Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Journal of Food Engineering 130 (2014) 36–44

Contents lists available at ScienceDirect

Journal of Food Engineering


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

Thermodynamic analysis of salting cheese process


J. Velázquez-Varela, P.J. Fito, M. Castro-Giráldez ⇑
Instituto Universitario de Ingeniería de Alimentos para el Desarrollo, Universidad Politécnica de Valencia, Camino de Vera s/n, 46022 Valencia, Spain

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

Article history: The manufacture of cheese is a complex process which involves several phases and biochemical transfor-
Received 21 October 2013 mations. All these variables have important influence over the performance, composition and quality of
Received in revised form 16 January 2014 cheese. Therefore, the optimization of the manufacturing process of this product is difficult. Salting
Accepted 25 January 2014
process involves complex phenomena which are hardly to explain, and one of its main problems is the
Available online 31 January 2014
difficulty to control de simultaneous water and salt fluxes into the product and to control the proteins
transformations throughout process. Water and salt transport during osmotic treatment of cheese in
Keywords:
saturated brine, were investigated by using a thermodynamic approach. Moisture, water activity, dielec-
Cheese
Salting process
tric properties and volume were measured in treated and reposed samples. A nonlinear irreversible ther-
Nonlinear irreversible thermodynamic modynamic model has been developed to determine the transport of water and ions through the
model structure of cheese, obtaining a phenomenological coefficient of 1.8  105 mol2 J1 s1 m2. This model
describes various phenomena that occur in salting cheese process, including the different driving forces.
Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction Cuartirolo Argentino cheese, and the value of effective diffusion


coefficient was 3.6  1010 m2 s1. Yanniotis and Anifantakis
Cheese is a colloidal system formed by three phases (protein, fat (1983) for feta cheese, the value of effective diffusion coefficient
and aqueous) in electrical equilibrium (gel conformation). There- was 2.3  1010 m2 s1. Hardy (1976) studied Camembert cheese and
fore, the analysis of microstructure in cheese could be necessary the value of effective diffusion coefficient was 2.54  1010 m2 s1.
to understand the mechanisms involved during salting process These estimates were obtained with Fick’s model, applying the
(Aguilera and Stanley, 1999; Fito et al., 2007). simplifications of this model. These simplifications allow its use only
Salting treatment is one operation commonly used in dairy in systems with ideal conditions, without changing the state variables.
industry to preserve and to conform the final quality of cheese. For this reason the values obtained are different in each experimental.
In osmotic treatments, native and added chemical compounds However, there are different mechanisms involved during the
fluxes are produced by chemical potential gradients (Castro-Girál- mass transfer of solutes and it is difficult to establish an individual
dez et al., 2010a). effect of each component due to the existence of strong interac-
The flux of solutes among the heterogeneous phases of the tions between them (Castro-Giráldez et al., 2010a; Fito et al.,
cheese matrix has been reported for salting cheese processes. This 2007; Lawrence and Gilles, 1982). It has been observed that the
process has been frequently described by the second Fick’s law in changes occurred in this treatment are produced by different
terms of effective diffusion coefficient; Floury et al. (2009) ob- driving forces, such as chemical and mechanical free energies
tained an effective diffusion coefficient of 2.81–3.43  1010 m2 s1 (Castro-Giráldez et al., 2011).
for model cheese systems. Baroni et al. (2003) studied this process Even though the salting treatment is not a novel operation, the
for Prato cheese, obtaining a value for effective diffusion coefficient mass transport phenomena involved are not yet fully understood.
of 1.64–4.25  1010 m2 s1; Turhan and Gunasekaran (1999) stud- Usually the models oversimplify the food system (Fito et al.,
ied white cheese, obtaining a value for effective diffusion coeffi- 2008) which is supposed to be isotropous, homogeneous and con-
cient of 2.1–3.4  1010 m2 s1; Guinee and Fox (2004) studied tinuous, with few components and phases (Crank, 1985). The ther-
Romano cheese, obtaining a value for effective diffusion coefficient modynamic has the capability to identify pathways of various
of 2.54–3.35  1010 m2 s1; Luna and Bressan (1986) studied processes, providing data of the mechanisms that are involved.
Various new approaches in thermodynamics have been developed
to understand the transport phenomena in systems. The theory of
non-equilibrium thermodynamics are based on the irreversible
⇑ Corresponding author. Tel.: +34 96 387 98 32.
character of fluxes and forces of an open system, and describe
E-mail address: marcasgi@upv.es (M. Castro-Giráldez).

http://dx.doi.org/10.1016/j.jfoodeng.2014.01.017
0260-8774/Ó 2014 Elsevier Ltd. All rights reserved.
J. Velázquez-Varela et al. / Journal of Food Engineering 130 (2014) 36–44 37

the transports in the processes coupled with chemical reactions considered to be surface aw. Water content was determined by
(Demirel, 2002a,b). AOAC Method 934.06 (2000). The volume was determined by image
The objective of this work was to describe the salting cheese analysis (CanonÒ Power Shot SX210 IS), using Adobe PhotoshopÒ
process by using a thermodynamic approach, taking into account software (Adobe Systems Inc., San Jose, CA, USA) in order to get
the structure of cheese matrix in water and salt transport the diameter and the thickness of the samples. Ion and cation
phenomena. content was carried out by means of an ion chromatograph
(Methrom Ion Analysis, Herisau, Switzerland), using an universal
2. Materials and methods standard column (Metrosep C2-150, 4.0  150 mm) along with an
eluent composed of tartaric acid (4.0 mmol/L) and dipicolinic acid
2.1. Raw material (0.75 mmol/L), equipped with electronic detectors. In every case,
the cheese samples were previously homogenized at 4200 rpm in
18 cheeses were proportionated by FORMATGERIA GRANJA RI- an Ultraturrax T25 for 5 min and centrifuged (J.P. Selecta S.A.,
NYA S.L. factory (Valencia, Spain). All the cheeses were fresh Tron- Medifriger-BL, Barcelona, Spain) at 8000 rpm for 20 min at 4 °C.
chon type, they were conducted under the same pressing Afterwards, 1 mL of supernatant was diluted with MilliÒ-Q water
conditions and with identical compositional characteristics. The in a 250 mL Erlenmeyer flask. The clarified extract was filtered
cheeses were provided unsalted and just after the pressing opera- through a 0.45 lm Millipore filter; 15 mL was used to analyze the
tion. Samples were obtained from the center of each cheese. Brine anion content. Measurements were taken in duplicate.
was prepared with sodium chloride (PRS-codex, PanreacÒ Química Representative fresh samples and reposed samples were used to
SA, Barcelona, Spain) and distilled water. obtain images of the microstructure by Cryo-SEM.

2.2. Salting operation 2.2.1. Dielectric properties measurement


Dielectric properties were measured by using an Agilent
Experimental flowchart is shown in Fig. 1, where it is possible to 85070E Open-ended coaxial probe connected to an Agilent
observe the salting treatment conditions using 25% (w/w) sodium E8362B vector network analyzer. The software of the network ana-
chloride brine at 4 °C. 72 cheese cylinders (30 mm diameter and lyzer calculates the dielectric constant and loss factor as a reflected
10 mm thick) were cut. Samples were immersed in a vessel con- signal function. The system was calibrated by using three different
taining the osmotic solution with continuous stirring, for 0, 10, types of loads: open circuit (air), short-circuit and 25 °C distilled
20, 30, 40, 50, 60, 90, 120, 180, 240, 360, 480, 720, 900 and water. Once the calibration was made, 25 °C distilled water was
1440 min; later they were equilibrated in an isothermal chamber measured again to check calibration suitability.
at 4 °C for 24 h. The dielectric properties were measured by attaching the probe
Mass was determined by using a Mettler Toledo Balance to the surface of the samples at 25 °C from 500 MHz to 20 GHz. The
(±0.0001) (Mettler-Toledo, Inc., USA). The surface water activity mean values of at least three replicates of the cheese samples are
was determined by a dew point hygrometer Decagon (AqualabÒ reported in this article.
series 3TE) with precision ±0.003. Also the water activity of the
solution (brine) was measured at each osmotic time. Measurements 2.2.2. Low temperature scanning electron microscopy (Cryo-SEM)
were done in structured samples (not minced), thus aw obtained is A Cryostage CT-1500C unit (Oxford Instruments, Witney, UK),
coupled to a Jeol JSM-5410 scanning electron microscope (Jeol, To-
kyo, Japan), was used. The sample was immersed in slush N2
(210 °C) and then quickly transferred to the Cryostage at 1 kPa,
where sample fracture took place. The sublimation (etching) was
carried out at 95 °C; the final point was determined by direct
observation in the microscope, working at 5 kV. Then, once again
in the Cryostage unit, the sample was coated with gold in vacuum
(0.2 kPa), applied for 3 min, with an ionization current of 2 mA.
The observation in the scanning electron microscope was carried
out at 15 kV, at a working distance of 15 mm and a temperature
6130 °C.

3. Results and discussion

Cheese samples were analyzed after the salting treatment,


where the samples show concentration profiles, being non-equili-
brated. After 24 h of repose, samples reached the equilibrium and
the measurements were repeated.
Salting operation produces mass transfer phenomena (Castro-
Giráldez et al., 2010a), and the behaviors involved in that transfer
are complex. In order to determine the transport of the major
chemical compounds involved in this complex process, as a first
step to understand those phenomena, mass variation can be used:

Mt  M0
DM j ¼ ð1Þ
M0

M t  xti  M0  x0i
DMji ¼ ð2Þ
Fig. 1. Diagram of the experimental procedure.
M0
38 J. Velázquez-Varela et al. / Journal of Food Engineering 130 (2014) 36–44

where M represents the mass (kg), x is the mass fraction (kg/kg), the mechanical energies relevant, especially at short times in which
superscript t is the treatment time and 0 is the initial time, the there are marked fluctuations in the volume (non-significant).
superscript j could be 0 h or 24 h and represent the non-equili- Water and ions fluxes can be understood as compression and
brated samples or the equilibrated samples, respectively. relaxation phenomena. The different flows promoted in osmotic
Fig. 2a shows that the largest loss of mass occurs during the first treatment can be calculated by mass variation, the processing time
200 min of treatment. From that salting time, the loss of mass is and samples surface (Eqs. (4) and (5)).
less pronounced. This mass lost is explained mainly by the dehy-
dration suffered by the sample during treatment (Fig. 2b), since DM 0hw  Mo
J 0h
w ¼ ð4Þ
both figures show similar trends especially during the early stages Dt  S  Mrw
of treatment. The curves show differences that can be attributed to
the movement of other chemical species and other driving forces of DM 0h
i  Mo
J 0h
i ¼ ð5Þ
transport such as mechanical or electrical. Dt  S  Mri
Fig. 3a shows the mass variations of sodium and chloride ions where J represents the molar flux (mol s1 m2), Dt is the process-
during the treatment. Massive inflows of these ions from the solu- ing time (s), S is the surface measured during treatment (m2), Mr is
tion into the interior of the sample can be observed caused by high the molecular weight of water, chloride, sodium and calcium, sub-
activity gradients. This gain occurs mainly during the early stages scripts i and w represent ions and water, respectively.
of treatment. In Fig. 5(a–d), the molar fluxes of water and ions are shown;
In Fig. 3b, the mass variation of Ca+2 during the salting treat- these fluxes are promoted by changes of chemical potentials
ment is represented. This figure represents the release of calcium (Castro-Giráldez et al., 2010b).
from the protein structure of the cheese into its liquid phase during When cheese is placed in brine, there is a movement of sodium
the salting process. This release presents a maximum at 120 min of and chloride molecules from the solution into the cheese, and
treatment. The calcium released into the cheese liquid phase will simultaneously an outflow of water and calcium from the cheese
move to the osmotic solution, driven by high activity gradients, to the solution, promoted by the gradients of chemical potentials
because there is no calcium in the solution. Therefore, both mech- (Fig. 5a–d).
anisms (the release of calcium and the calcium transport) are In Fig. 6, it is represented a surface property (surface water
coupled. activity) with regard to an average measurement of the system
In cheese, there are multiple dynamic equilibriums between (moisture in dry basis). Both measurements were studied just after
caseins with different forms of calcium and phosphate in the mi- the treatment and after the repose time (24 h). It is possible to
celles and in the solution. The micelles can change considerably observe an increase in the surface water activity caused by the
due to changes in the environment that can drag little caseins internal fluxes that occur in cheese during the repose time. With
and ionic calcium into the solution. Some authors have shown that the equilibrated data, it is possible to estimate the surface moisture
the salt used in cheese manufacture is probably responsible of the in treated samples (non-equilibrated samples).
calcium ion release to the cheese liquid phase by a sodium-calcium Fig. 7 numbers 1 and 2 show the structure of fresh cheese; it is
exchange in protein structure (Guinee and Fox, 1986, 2004). possible to appreciate the protein matrix of caseins with cavities
The variation of volume during the salting treatment (Fig. 4) containing fat globules and liquid phase, the numbers 3 and 4
was also studied (Eq. (3)). show that the pores between caseins were becoming smaller with
Vt  V0 the salting treatment. With the salting, the water is going out and
DV j ¼ ð3Þ the caseins begin to aggregate. The protein network seems to
V0
become more continuous and compact; the ions chloride and
where V represents the volume (m3), the superscript t is the treat- sodium go inside the cheese, having a great impact in the protein
ment time and 0 is the initial time, the superscript j could be 0 h or matrix and in the fat globules. The addition of salt increases the
24 h and represent the non-equilibrated samples or the equilibrated ionic strength and decreases the pKa of proteins, modifying also
samples, respectively. the surface free energy (surface tension) of the fat globules (Lopez
In the figure, it is possible to observe that the loss of volume is et al., 2007). In the number 5, it is possible to appreciate fat glob-
not exactly coupled to the mass changes of the samples, being the ules which are inside the protein network already compacted with

Δ Δ

(a) (b)
Fig. 2. (a) Overall mass variation during the salting treatment (non-equilibrated samples). (b) Water mass variation during the salting treatment (non-equilibrated samples).
J. Velázquez-Varela et al. / Journal of Food Engineering 130 (2014) 36–44 39

0.03 ∇
MNa+ ∇
∇ 0.0035 MCa+2
MCl-

0.0030

0.02 0.0025

0.0020

0.0015 0.004
0.03 0.004
0.01 0.003
0.003
0.02 0.0010
0.002 0.002
0.01
0.001 0.001
0.0005
0.00 0.000
0 100 200 0.000
0 100 200
0.00 0.0000
0 200 400 600 800 1000 1200 1400 0 200 400 600 800 1000 1200 1400
t (min) t (min)
(a) (b)
Fig. 3. (a) Mass variation of sodium and chloride ions during the salting treatment (non-equilibrated samples). (b) Mass variation of calcium cation during the salting
treatment (non-equilibrated samples).

!
2 I1=2
log c ¼ jzj A  þbI ð7Þ
1 þ aB  I1=2

where A, B, a, b are empirical adjustable parameters which values


can be observed in Table 1 (Langmuir, 1997).
The activity of each chemical species can be obtained by using
the Eq. (8):

ai ¼ ci  ci ð8Þ

3.1. Nonlinear thermodynamic approach

A nonlinear thermodynamic model has been developed for


describing the cheese salting process (Fig. 8). In order to under-
stand and to estimate the molecular transport through the inter-
Fig. 4. Volume variation during the salting treatment (non-equilibrated samples). face, it is necessary to determine the free energy variation across
the interface per mol of any chemical specie in motion. Gibbs free
no pores. The amount of salt affects the amount of dispersed fat energy variation has been determined by Eq. (9) (Demirel,
(fat globules) contained in the cheese (Everett et al., 2004). Finally 2002a,b):
the number 6, shows the network of protein fully compacted in a X
continuous system. dG ¼ SdT þ VdP þ Fdl þ wde þ li dni ð9Þ
The behavior of ions in the food liquid phase is mainly influ- i

enced by two factors: the attraction between ions of opposite where SdT corresponds to the thermic term and it is directly related
charge and the repulsion between ions of the same sign (Coulomb’s to heat fluxes, VdP and Fdl are the mechanical energies related to
law), and the influence of thermal agitation (Boltzmann distribu- the structural changes; where VdP corresponds to pressure varia-
tion law) to counteract the electric attractions and repulsions. tion and Fdl to the elongation force, finally de represents the effect
These effects depend on the total ionic composition of the food li- of the electric field induced by solved ions. The term Rilidni is the
quid phase which can be expressed by its ionic strength, I addition of the chemical potentials of all the compounds in the sys-
(mol L1): tem considering P, T and the molar concentration of the rest of the
compounds, constant.
1X
I¼ ½ci  z2i  ð6Þ The system is isotherm, then the entropic term can be
2
neglected.
where c is ionic molar concentration (mol L1) and z is the valence, X
subindex i represents the electrolytic chemical specie.
dG ¼ VdP þ Fdl þ wde þ li dni ð10Þ
i
The activity coefficient, c (dimensionless), can be estimated
from the extended Debye-Hückel law for heterogeneous system Because salted cheese is an electrolytic system, it is necessary to
(Atkins and De Paula, 2006): establish the electric effect in order to calculate the free energy.
40 J. Velázquez-Varela et al. / Journal of Food Engineering 130 (2014) 36–44

(a) 5.E-02 Jw (mol/m2 s)


(c) 3.E-03
JCl - (mol/m2 s)
4.E-02
4.E-02 2.E-03

3.E-02
2.E-03
3.E-02
2.E-02
1.E-03
2.E-02
1.E-02 5.E-04
5.E-03
0.E+00 0.E+00
0 200 400 600 800 1000 1200 1400 0 200 400 600 800 1000 1200 1400
t (min) t (min)

(b) 3.E-03 (d) 1.E-04


JNa+ (mol/m2 s) JCa+2 (mol/m2 s)

3.E-03 1.E-04

2.E-03 8.E-05

2.E-03 6.E-05

1.E-03 4.E-05

5.E-04 2.E-05

0.E+00 0.E+00
0 200 400 600 800 1000 1200 1400 0 200 400 600 800 1000 1200 1400
t (min) t (min)

Fig. 5. (a) Water, (b) Sodium, (c) Chloride and (d) Calcium fluxes during the salting treatment (non-equilibrated samples).

2.5 that the addition of calcium to the osmotic solution, at the same
Xw (kg w/kg dm )
concentration of native ionic calcium, reduces the calcium losses
and decreases the water and sodium transport (Luo et al., 2012).
2.0
This occurs because it changes the calcium chemical potential and
reduces the effect of calcium in the electric field.
1.5 The variation of electric potential (W) can be calculated from
the electric field intensity (E) generated by the flow of ions to the
medium (Eq. (12)):
1.0
E ¼ rw ð12Þ
0.5 The electric field can be associated with the current charges inten-
sity and the ionic conductivity of the medium (Ohm’s law). Thus,
w the current intensity of an electrolytic solution for an isothermal
0.0
0.75 0.8 0.85 0.9 0.95 1 system can be defined as follows (Eq. (13)):

Fig. 6. Relationship between the moisture in dry basis and the surface water
I ¼rE ð13Þ
activity for treated samples () and reposed 24 h ( ). Where (- -) correspond to the 2 1
where I is the current intensity (C m s ), r is the electric conduc-
estimation of GAB model for the treatment samples and ( ) for equilibrated
samples. tivity (S m1) and E is the electric field generated by the charges
(V m1).
The current intensity for an electrolytic system can be described
Therefore, the charge variation of the electric term (de) can be writ- by the next equation (Eq. (14)):
ten as follows: X
X I¼ zi  F  J i ð14Þ
i¼1
de ¼ zi  F  dni ð11Þ
This equation applied to the cheese system can be written as
where zi is the valence of each electrolyte (with sign), F is the Far- follows:
aday constant (96485.3415 C mol1) and ni is the variation of each
ion. In general, when a transport of ionic species with the same va- I ¼ F  ð1  J Cl þ 1  J Na þ 2  J Ca Þ ð15Þ
lence is produced, the charges cancel the electric field generated Ionic conductivity can be obtained from the loss factor at 500 MHz,
and this term is 0. In the case of cheese salting, the release of cal- because at this frequency the ionic conductivity is the major contri-
cium ions causes the inclusion of new charges to the medium which bution to the loss factor (Castro-Giráldez et al., 2010b):
produce a disequilibrium of the system, causing an electric field
intensity associated to the ions movement. A recent study showed r ¼ e00  e0  - ¼ e00  e0  2  p  f ð16Þ
J. Velázquez-Varela et al. / Journal of Food Engineering 130 (2014) 36–44 41

Fig. 7. Cryo-SEM pictures of cheese samples at different treatment times; where 1 and 2 correspond to fresh cheese, at 1000 and 5000, respectively; 3 and 4 treated
samples at 20 min; at 5000; 5 and 6 treated samples at 1440 min, at 1000 and 5000 respectively.

Table 1 5
Extended Debye-Hückel parameters used in the calculations (Langmuir, 1997). E (V/m)
t (min)
Chloride Sodium Calcium
0
A 0.5083 a 3.5  10 8 8
4  10 8
5  10 0 200 400 600 800 1000 1200 1400
B 0.3287 b 0.015 0.075 0.165
-5

0
0 20 40 60 80 100
-10
-5

-10
-15
-15

-20 -20

Fig. 9. Evolution of electric field throughout the process time.

!
X X axj
dG ¼ VdP þ Fdl þ zj  F  E  dnj þ RT ln dnj ð17Þ
j j
axþdx
j

Applying the variation of free energy to the interface between the


osmotic solution and the cheese surface, the following equation is
obtained:
!
X X aext
j
Fig. 8. Scheme representing the interface between cheese and osmotic solution. DG ¼ V DP þ F Dl þ zj  F  E  Dni þ RT ln int Dnj ð18Þ
i j
aj

Being i the electrolytic species and j all the chemical species.


where e00 is the loss factor (dimensionless), e0 is the vacuum permit-
Considering that the mol variation can be obtained from the
tivity, 8.854  1012 (F m1); - angular velocity (rad s1), f fre-
fluxes, from the average surface area between the initial and the
quency (Hz).
final surfaces and from the time, it can be written as follows:
Therefore, with Eq. (13), it is possible to estimate the electric
field produced by the ions flow. Fig. 9 shows the evolution of elec- Dnj ¼ Jj  S  Dt ð19Þ
tric field throughout the process time. This electric field is pro-
duced by an imbalance between the anions and cations in liquid Applying Eq. (18) without the mechanical energy terms, it is
medium. The responsible of the imbalance in charge is calcium obtained a partial free energy term (Eq. (20)):
!
ion, liberated from the protein matrix. X X aext
~¼ j
Eq. (10), developed specifically per mole of chemical specie in DG zj  F  E  Dni þ RT ln int Dnj ð20Þ
i j
aj
motion, can be expressed as follows:
42 J. Velázquez-Varela et al. / Journal of Food Engineering 130 (2014) 36–44

t (min)
0
0 200 400 600 800 1000 1200 1400 1600

- 10

- 20

- 30

- 40

- 50

(J)
- 60

Fig. 10. The evolution of the partial free energy (without mechanical terms) during Fig. 12. Evolution of the free energy during the salting treatment.
the treatment.

The evolution of the partial free energy during the treatment can be 20 VΔ P+FΔl (J)
observed in Fig. 10. 10
In order to understand the evolution of partial free energy t (min)
0
(Fig. 10) it is necessary to comeback to Fig. 3 (ions mass variation), 0 200 400 600 800 1000 1200 1400 1600
where it is possible to observe a short stage of decrease (around -10
0 20 40 60 80 100
60 min) inside the general trend of increase. In Fig. 10, a minimum -20 0

peak of the partial free energy can be appreciated at 60 min, coin- -30 -10
ciding with the minimum peak mentioned before in Fig. 3. There- -20
fore, in the ions transport, the terms of activity are the main -40
-30
driving forces in the first minutes of the treatment. Thus, it is pos- -50
-40
sible to calculate the chemical potentials of the different chemical -60 -50
species at constant pressure (excluding mechanical terms) (see Eq.
-70 -60
(21)).
-70
-80
~
DG
l~ j ¼ ð21Þ Fig. 13. Evolution of the mechanical terms of the free energy during the salting
Dnj treatment.

Applying the first relation of Onsager (Demirel, 2002a,b; Gekas,


2001) (Eq. (22)), the fluxes can be related with the engine of trans-
port by a constant called phenomenological coefficient. Fig. 11 0,003

shows the relation between the water fluxes with the chemical Ji (mol/m 2s)
0,0025
potential (without the mechanical terms).
0,002
J w ¼ L w  Dl
~w ð22Þ
0,0015
In the figure, two tendencies can be observed: at the beginning of
the treatment, the mechanical terms cannot be neglected because
0,001
cheese undergoes a compression process in the operation prior to
0,0005

0 100 200 300 400 500 600 700 800 0


0.000 -50000 -40000 -30000 -20000 -10000 0
Δμw (J/mol)
-0.005 Δ μi

-0.010 Fig. 14. Variation of the sodium (j) and chloride (h) fluxes with regard to the
respective chemical potentials.
-0.015

-0.020
mold, which causes an accumulation of mechanical energy that will
-0.025 be released during the salting process. From 120 min to 1440 min,
mechanical terms can be negligible and, applying the Onsager rela-
-0.030 Treatment time
tion, the phenomenological coefficient of water during salting is
-0.035 1:83  105 mol2 J1 s1 m2 (R2 = 0.9925).
Applying the phenomenological coefficient obtained to the
-0.040
Jw (mol w /s m 2 ) whole salting treatment, it is possible to calculate the water chem-
-0.045 ical potential and the free energy of the process (Eqs. (23) and
(24)).
Fig. 11. Variation of the water chemical potential (without mechanical terms) with
regard to water flux, where () corresponds to the samples affected by mechanical Jw
terms of the chemical potential, ( ) corresponds to the samples with negligible Dlw ¼ ð23Þ
Lw
mechanical terms of the chemical potential.
J. Velázquez-Varela et al. / Journal of Food Engineering 130 (2014) 36–44 43

DG ¼ lw  Dnw ð24Þ interaction between the liquid compounds in motion with the solid
structure, is bigger than that of the chloride. On the other hand, the
Fig. 12 shows the evolution of the free energy during the salting
sodium cation flows against the electric field gradient and the chlo-
treatment.
ride anion is promoted by it. For this reason, the chemical potential
Mechanical terms can be obtained from Eq. (25).
of chloride is bigger than that of the sodium and it produces more
~
V DP þ F Dl ¼ DG  DG ð25Þ accumulation of chloride mass in liquid phase (Fig. 3) than the
sodium mass.
Fig. 13 shows the evolution of the mechanical terms during the salt-
ing treatment.
4. Conclusions
Fig. 13 shows strong variations of the mechanical terms pro-
moted by the mechanical energy stored during the prior operation
Nonlinear irreversible thermodynamic model has been devel-
of molded pressing. Therefore, it can be considered that the
oped to determine the transport of water and ions through the
mechanical energy stored is returned in changes of the internal
structure of cheese, obtaining a phenomenological coefficient of
absolute pressure. Taking into account that the external pressure
1.8  105 mol2/J s m2. This model describes various phenomena
is the atmospheric, and that the mechanical terms obtained are
that occur in salting cheese process, including the different driving
negative, this means that subatmospheric pressures occur inside
forces, electrical and mechanical activity terms. The importance of
the cheese, favouring the rapid entry of external ions into the inte-
the role of the calcium ion has been demonstrated, not only in the
rior of cheese.
structural changes that occur at protein level, but also as a trans-
From Eq. (26), the chemical potential of the different chemical
port inductor due to his role in the generation of the electric field.
species can be calculated.
It was also shown that cheese undergoes strong contractions and
DG expansions that are transmitted fast to the driving forces, leading
lj ¼ ð26Þ to strong variations in the mass fluxes, especially at the beginning
Dnj
of treatment. In addition, osmotic dehydration also causes loss of
Figs. 14 and 15 show the relation between the chemical potential native components of cheese, mainly calcium.
and the fluxes of the different species in motion during the salting
treatment. Applying the Onsager relation, the phenomenological
Acknowledgements
coefficients of the different species can be calculated (Table 2).
Table 2 shows that the ionic species show lower phenomeno-
The authors acknowledge the financial support from the
logical coefficients than water, it is explained by the high interac-
Spanish Ministerio de Ciencia e Innovación throughout the
tion of electrolytic compounds with the charges of casein net,
project AGL2011-30096. Author Marta Castro-Giráldez thanks
compared with water. The calcium ions show the lowest phenom-
the Campus de Excelencia Internacional VLC/CAMPUS for their
enological coefficient, because the calcium flux depends on the
support.
calcium release by exchanging sodium in its position in the casein
Author J. Velázquez-Varela thanks the Consejo Nacional de
network. In this sense, sodium cation has an interaction with the
Ciencia y Tecnología (CONACyT) of México for their support.
protein matrix displacing the calcium, and producing an accumula-
tion of sodium in the solid structure. For this reason the phenom-
enological coefficient of sodium (Table 2), which describes the References

Aguilera, J.M., Stanley, D.W., 1999. Food structuring. In: Aguilera, J.M., Stanley, D.W.
(Eds.), Microstructural Principles of Food Processing and Engineering. Aspen
0,00008 Publication, Gaithesburg, Maryland, pp. 190–230.
JCa2+ (mol/m 2s) Atkins, P.W., De Paula, J., 2006. Chemical equilibrium. In: Atkins, P.W., De Paula, J.
0,00007 (Eds.), Physical Chemistry. Oxford University Press, Great Britain, pp. 200–240.
Baroni, A.F., Menezes, M.R., Adell, E.A.A., Ribeiro, E.P., 2003. Modeling of prato
0,00006 cheese salting: fickian and neural netwok approaches. In: Welti-Chanes, J.,
Velez-Ruiz, J.F., Barbosa-Canovas, G.V. (Eds.), Transport Phenomena in Food
0,00005 Processing. CRC Press, Boca Raton, USA.
Castro-Giráldez, M., Fito, P.J., Fito, P., 2010a. Application of microwaves dielectric
0,00004 spectroscopy for controlling pork meat (Longissimus dorsi) salting process. J.
Food Eng. 97 (4), 484–490.
0,00003 Castro-Giráldez, M., Botella, P., Toldrá, F., Fito, P., 2010b. Low-frequency dielectric
spectrum to determine pork meat quality. Innovative Food Sci. Emerg. Technol.
0,00002 11 (2), 376–386.
Castro-Giráldez, M., Fito, P.J., Dalla Rosa, M., Fito, P., 2011. Application of
0,00001 microwaves dielectric spectroscopy for controlling osmotic dehydration of
kiwifruit (Actinidia deliciosa cv Hayward). Innovative Food Sci. Emerg. Technol.
0 12, 623–627.
-250000 -200000 -150000 -100000 -50000 0 Crank, J., 1985. Mathematics of Diffusions. Oxford University Press, New York.
Δ μ Ca2+ Demirel, Y., 2002a. Fundamentals of equilibrium thermodynamics. In: Demirel, Y.
(Ed.), Nonequilibrium Thermodynamics. Transport and Rate Processes in
Physical, Chemical and Biological Systems. Elsevier Science & Technology
Fig. 15. Variation of the calcium (j) flux with regard to the chemical potential. Books, USA, pp. 1–52 (Chapter 1).
Demirel, Y., 2002b. Thermodynamics and biological systems. In: Demirel, Y. (Ed.),
Nonequilibrium Thermodynamics. Transport and Rate Processes in Physical,
Chemical and Biological Systems. Elsevier Science & Technology Books, USA, pp.
Table 2
541–598 (Chapter 11).
Phenomenological coefficients, the initial chemical potential and squared correlation
Everett, D.W., Rowney, M.K., Hickey, M.W., Roupas, P., 2004. Salt-induced structural
coefficients for each ion involved in transport. changes in Mozzarella cheese and the impact upon free oil formation in
ripening cheese. Le Lait. Diary Sci. Technol. 84, 539–549.
Chloride Sodium Calcium
Fito, P., LeMaguer, M., Betoret, N., Fito, P.J., 2007. Advanced food process engineering
Li Jmol
2
4.3  108 5.7  108 3.4  1010 to model real foods and processes: the ‘‘SAFES’’ methodology. J. Food Eng. 83
s m2
  4 4 (2), 173–185.
l0i J 1.08  10 1.04  10 8.66  106
mol Fito, P., Le Maguer, M., Betoret, N., Fito, P.J., 2008. Advanced food products and
R2 0.9119 0.9055 0.9435 process engineering (SAFES) I: concepts and methodology. In: Gutiérrez-López,
G.F., Barbosa-Cánovas, G.V., Welti-Chanes, J., Parada-Arias, E. (Eds.), Food
44 J. Velázquez-Varela et al. / Journal of Food Engineering 130 (2014) 36–44

Engineering Integrated Approaches, Food Engineering Series. Springer, USA, pp. Lawrence, R.C., Gilles, J., 1982. Factors that determine the pH of young Cheddar
117–137. cheese. N. Z. J. Dairy Sci. Technol. 17, 1–14.
Floury, J., Rouaud, O., Le Poullennec, M., Famelart, M.H., 2009. Reducing salt level in Lopez, C., Camier, B., Gassi, J.Y., 2007. Development of the milk fat microstructure
food: Part 2. Modelling salt diffusion in model cheese systems with regards to during manufacture and ripening of Emmental cheese observed by confocal
their composition. LWT – Food Sci. Technol. 42, 1621–1628. laser scanning microscopy. Int. Dairy J. 17, 235–247.
Gekas, V., 2001. Mass transfer modeling. J. Food Eng. 49, 97–102. Luna, J.A., Bressan, J.A., 1986. Mass-transfer during brining of Cuartirolo Argentino
Guinee, T.P., Fox, P.F., 1986. Transport of sodium chloride and water in Romano- cheese. J. Food Sci. 51, 829–831.
type cheese slices during brining. Food Chem. 19, 49–64. Luo, J., Pan, T., Guo, Y., Ren, F.Z., 2012. Effect of calcium in brine on salt diffusion and
Guinee, T.P., Fox, P.F., 2004. Salt in cheese: physical, chemical and biological aspects. water distribution of Mozzarella cheese during brining. J. Dairy Sci. 96, 824–
In: Fox, Patrick F., McSweeney, Paul L.H., Cogan, Timothy M., Guinee, Timothy P. 831.
(Eds.), Cheese Chemistry, Physics and Microbiology. Elsevier, Ltd Ireland, pp. Turhan, M., Gunasekaran, S., 1999. Analysis of moisture transfer in White cheese
207–259. during brining. Milchwissenschaft 54, 446–450.
Hardy, J., 1976. Etude de la diffusion du sel dans les fromages à pâte molle de type Yanniotis, S., Anifantakis, E., 1983. Diffusion of salt in dry-salted Feta cheese. In:
camembert. Comparaison du salage à sec et du salage en saumure, Ph.D. Thesis, Jowitt, R., Escher, F., Hallstrom, B., Meffert, H.F.T., Spiess, W.E.L., Vos, G. (Eds.),
Université Nancy, France. Physical Properties of Foods. Applied Science Publishers, London, UK.
Langmuir, D., 1997. Activity coefficients of dissolved species. In: McConnin, Robert
(Ed.), Aqueous Environmental Geochemistry. Prentice-Hall Inc., New Yersey, pp.
123–148.

You might also like