilibrium
770 amb ambient
771 sat saturation
772 rad radiation
773 bin binary

28 Page 28 of 54
774 var variable
775
776
777
778
779
780
781

t
782

ip
783
784

cr
785

us
786

787

an
788

789 References
M
790 Abonyi, B., Tang, J., & Edwards, C. 1999. Evaluation of Energy Efficiency and Quality Retention for the
791 Refractance Window™ Drying System. Research Report. Retrieved January 2008, from
792 <http://www.docstoc.com/docs/2913708/Draft-Report-on-the-Refractrance-WindowTM>.
d

793 Abul-Fadl, M., & Ghanem, T. 2011. Effect of Refractance Window (RW) drying method on quality
te

794 criteria of produced tomato powders as compared to the convection drying method. World
795 Applied Sciences Journal, 15(7), 953-965.
p

796 Adamov, D., & Savinich, V. 1979. Temperature dependence of polymer film emissivity. Translated from
ce

797 Inzhenerno-Fizicheskii Zhurnal, 37(2), 285-288.

798 Almeida, M., Torrance, K., & Datta, A. 2006. Measurement of optical properties of foods in near- and
799 mid-infrared radiation. International Journal of Food Properties, 9, 651–664.
Ac

800 Bolland, K. 2000. A new low temperature/short time drying process. Cereal Foods World, 45, 293-296.

801 Cadwallader, K., Moore, J., Zhang, Z., & Schmidt, S. 2010. Comparison of spray drying and Refractance
802 WindowTM drying tecnologies for the encapsulation of orange oil. In C.T. Ho, C. J. Mussinan, F.
803 Shahidi, & E. T. Contis, Recent Advances in Food and Flavor Chemistry: Food Flavors and
804 Encapsulation (p.p. 246-254). The Royal Society of Chemistry.

805 Capariño, O., Tang, J., Nindo, C., Sablani, S., Powers, J., & Fellman, J. (2012). Effect of drying methods
806 on the physical properties and microstructures of mango (Philippine ‘Carabao’ var.) powder.
807 Journal of Food Engineering, 111, 135–148.

29 Page 29 of 54
808 Choi, Y., Okos, M.R., 1986. Thermal properties of liquid foods ‒ review. In: Okos, M.R.(Ed.), Physical
809 and Chemical Properties of Food. American Society of Agricultural Engineers, St. Joseph, MI,
810 USA.

811 Datta, A.K., (2006). Hydraulic permeability of food tissues. International Journal of Food Properties,
812 9(4): 767–780.

t
813 Dongfeng, W., Zhuoyan, H., Kai, Y., Shanshan, C., & Xiaolin, Y. (2009). China Academic Journal

ip
814 Electronic Publishing House. Transactions of the CSAE 25(10). Recuperado el 2013, de
815 http://www.cnki.net

cr
816 Earle, R.L., (1983). Unit operations in food engineering, Pergamon Press, Oxford, 46 pp

817 Fonseca Caetano, V., Vinhas, G., Pimentel, M. d., & Ugulino de Araújo, M. (2013). Prediction of

us
818 mechanical properties of polyethylene terephthalate: using Infrared Spectroscopy and
819 Multivariate Calibration. Journal of Applied Polymer Science, 127 (5), 3441-3446.

820 Ghanem, T.H. (2010). Modeling of Refractance Window film dryer for liquids. Misr Journal of

an
821 Agricultural Engineering, 27 (2), 676-687

822 Guine, R.P.F.; Pinho, S.; Barroca, M.J. (2011). Study of the convective drying of pumpkin (Cucurbita
M
823 maxima). Food and Bioproducts Processing, 89, 422-428.

824 Gulati, T. & Datta, A.K. (2013). Enabling computer-aided food process engineering: property estimation
825 equations for transport phenomena-based models. Journal of Food Engineering, 116(2), 483-504.
d

826 Halder, A., Datta, A.K. & Spanswick, R.M. (2011). Water transport in cellular tissues during thermal
te

827 processing. AIChE Journal, 57, 2574-2588.

828 Halder, A., Dhall, A., & Datta, A. K. (2007). An improved, easily implementable, porous media based
p

829 model for deep-fat frying - Part I: Model development and input parameters. Food and
830 Bioproducts Processing, 85(C3), 209-219
ce

831 Incropera, F.P., DeWitt, D.P., (1990). Fundamentals of Heat and Mass Transfer, third ed. John Wiley &
832 Sons, New York.
Ac

833 Kaspar, K., Park, J., Mathison, B. B., M., S., & Boon, P. (2012). Processing of pigmented-flesh potatoes
834 (Solanum tuberosum L.) on the retention of bioactive compounds. International Journal of Food
835 Science and Technology, 47, 376–382.

836 Krimm, S. (1960). Infrared Spectra of High Polymers. Fortschr. Hochpolym. Forsch., 2, 51-172.

837 Lewis, M., 1987. Physical Properties of Foods and Food Processing Systems. VCH, Deerfield Beach, FL.

838 Magoon, R. (1986). Method and Apparatus for Drying Fruit Pulp and the Like. US Patent 4,631,837.

839 Mayor, L., Moreira, R., & Sereno, A.M. (2011). Shrinkage, density, porosity and shape changes during
840 dehydration of pumpkin (Cucorbita pepo L.) fruits. Journal of Food Engineering, 103, 29–37.

30 Page 30 of 54
841 McCabe, W., Smith, J., Harriott, P., (2005). Unit Operations of Chemical Engineering, seventh ed.
842 McGraw-Hill, Boston, MA.

843 Michailidis, P., Krokida, M., & Rahman, S. (2009). Data and models of density, shrinkage, and porosity
844 (Chap. 14). In S. Rahman, Food Properties Handbook, Second Edition (pp. 418-490). Boca
845 Raton, FL: Taylor & Francis Group, LLC.

t
846 Nayak, B., Berrios, J., Powers, J., Tang, J., & Ji, Y. (2011). Colored potatoes (solanum tubersum L.) dried

ip
847 for antioxidant-rich value-added foods. Journal of Food Processing and Preservation, 35, 571-
848 580.

cr
849 Ni, H., (1997). Multiphase moisture transport in porous media under intensive microwave heating, PhD
850 thesis, Cornell University.

us
851 Nindo, C., & Tang, J. (2007). Refractance WindowTM dehydration technology: a novel contact drying
852 method. Drying Technology, 25 (1), 37–48.

853 Nindo, C., Feng, H., Shen, G., Tang, J., & Kang, D. (2003a). Energy utilization and microbial reduction

an
854 in a new film drying system. Journal of Food Processing and Preservation, 27 (2) , 117–136.

855 Nindo, C., Sun, T., Wang, S., Tang, J., & Powers, J. (2003b). Evaluation of drying technologies for
856 retention of physical quality and antioxidants in asparagus (Asparagus officinalis, L.).
M
857 Lebensmittel Wissenschaft und Technologie, 36, 507–516.

858 Nindo, C., Tang, J., Cakir, E., & Powers, J. (2006). Potential of Refractance Window Technology for
d

859 Value Added Processing of Fruits and Vegetables in Developing Countries . ASABE Annual
860 International Meeting Sponsored. Portland, Oregon.
te

861 Nindo, C., Tang, J., Cakir, E., & Powers, J. (2006). Potential of Refractance Window Technology for
862 Value Added Processing of Fruits and Vegetables in Developing Countries. ASABE Annual
p

863 International Meeting Sponsored. Portland, Oregon.


ce

864 Nindo, C., Tang, J., Powers, J., & Bolland, K. (2004). Energy consumption during Refractance Window
865 evaporation of selected berry juices. International. Journal of Energy Research, 28 , 1089–1100.

866 Ochoa-Martínez, C., Quintero, P., Ayala, A., & Ortiz, M. (2012). Drying characteristics of mango slices
Ac

867 using the Refractance Window™ technique. Journal of Food Engineering, 109, 69–75.

868 Pavan, M., Schmidt, S., & Feng, H. (2012). Water sorption behavior and thermal analysis of freeze-dried,
869 Refractance Window-dried and hot-air dried açaí (Euterpe oleracea Martius) juice. LWT - Food
870 Science and Technology, 48, 75-81.

871 Ratti, C., & Mujumdar, A. (1995). Infrared Drying. En A. Mujumdar, Handbook of Industrial Drying
872 (Vol. Vol. 1, pages. 1–39). New York: Ed. Marcel Dekker.

873 Ruiz-López, I.I., Córdova, A.V., Rodŕiguez-Jimenes, G.C. & Garćia-Alvarado, M.A. (2004). Moisture
874 and temperature evolution during food drying: effect of variable properties. Journal of Food
875 Engineering, 63, 117–124.

31 Page 31 of 54
876 Sandu, C. (1986). Infrared radiative drying in food engineering: A process analysis. Biotechnology
877 Progress, 2, 109–119.

878 Siegel, R., & Howell, J. (2002). Thermal Radiation Heat Transfer (4th ed.). New York: Taylor & Francis.

879 Tech Service Hopewell, Tech-Service.Hopewell@usa.dupont.com

880 Topuz, A., Dincer, C., Özdemir, K., Feng, H., & Kushad, M. (2011). Influence of different drying

t
881 methods on carotenoids and capsaicinoids of paprika (Cv., Jalapeno). Food Chemistry , 129, 860–

ip
882 865.

cr
883 Topuz, A., Feng, H., & Kushad, M. (2009). The effect of drying method and storage on color
884 characteristics of paprika. LWT - Food Science and Technology, 42, 1667–1673.

us
885 Tsilingiris, P. (2003). Comparative evaluation of the infrared transmission of polymer films. Energy
886 Conversion and Management, 44, 2839–2856.

887 Vargaftik, N., (1975). Tables on thermophysical properties of liquids and gases, second ed. Hemisphere,

an
888 Washington, DC, 1975.

889 Vega-mercado, H., Góngora-Nieto, M., & Barbosa-Cánovas, G. (2001). Advances in dehydration of
890 foods. Journal of Food Engineering, 49, 271-289.
M
891 Warning, A., Dhall, A., Mitrea, D., & Datta, A.K. (2012). Porous media based model for deep-fat vacuum
892 frying potato chips. Journal of Food Engineering, 110, 428-440.
d

893
te

894

895
p

896
ce

897
Ac

898

899

900

901

902

903

32 Page 32 of 54
904

905

906

907

t
ip
908

cr
909

910

us
911

an
912

913 Table 1. Minimum total transmittance of Mylar®, occurring near the wavelengths of 6.5 and 14.5  m
M
Source temperature Mylar® film Total transmittance
for thermal radiation thickness (mm) (%)
60°C 0.1 15
d

0.2 5
te

90°C 0.1 16
0.2 7
p

914
ce

915

916
Ac

917

918

919

920

921

922

33 Page 33 of 54
923

924

925

926

t
ip
927

cr
928

929

us
930

an
931 Table 2. Transmittance of Mylar®, occurring near the wavelengths of 3, 4.7, 6 and 8  m from different
932 sources in literature
M
Wavelength,  Transmission (%)
(m) Dupont Krim (1960) Tsilingiris (2003)
3 88 80 63
4.7 90 85 45
d

6 75 55 4
8 0 10 0
te

933

934
p
ce

935

936
Ac

937

938

939

940

941

942

34 Page 34 of 54
943

944

945

946

t
ip
947

cr
948

949

us
950

an
951 Table 3. Mylar transmissivity at wavelengths corresponding to high absorptivity or low transmissivity

Wavelength,  Spectral transmissivity,  (%)


(m) Mylar® - air Mylar® - water Mylar® - food*
M
3 4.5 9.2 10.2
4.7 14.8 19.4 20.5
6 3.4 8.1 9.1
8 0 0 0
d

952 *considering food as a concentrated solution of sucrose at 75%


te

953
p

954
ce
Ac

35 Page 35 of 54
954

955 Table 4. Input parameters for the model


956 Parameter Symbol Value Units Source
957 Dimensions
958 puree 5x 104 x 0.3 m Nindo et al. (2003a)
959 slab 0.0125 x 0.06 m this work

t
ip
960 Optical properties of Mylar®
961 Relectivity

 ,1

cr
962 Air-Mylar® 0.0598 this work

963 Water-Mylar® ,2 0.0132 this work



us
964 Absorptivity Eqn. 3 Almeida et al. (2006)
965 Transmissivity  Eqn. 4
966 Density
967 Water w Eqn. 47 kg/m3 McCabe et al. (2005)

an
968 Vapor v Ideal gas kg/m 3

969 Air a Ideal gas kg/m3


M
970 Solid s 1533 kg/m3 Michailids et al. (2009)

971 Mylar®  mylar 1430 kg/m3


972 Specific heat capacity
973
d
Water c p, w Eqn. 39 J/kg  K Lewis (1987)

974 Vapor c p ,v Eqn. 41 J/kg  K Lewis (1987)


te

975 Air c p,a Eqn. 40 J/kg  K McCabe et al. (2005)

976 Solid c p ,s 1750 J/kg  K Ruiz-Lopez et al. (2004)


p

977 Mylar® c p,mylar 1600 J/kg  K


ce

978 Thermal conductivity


979 Water kw Eqn. 42 W/m  K Choi & Okos (1986)
980 Vapor kv 0.026 W/m  K Choi & Okos (1986)
981
Ac

Air ka 0.026 W/m  K Choi & Okos (1986)


982 Solid ks 0.05 W/m  K Ruiz-Lopez et al. (2004)
983 Mylar® kmylar 0.19 W/m  K
984 Viscosity
985 Water w 103 Pa.s McCabe et al. (2005)

986 Vapor & Air g 1.8 x 10


5
Pa.s McCabe et al. (2005)
987 Intrinsic permeability
988 Water kin,w 1017 m2

989 Air & Vapor kin, g Eqn. 44 m2 Warning et al. (2012)


990 Relative permeability
991 Water kr , w Eqn. 45 - Bear (1972)

36 Page 36 of 54
992 Vapor & air kr , g Eqn. 46 - Bear (1972)
993 Porosity
994 puree  0.90
995 slab  0.96
996 Capillary Diffusivity Dw,cap Eqn. 37 m 2 /s this work
997 Emissivity of water  0.95
998 m 2 /s

t
Binary diffusivity of vapor Dbin Eqn. 38 Vargaftik (1975)

ip
999 Heat transfer coefficient ht Eqn. 50 W/m .K2
Earle (1983)
1000 Mass transfer coefficient hm Eqn. 51 m/s Incropera & DeWitt (1990)
1001 Latent heat of vaporization  2260 kJ/kg

cr
1002 Water activity aw Eqn. 32 Molina Filho et al. (2011)
1003 Evaporation rate constant Kevap 1000 1/s Haldar et al. (2007)

us
1004

1005

1006

an
M
1007

1008
d

1009
te

1010
p

1011
ce

1012

1013
Ac

1014

1015

1016

1017

1018

37 Page 37 of 54
1019

1020 Figures

t
ip
cr
us
an
M
d
te

1021
p

1022 Figure 1. Refractive WindowTM drying showing suggested modes of heat transfer to the food from the
ce

1023 hot water bath. Convection and conduction heat at the bottom surface of Mylar® (plastic)
1024 sheet, conduction from the plastic sheet, thermal radiation from the hot water through the
1025 plastic sheet and convection at the top surface of food material.
Ac

1026

38 Page 38 of 54
t
ip
cr
us
1027

1028 Figure 2. Spectral transmissivity of Mylar® film, covering the wavelength ranges of (a) 2.5-7µm and (b)

an
1029 7-50 µm, respectively
1030 (http://usa.dupontteijinfilms.com/informationcenter/downloads/Optical_Properties.pdf)
1031
M
1032
1033
1034
1035
d

1036
p te
ce
Ac

39 Page 39 of 54
t
ip
cr
us
an
M
1037

1038 Figure 3. Current, literature-based understanding of the refractance window concept in RWTM drying.
d

1039 Shown are the two extremes: (a) water-Mylar-air (or food) system that would be representative
te

1040 of no product or a very dry product, and (b) water-Mylar-wet food product. Transmission of
1041 infrared through the Mylar sheet is expected to be much higher in (b) than (a).
1042
p

1043
1044
ce

1045
1046
1047
Ac

1048

40 Page 40 of 54
t
ip
cr
1049

1050 Figure 4a. Schematic and boundary conditions of the process (not drawn to scale)

us
1051

1052

an
M
d
p te
ce

1053
Ac

1054 Figure 4b. Schematic and boundary conditions of the process with the inclusion of a small air gap
1055 between the pumpkin slab and Mylar® sheet (not drawn to scale).
1056

1057

41 Page 41 of 54
t
ip
cr
us
an
M
d

1058
te

1059 Figure 5: Computed and measured (experimental data from Nindo et al. (2003a)) moisture content in the
1060 puree as function of drying time.
p
ce
Ac

42 Page 42 of 54
t
ip
cr
us
an
M
d

1061
te

1062 Figure 6: Computed and measured (experimental data from Nindo et al. (2003a)) temperature at the top
1063 surface of the puree sample at different times with and without radiation effects.
p
ce
Ac

43 Page 43 of 54
t
ip
cr
us
an
M
d

1064
te

1065 Figure 7: Computed temperatures at the top surface of puree sample heated at three different water bath
1066 temperatures (55 °C, 75 °C and 95 °C).
p
ce
Ac

44 Page 44 of 54
t
ip
cr
us
an
M
d

1067
te

1068 Figure 8: Evolution of effective thermal conductivity (keff) and gas saturation (Sg) within the puree
1069 sample with drying time during RWTM drying at water bath temperature of 75 °C. The gas
p

1070 phase comprises primarily of air that contributes towards low effective thermal conductivity of
1071 the puree sample
ce
Ac

45 Page 45 of 54
t
ip
cr
us
an
M
d

1072
te

1073 Figure 9: Computed conductive and radiative flux through the Mylar® sheet. Conduction is the prime
1074 mode of heat transfer with radiation playing a very little role in the RWTM drying process
p
ce
Ac

46 Page 46 of 54
t
ip
cr
us
an
M
d

1075
te

1076 Figure 10: Snapshots of the dried pumpkin slabs at different instants of time: (a) 0 min (b) 30 min (c) 1 h
1077 and (d) 5 h. One can clearly see the development of air spaces between the plastic sheet and
p

1078 food material with drying time


ce
Ac

47 Page 47 of 54
t
ip
cr
us
an
M
d
te

1079

1080 Figure 11: Computed and experimental moisture content (dry basis) of pumpkin slab for different air gap
p

1081 thickness.
ce
Ac

48 Page 48 of 54
t
ip
cr
us
an
M
d

1082
te

1083 Figure 12: Computed and experimental temperature profile at top and bottom surface of pumpkin slab for
1084 different air gap thickness. Note that one single value of air gap cap cannot capture the
p

1085 experimental curves


ce

1086

1087
Ac

1088

1089

1090

49 Page 49 of 54
t
ip
cr
1091

1092

us
1093 Figure 13: Snapshot of computed spatial variation of effective thermal conductivity (keff) after 1.5 h of
1094 drying. A thin, thermally resistive layer of very low thermal conductivity develops at the
1095 bottom surface of pumpkin slab that decreases the heat flux to the top surface

an
1096
1097
1098
1099
M
d
p te
ce
Ac

50 Page 50 of 54
t
ip
cr
us
an
M
d

1100
te

1101 Figure 14: Computed and experimental temperature profile at top and bottom surface of pumpkin slab for
1102 variable air gap thickness. The air gap thickness was varied linearly with drying time.
p
ce
Ac

51 Page 51 of 54
t
ip
cr
us
an
M
d

1103
te

1104 Figure 15: Computed and experimental moisture content (dry basis) of pumpkin slab for variable air gap
1105 thickness. The air gap thickness was varied linearly with drying time.
p
ce
Ac

52 Page 52 of 54
t
ip
cr
us
an
M
d

1106
te

1107 Figure 16: Sensitivity of the process to change in liquid (and corresponding gas permeabilities) using the
1108 total moisture loss as the variable of interest. The moisture loss history is plotted for three
p

1109 different sets of liquid water permeabilities: the original set and  50% of the original set.
ce
Ac

53 Page 53 of 54
t
ip
cr
us
an
M
d

1110
te

1111 Figure 17: Sensitivity of the process to change in mass transfer coefficient at the base using total moisture
1112 loss as the variable of interest. The moisture loss history is plotted for three different sets of
p

1113 hm,bottom : hm,top , hm,top / 5 and hm,top /10 .


ce
Ac

54 Page 54 of 54

You might also like