Sei sulla pagina 1di 15

Journal of Food Engineering 56 (2002) 1–15

www.elsevier.com/locate/jfoodeng

Behavior of plant tissue in osmotic solutions


a,*
Maria Aparecida Mauro , Debora de Queiroz Tavares b, Florencia Cecilia Menegalli c

a
Department of Food Engineering and Technology, Institute of Biosciences, Language, and Physical Sciences, Paulista State University (UNESP),
CEP 15054-000, S~ ao Jose do Rio Preto, S~
ao Paulo, Brazil
b
Department of Nutrition, Faculty of Food Engineering, State University of Campinas (UNICAMP), FEA, Caixa Postal 6121,
CEP 13081-970, Campinas, S~ ao Paulo, Brazil
c
Department of Food Engineering, Faculty of Food Engineering, State University of Campinas (UNICAMP), FEA, Caixa Postal 6121,
CEP 13081-970, Campinas, S~ ao Paulo, Brazil
Received 30 December 1999; received in revised form 11 February 2002

Abstract
The effect of the concentration of sucrose solutions on the cellular structure of potato tissue in equilibrium at 27 °C was studied.
Two different methods of investigation were used to determine the volume of the different phases composing the cellular tissue of the
potato when in equilibrium with the solutions, one based on data of the concentration itself and the overall volume of 2 mm slices
after 48 h at equilibrium, and the other on microscopic images of cells in thin slices of fresh tissue stained with neutral red after an
hour in equilibrium to show protoplasts, vacuoles and plasmolysis spaces. The results of these methods were compared with those
obtained by a predictive thermodynamic approach considering the semipermeability of cell membranes. Phase volume data obtained
from microscopic analysis were more similar to what was predicted by the theoretical model than those obtained by means of
composition measurement, where the long equilibrium time apparently led to the loss of semipermeability of the cell membranes,
since total volumes calculated without consideration of the cell membranes were similar to those measured. This suggests that the
length of time of osmotic dehydration brings about a change in cell structure and the consequent involvement of a different
mechanism in mass transfer.
Ó 2002 Elsevier Science Ltd. All rights reserved.

Keywords: Osmotic dehydration; Cellular structure; Plasmolysis; Potato; Sucrose

1. Introduction more easily, as has been reported for previously steam-


blanched banana slices (Mauro & Menegalli, 1995).
Food can be dehydrated osmotically by immersion in Temperature is one of the factors which is known to be
a concentrated solution containing one or more solutes. involved in the rupture of the integrity of plant tissues
When used for whole plant tissues, such dehydration is and membranes, for instance plasmatic membranes be-
characterized by the large-scale transfer of certain gin to suffer irreversible damage at 55 °C (Thebud &
compounds through the cell membrane, especially those Santarius, 1982). Time of exposure also seems to affect
of the solvent (such as water), whereas that of others, the behavior of such membranes as it happens with
especially the solute, is limited. This behavior is due to onions dehydrated in osmotic saline solutions: a slight
the differential permeability of these membranes (Bi- increase in moisture content was detected after an initial
dwell, 1974). This osmotic phenomenon is largely con- drop in a 2 h process (Baroni & Hubinger, 1997). It
trolled by the plasmalemma, which is the membrane seems probable that as time passes, the cell membranes
surrounding the protoplast (Nobel, 1991). When prior no longer provide an effective barrier for the solute,
treatment destroys the cellular structure, the tissue in- which is then free to penetrate into all parts of the cell.
volved looses its selectivity and modifies the osmotic The food industry exploits various processes of im-
process (Ponting, 1973). The solute can then penetrate mersion or drying at low temperatures which involve the
transfer of solutes into integral tissues. Dehydration by
*
Corresponding author. Tel.: 55-17-221-2200; fax: +55-17-224-
immersion in an osmotic solution has aroused special
8692. interest, since it can increase the quality of the final
E-mail address: cidam@eta.ibilce.unesp.br (M.A. Mauro). product when used in combination with other processes
0260-8774/02/$ - see front matter Ó 2002 Elsevier Science Ltd. All rights reserved.
PII: S 0 2 6 0 - 8 7 7 4 ( 0 2 ) 0 0 1 0 7 - 3
2 M.A. Mauro et al. / Journal of Food Engineering 56 (2002) 1–15

Nomenclature

A area, m2 x molar fraction


aw water activity X total moisture content on dry weight basis
B constant q density of untreated dry potatoes, kg/m3
C constant qpot density of potatoes in equilibrium with os-
csuc volumetric concentration of sucrose in os- motic solution, kg/m3
motic solution, kg/m3 N total moisture content divided by initial dry
CV coefficient of variance mass of potatoes
h thickness of slice, m Wm matrix potential, kg/ms2
0
K constant DM=M variation in mass in relation to total initial
M0 total mass of untreated potatoes, kg mass of potatoes
m0 total solids (mass) in untreated potatoes, kg lw chemical potential of water in system, kJ/
N number of observations kmol
P hydrostatic pressure, N/m2
Subscripts
P atm atmospheric pressure, N/m2
OS osmotic solution
r radius, m
st starch
R universal gas constant, kJ/kmol K
pr protein
R2 correlation coefficient
cel cellulose
S ratio of percentage of sucrose to total potato
v vacuole
mass
v cytoplasm
SG percentage of solid gain
j free space
Sos mass of sucrose expressed as percentage of
proto protoplasm
total mass of osmotic solution
cw cell wall
t time, s
T absolute temperature, K Superscripts
TS ratio of percentage of total solids to total d dry
potato mass 0 refers to reference state, either untreated po-
V volume, m3 tato material or pure water at the given
V partial molar volume, m3 /kmol temperature and atmospheric pressure
w mass fraction (dry weight basis)  refers to the pure component
WL water loss (percentage of water lost in rela-
tion to total initial mass)

such as convection drying, vacuum drying or freezing. the transfer of water during conventional drying of
Many studies have been conducted on osmotic drying of plants. Marcotte, Toupin, and Le Maguer (1991),
plant food products during the past three decades. Some Toupin, Marcotte, and Le Maguer (1989) and Yao and
of these have utilized models based on empirical transfer Le Maguer (1996, 1997a,b) also considered it in the
coefficients (Hawkes & Flink, 1978; Magee, Hassabal- mathematical models proposed for describing transport
lah, & Murphy, 1983), whereas others utilize more ex- during the osmotic dehydration of plant tissue.
acting analytical solutions employing the Fick equation A more detailed study of the volume occupied by the
(Beristain, Azuara, Cortes, & Garcia, 1990; Conway, various phases of cell tissue was conducted by Marcotte
Castaigne, Picard, & Vovan, 1983). However, none of and Le Maguer (1991). They proposed a model for the
these models include special suppositions about the presence of water in the different phases based on the
cellular structure of the tissue. semipermeability of the plasmatic membranes when
In general, transport mechanisms of the intact cell are working with potato tissue in equilibrium with various
still poorly understood, which makes the projection of aqueous sucrose solutions (5–60%). Using typical data
the osmotic dehydration process even more difficult. for the chemical composition of potatoes and equations
Studies and models considering the selectivity of cell relating equilibrium water content to the chemical po-
membranes on the transport of solvents and solutes in tential of water in each phase, they were able to predict
plant tissues have been few in number. Rotstein and the effect of different osmotic solutions on the volume
Cornish (1978) and Crapiste, Whitaker, and Rotstein occupied by each cell phase, including the intercellular
(1988a,b) considered the permeability of membranes in space between the protoplast and the cell wall formed
M.A. Mauro et al. / Journal of Food Engineering 56 (2002) 1–15 3

during cell plasmolysis. A comparison of the variation amino acids can be present in certain species (Nobel,
between theoretical volumes and those calculated on the 1991).
basis of experimental data showed that this model pro-
vided a reasonable fit. 1.2. Equilibrium of water sorption
The objective of the present paper was to study the
volume of the different plant cell phases when in equi- The prediction of equilibrium content of water in
librium with osmotic solutions, identifying these phases plant tissues can be made based on the equality of the
and quantifying their water content after relatively long chemical potential of the water in the internal phases of
periods of equilibrium, as well as investigating micro- the cell with that of the external osmotic solutions. The
scopic images of protoplasts and vacuoles of cells after estimation of the water contents in each phase requires a
only brief periods of equilibrium with a solution. The model which, in this case, is based on a rather simplified
results obtained by measurement were compared with description of cells. In the potato cell, the phases con-
those predicted using equilibrium data based on ther- sidered are the cell wall, the free space, the cytoplasm,
modynamic considerations of the equality of the chem- and the vacuole. The free space is that portion of the cell
ical potential of water in all phases. This revealed composed of the intercellular space, the interstices or
variations in tissue behavior as a function of time of pores in the cell walls, and the plasmolysis space, which
equilibrium. forms between the cell wall and the plasmalemma. When
the osmotic agent penetrates the cell wall, yet fails to
penetrate the plasmalemma.
1.1. Cell structure The chemical potential of water in a plant system in
relation to the standard state can be expressed by the
The polyhedral parenchyma cells of the potato have following equation (Crapiste & Rotstein, 1982):
approximately equal diameters on the different planes;
lw  l0w ¼ V w ðP  P atm Þ þ RT ln aw þ V w Wm ð1Þ
moreover, the intercellular spaces of potato tubercules
are limited, occupying only from 0.2–1% of the volume where lw represents the water potential of the system,
of the potato tissue. Tests forcing air, water and aqueous and l0w represents the reference state of pure water at the
solutions through potato tissue showed that these spaces same temperature and at atmospheric pressure; the first
are connected by long narrow passages (Woolley, 1962). term on the left-hand side of the equation represents the
The cellulosic plant cell wall lends firmness to the pressure potential within the cell, with (P  P atm ) equal
tissue, but is not the main barrier to the transfer of to hydrostatic pressure in excess of atmospheric pressure
substances into and out the cell because it contains nu- (Nobel, 1991) and V w the partial molar volume of the
merous relatively large interstices which make it per- water; the second term represents the water potential in
meable to water and small solute particles (Nobel, 1991). the cell due to dissolved solutes (osmotic potential), with
Carpita, Sabularse, Montezinos, and Delmer (1979) aw being the activity of the water in the system; the third
have estimated that the average diameter of the pores in term, Wm , is the so-called matric potential (Nobel, 1991)
plant cell walls is about 3.5 nm (35 A ), whereas sucrose due to the strong interaction between water and the
has an average diameter estimated at only 1 nm. large, irregular, porous surfaces of the solids.
All the other cell contents are found inside the cell The osmotic solution can be described by the fol-
wall. These contents as a whole are generally known as lowing equation:
the protoplast, which is separated by the plasmalemma
ðlw  l0w ÞOS ¼ RT ln awOS ð2Þ
(plasma membrane) from the cell wall.
The other major component of the protoplast is the where awOS represents the activity of the water in the
cytoplasm, which consists of the cytosol, a solution osmotic solution.
containing various organelles such as the chloroplasts, Under equilibrium conditions, these two equations
mitochondria, peroxisomes, and ribosomes, as well as are equal. Since each term can be made explicit as a
proteins and other macromolecules and structures function of the concentration of water, these equations
which influence the thermodynamic properties of water thus serve as a basis for the calculation of the water
(Nobel, 1991). content in each phase of the system.
In the mature cells of higher plants and many lower Due to the rigidity of the cell walls, fairly large hy-
(less advanced) ones, there is a large central space inside drostatic pressures can be built up inside plant cells
the protoplast, known as the vacuole. It is filled with when the water content is high. The differential perme-
water, and is surrounded by a membrane, the tonoplast. ability of the plasmalemma permits solvents to pass
This vacuole can occupy up to 90% of the volume of a easily through it, while solutes are limited to restricted
mature plant cell. Its aqueous solution contains princi- passage. Under normal circumstances, the internal
pally inorganic ions or organic acids as solutes, despite pressure of the cell pushes the plasmalemma firmly
the fact that relatively large quantities of sugars and against the elastic cell wall, causing the cell to expand
4 M.A. Mauro et al. / Journal of Food Engineering 56 (2002) 1–15

and leading to stress in the cell wall. Plant tissue in (William, 1970). Insoluble potato fibers were determined
equilibrium with pure water exhibits maximum turgid- by the enzymatic method (Asp, Johansson, Hallmer, &
ity. As the external solute concentration increases, hy- Siljestr€
om, 1983). For protein content, nitrogen content
drostatic pressure drops, and the entire cell shrinks until, was determined by the Kjeldahl method (Williams,
at some point, it looses its turgidity and P ¼ P atm . This is 1984). Ash was determined gravimetrically (Williams,
the point of incipient plasmolysis (Nobel, 1991). From 1984).
here on, an increase in the external solute concentration The density of osmotically dried potato slices was
leads to the formation of a space, filled with plasmo- determined by means of the equilibrium solution itself
lyzing solution, between the protoplast and the cell wall. using the technique of volume dislocation. After re-
moval from the solution and elimination of the excess
solution, 23.2 mm diameter plugs were cut from the slice
2. Materials and methods with a cork borer. Mass and volume were determined
and, since the diameter was known, it was possible to
2.1. Determination of equilibrium data calculate the thickness. The ratio between the thickness
at equilibrium and the initial thickness gave the
Post harvested potatoes (Bintje) were purchased on shrinkage coefficient for this specific dimension of the
the local market and stored at 5 °C for a maximum of slice. The remainder of the slice was then used to analyze
one week while the tests were being conducted. total solids and total and reducing sugars. However, in
Slices with an average thickness of 1.96 mm were cut order to determine the shrinkage coefficient more pre-
using specially-designed equipment and left to reach cisely, confirmatory tests were made with a 41.0 mm
equilibrium in various sucrose solutions (5%, 10%, 15%, cork borer. In addition, the ratio between the volume at
20%, 27%, 35%, 43%, and 50%) at 27 °C; 0.20–0.24% of equilibrium and the initial volume was determined by
potassium sorbate was added to avoid deterioration. independent tests, applying the liquid displacement
The tests for the various concentrations were conducted method to the whole slice.
in duplicate. Each potato slice was placed in a 500-ml
Erlenmeyer flask, with stopper, containing approxi- 2.3. Microscopic analysis
mately 550 g of solution (the ratio of solution to tu-
bercule was approximately 50:1). The Erlenmeyer flask Histological techniques were used to photograph
was immersed in a heated water bath, and samples were potato cells after osmotic treatment. Preliminary tests
collected after 24, 48 and 60 h (although little difference with various tissue fixation techniques and stains were
was observed between those left for 48 and 60 h, which conducted. Vital stains, i.e. stains with no short-term
suggests that after 48 h equilibrium had already been effect on cell physiology were selected: neutral red and
reached). After this period of time, the slices were re- acridine orange. The apparatus used to obtain the im-
moved and blotted dry. The samples were then used to ages with the neutral red was a standard light micro-
determine the chemical composition of the potato (total scope, whereas the images with acridine orange were
and reducing sugars and total solids), as well as the obtained with a fluorescent light microscope with wave-
density and the shrinkage coefficient. length filters from ultraviolet to red.

2.2. Analytic methods 2.3.1. Neutral red


Neutral red is a vital stain of relatively low molecular
The average composition of fibers, ash, starch and weight (288.783) which penetrates the vacuole of intact
proteins in dry, powdered potatoes was determined, protoplasts of plant cells. Its molecules have no electric
using quadruplicate measurements. Total solids for each charge, but the low pH inside the vacuoles transforms
of the fresh potato samples to be used in each of the them to an ionic state incapable of penetrating the
different equilibrium tests were also determined, using tonoplast, where the stain remains confined; neutral red
the average of three measurements. The content of total thus accumulates in the vacuole (Thebud & Santarius,
and reducing sugars was determined from the average of 1982), and is used to provide contrast to the protoplast.
ten different samples. It has been used to observe the plasmolysis of plant cells
Reducing sugars were determined with the colori- in various solutions (Carpita et al., 1979).
metric method of Somogyi–Nelson (Nelson, 1944; For staining with neutral red, potato juice was ob-
Somogyi, 1945), while total sugars were determined by tained by blending the raw potato in a centrifuge and
oxyreduction titration (William, 1970). Total solids were subjecting to it thermal inactivation and filtering. The
determined gravimetrically by the drying of the samples stain was then added to this juice. The stain used had a
in a vacuum oven at 60 °C until a constant weight was final concentration in the juice calculated to be 0.05%.
obtained. The starch was determined by acid hydrolysis The ideal staining time for good definition in photo-
(William, 1980), followed by oxyreduction titration graphs was found to be from 10–20 min. Slices were cut
M.A. Mauro et al. / Journal of Food Engineering 56 (2002) 1–15 5

manually and, after staining, were immersed in aqueous 3.1. Determination of phase volumes from equilibrium
sucrose solutions (5, 10, 20, 30, 40, and 50%) for a tests
minimum of 60 min. Slides were then made from pieces
of the potato slices accompanied by a drop of the su- The results of the measurements are given in Table 1.
crose solution and covered with the slide cover; these These include the percentage of total solids in relation to
slides were viewed immediately and photographed. The the total mass of the potato (TS), the percentage of
control slices were washed in potato juice for from 10– sucrose infused into the potato in relation to this total
30 min after dying and used to prepare slides with their mass (S), and the density of the potato (qpot ) as a
own juice. Photographs were made using colored 35 mm function of the concentration of the osmotic solution
film (ASA 100) and a blue filter. (Sos ) expressed as a percentage of the mass of sucrose in
relation to the total mass of the solution.

2.3.2. Acridine orange


Acridine orange is a basic fluorescent dye with an Table 1
absorption of 497 nm and a molecular weight of Average percentage of total solids (TS) and sucrose (S) in relation to
301.822. At a level of 0.01% and a pH between 3 and 8, total potato mass at equilibrium on a given sucrose concentration (Sos ).
acridine orange is a vital stain in plant tissues; vacuoles Average density of potato tissue (qpot ) in equilibrium with osmotic
solution
appear in shades of green or red (ConnÕs, 1990), while
the nucleus and cytoplasm appear orange (Munck, Sos (%, TS (%, CVa S (%, CVa qpot (g/ CVa
w/w) w/w) (%) w/w) (%) cm3 ) (%)
1989). The slices of Bintje potatoes in the present ex-
periment were stained for 20 min in this dye in a solution 0 15.33 0.94 0.12 71.38 1.0476 1.17
of 0.005% with potato juice (mass/volume); the same 5.2 21.78 8.81 4.02 0.25 1.0496 1.53
10.2 25.42 1.84 7.56 1.89 1.0673 1.63
procedure was used as with neutral red, with immersion 15.0 29.56 4.55 8.99 1.50 1.0880 1.18
in osmotic solutions for a minimum of 60 min. The 20.1 33.84 3.26 12.83 2.30 1.1082 1.52
cytosol and all of its inclusions fluoresced at a wave 26.9 40.56 2.15 19.41 0.36 1.1380 0.63
length of 480–500 nm. 34.8 46.27 0.66 26.00 3.10 1.1736 1.02
42.7 52.56 1.31 31.87 0.91 1.2006 0.91
49.7 58.49 1.07 37.26 1.02 1.2496 1.40
a
CV is the coefficient of variance resulting from the ratio between
2.3.3. Image measurement the standard deviation and the average measurement of the data. For
Retraction of the vacuole and protoplast were mea- determination of the standard deviation was used N  1 to represent
sured by a comparison of their areas in relation to the the N samples.
total cell area in the images obtained using neutral red.
These areas were drawn on parchment paper of homo-
geneous density, cut out, and weighed. Three or more
measurements of the images of the vacuoles and pro-
toplasts of each osmotic solution were made.

3. Models and results

Two different methods were used to evaluate the


volume of the different phases of the potato cell. One
based on experimental data of the concentration of
water and sucrose and the determination of overall
volumes as a function of the equilibrium concentration,
and the other based on photographs of cells submitted
to various osmotic treatments in which it was possible
to observe the retraction of the protoplasts and vacuoles
and the establishment of a plasmolysis space between
the plasmalemma and the cell wall. Phase volumes
determined by these two methods were compared to
those resulting from theoretical calculations based
Fig. 1. Unidimensional and volumetric shrinkage coefficients: com-
on the thermodynamic equilibrium of water which parison between experimental values and those calculated according to
considered the semipermeability of the membranes in- Eqs. (3) and (4); comparison between thickness shrinkage (Eq. (3)) and
volved. radius shrinkage (Eq. (5)).
6 M.A. Mauro et al. / Journal of Food Engineering 56 (2002) 1–15

Independent confirmatory tests were conducted to sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi


estimate the shrinkage of the thickness of the slices as a 0 ðV =V 0 Þ
ðr=r Þ ¼ ð5Þ
function of the concentration of the osmotic solution. ðh=h0 Þ
The ratio between the thickness of a slice in equilibrium
(h) and its initial thickness prior to osmotic treatment The radius shrinkage coefficient (r=r0 ) is plotted in
(h0 ) was considered to be an unidimensional coefficient Fig. 1. On comparing the (r=r0 ) and (h=h0 ), it can be
of shrinkage (h=h0 ) (Fig. 1). concluded that there is a remarkable selective shrinkage
The one-dimensional coefficient of shrinkage was in thickness over the area of the potato slice.
correlated with the percentage of sucrose in the osmotic Fig. 1 also shows the efficiency of the adjustments by
solution (Sos ) using a third-order polynomial, and the comparing the experimental results with those calcu-
non-linear model was fitted to the experimental data lated according to Eqs. (3) and (4). It can be seen that up
according to the steepest-descent method (Marquard, to concentrations of 10%, the reduction in volume is
1959). accentuated, but after this the loss diminishes, probably
due to the initiation of plasmolysis. However, there are
ðh=h0 Þ ¼ 6:73  106 ðSos Þ3 þ 6:07  104 ðSos Þ2 strong indications that degradation of the tissue has
occurred; hence, the polynomial would represent the
 1:83  102 ðSos Þ þ 9:92  101 ð3Þ volume as a function of the density of the cellulosic
materials of the cell wall in equilibrium with the osmotic
To evaluate the adjustment, the residual root mean solution without the influence of turgidity.
square (RMSR) was calculated (Daniel & Wood, 1980); The variation in mass (DM) in relation to total initial
this resulted in a value of 5.4%. mass of the potato (M 0 ) can be calculated as follows:
Also independent tests were conducted to estimate DM ðTS0  TSÞ þ ðS  S 0 Þ
the shrinkage of the whole material as a function of the ¼ ð6Þ
M0 ðTS  SÞ
concentration of the osmotic solution. The ratio be-
tween the volume of a slice in equilibrium (V) and its where TS represents the total solids percentage, TS0 the
initial volume prior to osmotic treatment (V 0 ) was initial percentage (15.33%), S the sucrose percentage and
considered to be a volumetric coefficient of shrinkage S 0 the initial percentage (0.12%).
(V =V 0 ) (Fig. 1). This volumetric coefficient was corre- The water loss (WL) (as a percentage of the total
lated to the percentage of sucrose in osmotic solution initial mass) was calculated as follows:
(Sos ) using a third-order polynomial:   
DM  
WL ¼ ð100  TSÞ 0
þ 1  100  TS0 ð7Þ
3
ðV =V 0 Þ ¼ 1:11  105 ðSos Þ þ 1:03  103 ðSos Þ
2 M
The percentage of sugar gain (SG) was defined as the
 3:11  102 ðSos Þ þ 9:99  101 ð4Þ ratio between the sucrose penetrating the potato tissue
and the total initial mass. This would be calculated as
The RMRS resulted in a value of 1.17%. follows:
The third-order polynomial provided an approxima-  
tion of the deformation behavior of the potato tissue, as DM
SG ¼ S þ 1  S0 ð8Þ
described by Marcotte and Le Maguer (1991). These M0
authors observed that once the stage of incipient
plasmolysis was reached, the total volume remained Table 2 presents the variation in final mass as a per-
approximately stable at larger concentrations until this centage of initial mass (DM=M 0 ), water loss (WL) and
reached a concentration corresponding to the critical
cell volume. This critical volume is the point after which Table 2
greater concentrations lead to a collapse of the cell Variation in mass in relation to total initial mass of potatoes (DM=M 0 ),
structure and a significant reduction in total volume. water loss (WL) and sugar gain (SG), all calculated from experimental
data as a function of osmotic concentration (Sos )
Since potato is a relatively isotropic material, due to
polyhedrical cells with relatively similar diameters in Sos DM=M 0  100 (%) WL (%) SG (%)
different planes, one would expect to find similar 0
shrinkage in all dimensions. For slices, however, a se- 5.2 14.83 18.13 3.30
10.2 14.85 21.17 6.32
lective shrinkage in one dimension over others was ob-
15.0 26.07 32.59 6.52
served. Also another one-dimensional coefficient was 20.1 27.60 36.77 9.17
calculated to describe the shrinking of the radius of the 26.9 28.10 41.94 13.84
slice, i.e., the ratio between the radius of a slice in 34.8 24.96 44.35 19.39
equilibrium and its initial radius prior to osmotic 42.7 26.50 49.80 23.30
49.7 28.34 54.92 26.58
treatment (r=r0 ).
M.A. Mauro et al. / Journal of Food Engineering 56 (2002) 1–15 7

sugar gain (SG), all calculated from experimental data


as a function of osmotic concentration (Sos ).
Cell membranes are only slightly permeable to su-
crose, but quite permeable to water; the cell wall, on the
other hand, is equally permeable to both water and su-
crose. Given this fact, for the calculation of free space
(which encompasses the interstitial spaces of the cell
wall, intercellular spaces, and plasmolysis spaces) it was
assumed that all free space is occupied by sucrose from
the osmotic solution so that, at equilibrium, the con-
centration in this space is the same as that of sucrose in
the osmotic solution (Marcotte & Le Maguer, 1991).
For the calculation of the volume of the free space oc-
cupied by the water–sucrose solution, the volumetric
concentration of sucrose in the osmotic solution (csuc ) is
considered to be as follows:
qos Sos
csuc ¼ ð9Þ
100
where qos is the density of the osmotic solution, which is Fig. 2. Comparison of results: (a) ratio of overall volume to initial
supposedly the same as that found in the free space. The volume (V =V 0 ) according to Eq. (4) (b) ratio of protoplast volume to
initial volume (Vproto =V 0 ) calculated on the basis of thermodynamic
density of aqueous sucrose solutions as a function of equilibrium of water (c) (Vproto =V 0 ) measured from microscopic images
temperature is taken from the literature (AOAC, 1984: and (d) (Vproto =V 0 ) predicted from experimental data measured over
52,008 and 52,020) for the corresponding concentra- long periods of time.
tions.
The volume occupied by aqueous sucrose solution in
the free spaces (Vj ), as a percentage of the overall vol-
ume of the untreated tissue (V 0 ) will thus be the fol-
lowing:
S
DM

Vj 100 M0
þ 1 q0pot
¼ ð10Þ
V0 csuc
where q0pot represents the density of the untreated po-
tato.
The volume occupied by the protoplasts (Vproto ),
where sucrose was unable to penetrate, is determined by
calculating the difference between the total volume of
the tissue (Eq. (4)) and the volume occupied by free
space (Eq. (10)):
Vproto V Vj
0
¼ 0 0 ð11Þ
V V V
Fig. 2 shows the volume of the protoplasts as a per-
centage of the initial volume (Vproto =V 0 ), and Fig. 3 that
of the free space (Vj =V 0 ), as a function of the different
osmotic concentrations. Fig. 3. Comparison of results: (a) ratio of overall volume to initial
Post harvested potatoes variety Bintje presented high volume (V =V 0 ) according to Eq. (4) (b) ratio of free space volume to
moisture content (15.33% TS, Table 1). Consequently, initial volume (Vj =V 0 ) calculated on the basis of thermodynamic
small concentrations of sucrose in osmotic solution equilibrium of water (c) (Vj =V 0 ) measured from microscopic images
caused large dehydration. It can be seen in Table 2 that and (d) (Vj =V 0 ) predicted from experimental data measured over long
periods of time.
the water loss for potatoes in equilibrium with 5.2% and
10.2% osmotic solution resulted in about 20% of water
loss. On the other hand, the solid gained was consider-
ably high for small concentrations of sucrose. It suggests branes, thus making the protoplast space also available
a substantial loss of the permeability of cellular mem- to sucrose entrance.
8 M.A. Mauro et al. / Journal of Food Engineering 56 (2002) 1–15

Table 3 It was assumed that the cell wall is constituted prin-


Chemical composition of Bintje potatoes cipally of cellulose, in a quantity equal to the fiber
Constituent Composition (kg/100 kg of dry matter) content, as determined experimentally. For the sorption
Starch 77.12 equilibrium between cellulose and water, Papadakis,
Proteins 10.5 Bahu, Mckenzie, and Kemp (1993) adjusted their data
Fiber 3.52 reasonably well using the correlation of Henderson
Sucrose 0.79 (1952):
Reducing sugars 0.26
Ash 2.83
  0:573
1 1
Others 4.98 Xcel ¼ 2:48 ln ð15Þ
T 1  aw
where Xcel is the water content retained by the dry mass
3.2. Estimation of phase volume on the basis of thermo- of the pure cellulose, aw is the relative humidity of the air
dynamic equilibrium or its water activity, and T is the absolute temperature
(K).
With the data about the initial composition of the Thus, the water potential of the cellulose can be made
potato and the curves of sorption for the different explicit as follows:
components, it was possible to calculate the water con-  
tent and the volume of each of the phases, considering lw  l0w cel ¼ RT ln aw ð16Þ
thermodynamic equilibrium between the phases and the On the other hand, since sucrose is not able to penetrate
composition of each of them. To do this, the chemical into the protoplast, this can be assumed to maintain the
potentials of the phases were equaled. This method sorption characteristics of the untreated potato. Thus,
made it possible to estimate the average volume of the the curve of sorption of the potato represents the re-
vacuoles, which was not possible from the experimental tention of water by the protoplast and the cell wall, ig-
data collected. noring that retained by the solution in the free space.
The average chemical composition of the Bintje po- Crapiste (1985) found adsorption isotherms for water
tatoes studied here is found in Table 3. The major retention of Huinkul potatoes at 20, 40 and 60 °C when
constituents of the total solids of the potato cell are no previous treatment was used. He adjusted his data to
starch and protein. These elements are present in the diverse models found in the literature and obtained the
cytoplasm, forming a colloidal solution with a very low best adjustment by the correlation of Halsey (1948) for
concentration of soluble solids. Both starch (st) and elevated relative humidity:
protein (pr) can be considered to be phases, each with a " #
matrix retaining a certain quantity of water; in equilib- B 1
rium, however, each has the same chemical potential for aw ¼ exp ð17Þ
RT ðX ÞC
water as the osmotic solution (Eq. (1)). This is also true
for the cellulosic material (cel) which constitutes the cell where X is the water content retained by the dry mass of
wall. Thus: the potato, R is the universal constant of gases, T is the
      absolute temperature (K) and B and C are constants.
RT ln awos ¼ lw  l0w st ¼ lw  l0w pr ¼ lw  l0w cel When an adsorption isotherm of a plant is deter-
ð12Þ mined, part of the cell structure is modified due to the
fact that the material is first dried; samples will, for
For starch, the curve of sorption at 25 °C given by example, be highly unlikely to present turgidity. The
Crapiste and Rotstein (1982) was used: pressure term of Eq. (1) was thus ignored in the calcu-
  lations.
lw  l0w st ¼ RT lnð1  expð53:4759Xst2:3015 ÞÞ ð13Þ Based on the water activity for each concentration in
the osmotic solution, it is possible to determine the
where Xst is the water content retained by the dry mass
water retained by the protoplast and the cell wall (N)
of the pure starch.
using Eq. (17) as well as the water content of each of the
For protein, the data of Bull (1944) were adjusted to
phases using Eqs. (12)–(16). The water activity of the
the highest level of aw (0.6–0.95), according to the cor-
osmotic solution was determined using the equation of
relation proposed by Crapiste and Rotstein (1982),
Norrish (1966):
which gives the following:
  awOS ¼ xw expðKx2s Þ ð18Þ
lw  l0w pr ¼ RT ð6:5932  103 Xpr2:10189 Þ ð14Þ
where xw and xs are the molar fractions of the water and
where Xpr is the water content retained by the dry mass sucrose in the osmotic solution.
of the pure protein; the correlation coefficient for the The constant utilized in Eq. (18) ðK ¼ 6:47Þ was de-
adjustment was R2 ¼ 0:99986. termined by Chirife, Fontan, and Benmergui (1980).
M.A. Mauro et al. / Journal of Food Engineering 56 (2002) 1–15 9

Table 4
Water activity (aw ), fraction of water retained by the protoplast plus cell wall, by the starch, by the protein, by the cellulosic matter, and by the
vacuole in relation to the dry mass of the untreated potato as a function of osmotic concentration (Sos )
Sos (%, w/w) aw N Nst Npr Ncel Nv
a
0 0.9987 5.5239 0.3111 0.2242 0.0098 4.9788
5.2 0.9971 3.2403 0.2944 0.1541 0.0091 2.7827
10.2 0.9938 2.0526 0.2775 0.1082 0.0084 1.6584
15.0 0.9902 1.5459 0.2663 0.0869 0.0080 1.1846
20.1 0.9858 1.2275 0.2568 0.0727 0.0076 0.8904
26.9 0.9787 0.9547 0.2458 0.0599 0.0072 0.6418
34.8 0.9680 0.7402 0.2342 0.0492 0.0068 0.4501
42.7 0.9533 0.5837 0.2226 0.0409 0.0063 0.3139
49.7 0.9356 0.4764 0.2121 0.0350 0.0059 0.2234
a
This value corresponds to the water activity of untreated potatoes, calculated on the basis of N0 .

The values of awOS as a function of the concentration of Vprotoþcw N 1


the osmotic solution (Sos ) are found in Table 4. ¼  þ d ð24Þ
m0 qw qpot
Under initial conditions, i.e., prior to treatment, the
water content retained by the initial dry mass of potato where qw is the density of pure water at 27 °C, qdpot is the
(N) is determined by means of the ratio of water content average density of the dry potato (calculated from the
to total solid content of untreated potatoes: composition given in Table 3 and the density of each
component), and m0 represents the mass of the initial
1  TS0 total solids of the untreated potato, which includes
N0 ¼ ð19Þ
TS0 protoplast and cell wall, considered to be constant
throughout the process of osmotic dehydration:
The retention of water mass by the vacuole in relation to
initial dry mass of the potato is determined by the fol- m0
¼ TS0 q0pot ð25Þ
lowing: V0
Nv ¼ N  Nst  Npr  Ncel ð20Þ The volume of the cytoplasm (Vv ) and the cell wall (Vcw )
are calculated as follows:
where Nst , Npr and Ncel are the retention of water mass by   " #
the starch, protein and cellulosic matter respectively, in Vv Nst w0st Npr w0pr
¼ þ þ þ ð26Þ
relation to the initial dry mass of potato, i.e.: m0 qw qst qw qpr
 
Nst ¼ Xst w0st ð21Þ Vcw Ncel w0cel
¼ þ ð27Þ
m0 qw qcel
Npr ¼ Xpr w0pr ð22Þ
where qst is the density of pure starch, which varies ac-
cording to moisture content (Nara, 1979) and qpr is the
Ncel ¼ Xcel w0cel ð23Þ density of pure protein (Kuntz & Kauzmann, 1974).
given that w0st is the initial starch content of the potato The difference between Eq. (24) and Eqs. (24)–(26)
(dry basis); w0pr is the initial protein content (dry basis); makes it possible to calculate the volume of the vacuole
and w0cel is the initial content of cellulosic matter (dry (Vv ):
basis), all considered to be constants (data in Table 3). Vv Vprotoþcw Vv Vcw
¼  0 0 ð28Þ
Table 4 presents the water activity of untreated po- m0 m0 m m
tatoes and that of the osmotic solution, as well as the Subtracting (Vprotoþcw =m0 ) from the total volume of the
quantity of water retained by the protoplast plus the cell cell (V =m0 ), with the latter calculated on the basis of
wall (N), by the starch (Nst ), by the protein (Npr ), by the experimental data (Eqs. (4) and (25)), thus leads to the
cellulosic matter (Ncel ) and by the vacuole (Nv ) in rela- obtention of the volume of the free space (Vj ), i.e.:
tion to the initial dry mass of the potato.
Based on the data obtained in Table 4, the volume of Vj V Vprotoþcw
¼  ð29Þ
the cell phases were calculated, i.e., the volume of the m0 m0 m0
protoplast, the cytoplasm, the cell wall, the vacuole, and Multiplying the results of Eqs. (26), (28), (29) by m0 =V 0
the free space. The volume of the protoplast and the cell (Eq. (25)) results in Vv =V 0 for the cytoplasm, Vv =V 0 for
wall (Vprotoþcw ) can be determined as the sum of the the vacuole, and Vj =V 0 for the free space. (Vproto =V 0 ) can
volume occupied by water plus the volume occupied by then be calculated by means of the difference between
the dry solids composing the potato: the values obtained from Eq. (24) and those from Eq.
10 M.A. Mauro et al. / Journal of Food Engineering 56 (2002) 1–15

3.3. Determination of phase volumes from microscopic


images of tissue

Images from the light microscope made it possible to


measure the volume of the vacuoles, protoplasts, and
free spaces of the cells in equilibrium with aqueous su-
crose solutions. Images of cells in equilibrium in various
osmotic concentrations were obtained using neutral red.
Fig. 6a shows the control (i.e. stained cells with no os-
motic treatment). These cells have large starch granules
(S) throughout because the vacuoles keep the cytoplas-
matic material around the edges. Various small lipid and
protein inclusions (I), impregnated with the stain were
also observed, as well as the cellulosic cell walls (CW)
confining the cells. Fig. 6b was taken of samples im-
mersed in 10% sucrose solution, while the sample in Fig.
6c was immersed in a 20% solution. Plasmolysis can be
seen in the photographs of the two experimental samples
("#), as this phenomenon leads to the separation of the
Fig. 4. Comparison of results: (a) ratio of overall volume to initial plasmalemma from the cell wall and the consequent
volume (V =V 0 ) according to Eq. (4) (b) ratio of vacuole volume to formation of a translucent space, filled with the aqueous
initial volume (Vv =V 0 ) calculated on the basis of thermodynamic equi- sucrose solution, between the wall and the protoplast.
librium of water and (c) (Vv =V 0 ) measured from microscopic images. The red area at the lower part of the cell (Fig. 6b and
c) corresponds to the cell vacuole (V); it is smaller in
relation to cell area in equilibrium with the more con-
centrated sucrose solution, as this greater concentration
caused the retraction of the vacuole through loss of
water and a consequent increase in free space.
Fig. 6d shows an accentuated area of plasmolysis ("#)
for cells in 40% sucrose solution. The dissolving of the
stain in the vacuole and its impregnation of the other
structural elements of the cytosol helps to define the
limits of the protoplasts (Fig. 6d) and enhances the
visibility of the vacuoles (Fig. 6e). Fig. 6f shows a
plasmolysed cell in 50% sucrose solution where the cy-
toplasmatic material covers partially the vacuole.
Acridine orange was used to confirm the presence of
plasmolysis space. This stain impregnates cytosolic ele-
ments and darkens the vacuoles (Fig. 7b), instead of
leaving them relatively evident, as occurred with neutral
red (Fig. 6b–e); Fig. 7a shows the control. The impreg-
nation of the cytosol elements with the acridine orange
leaves clear plasmolysis spaces (green) where the cytosol
contents are no longer present.
Retraction was measured as a ratio of protoplast area
Fig. 5. Comparison of results: (a) ratio of overall volume to initial to total cell area (Aproto =A), as well as that of the vacuole
volume (V =V 0 ) according to Eq. (4) (b) ratio of cytoplasm volume to area to the total cell area (Av =A), and these ratios are
initial volume (Vv =V 0 ) calculated on the basis of thermodynamic presented in Table 5 for the various osmotic concen-
equilibrium of water and (c) (Vv =V 0 ) measured from microscopic im- trations (Sos ). With 5% osmotic solutions no plasmolysis
ages.
was observed.
Assuming that the protoplast and vacuole follow the
shape of the cell, the areas were transformed into vol-
(27), multiplied by the results of Eq. (25). These ratios ume as follows:
represent the volumetric fractions of the phases in  3=2
Vproto Aproto
equilibrium for the various concentrations of osmotic ¼ ð30Þ
solution and are illustrated in Figs. 2–5. V A
M.A. Mauro et al. / Journal of Food Engineering 56 (2002) 1–15 11

Fig. 6. Slides of potato tissue with neutral-red-stained cytosol after immersion in osmotic sucrose solutions for 60 min: (a) cells prior to osmotic
treatment with no plasmolysis present (S, starch; I, lipid and protein inclusions; CW, cell wall), (b) plasmolysed cells ("#) in 10% sucrose solution (c)
plasmolysed cells ("#) in 20% sucrose solution, with vacuole (V) visible due to dissolving of the stain in its interior (d) plasmolysed cells ("#) in 40%
sucrose solution, with great retraction of protoplasts. Small-diameter inclusions are visible throughout the vacuole (V). (e) 50% sucrose solution,
showing a great impregnation of neutral red in vacuole (V) and (f) plasmolysed cell ("#) in 50% sucrose solution. The cytoplasmatic material covers
partially the vacuole.

Fig. 7. Slides of potato tissue stained with acridine orange prior to immersion in sucrose solutions: (a) cells prior to osmotic treatment with no
plasmolysis present and (b) plasmolysed cells ("#) in 43% sucrose solution showing extensive impregnation of acridine orange in the cytosol elements.
12 M.A. Mauro et al. / Journal of Food Engineering 56 (2002) 1–15

Table 5 sured, especially for the vacuoles. However, for the


Measured areas based on microscope images of protoplasts (Aproto =A) vacuoles, the comparison of the average of such mea-
and vacuoles (Av =A) for cells in equilibrium with different concentra-
tions of osmotic solution
surements with those calculated on the basis of ther-
modynamic relationships proves the results to be
Sos (%, w/w) ðAproto =AÞ CVa (%) ðAv =AÞ 100 CVa (%)
similar.
100 (%) (%)
The protoplast volume calculated on the basis of
10 70.75 10.85 49.17 17.84 thermodynamic equilibrium of water and that measured
20 66.18 14.83 33.02 24.00
30 59.54 2.40 22.65 32.45
using microscopic images (measured as a fraction of the
40 52.35 10.12 16.00 12.93 initial cell) show a good agreement only for 10 and 50%
50 47.89 13.11 13.97 18.11 sucrose solutions (Fig. 2); on the other hand, the vol-
a
CV is the coefficient of variance resulting from the ratio between umes determined experimentally using data based on
the standard deviation and the average measurement of the data. For longer exposure to an osmotic solution, although they
determination of the standard deviation was used N  1 to represent show a reasonable agreement in concentrations of 15%
the N samples.
or more, presented important discrepancies when os-
motic concentrations of 5 and 10% were involved.
Comparing the protoplast volume with that of the total
and cell shows that as the osmotic concentration increases,
 3=2 there is a corresponding relative reduction in total vol-
Vv Av
¼ ð31Þ ume. The free space, which represents the difference
V A between the total volume and that of the protoplast,
where A represents the total measured area of the cell at increases with greater concentrations of osmotic solu-
a given osmotic concentration, Aproto the protoplast area, tion, occupying more than half the total cell volume at
Av the vacuole area, V the total volume of the potato the highest concentrations. It is clear that the compari-
tissue at the given osmotic concentration, Vproto the son between protoplast volume and overall volume does
volume occupied only by protoplasts, and Vv the volume not provide a complete picture of cell volume prior to
occupied by the vacuoles. Since intercellular spaces in plasmolysis (concentrations below 10%). As long as the
these images are minimal, they were ignored. potato cell suffers no plasmolysis (in osmotic concen-
The difference between the total volume (V) and the trations up to 5%), the free space is extremely limited
protoplast volume (Vproto ) is approximately equivalent to and is reduced to the intercellular spaces and the pores
the volume occupied by free space (Vj ): in the cell wall, but once the permeability of the mem-
branes is lost and there is a penetration of sucrose into
Vj Vproto
¼1 ð32Þ the interior of the cell, this is no longer true. Use of a
V V 10% solution, for example, led to some plasmolysis.
To calculate the cytoplasm volume (Vv ), the volume of A similar comparison could not be made for vacu-
vacuoles was subtracted from that of the protoplasts: oles, however. Fig. 4 shows a comparison of only the
Vv Vproto Vv vacuole volume calculated on the basis of the thermo-
¼  0 ð33Þ dynamic equilibrium of water and that measured on the
V0 V0 V
basis of microscopic images, since the prediction of the
These ratios were determined by Eqs. (30)–(33) and were volume of this phase was impossible from experimental
multiplied by the corresponding V =V 0 (Eq. (4)) in order data taken over a longer period of time in the osmotic
to compare them with those determined from the theo- solution. The two methods of determining vacuole vol-
retical–predictive model (Figs. 2–5). ume produced good agreement, although the images
present a slightly lower volume. This figure shows the
great reduction in volume suffered by the vacuoles in
4. Discussion relation to the total cell volume.
The cytoplasm volume calculated on the basis of
The results of volumes obtained by the two experi- thermodynamic equilibrium was compared with that
mental methods are compared to those of data calcu- measured from microscope images (Fig. 5), with the two
lated on the basis of the theoretical–predictive model. data showing a low agreement. The fact that the cyto-
Remarkable discrepancies can be detected between one plasm volume is calculated on the basis of the image
of the methods and the theoretical model (Figs. 2 and 3). measurements of the protoplasts and vacuoles (Eq. (33))
These discrepancies have given rise to the hypothesis of and that vacuoles and whole cells have more regular
the degradation of membranes as a consequence of the shape than protoplasts of potato tissue (Fig. 6b–f) sug-
time of exposure. gests that the irregular shape of the protoplasts increase
An analysis of the data in Table 5 shows that the the errors in the measurements of these images. Since the
coefficients of variance are high when images are mea- cytoplasm volume remains approximately constant in
M.A. Mauro et al. / Journal of Food Engineering 56 (2002) 1–15 13

relation to initial volume, it seems to be little affected by


the water content. Thus, the phase most affected by the
volumetric reduction of the protoplast is the vacuole.
Analyzing the differences observed between the vol-
umes determined in different ways in Fig. 3, the same
behavior can be seen for all concentrations, i.e., this
figure shows that the free space calculated using exper-
imental data over time always occupies more space than
would be expected from calculation on the basis of
thermodynamic equilibrium of water. Such long-term
tests show an excessive presence of sucrose in the solids
in equilibrium with the various solutions. This could be
the result of the occurrence of degradation of cell
membranes as a function of the long periods of expo-
sition in the osmotic solutions, with the consequent
entrance of sucrose into other cell compartments.
Although this explanation seems reasonable, it is also
possible that the presence of the preservative agent po-
tassium sorbate might have affected the permeability of
the membranes, since the mechanism of action of this Fig. 8. Comparison of cell volume: (a) (V =V 0 ) determined by the ex-
compound is as yet unexplained (Dergal, 1981). To in- perimental data (Eq. (4)) and (b) (V =V 0 ) calculated on the basis of Eq.
vestigate this possibility, microscopic tests were made (34).
without the presence of the agent. These tests did not
show any obvious change in relation to permeability of Probably these results reflect an increase in experimental
the membranes to sucrose, since the vacuoles and pro- errors due to the sum of the terms in Eq. (34), all ob-
toplasts were perfectly visible for the various osmotic tained from experimental composition data.
solutions, although the contrast was greater with the In an attempt to determine how long potato tissue
presence of sorbate. maintains its integrity when exposed to osmotic solu-
If the membranes had lost their permeability, one tions, various consecutive microscopic observations
would expect to find basically starch, protein, water and were made during a period of 6 h; the tissue used was
sucrose inside the cell in equilibrium. Other solutes stained with neutral red and immersed in sucrose solu-
originally present in the cytoplasm and the vacuole tions of 5–50% concentration. After 2 h of exposure,
would leave the cells through the cell walls, which are evidences of degradation were found for samples im-
sufficiently porous for them to pass. Therefore one can mersed in concentrations of 30% or more, as indicated
calculate the total cell volume as approximately that of a by the weakening of the color of the vacuoles and pro-
homogeneous solution of sucrose, starch and protein, toplasts. On the other hand, 6 h were insufficient for
considering that the volumes are additive, i.e.: extensive effects on the semipermeability of the mem-
! branes, even those exposed to concentrations of 20% or

V 1  TS DM w0st w0pr less of sucrose, to take place. These results suggest that a
¼ 1 þ þ þ more detailed study would be useful to determine the
m0 TS0 qw M0 qst qpr
!  rate of degradation of tissue over time.
S DM
þ 1 þ ð34Þ
TS0 qsuc M0
5. Conclusions
where the first term on the right represents the volume of
water, the second the volume of the starch, the third that Experimental application of the theoretical model
of the protein, and the fourth the volume of the sucrose, (Eq. (12)), which considers that the semipermeability of
with qsuc being the density of pure sucrose. When the membranes, predicts approximately the volume of the
results of Eq. (34) are multiplied by those of Eq. (25) various cell phases from data about tissue composition,
(initial volume), one obtains the ratio of the total vol- the water activity of each osmotic concentration, and
ume to the initial volume (V =V 0 ), which can be com- the overall tissue volume as a function of the concen-
pared to the same ratio (V =V 0 ) determined by the tration of sucrose, as long as exposure to the osmotic
experimental data (Eq. (4)). Fig. 8 shows that these re- solution is limited in time.
sults agree reasonably well, but they are systematically Images of thin slices of fresh potato tissue and in
lower than those determined by the measurement of equilibrium in osmotic solutions for short periods of
volumes (about 4–9%), except for 10% sucrose solution. time presented better agreement with the model based
14 M.A. Mauro et al. / Journal of Food Engineering 56 (2002) 1–15

on thermodynamic equilibrium than tests involving Dergal, S. B., (1981) Quımica de los Alimentos. Mexico: Editorial
longer periods of equilibrium. Alhambra.
Halsey, G. (1948). Physical adsorption on non-uniform surfaces.
Apparently, extended exposure to osmotic solutions Journal of Chemical Physics, 16, 931–937.
in equilibrium led to the degradation of cellular struc- Hawkes, J., & Flink, J. M. (1978). Osmotic concentration of fruit slices
tures. As the osmotic dehydration process progress, prior to freeze dehydration. Journal of Food Processing and
changes occur in the tissues, thus altering the mecha- Preservation, 2, 265–284.
nisms involved in the transfer of matter within the cell; Henderson, S. M. (1952). A basic concept of equilibrium moisture.
Agricultural Engineering, 33, 29–31.
this alteration should consequently be considered in Kuntz, I. D., Jr., & Kauzmann, W. (1974). Hydration of proteins and
models of the phenomenon. The validity of the hy- polypeptides. Advances in Protein Chemistry, 38, 239–345.
pothesis of semipermeability of membranes during these Magee, T. R. A., Hassaballah, A. A., & Murphy, W. R. (1983).
processes depends on the time for which tissue is ex- Internal mass transfer during osmotic dehydration of apple slices in
posed to osmotic solution. sugar solutions. Irish Journal of Food Science and Technology, 7,
147–155.
Marcotte, M., & Le Maguer, M. (1991). Repartition of water in plant
tissues subjected to osmotic processes. Journal of Food Process
Acknowledgements
Engineering, 13, 297–320.
Marcotte, M., Toupin, C. J., & Le Maguer, M. (1991). Mass transfer in
The authors thank the UNESP Research Interna- cellular tissues. Part I: The mathematical model. Journal of Food
tionalization Program (PROINTER/PROPP) and Engineering, 13, 199–220.
CAPES-PICD Program for financial support. Marquard, D. W. (1959). Solution of nonlinear chemical engineering
models. Chemical Engineering Progress, 55(6), 65–70.
Mauro, M. A., & Menegalli, F. C. (1995). Evaluation of diffusion
coefficients in osmotic concentration of bananas (Musa cavendish
References Lambert). International Journal of Food Science and Technology,
30, 199–213.
Asp, N.-G., Johansson, C. G., Hallmer, H., & Siljestr€ om, M. (1983). Munck, L. (Ed.). (1989). Fluorescence analysis in food. Harlow:
Rapid enzymatic assay of insoluble and soluble dietary fiber. Longman Scientific & Technical.
Journal of Agricultural and Food Chemistry, 31, 476–482. Nara, S. (1979). On the relationship between specific volume and
Baroni, A. F., & Hubinger, M. D. (1997). Cinetica da desidratacß~ao de crystallinity of starch. Starch/St€arke, 31(3), 73–75.
cebola. Anais do XXIV Congresso Brasileiro de Sistemas Particu- Nelson, N. (1944). A photometric adaptation of the Somogyi method
lados (XXIV ENEMP) (vol. I, pp. 375–380). Brazil: Universidade for the determination of glucose. The Journal of Biological
Federal de Uberl^ andia. Chemistry, 153, 375–380.
Beristain, C. I., Azuara, E., Cortes, R., & Garcia, H. S. (1990). Mass Nobel, P. S. (1991). Physicochemical and environmental plant physiol-
transfer during osmotic dehydration of pineapple rings. Interna- ogy. San Diego: Academic Press Inc.
tional Journal of Food Science and Technology, 25, 576–582. Norrish, R. S. (1966). An equation for the activity coefficients and
Bidwell, R. G. S. (1974). Plant Physiology. New York: Macmillan equilibrium relative humidities of water in confectionery syrups.
Publishing Co. Inc. Journal of Food Technology, 1, 25–39.
Bull, H. B. (1944). Adsorption of water vapor by proteins. Journal of Papadakis, S. E., Bahu, R. E., Mckenzie, K. A., & Kemp, I. C. (1993).
the American Chemical Society, 66, 1499–1507. Correlations for the equilibrium moisture content of solids. Drying
Carpita, N., Sabularse, D., Montezinos, D., & Delmer, D. P. (1979). Technology, 11(3), 543–553.
Determination of the pore size of cell walls of living plant cells. Ponting, J. D. (1973). Osmotic dehydration of fruits––Recent modi-
Science, 205, 1144–1147. fications and applications. Process Biochemistry, 8, 18–20.
Chirife, J., Fontan, C. F., & Benmergui, E. A. (1980). The prediction Rotstein, E., & Cornish, A. R. H. (1978). Influence of cellular
of water activity in aqueous solutions in connection with interme- membrane permeability on drying behavior. Journal of Food
diate moisture foods. IV. aw prediction in aqueous non electrolyte Science, 43, 926–934.
solutions. Journal of Food Technology, 15, 59–70. Somogyi, M. (1945). A new reagent for the determination of sugars.
ConnÕs, H. J. (1990). Biological stains. St. Louis: Sigma Chemical The Journal of Biological Chemistry, 160, 61–68.
Company. Thebud, R., & Santarius, K. A. (1982). Effects of high-temperature
Conway, J., Castaigne, F., Picard, G., & Vovan, X. (1983). Mass stress on various biomembranes of leaf cells in situ and in vitro.
transfer considerations in the osmotic dehydration of apples. Plant Physiology, 70, 200–205.
Canadian Institute of Food Science and Technology Journal, 16, 25– Toupin, C. J., Marcotte, M., & Le Maguer, M. (1989). Osmotically-
29. induced mass transfer in plant storage tissues: A mathematical
Crapiste, G. H., & Rotstein, E. (1982). Prediction of sorptional model. Part I. Journal of Food Engineering, 10, 13–38.
equilibrium data for starch-containing foodstuffs. Journal of Food William, H. (1970). Official methods of analysis (11th ed.). Washington
Science, 47, 1501–1507. DC: Association of Official Analytical Chemists, Inc.
Crapiste, G. H., (1985). Fundamentals of drying of foodstuffs. PhD William, H. (1980). Official methods of analysis (13th ed.). Washington
thesis, Universidad Nacional del Sur, Argentina. DC: Association of Official Analytical Chemists, Inc.
Crapiste, G. H., Whitaker, S., & Rotstein, E. (1988a). Drying of Williams, S. (1984). Official methods of analysis (14th ed.). Arlington,
cellular material-I. A mass transfer theory. Chemical Engineering Virginia: Association of Official Analytical Chemists, Inc.
Science, 43(11), 2919–2928. Woolley, J. T. (1962). Potato tuber tissue respiration & ventilation.
Crapiste, G. H., Whitaker, S., & Rotstein, E. (1988b). Drying of Plant Physiology, 37(6), 793–798.
cellular material-II. Experimental and numerical results. Chemical Yao, Z., & Le Maguer, M. (1996). Mathematical modelling and
Engineering Science, 43(11), 2929–2936. simulation of mass transfer in osmotic dehydration processes. Part
Daniel, C., & Wood, F. S. (1980). Fitting equations to data. New York: I: Conceptual and mathematical models. Journal of Food Engi-
John Wiley & Sons. neering, 29, 349–360.
M.A. Mauro et al. / Journal of Food Engineering 56 (2002) 1–15 15

Yao, Z., & Le Maguer, M. (1997a). Mathematical modelling and Yao, Z., & Le Maguer, M. (1997b). Mathematical modelling and
simulation of mass transfer in osmotic dehydration processes. Part simulation of mass transfer in osmotic dehydration processes.
II: Simulation and model verification. Journal of Food Engineering, Part III: Parametric study. Journal of Food Engineering, 32,
32, 21–32. 33–46.

Potrebbero piacerti anche