You are on page 1of 12

food and bioproducts processing 9 7 ( 2 0 1 6 ) 88–99

Contents lists available at ScienceDirect

Food and Bioproducts Processing

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

Mass transfer modeling in osmotic dehydration:


Equilibrium characteristics and process dynamics
under variable solution concentration and
convective boundary

H. Pacheco-Angulo a , E. Herman-Lara b , M.A. García-Alvarado a ,


I.I. Ruiz-López c,∗
a Unidad de Investigación y Desarrollo en Alimentos, Departamento de Ingeniería Química y Bioquímica, Instituto
Tecnológico de Veracruz, Av. M.A. de Quevedo S/N. Col. Formando Hogar, C.P. 91860 Veracruz, Veracruz, México
b Coordinación de Posgrado e Investigación, Departamento de Ingeniería Química y Bioquímica, Instituto Tecnológico

de Tuxtepec, Av. Dr. Víctor Bravo Ahuja S/N. Col. 5 de Mayo, C.P. 68350, Tuxtepec, Oax. , México
c Facultad de Ingeniería Química, Benemérita Universidad Autónoma de Puebla, Av. San Claudio y 18 Sur. Ciudad

Universitaria, Puebla, Puebla, México

a r t i c l e i n f o a b s t r a c t

Article history: The aim of this study was to model both the dynamic and equilibrium mass transfer periods
Received 22 April 2015 for water, osmotic solute and food solids interchange between product and solution during
Received in revised form 10 an osmotic dehydration (OD) process. The OD model is able to represent situations where
November 2015 concentration of osmotic media changes during the process or where interfacial resistance
Accepted 25 November 2015 to mass transfer cannot be neglected. Water and solute are considered to move within the
Available online 8 December 2015 product by a diffusion mechanism based on Fick’s second law, while external convective
mass transfer is considered in the fluid. The state-space form of the model is analyti-
Keywords: cally solved for one-dimensional mass transfer in products with flat slab, infinite cylinders
Interfacial resistance and sphere geometries. The developed theory was applied to the analysis of equilibrium
Mass Biot number and OD dehydration curves of carrot slices obtained at 40 ◦ C in sodium chloride solutions
Mass transfer with and without stirring and different ratios between solution volume and product mass.
Osmotic dehydration Water and NaCl diffusivities were identified in the narrow ranges of 6.0–7.6 × 10−10 m2 /s
Variable solution concentration and 3.5–4.1 × 10−10 m2 /s, respectively, demonstrating the applicability of the proposed model
Equilibrium distribution under a wide range of operating conditions.
© 2015 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.

1. Introduction is removed from food toward the liquid media with a simul-
taneous solute gain by the product (Khan et al., 2008; Goula
Osmotic dehydration is a solid–liquid contact operation and Lazarides, 2012; Herman-Lara et al., 2013; Souraki et al.,
involving the immersion of food products, especially fruits 2013, 2014). If processing is performed for a long enough
and vegetables, in hypertonic solutions such as brines or time, then both water loss and solute gain reach a stationary
syrups. When the food is immersed in the solution, an osmotic state, where the driving potential for mass transfer between
pressure gradient is developed between the involved phases food and solution becomes zero (Sablani and Rahman, 2003;
originating a dynamic mass transfer period in which water Herman-Lara et al., 2013). This operation involves several mass


Corresponding author. Tel.: +52 222 2295500x7250.
E-mail addresses: irving.ruiz@correo.buap.mx, irvingisrael@hotmail.com (I.I. Ruiz-López).
http://dx.doi.org/10.1016/j.fbp.2015.11.002
0960-3085/© 2015 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.
food and bioproducts processing 9 7 ( 2 0 1 6 ) 88–99 89

Nomenclature
 dynamic viscosity (Pa s)
A surface area available for mass transfer (m2 )  dimensionless group related with product-to-
A state-space matrix osmotic media mass ratio
aw water activity density (kg/m3 )
Bi mass Biot number (dimensionless)
chemical potential (J/mol)
C concentration (g/L) dimensionless coordinate
d maximum inner diameter of flask (m) , dimensionless concentration in product: local
D diffusion coefficient (m2 /s) and average, respectively
F Faraday constant (96485.34 C/mol)
Fo Fourier number for mass transfer (dimension- Subscripts
less) 0 at beginning of the process
g gravitational constant (9.81 m2 /s) ◦ at infinite dilution
Gr mass Grashof number (dimensionless) f denotes food solids
kc convective mass transfer coefficient (m/s) e at equilibrium
Ke equilibrium partition coefficient (kg product/kg i at the product–solution interface
osmotic media) j denotes either water or osmotic solute
L, Lc characteristic lengths for diffusion and convec- k denotes node numbering
tion, respectively (m) o denotes osmotic media
m solute molality (mol solute/kg solvent) p denotes product
m mass in product (kg) s denotes osmotic solute
M mass in osmotic media (kg) t at time t
n number of nodes w denotes water
N number of components
n+ , n− valences of cation and anion, respectively
(dimensionless) transfer mechanisms; however, diffusion of water and solute
R gas constant (8.314 J/mol K) within product are usually considered the controlling factors,
Re Reynolds number (dimensionless) usually described by Fick’s second law (Kaymak-Ertekin and
Sc Schmidt number (dimensionless) Sultanoğlu, 2000; Rastogi and Raghavarao, 2004; Goula and
Sh Sherwood number (dimensionless) Lazarides, 2012; Souraki et al., 2012; da Silva et al., 2013, 2014;
t time (s) Rodríguez et al., 2013).
T temperature (K) Depending on modeling assumptions, non-steady-state
u axial or radial coordinate (m) diffusion equation may be solved under analytical or
U food solids-to-(osmotic solute + water) mass numerical techniques (Kaymak-Ertekin and Sultanoğlu, 2000;
ratio in product (kg/kg) Porciuncula et al., 2013; Rodríguez et al., 2013); however, the
v orbital velocity (1/s) use of analytical models remains widespread due to sev-
V volume (m3 ) eral practical advantages, including easier implementation
X mass fraction of a given component in product and lower computational effort. Analytical solutions for OD
(kg/kg product) process in well-stirred systems (i.e., those with negligible
X* mass fraction in product free of food solids external resistance to mass transfer) can be classified in
(kg/kg osmotic solute + water) two main groups based on the available amount of work-
X state-space vector ing solution (Crank, 1975). Most studies consider an infinite
Y mass fraction of a given component in solution volume of osmotic media (i.e., of constant concentration;
(kg/kg solution) Rastogi and Raghavarao, 2004; Rodríguez et al., 2013; Souraki
Y* mass fraction in solution free of food solids (kg et al., 2012, 2013, 2014), experimentally achieved by using
osmotic solute + water) mass ratios between solution and product of 10 and above
x1 , x2 coded temperature and time, respectively (Herman-Lara et al., 2013). The use of lower ratios between
W food solids-to-(osmotic solute + water) mass osmotic media and product may result in an appreciable
ratio in osmotic media (kg/kg) decrease of solute concentration during product impregna-
tion, affecting water loss and solute gain rates as well as
Greek letters final dehydration and impregnation levels at equilibrium.
˛ parameter defining product geometry Therefore, special analytical solutions considering a finite vol-
ı mass ratio between osmotic media and product ume of osmotic media (i.e., of variable concentration) should
ε volume fraction occupied by osmotic media (m3 be applied in these cases (Singh et al., 2007; Bellary et al.,
solution/m3 solution + product) 2011), or alternatively, resort to a numerical solution allow-
ϕ thermodynamic factor (dimensionless) ing the use of additional assumptions such as variable mass
 dimensionless concentration in solution of involved phases (Kaymak-Ertekin and Sultanoğlu, 2000). As
± mean ionic activity coefficient of solute (dimen- a diffusion-controlled process is presumed in the aforemen-
sionless) tioned models (Crank, 1975), these may not be applicable to
◦+ , ◦− limiting (zero-concentration) ionic conduc- describe experimental conditions where convective resistance
tances of cation and anion, respectively to mass transfer may occur (for example, processes conducted
(S m2 /mol) without or with mild stirring), which may be required in prod-
ucts with fragile structures.
90 food and bioproducts processing 9 7 ( 2 0 1 6 ) 88–99

Analytical solutions considering both diffusion and con- Similarly, the mass ratio between food solids and the sum
vection at product surface have also been developed and are of osmotic solute and water in product and osmotic media are
expressed in terms of a mass Biot number (Crank, 1975), whose defined as
value may be used to identify the controlling mechanism (da
Xf Yf
Silva et al., 2013, 2014). A well-known fact in mass trans- U= and W= (4)
fer operations is that any solute will not distribute equally Xs + Xw Ys + Yw
between two or more phases due to different chemical and
Above expressions are related with regular mass fractions
physical affinities (Corzo and Bracho, 2004; Toğrul and İspir,
either of water or osmotic solute by simply algebra (for j = s, w)
2008). Consequently, an equilibrium relationship is required
to completely characterize a multiphase system. In drying, for Xj∗
example, the equilibrium water contents in product and air Xj = (5)
1+U
are related through the sorption isotherm, and a mean parti-
tion constant should be used to obtain the correct mass Biot Yj∗
definition to apply the existing analytical solutions. In order Yj = (6)
1+W
to consider a convective boundary condition in the OD model,
some studies have considered that product and solution equil- 2.2. Description of mass transfer equilibrium
ibrate at the same concentration (da Silva et al., 2013, 2014).
However, the use of models with convective boundary in OD According to equilibrium thermodynamic theory (Gibbs, 1876),
processes is close to non-existent because: (i) equilibrium data a necessary and sufficient condition for mass transfer equilib-
are scarce and should be experimentally obtained, (ii) existing rium in OD process is given by,
analytical solutions are only applicable for solutions of infinite    
volume (Crank, 1975) and (iii) physical properties necessary to
j =
j (7)
product osmotic media
evaluate convective mass transfer coefficient (such as viscos-
ity and density of solution, or diffusivity of solute in osmotic or, equivalently, in terms of water activity (Barbosa-Cánovas
media) are not readily available in comparison with other well- and Vega-Mercado, 2010)
characterized systems such water vapor–air mixtures.
The objective of this study was to develop a rigorous model (aw )product = (aw )osmotic media (8)
to describe the simultaneous water and solute interchange
between product and solution during an OD process, applica- According to Ross (1975), water activity in a multi-
ble under any ratio between product and osmotic media and component system can be developed as the product sequence
any convective mass transfer condition. To fully achieve this of water activities of each involved solute in a series of binary
purpose the following topics were covered: (i) mathematical mixtures. In this case,
description of mass equilibrium during OD, (ii) development
of a general model for unsteady state mass transfer of dif- N
aw = ˘ awi = aws awf (9)
fusing substances, (iii) evaluation of diffusion coefficients of i=1
water and solute and (iv) assessment of dominant mass trans-
fer phenomena under different experimental conditions. The Components such as proteins and carbohydrates, etc., are
model was validated using experimental OD data of carrot all lumped in Eq. (9) in the mixture of food solids. Since
slices in NaCl solutions. the effect of a given solute on water activity depression
mainly depends upon its molar fraction (Barbosa-Cánovas
and Vega-Mercado, 2010), it is clear that high-molecular
2. Theory
weight substances originally present in product such as
proteins, oligo- and polysaccharides will have a negligible
2.1. Basic definitions effect on water activity depression and thus in mass trans-
fer equilibrium. On the other hand, the impact of mono and
In order to develop an equilibrium relationship for OD process, disaccharides such as fructose, glucose and sucrose on OD
it will be considered that product (p) and osmotic (o) media are equilibrium will depend on the original amount of these sub-
made up from water (w), osmotic solute (s) and a mixture of stances in product and their lixiviated fraction, and the type
solids (f) which are naturally present in the product and may and concentration of osmotic solute. However, their effect on
leach into the osmotic solution, such as carbohydrates, fat and water activity depression may be also considered negligible
proteins. According to this assumption, under most OD processes performed with brines or syrups,
as will be later demonstrated. Therefore, if food solids do not
Xf + Xs + Xw = 1 (1) contribute to the water activity depression (awf ≈ 1) and the
osmotic solution is not in saturation, then mass transfer equi-
librium will be reached when the compositions of the binary
Yf + Ys + Yw = 1 (2)
mixtures between water and osmotic solute (i.e., excluding
food solids) in involved phases satisfy (for j = s, w)
where, X and Y represent the mass fractions of a given compo-
∗ ∗
nent in product and osmotic media, respectively. Let us define Xje = Yje = (1 + Ue ) Xje = (1 + We ) Yje (10)
the fractions of osmotic solute or water free of leaching food
solids in involved phases as (for j = s, w) From Eq. (10) it is clear that mass compositions Xje and Yje
are related by (for j = s, w)
Xj Yj
Xj∗ = and Yj∗ = (3) Yje = Ke Xje (11)
Xs + Xw Ys + Yw
food and bioproducts processing 9 7 ( 2 0 1 6 ) 88–99 91

1 + Ue equilibrium relationship relating mass fraction of specie j in


Ke = (12)
1 + We osmotic solution with those of the other components in prod-
uct and may have multiple mathematical representations.
Two important results are derived from Eqs. (11) and (12)
under the given assumptions: 2.4. Dimensionless form of osmotic dehydration model

(i) Partition coefficient (Ke ) is the same for both water and In this study, OD model is solved in dimensionless form under
osmotic solute. the following assumptions:
(ii) Ke is always greater than unity (since Ue > We ).
(1) Product has a constant density.
Additionally, if Ke does not change with the concentration (2) Apparent diffusivities of both solute and water in product
of osmotic solution, then the following statements apply: are constants.
(3) Mass transfer occurs in one direction in products with flat
(i) Equilibrium compositions for water and osmotic solute slab, infinite circular cylindrical or spherical geometries.
distribute along a straight-line equation with zero inter- (4) Equilibrium relationship Eq. (19) is given by Eq. (11).
cept.
(ii) Mass fraction of water in osmotic media at equilibrium Thus, Eqs. (16)–(19) can be rewritten as (for j = s, w),
only depends on its corresponding fraction in product.
The same statement holds for solute. ∂ ∂2 ˛∂ j
j j
(iii) Unsteady-state equations for water and osmotic solute = + (20)
∂Foj ∂ 2 ∂
concentrations can be solved independently from each
other. dj  
= − (˛ + 1) Bij ji + j (21)
dFoj
The previous statements still apply if it is considered
that food solids are not lost during OD process (We = 0 and ∂ ji  
Ke = 1 + Ue ). By assuming a negligible net mass change between = −Bij ji + j (22)

involved phases, the following steady-state mass balance can
be written (for j = s, w) where, ˛ takes value 0, 1, or 2, for infinite slabs, infinite
cylinders or spheres, respectively, while the following dimen-
mp Xj0 + Mo Yj0 = mp Xje + Mo Yje (13) sionless variables and groups were introduced

Finally, the combination of Eqs. (11) and (13) produce the fol- Xj − Xje Yj − Yje Dj t u
j = ; j = ; Foj = ; = (23)
lowing key relationships Xj0 − Xje Yj0 − Yje L2 L

Yj0 − Yje mp kc o 1 mp AL
=− =−
1
(14) Bij = Ke   ; = ; = (˛ + 1) (1 − ε) (24)
Xj0 − Xje Mo ı Djp /L p Ke Mo Vop

Xj0 + ıYj0 Here u and L represent the mass transfer direction (axial or
Xje = (15)
1 + ıKe radial) and the characteristic length for diffusion, respectively.
It can be observed that solution of dimensionless Eqs. (20)–(22)
2.3. General model for osmotic dehydration process only depends on parameters ˛, Bij and . Eq. (22) satisfies the
following limits:
The following equation system describes the mass transfer of
specie j (j = 1, . . ., n) between disperse and continuous phases ∂ ji
= −Bij ji for  → 0 (25)
in an OD process (without chemical reactions): ∂
 
∂ p Xj    ji =0 for  → 0 , Bij → ∞ (26)
= ∇ · Djp ∇ p Xj (16)
∂t
  ji = −j for Bij → ∞ (27)
d εVop o Yj  
= kc A o Yji − Yj (17)
dt Eqs. (25)–(27) represent the boundary conditions for (i)
    convective boundary with infinite surrounding media, (ii)
−Djp ∇ p Xji = kc o Yji − Yj (18) diffusion-controlled process with infinite surrounding media
and (iii) diffusion-controlled process with finite surrounding
Yji = fj (Xki ) for k = 1, . . ., n (19) media, respectively. Thus, Eqs. (20)–(22) share the analytical
solutions provided by Crank (1975) in these circumstances.
where, Vop is the combined volume of product and osmotic For simplicity but without loss of generality, the subscript j
media, kc is the convective mass transfer coefficient, A is the is hereafter dropped from equations.
surface area available for mass transfer, ε is the volume frac-
tion occupied by osmotic media, and the subscript i indicates 2.5. State-space representation of OD model and
at the product–solution interface. Eqs. (16) and (17) represent analytical solution
both the concentration change of specie j within a homoge-
neous and isotropic solid by diffusion and in osmotic solution Eq. (20) can be expressed as an equivalent system of ordinary
by convection, respectively, while Eq. (18) express the mass differential equations by using a discrete representation of
flux between product surface and bulk solution. Eq. (19) is an spatial derivatives. If central finite differences are used, then
92 food and bioproducts processing 9 7 ( 2 0 1 6 ) 88–99

the following model is obtained for 1 ≤ k ≤ N, where k repre- same advantages and limitations of the analytical ones pro-
sents a point in space vided by Crank (1975) (such as constant partition coefficient,
constant product density, etc.), with the added benefit that it
d k k+1 −2 k+ k−1 ˛ k+1 − k−1 can move between limiting cases. However, product shrinkage
= + (28)
dFo ( )
2 (k − 1)  2 or variable diffusivity could be also considered in the current
model by updating matrix A during time.
Points outside of the boundary where k < 1 or k > N, i.e., k+1
for k = N (interface) and k−1 for k = 1 (center), can be estimated 3. Methods
by applying convective boundary Eq. (22) and a symmetry con-
dition (only half domain is solved). Thus, 3.1. Osmotic dehydration experiments

k+1 = k−1 − 2 Bi ( k +  ) for k=N (29) The developed theory was applied to the analysis of OD data
sets using carrot as the food model system. All experiments
k−1 = k+1 for k = 1 (30) were conducted in NaCl solutions at a temperature of 40 ◦ C.
Fresh carrots, approximately 10 cm-long and 2.7 cm-diameter,
Eqs. (21) and (28)–(30) can be combined to form the state- were locally purchased (Puebla, Pue., México) and processed
space representation the same day. Carrots were washed, dried with a cloth and
then mechanically cut into circular 0.3 cm-thick slices with-
dX out removing the peel. On average, 4 slices were obtained
= AX (31)
dFo from the central part of each root while remaining portions
were reserved for the initial water/dry solids analysis. Carrot
where the state-vector X ∈ RN+1 is defined as
slices were OD-processed in separated dilution bottles placed
  in a controlled-temperature water bath (±0.1 ◦ C), which were
XT = 1 2 ... N  (32)
allowed to equilibrate before introducing samples. Samples
were osmodehydrated in separated bottles to eliminate varia-
and the non-zero elements of state-space matrix A ∈
tions in the solution-to-product ratio and avoid the mixing of
R(N+1)×(N+1) are
solution (in free convection experiments) that are introduced
a (1, 1) = − 2
; a (1, 2) = 2 by sampling removal when all samples are processed in a sin-
( )2 ( )2
(33)
gle recipient.

2
⎫ The following variables were recorded from these exper-
a (i, i) = − ⎬ iments at the beginning and during OD process: water
( )2
for 1 < i < N (34) fraction in product (Xw ), volumetric NaCl concentration
a (i, i ± 1) =
2(i−1)±˛ ⎭
2(i−1)( ) 2 (g/L) in osmotic media (Cs ), product mass (mp ) and solution
mass (Mo ). Water fraction in product was determined by
a (N, N − 1) = 2
; a (N, N) = −
[2(N−1)+˛] Bi+2(N−1) oven-drying the samples until constant weight was attained
(35)
( )2 (N−1)( )2
(when mass change was less than 0.001 g over and 8-h
period). On the other hand, NaCl concentration in osmotic
−[2(N−1)+˛] Bi
a (N, N + 1) = (36) media was determined with a sodium ion meter (HI 931101,
(N−1)( )2
Hanna Instruments, Woonsocket, RI, USA). Corresponding
a (N + 1, N) = − (˛ + 1) Bi; a (N + 1, N + 1) = − (˛ + 1) Bi (37) NaCl fractions in osmotic media (Ys ) were calculated from
the volumetric solution concentration (Cs ) and its density
Eq. (31) is an homogeneous linear time-invariant system of ( o ). Several experiments were completed in order to fulfill
ordinary differential equations, which accepts the well-known different objectives as follows. A brief description of all
solution (Ogata, 2010) experiments is given in Table 1.

X (t) = eAFo X (0) (38) 3.2. Estimation of equilibrium relationship

Mean concentration of water and osmotic solute in prod- The first set of OD experiments was conducted to character-
uct can be estimated by applying a spatial-averaging of local ize the effect of brine concentration on the mass equilibrium
values (Whitaker, 1977) relationship between product and osmotic media by identify-
ing parameter Ke . Carrot slices were processed without stirring
1 1 under several brine concentrations (20, 40, 60, 80, 100, 120, 140,
˛
= dVp / dVp = d / ˛ d (39) 160, 180 and 200 g/L NaCl solution) with a fixed ratio between
Vp Vp 0 0
solution volume and product mass (Vo /mp ) of 4:1. In these
An explicit solution to Eq. (39) was developed in this study experiments, carrot slices were immersed in brines for a total
from the trapezoidal rule, giving the expression of 24 h. This time was considered long enough for reaching
mass equilibrium in the product within the proposed solu-

tions according to previous similar studies with other plant
(˛ + 1) 
N−1
˛ ˛ ˛
= 0 1 +2 [(k − 1)  ] k + [(N − 1)  ] N materials (Herman-Lara et al., 2013). Mass balances for water,
2
k=2 osmotic solute and food solids allowed the straightforward
(40) estimation of Xse , Ywe , Ue , We and Ke by assuming Xs0 = 0 (prod-
uct initially has a negligible amount of osmotic solute) and
The state-space solution in Eq. (38) corresponds to a con- Yf0 = 0 (osmotic solution does not have food solids at the begin-
stant state-matrix A. In this case, the given solution shares the ning of the OD process).
food and bioproducts processing 9 7 ( 2 0 1 6 ) 88–99 93

Table 1 – Summary of all experiment sets.


Experiment set Purpose Experimental conditionsa

Stirring (rpm) Cs (g/L) Vo /mp (L/kg)

1 Evaluate equilibrium coefficients 0 20–200 4:1


2 Estimate diffusion coefficients 190 100 20:1
3 Appraise external resistance to mass transfer 0 100 4:1
4 Appraise external resistance to mass transfer 0 100 20:1
5 Appraise external resistance to mass transfer 0 150 20:1

a
All experiments were conducted at 40 ◦ C.

3.3. Estimation of effective water and solute terms of mass Grashof (Gr) and Schmidt (Sc) numbers. How-
diffusivities in carrot slices ever, under forced convection conditions, the mass Grashof
(Gr) is replaced by Reynolds (Re) number. These dimension-
Carrot slices were processed with a single brine concentration less groups are given below (Geankoplis, 1993; Büchs et al.,
(100 g/L NaCl solution) in separated dilution bottles (50 mm of 2000):
inner diameter) with a fixed Vo /mp ratio of 20:1. Bottles were
then placed in a thermal bath with orbital stirring (190 rpm). L3c sw g sw vd2 sw sw Lc kc
The Vo /mp ratio was selected to avoid a significant dilution Gr = , Re = , Sc = , Sh =
2sw sw sw Dsw Dsw
of osmotic solution, while the high stirring speed was used (41)
to allow for a negligible external resistance to mass trans-
fer (Herman-Lara et al., 2013). Samples were withdrawn from The Reynolds number (Re) definition in Eq. (41) is a special
the solution at predefined immersion times (5, 10, 15, 20, 25, form given by Büchs et al. (2000) for rotary shaking systems.
30, 40, 50, 60, 80, 100, 120, 150, 180, 210 and 250 min), gently The following correlations have been recommended for mass
blotted dry with a paper towel to remove adhering osmotic transfer around flat surfaces under natural (Geankoplis, 1993)
solution. Product and osmotic solution were analyzed for their and forced convection (Basmadjian, 2004), respectively
water/total solids content and NaCl as previously described.
Experimental OD curves (Xs vs t, Xw vs t) were fitted to OD Sh = 0.54(GrSc)1/4 (42)
model formed by Eqs. (38) and (40) in order to estimate Ds
and Dw , where carrot samples were considered to be slabs
(˛ = 0) due to their dimensions. Parameter  was calculated Sh = 0.66Re1/2 Sc1/3 (43)
from experimental variables Ke , Vo , mp0 and Cs0 as well as
Convective mass transfer coefficients were calculated
corresponding physical property o , estimated by solving Eq.
under studied experimental conditions and further used to
(A6). Equilibrium compositions were estimated with Eq. (15).
estimate a theoretical Bi number according to the definition
Osmotic solute fraction in product (Xs0 ) and food solids frac-
given in Eq. (24) and properties summarized in Appendix A.
tion in solution (Yf0 ) were assumed to be negligible in the
same way as during the estimation of equilibrium relation-
ship. In this case, two scenarios were considered; OD model 3.6. Data analysis
was solved with (i) Bis = Biw = 1000 in order to consider a
diffusion-controlled process and (ii) Bis and Biw were simul- Nonlinear regression (based on ordinary least squares) was
taneously estimated with Ds and Dw . used to estimate Ds , Dw , Bis and Biw according to the described
experiments (Table 1). Diffusion coefficients Ds and Dw were
3.4. Appraisal of external resistance to mass transfer individually fitted to experimental solute or water fraction
under no stirring conditions evolution curves from experiment set 2 by assuming a negligi-
ble external resistance to mass transfer (Bis and Biw were set to
Carrot slices were processed without stirring under the fol- 1000) or a convective boundary (Bis and Biw were also fitted dur-
lowing conditions: (i) 100 g/L NaCl solution with a Vo /mp ratio ing regression procedure). External resistance to mass transfer
of 20:1, (ii) 100 g/L NaCl solution with a Vo /mp ratio of 4:1 and in experiments conducted without stirring was appraised by
(iii) 150 g/L NaCl solution with a Vo /mp of 20:1 (experiment sets simultaneously fitting Bis to experimental solute and water
3–5 in Table 1). Remaining experimental details are the same fraction evolution curves from experiment sets 3–5, where
as above. Experimental OD curves were fitted to the proposed corresponding value for water transport was calculated as
model in order to estimate Bis and Biw using the values for Ds Biw = Ds Bis /Dw with Ds and Dw values obtained from exper-
and Dw determined as detailed in Section 3.3. Moreover, these iment set 2. In addition, Ds and Dw were also estimated by
data were also used to evaluate Ds and Dw , where Bis and Biw fitting experiment sets 3–5 with mass Biot numbers calcu-
were determined in advance with the aid of empirical mass lated with empirical mass transfer relationships. A summary
transfer correlations. of all properties used in unsteady-state model solution is pre-
sented in Table 2. The fitness quality of the identified model
3.5. Estimation of mass Biot number from empirical was quantified by the generalized determination coefficient
relationships (R2 ) and statistical significance of parameter estimates was
evaluated through their 95% confidence intervals (95% CI).
The value of external mass transfer coefficient kc can be cal- Nonlinear regression procedures as well as statistical analysis
culated from empirical correlations for the Sherwood number were performed using Matlab Statistics Toolbox 7.3 (Math-
(Sh). For natural convection, these expressions are given in Works Inc., Natick, MA, USA).
94 food and bioproducts processing 9 7 ( 2 0 1 6 ) 88–99

Table 2 – Summary of variables used during solution of unsteady-state OD modela .


Experiment set Xw0 (g/g)b Cs0 (g/L)b o (g/L)c Ys0 (g/g)d e Bij (j = s, w)

2 0.903 100.0 1058.1 0.094 0.0436 Estimated from experimental data


or set to 1000
3 0.904 100.3 1058.2 0.094 0.2181 Estimated from experimental data
4 0.904 100.3 1058.2 0.094 0.0436 or calculated with Eqs. (41)–(43)
5 0.909 150.2 1089.5 0.137 0.0424 and (A1)–(A6)

a
In all experiment sets: Xs0 = 0, Yf0 = 0, Ke = 1.08, ˛ = 0, N = 50, Xse and Xwe were calculated with Eq. (15).
b
Value determined during experiments.
c
Calculated by solving Eq. (A6).
d
Calculated from Cs0 .
e
Calculated with Eq. (24).

4. Results and discussion concentrations of about 0.15 g NaCl/g solution and above (del
Valle and Nickerson, 1967; Corzo and Bracho, 2004). It is readily
4.1. Equilibrium and distribution data observable that these water and solute data distribute along a
straight-line with approximately the same slope in both cases,
Eq. (15) provides a simple way to estimate final dehydration or confirming the fact that a single Ke value is able to describe the
impregnation levels in product based on experimental condi- mass equilibrium of these species according to the theoreti-
tions and equilibrium characteristics of product and osmotic cal analysis. In this case, Ke was identified as 1.083 with 95%
solute. This equation reduces to that previously reported by CI ranging from 1.075 to 1.092, i.e., the distribution constant
Beristain et al. (1990) if Ke = 1, i.e., by considering that osmotic is significantly higher than unity, which was an anticipated
media and product equilibrate at the same composition. How- result from the developed theory (Section 2.2).
ever, in some cases the model proposed by Beristain et al. It was considered that OD process reaches equilibrium
(1990) has failed to provide a satisfactory prediction of experi- when Xe∗ = Ye∗ , as stated in Eq. (10). A necessary condition to
mental mass equilibrium values (Parjoko et al., 1996). A quick validate this assumption is that food solids do not contribute
inspection of Eq. (15) reveals that overestimation of Xje will to water activity depression. On average, carrots slices lost
occur for increasing values of the difference Ke − 1, a fact that about the 33.6 ± 13.3% of their initial solids content. Sucrose,
was previously observed by Parjoko et al. (1996) and is now glucose and fructose account for about the 53% of the dry
explained by the lack of a proper equilibrium term in the matter content of carrot (Soujala, 2000; Nyman et al., 2005),
model reported by Beristain et al. (1990). the rest being both soluble and insoluble fiber, proteins and
Equilibrium data for water, NaCl and leached food solids fat. According to Nyman et al. (2005), sucrose, glucose and
are summarized in Table 3. According to these data, Ke values fructose represent the 3.8%, 1.25% and 1.14% of the fresh
did not exhibit a significant change with osmotic media con- weight of carrot, respectively. By assuming that all lixiviated
centration (p < 0.05) (Fig. 1). Thus, Ke can be simply estimated solids are sucrose, glucose and fructose in their initial pro-
as the slope of the Yje vs Xje plot simultaneously consider- portions, estimated differences in water activity calculated
ing data for both NaCl and water, as shown in Fig. 2. An with Ross equation and Raoult’s law are less than 0.1% than
excellent agreement was found between experimental and that obtained by neglecting these components. A compara-
predicted results (R2 = 0.998). Appreciable deviations of this ble result was obtained in product where differences in water
linear behavior have been reported by other authors for NaCl activity with and without considering sucrose, glucose and
fructose in the calculations were lower than 0.5%. Overall, pre-
dicted differences in water activity between both phases at

Fig. 1 – Effect of NaCl fraction in osmotic media on


equilibrium partition coefficient for osmodehydrated carrot
at 40 ◦ C. Experimental data were calculated with Eq. (12). Ke
in continuous line is the slope of the plot Yje vs Xje (for Fig. 2 – Distribution data for NaCl and water during osmotic
j = s, w). dehydration of carrot at 40 ◦ C.
food and bioproducts processing 9 7 ( 2 0 1 6 ) 88–99 95

Table 3 – Equilibrium dehydration and impregnation levels of osmotic dehydration of carrot slicesa .
Cs0 (g/L) Ue × 102 (g/g) We × 103 (g/g) Xse × 101 (g/g) Yse × 101 (g/g) Xwe × 101 (g/g) Ywe × 101 (g/g) Ke (g/g)

20 9.7 ± 0.8 3.9 ± 1.0 0.05 ± 0.01 0.17 ± 0.00 9.07 ± 0.07 9.79 ± 0.01 1.09 ± 0.01
40 7.9 ± 0.6 7.6 ± 0.3 0.32 ± 0.04 0.29 ± 0.01 8.95 ± 0.04 9.64 ± 0.01 1.07 ± 0.01
60 7.5 ± 0.0 8.5 ± 0.7 0.52 ± 0.04 0.41 ± 0.01 8.79 ± 0.04 9.51 ± 0.01 1.07 ± 0.00
80 6.5 ± 0.9 10.2 ± 1.6 0.76 ± 0.05 0.52 ± 0.01 8.62 ± 0.03 9.38 ± 0.01 1.06 ± 0.01
100 7.6 ± 0.7 9.2 ± 2.4 0.88 ± 0.12 0.68 ± 0.03 8.41 ± 0.09 9.22 ± 0.01 1.07 ± 0.01
120 7.8 ± 0.8 8.9 ± 3.0 0.98 ± 0.13 0.84 ± 0.03 8.30 ± 0.11 9.07 ± 0.03 1.07 ± 0.01
140 7.7 ± 3.1 7.0 ± 3.0 1.05 ± 0.09 1.00 ± 0.02 8.24 ± 0.19 8.93 ± 0.01 1.07 ± 0.03
160 4.8 ± 1.6 11.6 ± 0.8 1.40 ± 0.07 1.11 ± 0.01 8.14 ± 0.18 8.78 ± 0.01 1.04 ± 0.02
180 6.8 ± 2.4 7.7 ± 1.6 1.41 ± 0.08 1.25 ± 0.02 7.95 ± 0.14 8.67 ± 0.00 1.06 ± 0.03
200 9.6 ± 0.8 3.2 ± 0.4 1.34 ± 0.11 1.43 ± 0.11 7.78 ± 0.13 8.54 ± 0.03 1.09 ± 0.01
psdb 1.5 1.8 0.08 0.04 0.12 0.01 0.02

Temperature = 40 ◦ C, Vo /mp = 4 L/kg, slice diameter = 2.7 cm, slice thickness = 0.3 cm.
a
Data expressed as means ± standard deviation of three independent experiments.
b
Pooled standard deviation.

equilibrium were in the order of the 1.34%. These results give these cases, deviations from the theoretical behavior may be
support to the theoretical analysis presented in Section 2.2. due to an increased mass transfer resistance caused by higher
Several authors have used partition constants, such as that syrup viscosities and solute deposition on product surface,
Xje
defined by Eq. (11), albeit in their reciprocal form ( K1e = Yje ),
which may lead to a pseudo-equilibrium state.
to describe OD equilibrium (del Valle and Nickerson, 1967;
Parjoko et al., 1996). These values were transformed when 4.2. Dehydration and impregnation curves
necessary to allow a direct comparison with our results. For
example, del Valle and Nickerson (1967) obtained distribution Fig. 3 shows the evolution of solute and water fractions
constants for NaCl in swordfish muscle ranging from 1.09 to during OD of carrot slices under forced convection and the
1.43 (5–37 ◦ C, 0.01–0.31 g/g NaCl solution). Sarang and Sastry ratio between solution volume and product mass (experiment
(2007) reported distribution coefficients for NaCl from 1.04 to set 2, Table 1). These data were used to evaluate diffusion
1.09 during OD of Chinese water chesnut (25–80 ◦ C, 50–100 g/L coefficients assuming both a negligible external resistance to
NaCl solution). A similar study was conducted by Corzo and mass transfer (Bis and Biw were set to 1000) or a convective
Bracho (2004), reporting separated distribution coefficients for boundary (Bis and Biw were also fitted during regression pro-
water (from 1.60 to 2.00) and NaCl (from 1.28 to 1.89) during cedure). By assuming a diffusion-controlled process, water
osmotic dehydration of sardine sheets (32–38 ◦ C, 0.15–0.27 g/g and NaCl diffusivities were estimated as 6.0 × 10−10 m2 /s
NaCl solution). Corzo and Bracho (2004) considered that and 4.1 × 10−10 m2 /s, respectively (Table 4), allowing a good
reproduction of experimental dehydration and impregnation
mpe = mp0 − (mw0 − mwe ) + (mse − ms0 ) (44) curves (R2 > 0.96, Fig. 3). These diffusion coefficients remained
mostly unaffected (values changed between 4% and 6% of
their original value) when Bis and Biw were also estimated
ignoring the lixiviation of food solids and, according to Eq.
with data from the same experiment achieving a similar
(5), leading to the underestimation of Xse and Xwe and to the
fitness quality (R2 = 0.97) with overlapping model curves.
overestimation of Kse and Kwe . Our research group proposed a
However, while estimated Biot numbers were consistent with
method to estimate Kse from total solids gain data by estimat-
a diffusion-controlled process (Bis = 74.4 and Bis = 50.7), their
ing Xse as (Herman-Lara et al., 2013)
95% CI indicated a non-significant parameter estimation.

Xse = 1 − Xf 0 − Xwe (45)

This procedure, while neglecting leaching of food solids


as well, leads to an overestimation of Xse and the underes-
timation of Kse . This fact may explain some distribution for
NaCl and water during OD of radish (40 ◦ C, 0.05–0.25 g/g NaCl
solution) with values as low as 0.66 and 0.88, respectively
(Herman-Lara et al., 2013).
Several studies have also investigated the distribution
coefficients of diffusing substances during OD process in sugar
solutions (Parjoko et al., 1996; Rahman et al., 2001; Sablani
et al., 2002; Toğrul and İspir, 2008; Ruiz-López et al., 2011b).
However, in all cases osmotic solute has been grouped with
solids originally present in foods. Thus, a direct comparison
with the results obtained in this work is not attainable. On the
other hand, distribution coefficients for water have shown a
different trend to that expected in this study, with Kwe values Fig. 3 – Experimental and fitted evolution of water and
very often lower than unity, especially at high syrup concen- sodium chloride mass fractions in product during osmotic
trations, indicating that water cannot be eliminated even if dehydration of carrot slices using a 10% NaCl solution at
osmotic media provides the driving force for mass transfer. In 40 ◦ C with stirring (190 rpm) and a high Vo /mp ratio (20:1).
96 food and bioproducts processing 9 7 ( 2 0 1 6 ) 88–99

Table 4 – Estimated diffusion coefficients and mass Biot numbers during OD of carrot slices.
Cs0 (g/L) Stirring (rpm) Vo /mp (L/kg) Parameters Valuea 95% CIb R2 Predicted Bic

100 190 20:1 Dw (m2 /s)d 6.0 5.4/6.6 0.966 Biw = 87.9
Ds (m2 /s)d 4.1 3.8/4.5 0.976 Bis = 126.7
Dw (m2 /s) 6.3 4.5/8.2 0.972e
Ds (m2 /s) 4.3 3.4/5.3

Bis 74.4 266.2/414.9
(Biw ) 50.7

100 0 4:1 Bis 26.4 0.9/53.7 0.862e,f Bis = 21.3
(Biw ) 18.3 Biw = 14.8
Dw (m2 /s)d 7.6 5.3/10.0 0.787
Ds d 3.5 3.3/3.9 0.991

100 0 20:1 Bis 89.0 113.4/291.4 0.932e,f Bis = 27.0
(Biw ) 61.7 Biw = 18.8
Dw (m2 /s)d 7.0 5.5/8.4 0.891
Ds (m2 /s)d 3.7 3.6/3.9 0.991

150 0 20:1 Bis 4148 4.1E6/4.2E6 0.925e,f Bis = 28.3
(Biw ) 2877 Biw = 19.7
Dw (m2 /s)d 6.1 4.9/7.2 0.878
Ds (m2 /s)d 3.6 3.3/3.9 0.975

a
Diffusivity values × 1010 .
b
Bold numbers indicate non-significant parameter estimates at the given confidence level.
c
Estimated from empirical correlations.
d
Diffusivity values estimated by considering a negligible external resistance to mass transfer (Bis = Biw = 1000).
e
Parameters estimated by simultaneously fitting Xs and Xw curves during nonlinear regression with Biw = Ds Bis /Dw .
f
Dw and Ds values were obtained from experiment set 2.

Empirical correlations allowed the estimation of Biot number outside product (Figs. 4 and 5). In this case, minimum Bi num-
for water and NaCl transport under these experimental bers for water and solute transfer were identified as 18.30
conditions as 87.9 and 126.7, respectively. In both cases, and 26.38 (Table 4), respectively. These values, obtained with
these values indicate a strictly diffusion-controlled process, the lowest Vo /mp ratio (4:1) (experiment set 3, Table 1), are
which is reached for practical purposes when mass Biot near the limit of a diffusion-controlled process. On the other
number is higher than 30 or 40 in analytical solutions for hand, the analysis of remaining experimental data produced
invariable properties of continuous phase (Córdova et al., mass Biot numbers estimations well in the zone of a diffusion-
1996; Yanniotis, 2008; Ruiz-López et al., 2011a). Under such controlled process (Biw > 61, Bis > 88). However, under natural
conditions, theoretical model curves overlap no matter the convection conditions, all regression procedures produced
mass Biot value used to generate them. In this case, as long non-significant estimations of Bi numbers (p = 0.05). The Bi
as Bi > 30 (process is diffusion-controlled) water and solute values estimated with empirical correlations allow gaining
diffusivities will be estimated free of convective mass transfer a better insight into the mass transfer phenomena and the
effects, but any Bi value will produce a good fit of experimental lack of significance of regression parameters. Under natural
data, explaining the lack of significance of this parameter. convection conditions Bi values are in the ranges of 14.8–19.7
OD experiments conducted without stirring were pre- and 21.3–28.3 for water and NaCl transfer, respectively. Again,
sumed to show an important resistance to mass transfer these values are near the limit of a diffusion-controlled

Fig. 4 – Effect of solution concentration on dehydration Fig. 5 – Effect of solution mass-to-product volume ratio on
(j = w) and impregnation (j = s) curves of carrot slices (40 ◦ C, dehydration (j = w) and impregnation (j = s) curves of carrot
no stirring and Vo /mp = 20 : 1 L/kg). slices (40 ◦ C, no stirring and 10% NaCl).
food and bioproducts processing 9 7 ( 2 0 1 6 ) 88–99 97

Fig. 6 – Effect of mass Biot number on dehydration (j = w) Fig. 7 – Effect of Vo /mp ratio on dilution of osmotic media.
and impregnation (j = s) curves of carrot slices (40 ◦ C, 10%
NaCl and a Vo /mp ratio of 20:1). Arrows indicate the
et al., 2006). Comparable results were obtained during OD
direction in which Bi is increased. Numbers refer to the Bi
of cherry tomato (25 ◦ C, 0.10 and 0.25 g solute/g solution)
values used in model solution.
with water and NaCl diffusivities values in the ranges of
1.2–1.8 × 10−9 and 0.2–0.5 × 10−9 m2 /s (Azoubel and Murr,
2004). Sarang and Sastry (2007) reported NaCl diffusivities in
process where a change in Bi value will not alter the model
the range of 7.5–18.5 × 10−10 during OD of Chinese water ches-
behavior in an important way. This fact jointly with experi-
nut (25–80 ◦ C, 50–100 g/L NaCl solution).
mental data dispersion did not allow a reliable identification
There are few reports regarding the estimation of Bi number
of Bi. In fact, the rate at which mass transfer occurs under
during OD of foods. Selected studies include that of da Silva
natural convection conditions is practically the same to that
et al. (2013), who performed the OD of coco parallelepipeds
observed under forced convection conditions, causing exper-
(6 mm × 10 mm × 31 mm) in sucrose solutions under forced
imental data to overlap, as shown in Fig. 6. Moreover, this
convection conditions (magnetic stirring at 60 Hz, 35 ◦ Bx, 40 ◦ C)
figure also exhibits the minimal separation between simulated
and reported fitted values as Biw = 7.75 and Bis = 5.00. In a later
OD curves obtained under different Bi values ranging from
work, da Silva et al. (2014) studied the OD of pineapple cubes
10 to 1000, which is well between the observed experimental
(side length of 15 mm) in sucrose solutions without stirring
dispersion. In this way, diffusion coefficients could be also esti-
(40 and 70 ◦ Bx, 30 ◦ C) estimating Bi values in the ranges of
mated from experiments conducted without stirring as shown
1.5–2.3 and 0.62–0.95 for water and solute transport, respec-
in Table 4. This time, water diffusivities were estimated in
tively. Unlike current study, Bi values for solute transfer were
the range of 6.1–7.6 × 10−10 m2 /s, while corresponding values
smaller than those estimated for water; however, it should be
for NaCl are between 3.5 × 10−10 and 3.7 × 10−10 m2 /s. In this
highlighted that da Silva et al. (2013, 2014) analyzed solute
case, the 95% CI for these values overlap with those previously
transfer as total solids and did not report a statistical analy-
calculated under forced convection conditions.
sis of estimated parameters to allow a further comparison of
Figs. 4 and 5 allow observing the effect of solution con-
observed differences with values obtained in this study.
centration and solution-to-product ratio on impregnation and
dehydration curves. From Eq. (15) it is clear that the use of
a higher solute concentration (Ys0 ) will led to an increase 5. Conclusions
of water loss and solute gain at equilibrium. On the other
hand, decreasing Vo /mp (and thus Mo /mp ) will led to a reduc- Proposed models allowed the satisfactory description of both
tion of water loss and solute gain at equilibrium. Moreover, dynamic and equilibrium mass transfer periods for water,
as Mo /mp → ∞, Xje → Yj0 /Ke and when Mo /mp → 0, Xje → Xj0 . osmotic solute and food solids interchange between prod-
All these cases can be predicted with current equilibrium and uct and solution during an osmotic dehydration (OD) process.
unsteady-state models. A notable feature of proposed model The mathematical description of OD equilibrium helped to
is that, in addition to describing water and solute fractions explain differences between previous studies, describing the
in the product, it also allows the prediction of osmotic media necessary adjustments to eliminate them and providing well-
concentration evolution, which may change in a significant founded bases for further investigations. A better insight of
way especially in experiments conducted with a low Vo /mp mass transfer phenomena occurring during OD of carrot slices
ratio. For example, by lowering the Vo /mp ratio from 20:1 to in NaCl solutions was gained, where convection was almost
4:1, the final dilution of osmotic media increases from 4% to negligible even in non-stirred osmotic media. Further stud-
18% (Fig. 7). ies are required to evaluate governing mechanism for mass
Estimated diffusivities for water were smaller than the cor- transfer in other systems.
responding NaCl values, which may be very likely due to its
lower molecular weight as reported elsewhere (Panagiotou Acknowledgments
et al., 1999). Diffusivity values are comparable to those
obtained during OD of other foodstuffs. For example, water The authors wish to thank the Consejo Nacional de Cien-
and NaCl diffusivities during OD of potato (25–55 ◦ C, 0.10–0.18 g cia y Tecnología (CONACYT) for providing financial support
solute/g solution) have been estimated in the ranges of through project 130011. Hermelinda Pacheco Angulo acknowl-
8.7–12.3 × 10−10 and 8.2–12.2 × 10−10 m2 /s, respectively (Khin edges her doctoral scholarship from CONACYT. Miguel Ángel
98 food and bioproducts processing 9 7 ( 2 0 1 6 ) 88–99

García Alvarado acknowledges his sabbatical financial support References


from CONACYT.
Azoubel, P.M., Murr, F.E.X., 2004. Mass transfer kinetics of osmotic
Appendix A. Properties used in empirical dehydration of cherry tomato. J. Food Eng. 61 (3), 291–295.
prediction of mass Biot numbers Barbosa-Cánovas, G.V., Vega-Mercado, H., 2010. Dehydration of
Foods. Chapman & Hall, Lexington, Kentucky, USA.
Basmadjian, D., 2004. Mass Transfer—Principles and
Diffusion coefficient for NaCl–water mixture. Poling et al. (2001) Applications. CRC Press, Boca Raton, Florida, USA.
proposed the following equation to describe the effect of Bellary, A.N., Sowbhagya, H.B., Rastogi, N.K., 2011. Osmotic
solute concentration on mutual diffusion coefficients in elec- dehydration assisted impregnation of curcuminoids in
trolyte solutions of a single salt: coconut slices. J. Food Eng. 105 (3), 453–459.
Beristain, C.I., Azuara, E., Cortés, R., García, H.S., 1990. Mass
w transfer during osmotic dehydration of pineapple rings. Int. J.
Dsw = D◦sw ϕ (A1)
sw Food Sci. Technol. 25 (5), 576–582.
Büchs, J., Maier, U., Milbradt, C., Zoels, B., 2000. Power
where, the diffusion coefficient at infinite solution (D◦sw ) and consumption in shaking flasks on rotary shaking machines: I.
Power consumption measurement in unbaffled flasks at low
thermodynamic factor (ϕ) are given by (Poling et al., 2001)
viscosity. Biotechnol. Bioeng. 68 (6), 589–593.
Córdova, A.V., Ruiz-Cabrera, M.A., García-Alvarado, M.A., 1996.
RT 1/n+ + 1/n−
D◦sw = (A2) Analytical solution of mass transfer equation with interfacial
F2 1/◦+ + 1/◦− resistance in food drying. Dry. Technol. 14 (7–8),
1815–1826.
∂ ln ± m ∂± Corzo, O., Bracho, N., 2004. Effects of brine concentration and
ϕ =1+m =1+ (A3)
∂m ± ∂m temperature on equilibrium distribution coefficients during
osmotic dehydration of sardine sheets. LWT—Food Sci.
with + = 5.01 × 10−3 and − = 7.63 × 10−3 S/m2 mol for Na+ Technol. 37 (4), 475–479.
cation (n+ = 1) and Cl− anion (n− = 1), respectively (Pitzer et al., Crank, J., 1975. The Mathematics of Diffusion, Second ed. Oxford
University Press, Oxford, England.
1984).
da Silva, W.P., do Amaral, D.S., Duarte, M.E.M., Mata, M.E.R.M.C., e
Mean ionic activity coefficient ( ± ) of solute. Tabulated data
Silva, C.M.D.P.S., Pinheiro, R.M.M., Pessoa, T., 2013. Description
(100 kPa) in the ranges of 0.1–6 molal NaCl and 0–100 ◦ C pro- of the osmotic dehydration and convective drying of coconut
vided by Pitzer et al. (1984) were fitted in this study with (Cocos nucifera L.) pieces: a three-dimensional approach. J.
predictions within the ±4% of the original data (R2 = 0.9895): Food Eng. 115 (1), 1231–2131.
da Silva, W.P., e Silva, C.M.D.P.S., Lins, M.A.A., Gomes, J.P., 2014.
Osmotic dehydration of pineapple (Ananas comosus) pieces in
cubical shape described by diffusion models. LWT—Food Sci.
± = 0.7358 + 0.2170x2 − 0.0591x12 − 0.0472x12 x2 − 0.0303x1 x22
Technol. 55 (1), 1–8.
+ 0.0098x13 − 0.0972x23 + 0.1393x24 (A4) del Valle, F.R., Nickerson, J.T.R., 1967. Studies on salting and
drying fish. I. Equilibrium considerations in salting. J. Food Sci.
32 (2), 173–179.
where, x1 = [T (◦ C) − 50] /50 and x2 = (m − 3.05) /2.95.
Geankoplis, C.J., 1993. Transport Processes and Unit Operations,
Dynamic viscosity of solvent (w ) and solution (sw ). Tabulated Third ed. Prentice-Hall International, New Jersey, USA.
data (100 kPa) in the ranges of 0–6 molal NaCl and 20–100 ◦ C Gibbs, J.W., 1876. On the equilibrium of heterogeneous
provided by Kestin et al. (1981) were fitted in this study with substances. Transactions of the Connecticut Academy of Art
predictions within the ±3% of the original data (R2 = 1.0000): and Sciences, 3 (5). Yale University Press, New Haven, USA, pp.
108–248.
Goula, A.M., Lazarides, H.N., 2012. Modeling of mass and heat
sw × 103 = 0.6487 − 0.3626x1 + 0.2179x2 − 0.1289x1 x2
transfer during combined processes of osmotic dehydration
+ 0.1745x12 + 0.0466x22 + 0.0782x12 x2 − 0.0499x1 x22 and freezing (osmo-dehydro-freezing). Chem. Eng. Sci. 82,
52–61.
− 0.1094x13 + 0.0092x23 − 0.0367x13 x2 Herman-Lara, E., Martínez-Sánchez, C.E., Pacheco-Angulo, H.,
Carmona-García, R., Ruiz-Espinosa, H., Ruiz-López, I.I., 2013.
+ 0.0300x12 x22 + 0.0491x14 (A5) Mass transfer modeling of equilibrium and dynamic periods
during osmotic dehydration of radish in NaCl solutions. Food
where, x1 = [T (◦ C) − 60] /40 and x2 = (m − 3) /3. Eq. (A5) Bioprod. Process. 91 (3), 216–224.
becomes w when x2 = −1. Kaymak-Ertekin, F., Sultanoğlu, M., 2000. Modelling of mass
transfer during osmotic dehydration of apples. J. Food Eng. 46
Density of solvent ( w ) and solution ( sw ). Tabulated data
(4), 243–250.
(100 kPa) in the ranges of 0–6 molal NaCl and 0–100 ◦ C pro-
Khan, M.A.M., Ahrné, L., Oliveira, J.C., Oliveira, F.A.R., 2008.
vided by Pitzer et al. (1984) and Geankoplis (1993) were fitted Prediction of water and soluble solids concentration during
in this study with predictions within the ±3% of the original osmotic dehydration of mango. Food Bioprod. Process. 86 (1),
data (R2 = 1.0000): 7–13.
Kestin, J., Khalifa, H.E., Correia, R.J., 1981. Tables of the dynamic
sw = 1093.4898 − 26.8911x1 + 94.6352x2 − 2.5050x1 x2 and kinematic viscosity of aqueous NaCl solutions in the
temperature range 20–150 ◦ C and the pressure range
− 3.8475x12 − 8.9248x22 + 3.2483x12 x2 + 1.3048x1 x22 0.1–35 MPa. J. Phys. Chem. Ref. Data 10 (1), 71–87.
Khin, M.M., Zhou, W., Perera, C.O., 2006. A study of the mass
+ 0.7381x13 + 1.4752x23 − 0.9954x13 x2 − 0.5638x1 x23 transfer in osmotic dehydration of coated potato cubes. J.
Food Eng. 77 (1), 84–95.
− 1.3408x12 x22 − 0.3065x14 − 0.3796x24 (A6) Nyman, E.M.G.-L., Svanberg, S.J.M., Andersson, R., Nilsson, T.,
2005. Effects of cultivar, root weight, storage and boiling on
where, x1 = [T (◦ C) − 50] /50 and x2 = (m − 3) /3. Eq. (A6) carbohydrate content in carrots (Daucus carota L). J. Sci. Food
becomes w when x2 = −1. Agric. 85 (3), 441–449.
food and bioproducts processing 9 7 ( 2 0 1 6 ) 88–99 99

Ogata, K., 2010. Modern Control Engineering, Fifth ed. Prentice (Averrhoa carambola L.) in sugar solutions. J. Food Eng. 104 (2),
Hall, New Jersey, USA. 218–226.
Panagiotou, N.M., Karathanos, V.T., Maroulis, Z.B., 1999. Effect of Sarang, S., Sastry, S.K., 2007. Diffusion and equilibrium
osmotic agent on osmotic dehydration of fruits. Dry. Technol. distribution coefficients of salt within vegetable tissue: effects
17 (1–2), 175–189. of salt concentration and temperature. J. Food Eng. 82 (3),
Parjoko, Rahman, M.S., Buckle, K.A., Perera, C.O., 1996. Osmotic 377–382.
dehydration kinetics of pineapple wedges using palm sugar. Sablani, S.S., Rahman, M.S., Al-Sadeiri, D.S., 2002. Equilibrium
LWT Food Sci. Technol. 29 (5–6), 452–459. distribution data for osmotic drying of apple cubes in
Pitzer, K.S., Pelper, J.C., Busey, R.H., 1984. Thermodynamic sugar-water solution. J. Food Eng. 52 (2), 193–199.
properties of aqueous sodium chloride solutions. J. Phys. Sablani, S.S., Rahman, M.S., 2003. Effect of syrup concentration,
Chem. Ref. Data 13 (1), 1–102. temperature and sample geometry on equilibrium
Poling, B.E., Prausnitz, J.M., O’Connell, J.P., 2001. The Properties of distribution coefficients during osmotic dehydration of
Gases and Liquids, Fifth ed. McGraw-Hill, New York, USA. mango. Food Res. Int. 36 (1), 65–71.
Porciuncula, B.D.A., Zotarelli, M.F., Carciofi, A.M., Laurindo, J.B., Singh, B., Kumar, A., Gupta, A.K., 2007. Study of mass transfer
2013. Determining the effective diffusion coefficient of water kinetics and effective diffusivity during osmotic dehydration
banana (Prata variety) during osmotic dehydration and its use of carrot cubes. J. Food Eng. 79 (2), 471–480.
in predictive models. J. Food Eng. 119 (3), 490–496. Soujala, T., 2000. Variation in sugar content and composition of
Rahman, M.S., Sablani, S.S., Al-Ibrahim, M.A., 2001. Osmotic carrot storage roots at harvest and during storage. Sci. Hortic.
dehydration of potato: equilibrium kinetics. Dry. Technol. 19 85 (1–2), 1–19.
(6), 1163–1176. Souraki, B.A., Ghaffari, A., Bayat, Y., 2012. Mathematical modeling
Rastogi, N.K., Raghavarao, K.S.M.S., 2004. Mass transfer during of moisture and solute diffusion in the cylindrical green bean
osmotic dehydration determination of moisture and solute during osmotic dehydration in salt solution. Food Bioprod.
diffusion coefficients from concentration profiles. Food Process. 90 (3), 64–71.
Bioprod. Process. 82 (1), 44–48. Souraki, B.A., Ghavami, M., Tondro, H., 2013. Mass transfer during
Rodríguez, M.M., Arballo, J.R., Campañone, L.A., Cocconi, M.B., osmotic dehydration of green bean in salt solution: a
Pagano, A.M., Mascheroni, R.H., 2013. Osmotic dehydration of polynomial approximation approach. Food Bioprod. Process.
nectarines: influence of the operating conditions and 91 (3), 257–263.
determination of the effective diffusion coefficients. Food Souraki, B.A., Ghavami, M., Tondro, H., 2014. Correction of
Bioprocess Technol. 6 (10), 2708–2720. moisture and sucrose effective diffusivities for shrinkage
Ross, K.D., 1975. Estimation of water activity in intermediate during osmotic dehydration of apple in sucrose solution. Food
moisture foods. Food Technol. 29 (3), 26–34. Bioprod. Process. 92 (1), 1–8.
Ruiz-López, I.I., Ruiz-Espinosa, H., Guevara-Luna, M.L., Toğrul, I.T., İspir, A., 2008. Equilibrium distribution coefficients
García-Alvarado, M.A., 2011a. Modeling and simulation of heat during osmotic dehydration of apricot. Food Bioprod. Process.
and mass transfer during drying of solids with hemispherical 86 (4), 254–267.
shell geometry. Comput. Chem. Eng. 35 (2), 191–199. Yanniotis, S., 2008. Solving Problems in Food Engineering.
Ruiz-López, I.I., Ruiz-Espinosa, H., Herman-Lara, E., Springer, New York, USA.
Zárate-Castillo, G., 2011b. Modeling of kinetics, equilibrium Whitaker, S., 1977. Toward a diffusion theory of drying. Ind. Eng.
and distribution data of osmotically dehydrated carambola Chem. Fundam. 16 (4), 408–414.

You might also like