You are on page 1of 10

Journal of Food Engineering 144 (2015) 138147

Contents lists available at ScienceDirect

Journal of Food Engineering


journal homepage: www.elsevier.com/locate/jfoodeng

Drying of shrinkable food products: Appraisal of deformation behavior


and moisture diffusivity estimation under isotropic shrinkage
B. Ortiz-Garca-Carrasco a, E. Yaez-Mota a, F.M. Pacheco-Aguirre a,b, H. Ruiz-Espinosa a,
M.A. Garca-Alvarado b, O. Corts-Zavaleta b, I.I. Ruiz-Lpez a,
a
Colegio de Ingeniera en Alimentos, Facultad de Ingeniera Qumica, Benemrita Universidad Autnoma de Puebla, Av. San Claudio y 18 Sur, Ciudad Universitaria, Puebla,
Puebla, Mexico
b
Unidad de Investigacin y Desarrollo en Alimentos, Departamento de Ingeniera Qumica y Bioqumica, Instituto Tecnolgico de Veracruz, Av. M.A. de Quevedo 2779, Col. Formando
Hogar, C.P. 91860 Veracruz, Veracruz, Mexico

a r t i c l e

i n f o

Article history:
Received 20 May 2014
Received in revised form 21 July 2014
Accepted 28 July 2014
Available online 7 August 2014
Keywords:
Image analysis
Mass transfer
Shape change
Water diffusivity

a b s t r a c t
A new methodology based on image analysis was proposed to estimate the simultaneous size reduction
(shrinkage) and shape change (deformation) during food drying. Potato strips (9.525 mm 
9.525 mm  80 mm) were used as model system and subjected to convective drying at 50, 60, 70 and
80 C with an air velocity of 2 m/s. Developed protocol was used to analyze the shrinkage-deformation
behavior occurring in minor product dimensions, considered the dominant directions in mass transfer.
To this purpose, 2D perpendicular slices were obtained from original 3D product and their digital images
were processed to evaluate the changes in contour shape, perimeter, and cross-sectional and specic
areas of samples. Product contours were averaged to extract relevant deformation characteristics of dried
samples. Drying and shrinkage data were further used to estimate variable water diffusivities in product
with a previously reported analytical solution for shrinking solids, which was extended to allow for 2D or
3D mass transfer. Studied responses were successfully described as a function of free moisture fraction
(R2 > 0.84). It was demonstrated that shrinkage-deformation behavior was not affected by drying temperature under the tested conditions (p < 0.05). The analysis of averaged contour shapes showed that,
although shrinkage occurs from the beginning of drying, deformation appears at the nal stages, when
the free moisture fraction is below 0.2. Mean water diffusivities were estimated in the range of
3.045.36  1010 m2/s for studied drying temperatures.
2014 Elsevier Ltd. All rights reserved.

1. Introduction
Food products are complex systems which ordinarily require
the reduction of their water content in order to extend their
shelf-life. In this regard, convective drying with hot air is the most
widely used technique for the production of dehydrated foodstuffs.
Air drying involves several heat and mass transfer mechanisms;
yet, water diffusion within the product described by Ficks second
law is usually considered the controlling factor for modeling and
simulation purposes (da Silva et al., 2014; Pacheco-Aguirre et al.,
2014). This operation is accompanied by several changes in food,
but shrinkage and deformation, besides color, are the most evident.
Thus, the inclusion of food shrinkage is paramount for making a

Corresponding author. Tel.: +52 222 2295500x7250.


E-mail
addresses:
(I.I. Ruiz-Lpez).

irvingisrael@gmail.com,

http://dx.doi.org/10.1016/j.jfoodeng.2014.07.022
0260-8774/ 2014 Elsevier Ltd. All rights reserved.

irving.ruiz@correo.buap.mx

reliable estimation of diffusion coefcients and its description


remains an actual subject (Souraki and Mowla, 2008; GarcaPrez et al., 2012; Ruiz-Lpez et al., 2012).
In order to gather the relevant information to complement a
drying model, the appraisal of product shrinkage can be achieved
through universal methodologies including direct measurement
of product dimensions or volume displacement techniques
(Panyawong and Devahastin, 2007; Yan et al., 2008; Garca-Prez
et al., 2012; Ponkham et al., 2012; Curcio and Aversa, 2014). In
any case, dimensional changes are lumped in a single variable,
such as thickness or volume. Nevertheless, foods not only shrink
during drying but also may suffer a considerable deformation
(i.e., shape change), which can affect both the visual appeal and
packing properties of product. In addition, geometry deformation
may reect the collapse of microscopic porous structure of food,
leading to a drastic change in its physical properties (Ruiz-Lpez
et al., 2012). Both assessment and prediction of food deformation

B. Ortiz-Garca-Carrasco et al. / Journal of Food Engineering 144 (2015) 138147

139

Nomenclature
a
A
A
Bim
C
D
d0
d1
e
H
hm
K
k1, k2
L
mp
n
n
N
n1, n2
P
b P
P, P,
S
t
T
u
v
V
X

specic area (m1)


afne transformation matrix
cross-sectional area of product (m2)
Biot number for mass transfer (dimensionless)
number of contours to be averaged
effective diffusivity of water in food (m2/s)
parameter for moisture diffusivity equation (m2/s)
parameter for moisture diffusivity equation (m2/sC)
basis vector of the orthogonal coordinate system
absolute humidity of drying air (kg water/kg dry air)
external mass transfer coefcient (m/s)
water partition ratio between gas and solid phases
parameters for moisture diffusivity equation (dimensionless)
product length (m)
product mass (kg)
power constant (dimensionless)
normal unit vector
number of points on a product contour
parameters for moisture diffusivity equation (dimensionless)
perimeter of product contour (m)
edge coordinates of product contour: original, mirrored
and averaged, respectively
product surface (m2)
drying time (s)
cyclic order rotation of a given contour
moisture content (kg water/kg dry solids)
humid volume of drying air (m3 humid air/kg dry air)
product volume (m3)
characteristic length for water diffusion along x-axis (m)

represents a challenging task both numerically and experimentally. Thus, several studies have focused on the characterization
of macroscopic (external) shrinkage-deformation (SD) characteristics of food products, mainly through using novel image analysis
techniques.
Common properties obtained from digital images related with
product SD behavior include projected area and its corresponding perimeter, as well as selected sample dimensions which
are further used to evaluate secondary indices such as elongation, fractal dimensions, roundness, etc. (Campos-Mendiola
et al., 2007; Yan et al., 2008; Yadollahinia and Jahangiri, 2009;
Yadollahinia et al., 2009; Khazaei et al., 2013). Recently, volume
changes in dried products have been calculated using dimensions obtained from lateral and top digital images including
dual-camera setups (Sampson et al., 2014). However, while the
aforementioned characteristics are related to product quality
indices, reported methodologies do not pursue a further applicability of SD data in the modeling and simulation of drying
processes. For example, projected area is measured in nondominant mass transfer directions (top area in at slices) or is
affected by product bending in lateral views of samples.
Product deformation during drying imposes an additional difculty: no food sample shrinks and deform in the same way, even
under the most controlled conditions. Thus, the use of SD data
for advanced process simulations would require the development
of new protocols to extract relevant descriptors of product behavior. The main objective of this study is to develop and validate a
new methodology based on image analysis to estimate the simultaneous SD during food drying and obtain the representative patterns of this behavior using potato as food model. Besides,

characteristic length for water diffusion along y-axis


(m)

Greek letters
d
distance criterion between two contours
Dx
horizontal translation (m)
Dy
vertical translation (m)
c
denotes the c th cyclic order rotation of coordinates in
product contour
/
rotation angle (rad)
j
height-to-width ratio (dimensionless)
q
volumetric concentration of dry solids (kg dry solids/m3
product)
h
modied Fourier number for mass transfer in shrinkable
products with time-dependent diffusivity (dimensionless)
n
axial coordinate along x-axis (dimensionless)
w, W
free moisture fraction (dimensionless): local and averaged, respectively
f
axial coordinate along y-axis (dimensionless)
Subscripts
0
at the beginning of the drying process
a
for specic area
A
for cross-sectional area
e
at equilibrium
i
at the air-product interface
P
for perimeter
r
any reference product contour

generated data are further used to evaluate water diffusivity in


food system corrected for product shrinkage.
2. Methodology
2.1. Drying experiments
Two sets of air-drying experiments were conducted in order to
obtain the SD behavior of potato strips as a function of their moisture
content. Fresh, well-graded potatoes were locally purchased
(Puebla, Pue., Mxico) and dried the same day. Potatoes were
washed, dried with a cloth and sliced with a vegetable chipper
(9.525 mm-square openings) to produce strips that were further
cut to their desired length (80 mm). On average, 12 regular slices
were obtained from each tuber while remaining portions were
reserved for the initial water/dry solids analysis. Drying experiments were conducted with 20 potato strips placed at on a stainless
steel welded mesh open tray (dimensions: 0.25 m  0.20 m, openings: 4.5 mm  5.0 mm, wire diameter: 0.7 mm) in a tunnel dryer
(Armeld UOP8, Ringwood, UK) with airow parallel to the longest
product dimension. Samples were dried at 50, 60, 70 and 80 C for
about 330470 min with an air velocity of 2 m/s. Drying curves were
obtained in the rst experiment set, where moisture evolution was
calculated by continuously recording the weight of the product
throughout the process with trays that are carried on a support
frame that was in turn attached to a digital balance mounted above
the tunnel. A schematic view of the experimental setup was recently
presented in Pacheco-Aguirre et al. (2014). Moisture content in
product was expressed as the dimensionless free moisture fraction
W (the removable water portion left in product) according to

140

B. Ortiz-Garca-Carrasco et al. / Journal of Food Engineering 144 (2015) 138147

u  ue
mp  mpe

u0  ue mp0  mpe

These data were used to estimate the required time to approximately achieve a specic moisture content in product at each drying temperature (from W = 0.1 to W = 0.9 in 0.1 increments) in
order to obtain SD data regularly spaced over this variable in the
second experiment set.
With the purpose of evaluating the SD product behavior, groups
with 5 samples each were formed and dried for the predened
times. Then, a single transversal slice (perpendicular to the largest
dimension) of about 1 mm-thick was cut with a sharp blade from
the central part of the strip. A schematic view of samples along
with its relevant characteristics is shown in Fig. 1. Digital images
of resulting slices were immediately taken. Remaining product
portions were analyzed for their moisture content. The aforesaid
procedure was also applied to fresh (W = 1) and equilibrium-dried
(W = 0) samples. A total of ve slices were obtained for each moisture content-temperature combination.
Required moisture contents were determined by oven-drying
(Binder ED 53, Germany) the samples at 105 C until constant mass
weight (when mass change was less than 0.001 g over an 8 h period). Initial moisture content of product was 84.5 3.1 g water/
100 g product (mean s.d.).
2.2. Image acquisition
Potato slices jointly with a reference object of known dimensions (a black-anodized metal washer of 0.59 cm-diameter) were
placed on a blue paper sheet to provide plenty contrast for background extraction, and their digital images were acquired (Coolpix
L810, Nikon Corp., Japan). Digital camera was positioned with its
sight line normal to the supporting base. Illumination was provided through ordinary 18 W uorescent lamps without special
specications as color standardization was not needed between
images. Images were taken with the maximum available resolution
(4608  3456 pixels) in macro mode with a focal distance of about
0.10 m, using automatic settings. Digital images were stored in
JPEG format. The schematic view of the experimental image acquisition setup is shown in Fig. 2.
2.3. Image analysis
Two rectangular portions containing either the potato slice or
the reference object were manually selected from the original
image for their subsequent handling. Color information in these
images was transformed to the CIELAB color space for their analysis. Thus, every pixel was represented as a vector of color components L*, a* and b*. Afterward, color data were quantized using

k-means clustering algorithm (Press et al., 2007) and reduced to


3 dominant color descriptors, which were enough to retain the
image details in product border. This operation facilitated both
background extraction of image with a foreground mask and estimation of the pixel fraction corresponding to product. Quantized
image without background was transformed to grayscale and
nally coordinates of product boundary were obtained from this
image. A total of 400 points were used to describe each product
contour. Fig. 3 shows the image analysis steps used to determine
product deformation. On a separated procedure the relationship
between pixel number and real dimensions of reference object
was obtained, allowing the estimation of the slice area A. Boundary
coordinates were further used to estimate contour perimeter P as
the cumulative sum of the Euclidian distance between consecutive
points. All image analysis operations were performed with the
Matlab Image Processing Toolbox 7.0 (Matlab R2010a, MathWorks
Inc., Natick, MA, USA).
2.4. Modeling of shrinkage characteristics
By assuming a constant cross-sectional area (A) along major
dimension (L), the volume (V) can be estimated as the quantity
AL (Fig. 1). On the other hand, total surface of sample (S) can be
expressed as the sum of lateral (PL) and minor (2A) faces. If
PL  2A, then specic area can be estimated with the simple
expression

S PL 2A PL P

V
AL
AL A

In addition, changes in slice perimeter and cross-sectional area


are proportional to changes in mass transfer surface and product
volume if shrinkage is considered negligible along major dimension. Thus,

S
PL
P


S0 P 0 L 0 P 0
V
AL
A


V 0 A0 L0 A0

3
4

The following relationships are proposed in this study to relate


dimensional changes of product with moisture content:

A
DA 1  DA WnA
A0
P
DP 1  DP WnP
P0
na
a
1 Da ekW
a0

5
6
7

Eqs. (5) and (6) have been used to express the shrinkage behavior of food products during drying (Zielinska and Markowski, 2010;
Garca-Prez et al., 2012; Ruiz-Lpez et al., 2012). Here, parameters
nA and nP both control the deviations from the straight-line behavior, while DA and DP represent the nal fractions of cross-sectional
area and perimeter at the end of drying, respectively. On the other
hand, Eq. (7) is an exponential decay model, similar to Pages
model (Ruiz-Lpez et al., 2012), where Da controls the relative
increase of specic area and k and na dene the decaying rate
and shape of the curve, respectively.
2.5. Mean SD behavior

Fig. 1. Schematic view of potato strips with their initial geometry (right square
prism) and relevant characteristics.

Since all dried samples shrink and deform in a different way it is


desirable to develop a strategy for appraisal of relevant and common
characteristics of this behavior. In this study, the contours were
combined to obtain a mean deformation prole. However, contour
coordinates cannot be simply averaged after their estimation.

B. Ortiz-Garca-Carrasco et al. / Journal of Food Engineering 144 (2015) 138147

141

Fig. 2. Schematic view of experimental setup for image acquisition and sample preparation: (1) digital camera, (2) uorescent lamps, (3) reference object, (4) product slice,
(5) contrast background, (6) adjustable support, (7) dried product strip and (8) sharp blade.

Fig. 3. Image analysis steps used to determine product deformation: (a) original image, (b) simplied image with three color clusters, (c) gray-scale image after removing
non-product color clusters and (d) product contour. Image corresponds to an equilibrium-dried potato strip at 50 C (440 min). Initial contour shape is a square. (For
interpretation of the references to colour in this gure legend, the reader is referred to the web version of this article.)

Instead, they need to be translated and aligned with respect to


some reference point as both position and orientation of objects
in digital images changes with every shot. Otherwise, SD contour
characteristics would cancel or produce unexpected results when
averaged. Thus, the following algorithm was developed in order
to align contours. Let us consider that Pi contains the ordered
homogeneous coordinates, in clockwise or anticlockwise direction,
of every point on contour i. Thus,

xi

xi;1

...

xi;N

...

6 7 6
Pi 4 yi 5 4 yi;1
1

7
. . . yi;N 5

The alignment of contour Pi with respect to contour Pr involves the


translation of Pi coordinates in the amounts Dx and Dy, as well as
their rotation in the angle / by means of the afne transformation

P0i APi

where A is both a clockwise rotation matrix and a translation


matrix

3
cos / sin / Dx
6
7
A 4  sin / cos / Dy 5
0
0
1

10

The parameters Dx, Dy and / should minimize some distance


criterion between Pi and Pr. However, the straight comparison of
these contours relates point pairs (xi,1, yi,1) with (xr,1, yr,1), (xi,2, yi,2)
with (xr,2, yr,2) and in general form (xi,k, yi,k) with (xr,k, yr,k), which
not necessarily produce the best results. Consequently, both the
cyclic order rotation and mirrored projection of Pi were considered
to render the best overlap with reference contour. The cyclic order
rotation Ti,c of contour Pi was dened as

142

B. Ortiz-Garca-Carrasco et al. / Journal of Food Engineering 144 (2015) 138147

xi;c

xi;c

6
7 6
Ti;c 4 yi;c 5 4 yi;c
1

xi;N

xi;1

. . . yi;N

yi;1

...
...

...

xi;c1

7
. . . yi;c1 5

11

...

An Euclidian distance criterion (/c) between Pr and Ti,c is given


by
T

/c xr  xi;c xr  xi;c yr  yi;c yr  yi;c

12

which correspond to the squared sum of x and y distances


between both contours. Parameters Dx, Dy, / and c achieving the
best overlap between Pi and Pr (those minimizing /) can be estimated by means of nonlinear regression analysis. This procedure
is repeated with the mirrored contour

3 2
^i
maxxi  xi;1
x
6
bi 6
^i 7
P
yi;1
4y
54
1
1

. . . maxxi  xi;N
...
...

yi;N
1

3
7
5

13

Only the best overlap of contour Pi (achieved using the original


or mirrored image) with respect to Pr, designated as Pi?r, is considered in the rest of the analysis. Finally, contour coordinates can be
averaged to obtain a single representative image of product deformation as

Pr

Pi!r

14

Preliminary results indicated that any contour can be chosen as


the reference since the same nal image is obtained in all cases.
The adequacy of averaged contour to represent the global SD
behavior of product can be estimated by calculating the evolution
of its enclosed area which should be as close as possible to the
mean value of original contours at each moisture content-temperature combination. Please notice that it is expected that both
perimeter and specic area of averaged contour slightly deviate
from the mean values of original contours, especially at low
moisture contents, since the averaging procedure smooths out
roughness and irregularities developed in product border. Fig. 4
shows the steps used to determine the representative SD behavior
of potato strips.

diffusion model and its boundary condition are written as


(Ruiz-Lpez et al., 2012)





@u
@
@u
@
@u

D
D
@t @x
@x
@y
@y
hm
@ui
@ui
H  H n  Dq
ex  n  Dq
ey
v i
@x
@y

By assuming (i) negligible heat transfer, (ii) constant properties


of drying air, (iii) constant volumetric concentration of dry solids,
(iv) uniform initial distribution of water within the solid and (v)
two-dimensional mass transfer in product, the unsteady-state

16

Let us consider that both diffusion coefcient D and characteristic lengths for diffusion X and Y are explicit drying time functions,
i.e., D = D(t), X = X(t) and Y = Y(t). Then, Eqs. (15) and (16) can be
rewritten as

@w @ 2 w 1 @ 2 w

@h @n2 j2 @f2
@w
1 @wi
e
Bim twi n  i ex  n 
@n
j @f y
u  ue
w
u0  ue
Dt
@t
@h
Xt2
1
1
@n
@x; @f
@y
Xt
Yt
Yt
1=v
hm
j
; Bim t K
;
Xt
q Dt=Xt

17
18
19
20
21
K

dHi
dX i

22

The term K in Eq. (22) is the local slope of equilibrium curve (an
instantaneous partition coefcient). The variable transformations
in Eqs. (20) and (21) allowed expressing the original problem for
moisture diffusion in a shrinkable body with variable diffusivity
as the simpler case of mass transfer for constant D in a nonshrinkable product (Ruiz-Lpez et al., 2012). However, the
unknown variable nature of Bim(t) in boundary condition (18) represents a difculty for providing an analytical solution to the aforementioned problem. Ruiz-Lpez et al. (2012) demonstrated that if
the internal resistance to mass transfer by diffusion is accepted as
the only mechanism controlling drying rate throughout the drying
process then existing solutions could be used to solve Eqs. (17) and
(18) considering product shrinkage and variable water diffusivity.
Under this assumption boundary condition (18) is simplied to

wi 0
2.6. Modeling of drying data

15

23

The procedure to both evaluate water diffusivities and describe


drying kinetics of shrinkable food products proposed by
Ruiz-Lpez et al. (2012) is now extended to products with mass
transfer in more than one direction. It should be noticed that the
following procedure is only applicable for products with isotropic
shrinkage (solid shrinks in the same proportions in involved mass

Fig. 4. Contour manipulation steps used to determine a representative image of product deformation: (a) original, (b) aligned and (c) averaged contours (black contour was
chosen as the reference one). Contours correspond to an equilibrium-dried potato strip at 70 C.

143

B. Ortiz-Garca-Carrasco et al. / Journal of Food Engineering 144 (2015) 138147

Fig. 5. Effect of moisture content on relative cross-sectional area of dried potato


strips.

Fig. 7. Effect of moisture content on relative specic area of dried potato strips.

A
XY
jX 2

A0 X 0 Y 0 jX 20

or X X 0

p
A=A0

26

In this study, Eq. (24) was numerically solved for experimental

W values with bisection algorithm, while derivatives dh/dt in Eq.


(25) were estimated using nite differences with second-order
accuracy. Estimated diffusivities were further used to evaluate its
dependence on moisture content and drying temperature with
the model

DW d0 d1 T1  expk1 Wn1  expk2 Wn2

27

Mean water diffusivities for every drying temperature were


estimated from the numerical integration of instantaneous values
using trapezoidal rule according to

D
Fig. 6. Effect of moisture content on relative perimeter of dried potato strips.

transfer directions, that is, j is constant). The analytical solution


for Eq. (17) and (23) with mass transfer in x and y directions can
be obtained from the well-known at-slab solution and the superposition principle to obtain

"
8

1
X

p2
n1

"
8

1
X

p2
n1

1
2n12

1
2n12

exp

2 2
p
 2n1
4


2 2
p
exp  2n1
4

#


 ...

#


24

2.6.1. Moisture diffusivity estimation


Drying data, in the form of W vs. t and X vs. t (or W), can be used
for moisture diffusivity estimation using the following procedure:
(i) calculate h for every W value in the drying curve by solving Eq.
(24), (ii) calculate the derivative dh/dt and (iii) estimate moisture
diffusivities from Eq. (20) at every experimental t as

dh
dt

W2

Z
DWdW

W1

W2

dW

28

W1

2.6.2. Drying simulation


If the dependence of both D and X on time or moisture content
are known, for example with Eqs. (26) and (27), then drying curves
can be simulated from Eq. (20) by solving the initial value problem,

dt Xt2 XW2

dh
Dt
DW

with th 0 0

29

j2

Eq. (24) allows the calculation of W for a given drying time


expressed as variable h (or the inverse problem).

Dt Xt2

25

In our case, as cross-sectional area can be calculated with


Eq. (5), then

where W is calculated from h in each iteration with Eq. (24), and W


is further used to evaluate both D(W) and X(W) .
Generalization of the just described procedures to 3D mass
transfer or other coordinate systems should be evident.

2.7. Data analysis


The tness quality of identied models was quantied by the
determination coefcient (R2) and statistical signicance of parameter estimates was evaluated through their 95% condence intervals (95% CI). Numerical procedures, nonlinear regression (based
on ordinary least squares) and statistical analyses were performed
with the Matlab software and its Statistics Toolbox 7.3 (Matlab
R2010a, MathWorks Inc., Natick, MA, USA).

144

B. Ortiz-Garca-Carrasco et al. / Journal of Food Engineering 144 (2015) 138147

Table 1
Regression parameters for shrinkage models.
Response
Cross-sectional area
b

Perimeter

Specic area

a
b
c

Parameters

Value

95% CI

R2

DA
nA
DP
nP
Da
k
na

0.1733
0.7222
0.6390
0.9581
2.2641
5.8974
1.1257

0.1469/0.1997
0.6732/0.7712
0.6233/0.6547
0.8625/1.0537
2.1730/2.3553
4.8650/6.9297
1.0002/1.2512

0.9320
0.8460
0.9226

A=A0 DA 1  DA WnA .
P=P 0 DP 1  DP WnP .
na
a=a0 1 Da ekW .

3. Results and discussion


3.1. Shrinkage characteristics of potato strips
Figs. 57 show the dependence of cross-sectional area, perimeter and specic area of slices on moisture content. Initial values of
these variables were determined by image analysis as
86.5779 mm2, 37.4430 mm and 0.4325 mm1, respectively. These
values were slightly different than those expected from size of
chipper openings (9.525 mm), corresponding to 90.7256 mm2,
38.1 mm and 0.4199 mm1. Real product metrics produce initial
values for volume (V0), and surfaces of both lateral (P0L0) and
minor faces (2A0) of 6926.232 mm3, 2995.44 mm2 and
173.1558 mm2, respectively. Therefore, square faces of potato
strips amount to less than 6% of the total available surface for mass
transfer, validating the assumption made in the development of Eq.
(2) that P0L0  2A0 (by little over 17 times). As expected, perimeter
and cross-sectional area of samples decreased with moisture content, while the specic area exhibited the opposite behavior. No
signicant effect of drying temperature was observed on shrinkage
characteristics of potato (p < 0.05), which is in agreement with previous studies (Hassini et al., 2007). Thus, Eqs. (5)(7) were tted
over the entire temperature domain. Regression coefcients of
the models used for the mathematical description of these data
are presented in Table 1. A satisfactory reproduction of these
responses was obtained in all cases (R2 > 0.84).

As demonstrated in Eq. (4), changes in cross-sectional area of


the studied geometry are related with changes in product volume.
Drying of potato strips resulted in samples with about 17% of their
original size (Fig. 5). This shrinkage value is comparable to those
found in other studies, regardless of the sample geometry and
the method used to estimate it. For example, Wang and Brennan
(1995) and Hassini et al. (2007) reported potato slabs
(10 mm  45 mm  20 mm) shrinking up to 1720% of their original volume when dried at 4085 C with air velocities ranging from
0.5 to 4 m/s. Lozano et al. (1983) also reported a nal shrinkage
ratio of approximately 19% for potato cylinders (1 cm in diameter,
4 cm long) dried at 40 C with an air velocity of 1 m/s. Hassini et al.
(2007) determined volume of potato slabs using direct local measurements of length, thickness and width, while Lozano et al.
(1983) and Wang and Brennan (1995) used volume displacement
techniques with toluene and water, respectively. According to statistical analysis, the dependence of cross-sectional area on moisture content shows a noticeable deviation from straight line
behavior (nA 1, p < 0.05), indicating that dimensional changes of
product not only depend on volume of evaporated water, but also
on the collapse resistance of cell structure. As shown in Fig. 5, the
size of potato strips was initially reduced in about 40% when 60% of
available water was eliminated, but a comparable size reduction
was observed thereafter, when a smaller amount of water was
evaporated (the remaining 40%). We hypothesize that as water is
eliminated from product cell turgor is gradually loss, but rigidity
of cell walls offers a resistance against shrinkage. However, as drying proceeds, the continuous water ux across membranes might
rupture them once a critic moisture content is reached, debilitating
the inner structure of product and causing a pronounced shrinkage
(0.1 6 W 6 0.4). Finally, at very low moisture contents water could
be removed with minimum product shrinking if collapse of cell
structure is not complete, causing the development of an air-lled
porous network, which reects as a subtle tail of data in plot at
W < 0.1. Shrinkage deviations from the straight-line behavior have
been well-documented in several studies with current product
(Lozano et al., 1983; Ratti, 1994; Wang and Brennan, 1995;
Hassini et al., 2007).
According to Eq. (3), changes in contour perimeters are related
to changes in available surface for mass transfer. Surface area of

Fig. 8. Final edge deformation of slices cut from dried potato strips.

B. Ortiz-Garca-Carrasco et al. / Journal of Food Engineering 144 (2015) 138147

145

Fig. 9. Mean edge deformation of slices cut from dried potato strips (80 C). Inner numbers represent the free moisture fraction reached in product/elapsed drying time (min).

Fig. 10. Comparison of cross-sectional area estimated from original and averaged
contours.

potato strips decreased with moisture content up to about 64% of


its initial value (Fig. 6), with no signicant departure from the
straight-line behavior (nP = 0.9581, 95% CI = 0.8625/1.0537). Image
analysis has been previously used to estimate the surface area evolution of potato slices. Selected studies include those from
Campos-Mendiola et al. (2007) where the projected lateral area
of air-dried (55 C, 1.7 m/s) circular potato slices (2.5 mmthickness, 40 mm-diameter) was measured, observing a signicant
increase of this variable, caused by sample bending during process.
In later studies, Yadollahinia and Jahangiri (2009) and Yadollahinia
et al. (2009) followed the evolution of the upper area of
circular slices (10 mm-thickness, 35 mm-diameter) during drying
(6080 C, 0.51 m/s), reporting nal reductions between 50%
and 65% of the original values. These authors reported a marked
deviation from the straight-line behavior for this response for
W < 0.1, caused by upward bending of product developing an
irregular shape and affecting the measured area.

Fig. 11. Comparison of perimeter estimated from original and averaged contours.

Product volume lowered faster than available surface for mass


transfer, as expected from its proportionality with linear dimensions raised to the third, during drying causing a 3.26-fold increase
of the specic area (Fig. 7), with a near rst-order exponential
behavior (na 1, p < 0.05). Comparable results were obtained by
Ratti (1994) during drying (4060 C, 15 m/s) of potato disks
(4 cm-diameter, 0.9 cm-thickness) and cylinders (1 cm-diameter,
5 cm-length), reporting 2.18 and 2.95-fold increments in this
variable, respectively, where sample volume was determined by
liquid displacement and surface area was roughly estimated from
product dimensions.
3.2. Deformation characteristics of potato strips
Deformation and shrinkage characteristics of the cross section
of potato strips were successfully estimated with the proposed
methodology. It was found that all samples suffered a similar size

146

B. Ortiz-Garca-Carrasco et al. / Journal of Food Engineering 144 (2015) 138147

for an excellent reproduction of cross-sectional area from original


data; consequently, size reduction estimated from averaged contours is representative of observed behavior in individual samples.
On the other hand, perimeter is slightly underestimated (Fig. 11) at
lower moisture contents (W < 0.3, < 12% of relative difference)
when calculated from averaged contours since this is smoothed
during averaging procedure (Fig. 4). Consequently, the specic area
is also underestimated (<10% of relative difference, Fig. 12). Nevertheless, for simulation purposes observed discrepancies would not
translate in notably different results, as numerical convergence
would be reached before mesh neness had to be increased to
unpracticals level to replicate surface roughness. The use of mean
SD behavior data in the modeling and simulation of drying process
will be demonstrated in the second part of this study.
3.3. Moisture diffusivity and process simulation
Fig. 12. Comparison of specic area estimated from original and averaged contours.

reduction and shape change along process regardless of drying


temperature. A comparison of the contour shapes at the end of drying is presented as example in Fig. 8 (for W = 0), where it can be
veried that product surface developed irregularities or roughness
with drying, which is in agreement with other studies (CamposMendiola et al., 2007). It should be emphasized that shrinkage
and deformation behavior was unique for each sample, but they
clearly exhibit a repetitive pattern. Thus, the use of a mean deformation prole is desirable for a representative description of product shape changes during drying, both mathematically and
qualitative wise. For all studied temperatures, product reduced
its dimensions without a signicant shape change up to a free
moisture content of 0.2, with an important deformation occurring
thereafter, mainly manifested in a preferential contraction of the
middle section of contour edges toward the sample center, as
evidenced in Fig. 9 for product dried at 80 C. The estimation of relevant SD characteristics of product is also desirable for simulation
of drying process, as the effect of product deformation on water
diffusivity could be evaluated without requiring particular data
of a specic sample, especially if drying is conducted with several
slices. Thus, it is important to determine if an averaged contour
retains some characteristics of those from which it was obtained.
Figs. 1012 show a comparison of shrinkage properties (crosssectional area, perimeter and specic area) from original and
averaged contours. As shown in Fig. 10, averaged contours allowed

Drying curves from the rst experiment set are presented, in


both linear and semi logarithmic representations, in Fig. 13. These
data were used to evaluate moisture diffusivities plotted in Fig. 14.
As expected, the use of higher drying temperatures caused a significant increase in water diffusivity values (p < 0.05). Water diffusivity behavior can be divided in three sections. In the rst region, the
gradual increase of water mobility from the beginning of the
drying process up to W  0.6 indicates that product temperature
is in a transitory state (preheating period) from initial product
temperature to air temperature. This unsteady-thermal state is followed by an approximately constant water diffusivity period for
0.2 6 W 6 0.6. From then on, deceleration in water mobility might
be caused by cell structure collapse as suggested by the severe
product deformation observed at these moisture levels. Similar
trends have been previously reported during convective drying of
several foodstuffs including chayote, sh muscle, grapes and fresh
and pre-osmosed carrot cubes (Azzouz et al., 2002; Pinto and
Tobinaga, 2006; Singh and Gupta, 2007; Ruiz-Lpez et al., 2012).
Consequently, any constant diffusivity model would fail to reproduce the whole experimental behavior when product preheating
or severe shrinkage/deformation cannot be neglected. In this case,
the suggested procedure allows the objective identication of
these three regions. Regression parameters for moisture diffusivity
equation (26) are shown in Table 2. A good reproduction of experimental behavior was achieved with proposed model (R2 = 0.9404).
Statistical analysis revealed that all constants were signicant
(p < 0.05), thus the model is structurally identiable with current
data (i.e., its parameters can be uniquely estimated).

Fig. 13. Experimental (dots) and predicted (lines) potato drying curves in linear (left) and semi logarithmic representations (right).

B. Ortiz-Garca-Carrasco et al. / Journal of Food Engineering 144 (2015) 138147

147

Acknowledgments
The authors wish to thank the Consejo Nacional de Ciencia y
Tecnologa (CONACYT) for providing nancial support through
project 130011. Betzabeth Ortiz-Garca-Carrasco and Estefana
Yaez-Mota acknowledge their undergraduate scholarship from
CONACYT. Francisco Manuel Pacheco-Aguirre acknowledges his
doctoral scholarship from CONACYT.
References

Fig. 14. Water diffusivities as a function of moisture content during drying of


potato strips.

Table 2
Regression parameters for water diffusivity modela.
Parameter
10

d0  10 (m /s)
d1  1011 (m2/sC)
k1(dimensionless)
n1(dimensionless)
k2(dimensionless)
n2(dimensionless)
a

Value

95% CI

1.2333
1.0246
7.8981
0.6046
2.9294
7.1208

2.0408/0.0426
0.8903/1.1590
4.0639/11.7323
0.4932/0.7161
2.4228/3.4361
5.9278/8.3138

D d0 d1 T1  expk1 Wn1  expk2 Wn2 .

Mean water diffusivities were estimated as 3.04  1010,


3.61  1010, 4.57  1010 and 5.36  1010 m2/s for drying
processes at 50, 60, 70 and 80 C, respectively. Moisture diffusivities are comparable to those reported by other authors where
values have been also corrected for product shrinkage. For potato,
Hassini et al. (2007) reported diffusivity values ranging from
3.55  1010 to 1.92  109 m2/s (4085 C). For other vegetable
products, such as carrot, chayote and mango, values were between
2.58  10101.72  109 m2/s (6090 C), 4.44  10108.60 
1010 m2/s (4070 C) and 2.61  10101.30  109 m2/s (40
70 C), respectively (Dissa et al., 2008; Zielinska and Markowski,
2010; Ruiz-Lpez et al., 2012). As shown in Fig. 13, identied diffusivities jointly with shrinkage data allowed for an excellent reproduction of drying curves at all moisture contents (R2 > 0.99).
4. Conclusions
The proposed image analysis methodology allowed estimating
representative shrinkage-deformation characteristics of dried
potato strips along dominant mass transfer directions. Shrinkagedeformation characteristics determined by current protocols were
successfully applied in the estimation of water diffusivities and
drying simulation and results were comparable with other studies
where different methodologies have been used. A previously
reported one-dimensional drying model for shrinkable solids was
extended to consider two- or three-dimensional mass transfer,
with a remarkable reproduction of drying curves. It was demonstrated that individual shrinkage-deformation characteristics of
dried products could be combined in a single representative prole
which could be used in detailed simulations, as will be presented
in a following document currently in preparation.

Azzouz, S., Guizani, A., Jomaa, W., Belghith, G., 2002. Moisture diffusivity and drying
kinetic equation of convective drying of grapes. J. Food Eng. 55 (4), 323330.
Campos-Mendiola, R., Hernndez-Snchez, H., Chanona-Prez, J.J., Alamilla-Beltrn,
L., Jimnez-Aparicio, A., Fito, P., Gutirrez-Lpez, G.F., 2007. Non-isotropic
shrinkage and interfaces during convective drying of potato slabs within the
frame of the systematic approach to food engineering systems (SAFES)
methodology. J. Food Eng. 83 (2), 285292.
Curcio, S., Aversa, M., 2014. Inuence of shrinkage on convective drying of fresh
vegetables: a theoretical model. J. Food Eng. 123, 3649.
da Silva, W.P., Hamawand, I., e Silva, C.M.D.P.S., 2014. A liquid diffusion model to
describe drying of whole bananas using boundary tted coordinates. J. Food
Eng. 137, 3238.
Dissa, A.O., Desmorieux, H., Bathiebo, J., Koulidiati, J., 2008. Convective drying
characteristics of Amelie mango (Mangifera Indica L. cv. Amelie) with
correction for shrinkage. J. Food Eng. 88 (4), 429437.
Garca-Prez, J.V., Ozuna, C., Ortuo, C., Crcel, J.A., Mulet, A., 2012. Modeling
ultrasonically assisted convective drying of eggplant. Drying Technol. 29 (13),
14991509.
Hassini, L., Azzouz, S., Peczalski, R., Belghith, A., 2007. Estimation of potato moisture
diffusivity from convective drying kinetics with correction for shrinkage. J. Food
Eng. 79 (1), 4756.
Khazaei, N.B., Tavakoli, T., Ghassemian, H., Khoshtaghaza, H., Banakar, A., 2013.
Applied machine vision and articial neural network for modeling and
controlling of the grape drying process. Comput. Electron. Agric. 98, 205213.
Lozano, J.E., Rotstein, E., Urbicain, M.J., 1983. Shrinkage, porosity and bulk density of
foodstuffs at changing moisture contents. J. Food Sci. 48 (5), 14971502.
Pacheco-Aguirre, F.M., Ladrn-Gonzlez, A., Ruiz-Espinosa, H., Garca-Alvarado,
M.A., Ruiz-Lpez, I.I., 2014. A method to estimate anisotropic diffusion
coefcients for cylindrical solids: application to the drying of carrot. J. Food
Eng. 125, 2433.
Panyawong, S., Devahastin, S., 2007. Determination of deformation of a food
product undergoing different drying methods and conditions via evolution of a
shape factor. J. Food Eng. 78 (1), 151161.
Pinto, L.A.A., Tobinaga, S., 2006. Diffusive model with shrinkage in the thin-layer
drying of sh muscles. Drying Technol. 24 (4), 509516.
Ponkham, K., Meeso, N., Soponronnarit, S., Siriamornpun, S., 2012. Modeling of
combined far-infrared radiation and air drying of a ring shaped-pineapple with/
without shrinkage. Food Bioprod. Process. 90 (2), 155164.
Press, W.H., Teukolsky, S.A., Vetterling, W.T., Flannery, B.P., 2007. Numerical
Recipes. The Art of Scientic Computing. Cambridge University Press, New York,
USA, Third edition.
Ratti, C., 1994. Shrinkage during drying of foodstuffs. J. Food Eng. 23 (1), 91105.
Ruiz-Lpez, I.I., Ruiz-Espinosa, H., Arellanes-Lozada, P., Brcenas-Pozos, M.E.,
Garca-Alvarado, M.A., 2012. Analytical model for variable moisture diffusivity
estimation and drying simulation of shrinkable food products. J. Food Eng. 108
(3), 427435.
Sampson, D.J., Chang, Y.K., Rupasinghe, H.P.V., Zaman, Q.U.Z., 2014. A dual-view
computer-vision system for volume and image texture analysis in multiple
apple slices drying. J. Food Eng. 127, 4957.
Singh, B., Gupta, A.K., 2007. Mass transfer kinetics and determination of effective
diffusivity during convective drying of pre-osmosed carrot cubes. J. Food Eng.
79 (2), 459470.
Souraki, B.A., Mowla, A., 2008. Axial and radial moisture diffusivity in cylindrical
fresh green beans in a uidized bed dryer with energy carrier: modeling with
and without shrinkage. J. Food Eng. 88 (1), 919.
Wang, N., Brennan, J.G., 1995. Changes in structure, density and porosity of potato
during dehydration. J. Food Eng. 24 (1), 6176.
Yadollahinia, A., Jahangiri, M., 2009. Shrinkage of potato slice during drying. J. Food
Eng. 94 (1), 5258.
Yadollahinia, A., Lati, A., Mahdavi, R., 2009. New method for determination of
potato slice shrinkage during drying. Comput. Electron. Agric. 65 (2), 268274.
Yan, Z., Sousa-Gallagher, M.J., Oliveira, F.A.R., 2008. Shrinkage and porosity of
banana, pineapple and mango slices during air-drying. J. Food Eng. 84 (3), 430
440.
Zielinska, M., Markowski, M., 2010. Air drying characteristics and moisture
diffusivity of carrots. Chem. Eng. Process. 49 (2), 212218.

You might also like