shrinkage density porosity shape changes drying pumpkin

Upload: liliana-seremet

Post on 24-Feb-2018

216 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/24/2019 Shrinkage Density Porosity Shape Changes Drying Pumpkin

    1/9

    Shrinkage, density, porosity and shape changes during dehydration

    of pumpkin (Cucurbita pepo L.) fruits

    L. Mayor a,, R. Moreira b, A.M. Sereno c

    a Instituto Universitario de Ingeniera de Alimentos para el Desarrollo, Universidad Politcnica de Valencia, Camino de Vera, s/n, 46022 Valencia, Spainb Departamento de Enxeara Qumica, Universidade de Santiago de Compostela, Ra Lope Gmez de Marzoa s/n, E-15782 Santiago de Compostela, Spainc REQUIMTE, Department of Chemical Engineering, Faculty of Engineering, University of Porto, Rua Dr. Roberto Frias s/n, 4200-465 Porto, Portugal

    a r t i c l e i n f o

    Article history:

    Received 2 March 2010

    Received in revised form 27 August 2010

    Accepted 29 August 2010

    Available online 8 October 2010

    Keywords:

    Air drying

    Bulk density

    Modelling

    Osmotic dehydration

    Particle density

    Shape factors

    a b s t r a c t

    The aim of this work was to study the changes in volume, density, porosity and shape factors of pumpkin

    tissue during osmotic dehydration (OD) and air drying (AD). Pumpkin cylinders with length/diameter

    ratio of 5/3 were used. OD experiments were carried out with solutions of sucrose, sodium chloride

    and mixtures of both solutes at different temperatures. AD experiments were conducted at 70 C. Volume

    of samples decreased linearly with weight reduction (WR). Bulk density varied in a restricted range (5

    13%) during dehydration and for all the methods maximum values were found. Particle density increased

    during both processes. Porosity increased at advanced degrees of dehydration, showing a minimum value

    at the beginning of OD and AD. The proposed models to evaluate shrinkage, bulk and particle densities

    and porosity from WR were satisfactorily applied. Image analysis showed that shrinkage of samples dur-

    ing OD was isotropic. Pumpkin cylinders increased elongation and decreased roundness and compactness

    during osmotic dehydration.

    2010 Elsevier Ltd. All rights reserved.

    1. Introduction

    The knowledge of physicochemical properties of food materials

    is important for an adequate design of food operations as well as

    for the control and improvement of the quality of the final product

    (Rahman, 2005).

    Particularly in dehydration processes, the heat and mass trans-

    fer flows can modify physicochemical properties of the material

    such as chemical composition (McLaughlin and Magee 1998),

    mechanical properties (Lewicki and Lukaszuk, 2000) and volume

    and porosity. The quality of the dehydrated product depends on

    the extension of these changes. Regarding to the changes in vol-

    ume and porosity, high shrinkage and low porosity lead to prod-

    ucts with poor rehydration capability (McMinn and Magee,

    1997). Furthermore, the changes in volume and dimensions must

    be considered for mass transfer modelling during dehydration

    (Simal et al., 1998; Khalloufi et al., 2009). In fact, experimental

    shrinkage data during air drying (AD) of foods is relatively abun-

    dant (Mayor and Sereno, 2004).

    Knowledge of the bulk density of food materials is an important

    parameter in storage, transport, mixing and packaging operations

    (Rahman, 2005). Heat and mass transfer in solids depend on den-

    sity and porosity values (Rahman, 2001). Textural and sensorial

    properties of foods are also related to density and porosity

    (Rahman, 2001). Porosity is related to the chemical stability of

    dried products; degradation of sugars (White and Bell, 1999) and

    lipid oxidation (Shimada et al., 1991) depended on this property.

    Food shape is one of the main quality attributes perceived by

    the consumer (Fernandez et al., 2005). Drying not only causes vol-

    ume changes but also may cause changes in shape. In this sense,

    product deformation is not fully described by the evaluation of

    volumetric shrinkage (Panyawong and Devahastin, 2007).

    Osmotic dehydration (OD) is a non-thermal processthat consists

    in the immersion of a food material in a hypertonic solution. The

    difference of the chemical potential between the material and the

    solution promotes two main fluxes: the outcome of water from

    the material to the osmotic solution, and the income of soluble sol-

    ids from the osmotic solutionto the material. As osmotic agents are

    often used sugars (sucrose or glucose) and salts (sodium chloride).

    Volume changes during OD are mainly due to compositional

    changes and mechanical stresses associated to mass fluxes. These

    changes have been analyzed as variations in the volumes of solid,

    liquidand gas phases of the food material during the process (Barat

    et al., 2001), and have been correlated with changes in moisture

    content and WR (Moreira and Sereno, 2003), or with WL (Nieto

    et al., 2004). These three aforementioned works studied shrinkage

    phenomena during OD of apples. Volumetric shrinkage during OD

    of other food products has also been reported, as the case of

    strawberries (Viberg et al., 1998), mangos (Giraldo et al., 2003)

    and tomatoes (Souza et al., 2007; Bui et al., 2009). In spite of these

    0260-8774/$ - see front matter 2010 Elsevier Ltd. All rights reserved.doi:10.1016/j.jfoodeng.2010.08.031

    Corresponding author.

    E-mail address: [email protected](L. Mayor).

    Journal of Food Engineering 103 (2011) 2937

    Contents lists available at ScienceDirect

    Journal of Food Engineering

    j o u r n a l h o m e p a g e : w w w . e l s e v i e r . c o m / l o c a t e / j f o o d e n g

    http://dx.doi.org/10.1016/j.jfoodeng.2010.08.031mailto:[email protected]://dx.doi.org/10.1016/j.jfoodeng.2010.08.031http://www.sciencedirect.com/science/journal/02608774http://www.elsevier.com/locate/jfoodenghttp://www.elsevier.com/locate/jfoodenghttp://www.sciencedirect.com/science/journal/02608774http://dx.doi.org/10.1016/j.jfoodeng.2010.08.031mailto:[email protected]://dx.doi.org/10.1016/j.jfoodeng.2010.08.031
  • 7/24/2019 Shrinkage Density Porosity Shape Changes Drying Pumpkin

    2/9

    works, few experimental data on shrinkage during OD of fruits and

    vegetables are found in the literature, and the variety of food mate-

    rials studied is restricted.

    Pumpkin (Cucurbita pepo L.) is a seasonal crop, which has been

    used for human and animal feed. OD of pumpkin, as a single or

    combined process, can be an alternative way to develop new food

    products (Kowalska et al., 2008; Konopacka et al., 2010). OD kinet-

    ics of pumpkin were already determined under several experimen-

    tal conditions by using binary solutions of sucrose and glucose

    (Kowalska et al., 2008) or sodium chloride (Mayor et al., 2006)and ternary solutions of sucrose and sodium chloride (Mayor

    et al., 2007).

    The aim of this work was to present experimental data on

    changes in volume, bulk and particle densities and porosity during

    OD of pumpkin fruits. The obtained results were compared with

    the changes observed during air drying (the most conventional

    method of dehydration) of this food material. Models to estimate

    these physical properties during dehydration were proposed.

    Dimensional and shape changes were also studied by means of im-

    age analysis.

    2. Materials and methods

    2.1. Sample preparation

    Pumpkin fruits ( C. pepo L.) were purchased from a local pro-

    ducer, and stored at 1520 C in a chamber until processing. Pump-

    kins with similar initial moisture content (9597 kg water/100 kg

    product) and soluble solids (24 Brix) were selected for the exper-

    iments. Cylinders (25 mm length, 15 mm diameter) from the

    parenchyma tissue were obtained employing a metallic cork borer

    and a cutter. In order to obtain a good structural and compositional

    homogeneity in the samples, the cylinders were taken from the

    middle zone of the mesocarp, parallel to the major axis of the fruit.

    2.2. Dehydration experiments

    Pumpkin cylinders were dehydrated with sucrose solutions atdifferent concentrations and temperatures selected using a uni-

    form shell design (Doehlert, 1970), as observed inTable 1. These

    solutions were prepared with distilled water and commercial

    sucrose. The cylinders were put in baskets, which were fully

    immersed into stirred glass vessels (diameter 15 cm, height

    25 cm) containing the osmotic solution. Then, the vessels were

    hermetically closed to avoid water evaporation during the osmotic

    treatments. Agitation was conducted using a magnetic stirrer; the

    speed was chosen according to the kinematic viscosity of the

    osmotic solution to obtain a constant Reynolds number (ca.

    3000) (Mayor et al., 2006). Reynolds number was calculatedaccording Eq.(1)(Perry and Green, 1999)

    ReNd

    2

    m1

    where kinematic viscosities of the osmotic solutions were obtained

    from other authors (Chenlo et al., 2002), and ranged from

    0.94 106 up to 37.5 106 m2/s.

    The weight ratio of osmotic solution to pumpkin cylinders was

    20:1 to maintain a constant concentration of the osmotic solution

    during OD. Thermoregulation was obtained by means of a thermo-

    static bath (0.2 C). Some dehydrated samples were removed

    from the vessels at different process times, immediately they were

    gently blotted with paper to remove the excess of osmotic solution

    and kept in plastic boxes till experimental determinations.The same experimental procedure was followed in the OD

    experiments with sodium chloride solutions at concentrations of

    Nomenclature

    a parameter of Eqs.(11) and (12)AD air dryingARD average relative deviationb parameter of Eq.(11)c parameter of Eq.(11)

    C compactnessd diameter of the stirrer (m)D diameter (m)E elongationL length (m)Lm length of major axis (m)m sample mass (kg)N number of revolutions per second (s1)NMC normalized moisture contentOD osmotic dehydrationp perimeter (m)R roundnessR2 coefficient of determinationRe Reynolds numbers solid mass (kg)

    S surface area (m2)SG solids gain (kg/kg)

    suc sucroseT temperature (C)V volume (m3)WL water loss (kg/kg)WR weight reduction (kg/kg)

    X variable of Eq.(11)Y coded variable

    Greek symbolse porositym kinematic viscosity (m2/s)q density (kg/m3)

    Subscriptsb bulkis insoluble solidso initialp particless initial soluble solidssuc sucrosew water

    Table 1

    Experimental design for dehydration of pumpkin with sucrose solutions.

    Coded experimental plan Actual experimental plan

    Y1 Y2 Sucrose (kg/100 kg) T(C)

    1 0 30 25

    0 0 45 25

    1 0 60 25

    0.5 0.866 37.5 12

    0.5 0.866 37.5 38

    0.5 0.866 52.5 12

    0.5 0.866 52.5 38

    30 L. Mayor et al./ Journal of Food Engineering 103 (2011) 2937

  • 7/24/2019 Shrinkage Density Porosity Shape Changes Drying Pumpkin

    3/9

    5, 10 and 20 kg/100 kg at 25, 38 and 12 C, respectively; and with

    ternary solutions 3.75% NaCl58% sucrose and 7.5% NaCl45% su-

    crose at 25 C.

    AD of fresh samples was performed in an oven at 70 C under

    natural convection and an air relative humidity of 6 2% (deter-

    mined by wet bulb thermometer). A vent in the oven avoided

    humidity build-up during drying. Pumpkin cylinders were put on

    a plastic wire net to allow heat and mass transfer through thewhole surface of the samples. At different process times some cyl-

    inders were removed from the oven and kept in plastic boxes till

    experimental determinations.

    2.3. Experimental determinations

    At each selected process time, four samples were used for gravi-

    metric determinations. The bulk volume (V) of each cylinder was

    calculated from the resultant buoyant force of the sample when

    immersed inn-heptane (Lozano et al., 1980). After this measure-

    ment, particle volume (Vp) was obtained with a gas pycnometer

    specially designed to measure this property in moist products

    (Sereno et al., 2007). Sample porosity (e) was calculated by Eq.(2)

    e VVpV

    1 VpV

    2

    After that, the same samples used for volume measurements

    were weighed to determine its weight reduction (WR), Eq. (3),

    and the solids gain (SG), Eq. (4), was evaluated after vacuum drying

    at less than 104 Pa at 70 C till constant weight (AOAC, 1984)

    according to:

    WRmom

    mo3

    SGs somo

    4

    Water loss (WL) and normalized moisture content (NMC) were

    determined by Eqs.(5) and (6), respectively:

    WL SG WR 5

    NMC m smomosom

    6

    Particle (qp) and bulk (qb) densities were calculated from Eqs.

    (7) and (8), respectively:

    qp m

    Vp7

    qb m

    V 8

    Since the soluble solids of pumpkin flesh are mostly sugars, sol-

    uble solids of fresh samples were determined by refractometry

    (Abbe-3L refractometer, Bausch and Lomb, Rochester, NY, USA) at

    20

    C. The clear juice was extracted by manually pressing rawpumpkin flesh (ca. 2 g) between to plastic discs (diameter 3 cm)

    and analyzed directly in the refractometer. Insoluble solids were

    obtained from a mass balance with the values of soluble solids

    and total solids in fresh material.

    Image analysis was performed in samples osmodehydrated

    with 60% sucrose solutions at 25 C. Six samples were removed

    at different process times for the analysis of size and shape param-

    eters. From each sample, one rectangular slab of ca. 0.51 mm of

    thickness was gently cut parallel to the height of the cylinders at

    the maximum section area with a razor blade. One face of the slab

    was stained with a solution of methylene blue 0.1% (Mayor et al.,

    2005) during 15 s. After that, samples were ready for observationunder the stereomicroscope (Olympus SZ-11, Tokyo, Japan) work-

    ing in transmitted light mode. A digital video camera (Sony SSC-

    DC50AP, Tokyo, Japan) was attached to the microscope and con-

    nected to a computer. Image acquisition was carried out with an

    interface (PCTV videocard, Pinneacle Systems GmbH, Munich,

    Germany). Images were calibrated with a stage micrometer of

    2 mm length and divisions of 0.01 mm intervals (Leitz Wetzlar,

    Germany). Image analysis of the isolated sample contour was

    performed using free software UTHSCSA Image Tool v.2.0 (Health

    Science Centre, University of Texas, San Antonio, TX). Several geo-

    metrical parameters of the samples were analyzed (Lewicki and

    Pawlak, 2003; Mayor et al., 2005): surface area (S); perimeter of

    the contour (p), length of the major axis (Lm), defined as the length

    of the longest line that can be drawn through the object; length of

    the minor axis, defined as the length of the longest line that can be

    drawn through the object perpendicular to the major axis; elonga-

    tion (E), defined as the ratio of Lm to the length of the minor axis;

    roundness (R) and compactness (C) defined by the Eqs. (9) and (10),

    respectively:

    R4 pS

    p2 9

    C

    ffiffiffiffiffi4 Sp

    q

    Lm 10

    Rand Cparameters give an idea of the circularity of the object. Both

    shape factors range from 0 to 1; when the value is one, the object is

    a perfect circle, when their value decrease the object becomes less

    circular and less round.

    In order to determine the linear dimensions, average values of

    five measurements at different zones of the cylinders were per-

    formed for diameter (D) and length (L).Table 2shows a summary

    of the experimental determinations performed in this work, as well

    as the sampling times selected for each experiment.

    3. Results and discussion

    3.1. Shrinkage during dehydration

    Fig. 1shows experimental shrinkage data of pumpkin cylinders

    versus WL, WR and NMC, respectively. Plots on the left correspond

    to OD with sucrose solutions at different conditions, whereas plotson the right correspond to OD with different osmotic agents and

    AD. Osmodehydrated samples shrank up to 27% of their initial vol-

    ume, depending on the process conditions used. Air dried samples

    shrank at the end of drying up to 5% of the initial volume.

    Table 2

    Summary of experimental determinations.

    Dehydration treatment Process conditions Experimental determinations Sampling times (h)

    Osmotic dehydration Sucrose,Table 1 a, b, c 0, 0.08, 0.25, 0.5, 1, 1.5, 2, 2.5, 3, 4, 6 and 9

    60% Sucrose, 25 C a, d 0, 0.5, 1, 3, 6 and 9

    NaCl a, b 0, 0.08, 0.5, 1, 2, 2.5, 3, 4, 5, 6 and 8

    NaCl/sucrose a, b 0, 0.08, 0.25, 0.5, 1, 1.5, 2, 2.5, 3, 3.5, 4, 5 and 6

    Air drying Natural convection (oven drying) at 70 C a, b, c 0, 0.5, 1, 1.5, 2, 2.5, 3, 3.5, 4, 5, 6 and 8

    (a) Kinetic parameters (water loss, solids gain, weight reduction, normalized moisture content); (b) bulk volume; (c) particle volume; and (d) image analysis.

    L. Mayor et al. / Journal of Food Engineering 103 (2011) 2937 31

    http://-/?-http://-/?-
  • 7/24/2019 Shrinkage Density Porosity Shape Changes Drying Pumpkin

    4/9

    InFig. 1a can be observed a linear decrease of volume with WL

    during OD. The decrease is more accentuated in the case of NaCl

    solutions, followed by sucrose solutions and NaCl/sucrose solu-

    tions. No effect of process conditions concentration and tempera-

    ture is observed for the same osmotic agent, as shown in Fig. 1a

    for sucrose solutions. Nieto et al. (2004), observed a linear decrease

    of sample volume with WL during OD of apple; the decrease was

    more accentuated at the same WL for the samples osmodehydrat-

    ed with sucrose solutions compared with glucose solutions due to

    the corresponding lower SG value.Mavroudis et al. (1998) during

    OD of apples (var. Granny Smith) with sucrose solutions also ob-

    served a linear decrease of volume with the decrease of water in

    the material; they observed no effect of process temperature onshrinkage.

    Air dried samples also showed a linear volume decrease with

    WL, but in this case the decrease is higher than the shrinkage of

    OD samples. At the same WL, the SG during OD decreases the vol-

    ume reduction promoted by the water removal. Volume decreases

    linearly with increasing WR (Fig. 1b), independently on the process

    conditions and dehydration methods. Moreira and Sereno (2003)

    during OD of apple with sucrose solutions found linear relation-

    ships between WR and shrinkage independently on concentration,

    temperature and hydrodynamic conditions of the osmotic solu-

    tions. During AD, this linear behaviour of shrinkage against WR

    (and consequently against WL) is often reported (Lozano et al.,

    1983; Zogzas et al., 1994). Shrinkage of osmodehydrated and air

    dried samples shows a non-linear decrease with moisture content(Fig. 1c). This decrease is faster for pumpkins submitted to AD, due

    0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.00.0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1.0

    1.1

    V/Vo

    WL (kg/kg)

    30% suc, 25C45% suc, 25C

    60% suc, 25C37.5% suc, 12C

    37.5% suc, 38C

    52.5% suc, 12C52.5% suc, 38C

    Eq. (11)

    0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.80.0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1.0

    1.1

    V/V

    o

    WR (kg/kg)

    30% suc, 25C45% suc, 25C

    60% suc, 25C

    37.5% suc, 12C37.5% suc, 38C

    52.5% suc, 12C

    52.5% suc, 38C

    Eq. (11)

    1.0 0.9 0.8 0.7 0.6 0.50.0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1.0

    1.1

    V/Vo

    NMC

    30% suc, 25C

    45% suc, 25C

    60% suc, 25C37.5% suc, 12C

    37.5% suc, 38C

    52.5% suc, 12C

    52.5% suc, 38C

    Eq. (11)

    0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.00.0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1.0

    1.1

    V/Vo

    WR (kg/kg)

    OD sucrose solutions

    OD NaCl/sucrose solutions

    OD NaCl solutions

    Air drying

    Eq. (11), all treatments

    0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.00.0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1.0

    1.1

    V/Vo

    WL (kg/kg)

    OD, sucrose solutions

    OD, NaCl/sucrose solutionsOD, NaCl solutions

    Air drying

    (a)

    (b)

    (c)

    1.0 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0.00.0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1.0

    1.1

    V/Vo

    NMC

    OD, sucrose solutionsOD, NaCl/suc solutions

    OD, NaCl solutionsAir drying

    Fig. 1. Shrinkage during dehydration of pumpkin cylinders versus (a) water loss, (b) weight reduction, and (c) normalized moisture content. Left figures correspond to

    osmotic dehydration with sucrose solutions, whereas right figures correspond to osmotic dehydration with different solutions and air drying.

    32 L. Mayor et al./ Journal of Food Engineering 103 (2011) 2937

  • 7/24/2019 Shrinkage Density Porosity Shape Changes Drying Pumpkin

    5/9

    to the aforementioned explanations given for the analysis of

    shrinkage against WL.

    Variation in the volume of the gas phase inside the vegetable

    tissue also can contribute to increase or decrease the theoretical

    shrinkage produced by the removal of water. An ideal shrinkage

    could be defined as the shrinkage due to mass fluxes (WL and

    SG), assuming that the volume of the initial solids of the material

    remains constant. Considering volume additivity, the changes involume were calculated from experimental data of WL and solute

    gain. The density values of water, sucrose and NaCl at 25 C, are

    shown in Table 3. The assumption that volumes are additive is

    acceptable for the calculations, since excess volume of binary solu-

    tions of sucrose and NaCl and ternary NaCl/sucrose solutions ob-

    tained comparing experimental data from other authors is near

    to 1%, rarely exceeding 2% (Chenlo et al., 2002; Lide 2005).

    Fig. 2shows the ideal shrinkage of pumpkin cylinders dehy-

    drated with binary solutions of sucrose and NaCl and air dried ver-

    sus the experimental shrinkage data. It is observed that, for each

    dehydration method, experimental shrinkage is higher than the

    ideal shrinkage given by the mass fluxes during dehydration.

    These differences increase at high degrees of shrinkage, and conse-

    quently with the volume of removed water, and are related to the

    decrease of the air volume and collapse of the material ( Khalloufi

    et al., 2009).Barat et al. (2001), during OD of apples with sucrose

    solutions, observed that the decrease of total volume was higher

    than the decrease of the liquid phase volume in the samples. Sev-

    eral phenomena can promote the collapse, such as capillary forces

    caused by the water removal and loss of turgor pressure in the cells

    (Prothon et al., 2003).

    Shrinkage of pumpkin during dehydration can be correlated

    with the kinetic parameters, X, by means of empirical polynomial

    equations:

    V

    Vo1 aXbX

    2cX3 11

    whereXcan be WL, WR, or NMC. Eq. (11) was fitted to experimental

    shrinkage data obtained in this work. The corresponding fits were

    carried out considering each osmotic agent alone, considering all

    the osmotic treatments together, AD alone, and all the dehydration

    treatments together (AD and OD). Table 4shows the values of the

    parameters after fitting (only were considered acceptable the fit-

    tings with ARD < 10% andR2 > 0.9).Fig. 1shows some of these fit-

    tings, as examples. For WL and WR the linear fit is satisfactory in

    the case of OD treatments and AD, separately. For all the dehydra-

    tion methods together only WR fit (linear) is acceptable. For NMC,

    cubic models are acceptable (except for AD and for the global anal-

    ysis of all treatments). These equations are useful because allow the

    prediction of shrinkage data independently on the process condi-

    tions (concentration and temperature) used.

    3.2. Bulk density, particle density and porosity

    Table 5 shows the values of some physicochemical properties of

    raw pumpkin parenchymatic tissue obtained in this work. Particle

    and bulk densities and porosity present the typical variability ob-

    served in other vegetables; for this reason, the changes in these

    properties during dehydration are presented as reduced values.

    The change of bulk density during OD with sucrose or NaCl

    solutions was restricted (ca. 5%). Higher change was obtained for

    ternary NaCl/sucrose solutions (ca. 10%) and AD (ca. 13%). Similar

    trend is observed for all the treatments during dehydration; bulk

    density initially increases, reaching a maximum value and then de-creases or fluctuates till the end of the process (data not shown).

    Table 3

    Density,q , for fresh and dehydrated pumpkin components at 25 C.

    Component q (kg/m3) Reference

    Water 997 Lide (2005)

    Fructose 1665 Lide (2005)

    Glucose 1562 Lide (2005)

    NaCl 2170 Lide (2005)

    Sucrose 1581 Lide (2005)

    Cellulose 1550 Lozano et al. (1980)

    0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

    0.0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1.0

    Idealvolume

    loss/Vo

    (Vo-V)/Vo

    O.D. Sucrose solutions

    O.D. NaCl solutions

    Air drying

    Diagonal

    Fig. 2. Ideal volume loss versus actual volume loss during osmotic dehydration of

    pumpkin fruits with binary sucrose and NaCl solutions and air drying.

    Table 4

    Parameters of Eq.(11)for osmotic dehydration (OD) and air drying.

    X a b c R2 ARD (%)

    OD sucrose solutions

    WL (linear) 0.88 0.99 1.65

    WR (linear) 1.01 0.99 1.44

    NMC (cubic) 1.74 0.18 1.94 0.97 5.26

    OD NaCl solutions

    WL (linear) 0.94 0.96 2.40

    WR (linear) 1.09 0.99 1.19

    NMC (cubic) 16.04 37.82 21.74 0.91 6.38

    OD NaCl/sucrose solutions

    WL (linear) 0.82 0.99 1.18

    WR (linear) 1.02 0.99 1.96

    NMC (cubic) 2.17 1.50 0.66 0.99 2.67

    All OD treatments

    WL(linear) 0.87 0.99 3.13

    WR(linear) 1.02 0.99 1.81

    NMC (cubic) 1.74 0.22 1.49 0.91 8.09

    Air drying

    WL, WR (linear) 1.00 0.99 2.86

    All dehydration methods

    WR (linear) 1.02 0.99 2.43

    WL= water loss, WR= weight reduction, NMC = normalized moisture content,

    R2 = coefficient of determination, ARD= average relative deviation.

    Table 5

    Some physicochemical properties of raw pumpkin parenchymatic tissue.

    Property Average value Range

    Moisture content (%) 95.57 [94.4496.92]

    Soluble solids (%) 3.22 [2.143.63]

    Insoluble solids (%) 1.21 [0.781.97]

    Bulk density (kg/m3) 890 [860920]

    Particle density (kg/m3) 1040 [10031070]

    Porosity (%) 14.79 [10.2218.18]

    L. Mayor et al. / Journal of Food Engineering 103 (2011) 2937 33

  • 7/24/2019 Shrinkage Density Porosity Shape Changes Drying Pumpkin

    6/9

    Nieto et al. (2004), during OD of apples with sucrose and glu-

    cose solutions, observed an increase of bulk density of apple sam-

    ples in the beginning of the treatments (around the first hour of

    treatment) for both osmotic agents, and then bulk density fluctu-

    ated until the end of the processes. During AD of different fruits

    and vegetables, Lozano et al. (1983) and Krokida and Maroulis

    (1997)observed that the change in bulk density was different for

    each food material tested. In some cases bulk density increasedduring dehydration (banana, carrot), in other cases decreased

    (apple), and still in other cases initially increased, reached a max-

    imum value and then decreased (sweet potato, garlic). This differ-

    ent behaviour could be associated with different physicochemical

    characteristics of the raw material, such as chemical composition,

    initial porosity or the presence of soft/rigid structures, which can

    lead to different type of stresses during processing (Rahman,

    2001). It is reasonable to think that for OD the characteristics of

    the raw material also can influence in the change in bulk density

    as observed for AD of vegetables. The type and amount of infused

    solids are other important factors in the change of bulk density

    during OD.

    Bulk density changes can be predicted by means of Eq. (12):

    qb

    qbo

    1 WR

    V=Vo

    1 WR

    1 aWR 12

    where a is the corresponding regression coefficient of Eq. (11)

    (Table 4), for shrinkage and WR correlations for each osmotic agent

    and for AD. Average and maximum relative deviations between

    experimental and predicted density values were, respectively,

    2.4% and 5.6% for sucrose solutions; 1.3% and 4.8% for NaCl solu-

    tions; 3.2% and 6.0% for NaCl and sucrose solutions, and 2.9% and

    6.1% for AD.

    Fig. 3 shows the changes in normalized particle density (qp/qpo)

    and normalized porosity (e/eo) against WR. Similar trends of bulk

    and particle densities and porosity were observed with WL and

    WR of samples. For NMC the observed trend was not as clear as

    in the case of WL and WR (data not shown).

    Particle density (Fig. 3a) increases slowly at the beginning of ODwith sucrose solutions, but above WR = 0.5 the increase is more

    pronounced. No significant differences are observed among the

    process conditions tested. Particle density increases by the compo-

    sitional change of the wet solid matrix during dehydration. Ini-

    tially, water content is high, but during dehydration the content

    of more dense substances (sucrose, cellulose) increases leading to

    the increase of particle density. For air dried samples, the behav-

    iour is similar; initially particle density increases very low and

    above WR = 0.6 increases more pronouncedly up to a value around

    40% higher at the end of the process. During ODof apple with sugar

    solutions, a progressive increase of particle density along the pro-

    cess was observed (Nieto et al., 2004). For vegetables, the same

    behaviour of particle density along drying was also observed

    (Krokida and Maroulis, 1997).

    The modelling of particle density, Eq. (7), during dehydration

    can be performed from the composition of the material. The parti-

    cle volume can be calculated from the masses and densities of each

    component. For pumpkin parenchyma, fresh material is composed

    by water, insoluble and soluble solids, and gas phase. So the parti-

    cle volume (without the gas phase) can be defined as

    Vpo VwVisVss 13

    In terms of masses of the components and densities, Eq.(13)can be

    rewritten as

    Vpomwqw

    misqis

    mssqss

    14

    For osmodehydrated pumpkin with sucrose and NaCl solutions,

    the gained solids must be taken into account, so the particle vol-ume can be defined, as

    Vp mwqw

    misqis

    mssqss

    msucqsuc

    mNaClqNaCl

    15

    Lozano et al. (1980)considered the insoluble solids of apple tis-

    sue as cellulose. In this work the same assumption was considered.

    Soluble solids in pumpkin are mainly fructose and glucose in the

    same proportion, so an average value of the densities of fructose

    and glucose was taken as the density of initial soluble solids. Den-

    sity values used in the calculations are shown inTable 3.Predictedvalues of particle volume for fresh and dehydrated pumpkin fruits

    with sucrose solutions and AD were compared with experimental

    data obtained with the gas pycnometer. This predictioncan be con-

    sidered adequate, leading to a relative deviation of the predicted

    values of 2.63% on average.

    Finally, normalized particle density can be obtained by means

    of Eq.(16)

    qp

    qpo

    1 WR

    Vp=Vpo16

    where Vpo and Vp are obtained by means of Eqs. (14) and (15),

    respectively.

    Average relative deviation (ARD) between experimental data

    and predicted values obtained with Eq.(15)was 2.02%, indicating

    that the model gives a good prediction of experimental data. Fig. 3a

    shows predicted values of normalized particle density for OD with

    60% sucrose solutions at 25 C and AD at 70 C.

    Porosity of dehydrated pumpkin with sucrose solutions (Fig. 3b)

    slightly decreases up to intermediate WR values (ca. 0.4); after that

    point porosity increases till the end of the process. No effect of the

    process conditions on porosity trend with WR or WL is observed.

    For AD the behaviour is similar; at the beginning of the process

    porosity fluctuates but above WR = 0.5, porosity increases and at

    the end of the process almost triplicates (e/eo= 2.8, WR = 0.95) its

    initial value.

    Mavroudis et al. (1998) observed an increase of porosity in

    osmodehydrated apples at the end of the process. Giraldo et al.

    (2003) showed that during OD of mango porosity of dehydratedsamples initially decreased and after that increased. The initial de-

    crease of sample porosity can be explained by the fast initial

    impregnation of the tissue with the osmotic solution, which pene-

    trates into the external pores by capillary forces and other mass

    transfer mechanisms. The accumulation of sucrose in the external

    surface of the material generates a dense layer that hinders the fur-

    ther penetration of the osmotic solution and simultaneously min-

    imizes the gas flow from the material to the solution. These

    combined phenomena increase the food porosity.

    Porosity during AD of foodstuffs can follow different behaviours

    (Lozano et al., 1983; Krokida and Maroulis, 1997); in some cases it

    decreases (sweet potato); in others it initially decreases and then

    increases (pear) and still in other cases it increases during the

    whole drying process (apple, banana). As commented for bulk den-sity change, this different behaviour can be associated with the ini-

    tial structural and compositional characteristics of the raw

    material, as well as the process conditions.

    Fig. 4shows the changes in total volume, particle volume and

    air volume during OD of pumpkin with 60% sucrose solutions

    and AD. In the initial stage (up to WR = 0.5) in both treatments

    the three volumes decrease during dehydration and the relative

    decrease of air volume is the highest. The relative decrease of par-

    ticle volume and total volume is practically the same for OD. For

    AD, total volume decreases more than particle volume. Above

    WR = 0.5, in OD the particle volume starts to decrease in percent-

    age more than total volume, and air volume fluctuates and practi-

    cally remains constant until the end of the process. In the same

    range for dried samples, the air phase volume remains constantup to WR = 0.7 and then decreases until the end of drying; total

    34 L. Mayor et al./ Journal of Food Engineering 103 (2011) 2937

  • 7/24/2019 Shrinkage Density Porosity Shape Changes Drying Pumpkin

    7/9

    volume decreases less than particle volume like in OD samples. In

    this way, it seems that the gas phase is better retained during OD

    than in AD. As commented before, the formation of a dense layer of

    osmotic agent in the material surface can be the cause of this

    phenomenon.

    Porosity, Eq. (2), can be predicted by means of Vp calculated

    from Eqs.(14) and (15), andVwith Eq.(11). ARD in the predictionof porosity values was 14.5%.Fig. 3b shows experimental and pre-

    dicted values of normalized porosity for air dried and osmotic

    dehydrated pumpkin with 60% sucrose solutions.

    3.3. Sample shape analysis

    Fig. 5shows the contour of cylinders dehydrated with sucrose

    solutions (60%, 25 C), at different process times. It is observed asize decrease during the process. The shape also changes during

    0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    1.2

    1.4

    1.6

    1.8

    2.0

    WR (kg/kg)

    30% suc, 25C 45% suc, 25C

    60% suc, 25C 37.5% suc, 12C

    37.5% suc, 38C 52.5% suc, 12C

    52.5% suc, 38C Predicted 60% suc

    0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

    0.8

    0.9

    1.0

    1.1

    1.2

    1.3

    WR (kg/kg)

    30% suc, 25C 45% suc, 25C

    60% suc, 25C 37.5% suc, 12C

    37.5% suc, 38C 52.5% suc, 12C

    52.5% suc, 38C Predicted, 60% suc

    (a)

    (b)

    0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.00.8

    0.9

    1.0

    1.1

    1.2

    1.3

    1.4

    1.5

    WR (kg/kg)

    OD sucrose solutions

    Air drying

    Predicted, air drying

    0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.00.0

    0.5

    1.0

    1.5

    2.0

    2.5

    3.0

    WR (kg/kg)

    OD sucrose solutions

    Air drying

    Predicted, air drying

    Fig. 3. Changes in particle density (a) and porosity (b) during dehydration of pumpkincylinders versus weight reduction. Left figures correspond to osmotic dehydration with

    sucrose solutions at different process conditions, whereas figures on the right correspond to osmotic dehydration with sucrose solutions and air drying.

    0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

    0.0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1.0

    1.1

    V/Vo

    WR (kg/kg)

    Total volume

    Particle volume

    Air volume

    (a) (b)

    0.0 0.2 0.4 0.6 0.8 1.0

    0.0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1.0

    1.1

    V/Vo

    WR (kg/kg)

    Total volume

    Particle volume

    Air volume

    Fig. 4. Relative volume changes for total volume, particle volume and air volume during dehydration of pumpkin cylinders in (a) 60% sucrose solutions at 25 C and (b) air

    drying at 70 C.

    L. Mayor et al. / Journal of Food Engineering 103 (2011) 2937 35

  • 7/24/2019 Shrinkage Density Porosity Shape Changes Drying Pumpkin

    8/9

    dehydration; shrinkage is more accentuated at mid-length and

    mid-thickness of the cylinder, whereas at the edges shrinkage is

    less pronounced; this corner effect is clearly observed in the

    most dehydrated cylinder (Fig. 5f).Del Valle et al. (1998)also re-

    ported this effect during OD of apple cylinders, and Mulet et al.(2000)during AD of potato cubes.

    As observed inFig. 6, no significant differences were found in

    the decrease of length and diameter during the process, indicating

    an isotropic shrinkage. Trujillo et al. (2007) found similar results

    during the AD of beef meat discs (L/D 0.25). However, Mulet

    et al. (2000)showed an anisotropic shrinkage in the AD of potato

    (L/D 4.6) and cauliflower stem (L/D 1.7) cylinders, where the

    diameter decrease was higher than the length decrease for bothproducts. In the case of potato cylinders, the authors suggested

    that the high L/D ratio favoured the formation of an inner core

    along the axis length maintaining the shape along this axis. In

    the case of cauliflower stems, the presence of oriented fibres made

    the product stiffer in a preferential orientation and the shrinkage

    was more pronounced in the diameter. Based on those results,

    two important factors can affect the shrinkage isotropicity: the

    existence of preferential pathways of mass transfer (due to geo-

    metric and structural features) and the homogeneity of the mate-

    rial structure (due to structural features).

    Fig. 7shows the relative changes in the shape factors as a func-

    tion of WR. The average initial values of the shape factors were

    1.657, 0.670 and 0.749 for elongation, roundness and compactness,

    respectively. Elongation slightly increases, whereas roundness andcompactness decrease during dehydration. It is often reported a

    decrease of roundness during dehydration of foods, as in the case

    of AD of apple discs (Mayor et al., 2005; Fernandez et al., 2005)

    or apricot cubes (Riva et al., 2005). The tissue suffers deformations

    as a consequence of the water removed in the material; in this way

    roundness and compactness decrease during dehydration. Elonga-

    tion increases mainly due to the corner effect that reduces the

    minor axis length (minimum value of the diameter of the cylinder)

    but maintains the major axis length (distance between two oppo-

    site corners of the cylinder).

    4. Conclusions

    Shrinkage of samples was observed for all the dehydrationmethods studied. Independently on the process conditions (con-

    centration and temperature of the osmotic solution), volume de-

    creased linearly with WL and WR during OD. Air dried samples

    showed the same behaviour but shrinkage was more accentuated

    for the same WL.

    Bulk density varied in a restricted range during dehydration;

    differences between maximum and minimum values were around

    5% for OD with binary sucrose and NaCl solutions, 10% for ternary

    NaCl/sucrose solutions and 13% for AD. For all the dehydration

    methods, bulk density initially increased, then reached a maximum

    value and after that decreased or fluctuated till the end of the

    process.

    For all the treatments, particle density increased slowly at the

    beginning of the process, and at certain WR value (0.5 for OD withsucrosesolutionsand 0.6for AD)the increasewas morepronounced.

    Fig. 5. Changes in shape and size during osmotic dehydration of pumpkin cylinders

    in 60% sucrose solutions at 25 C, at different process times (h). (a) 0, (b) 0.5, (c) 1,

    (d) 3, (e) 6, and (f) 9. The horizontal line at the bottom of images corresponds to

    2 mm.

    0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    1.2

    Normalizeddimension

    WR (kg/kg)

    Diameter

    Length

    Fig. 6. Relative changes in dimensions of osmodehydrated pumpkin cylinders (60%sucrose, 25 C) versus weight reduction.

    0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    1.2

    1.4

    Normalizedsh

    apefactor

    WR (kg/kg)

    Elongation

    Roundness

    Compactness

    Fig. 7. Changes in shape factors during osmotic dehydration (60% sucrose, 25C) of

    pumpkin cylinders versus weight reduction.

    36 L. Mayor et al./ Journal of Food Engineering 103 (2011) 2937

  • 7/24/2019 Shrinkage Density Porosity Shape Changes Drying Pumpkin

    9/9

    The change in porosity also followed the same behaviour during

    OD and AD. Initially, porosity decreased (up to WR = 0.4 and 0.5 for

    OD and AD, respectively) and then porosity started to increase till

    the end of the treatment, increasing twofold and threefold the

    initial value for OD with sucrose solutions and AD, respectively.

    The proposed models to evaluate shrinkage, bulk and particle

    densities and porosity from WR were satisfactorily applied.

    Experimental results indicated that shrinkage of pumpkin cylin-ders was isotropic. During dehydration, elongation increased and

    roundness and compactness decreased.

    Acknowledgement

    The author Luis Mayor wishes to acknowledge SFRH/BD/3414/

    2000 PhD Grant to Fundaao para a Cincia e a Tecnologia,

    Portugal.

    References

    AOAC, 1984. Official Methods of Analysis, 14th ed. Association of Official AnalyticalChemists, Gaithersburg, MD.

    Barat, J.M., Fito, P., Chiralt, A., 2001. Modeling of simultaneous mass transfer andstructural changes in fruit tissues. Journal of Food Engineering 49, 7785.

    Bui, H.T., Makhlouf, J., Ratti, C., 2009. Osmotic dehydration of tomato in sucrosesolutions: Ficks law classical modelling. Journal of Food Science 74, E250E258.

    Chenlo, F., Moreira, R., Pereira, G., Ampudia, A., 2002. Viscosities of aqueoussolutions of sucrose and sodium chloride of interest in osmotic dehydrationprocesses. Journal of Food Engineering 54, 347352.

    Del Valle, J.M., Cuadros, T.R.M., Aguilera, J.M., 1998. Glass transitions and shrinkageduring drying and storage of osmosed apple pieces. Food Research International31, 191204.

    Doehlert, D.H., 1970. Uniform shell designs. Applied Statistics 19, 231239.Fernandez, L., Castillero, C., Aguilera, J.M., 2005. An application of image analysis to

    dehydration of apple discs. Journal of Food Engineering 67, 185193.Giraldo, G., Talens, P., Fito, P., Chiralt, A., 2003. Influence of sucrose solution

    concentration on kinetics and yield during osmotic dehydration of mango.Journal of Food Engineering 58, 3343.

    Khalloufi, S., Almeida-Rivera, C., Bongers, P., 2009. A theoretical model and itsexperimental validation to predict the porosity as a function of shrinkage andcollapse phenomena during drying. Food Research International42, 11221130.

    Konopacka, D., Seroczynska, A., Korzeniewska, A., Jesionkowska, K., Niemirowicz-

    Szczytt, K., Plocharski, W., 2010. Studies on the usefulness ofCucurbita maximafor the production of ready-to-eat dried vegetable snacks with a highcarotenoid content. LWT Food Science and Technology 43, 302309.

    Kowalska, H., Lenart, A., Leszczyk, D., 2008. The effect of blanching and freezing onosmotic dehydration of pumpkin. Journal of Food Engineering 86, 3038.

    Krokida, M.K., Maroulis, Z.B., 1997. Effect of drying method on shrinkage andporosity. Drying Technology 15, 24412458.

    Lewicki, P.P., Lukaszuk, A., 2000. Effect of osmotic dewatering on rheologicalproperties of apple subjected to convective drying. Journal of Food Engineering45, 119126.

    Lewicki, P.P., Pawlak, G., 2003. Effect of drying on microstructure of plant tissue.Drying Technology 21, 657683.

    Lide, D.R., 2005. CRC Handbook of Chemistry and Physics, 85th ed. CRC Press, BocaRaton, FL.

    Lozano, J.E., Rotstein, E., Urbicain, M.J., 1980. Total porosity and open-pore porosityin the drying of fruits. Journal of Food Science 45, 14031407.

    Lozano, J.E., Rotstein, E., Urbicain, M.J., 1983. Shrinkage, porosity and bulk density offoodstuffs at changing moisture contents. Journal of Food Science 48, 14971502 (1553).

    Mavroudis, N.E., Gekas, V., Sjholm, I., 1998. Osmotic dehydration of apples.Shrinkage phenomena and the significance of initial structure on mass transferrates. Journal of Food Engineering 38, 101123.

    Mayor, L., Sereno, A.M., 2004. Modelling shrinkage during convective drying of foodmaterials: a review. Journal of Food Engineering 61, 373386.

    Mayor, L., Silva, M.A., Sereno, A.M., 2005. Microstructural changes during drying ofapple slices. Drying Technology 23, 22612276.

    Mayor, L., Moreira, R., Chenlo, F., Sereno, A.M., 2006. Kinetics of osmoticdehydration of pumpkin with sodium chloride solutions. Journal of FoodEngineering 74, 253262.

    Mayor, L., Moreira, R., Chenlo, F., Sereno, A.M., 2007. Osmotic dehydration kineticsof pumpkin fruits using ternary solutions of sodium chloride and sucrose.Drying Technology 25, 17491758.

    McLaughlin, C.P., Magee, T.R., 1998. The determination of sorption isotherm andisosteric heats of sorption of potatoes. Journal of Food Engineering 35, 267280.

    McMinn, W.A.M., Magee, T.R.A., 1997. Physical characteristics of dehydratedpotatoes part II. Journal of Food Engineering 33, 4955.

    Moreira, R., Sereno, A.M., 2003. Evaluation of mass transfer coefficients andvolumetric shrinkage during osmotic dehydration of apple using sucrosesolutions in static and non-static conditions. Journal of Food Engineering 57,2531.

    Mulet, A., Garcia-Reverter, J., Bon, J., Berna, A., 2000. Effect of shape on potato andcauliflower shrinkage during drying. Drying Technology 18, 12011219.

    Nieto, A.B., Salvatori, D.M., Castro, M.A., Alzamora, S.M., 2004. Structural changes inapple tissue during glucose and sucrose osmotic dehydration: shrinkage,porosity density and microscopic features. Journal of Food Engineering 61,269278.

    Panyawong, S., Devahastin, S., 2007. Determination of deformation of a foodproduct undergoing different drying methods and conditions via evolution of ashape factor. Journal of Food Engineering 78, 151161.

    Perry, R.H., Green, D.W., 1999. Perrys Chemical Engineers Handbook, seventh ed.McGraw-Hill Companies, New York.

    Prothon, F., Ahrne, L., Sjoholm, I., 2003. Mechanisms and prevention of plant tissuecollapse during dehydration: a critical review. Critical Reviews in Food Scienceand Nutrition 43, 447479.

    Rahman, M.S., 2001. Toward prediction of porosity in food during drying: a briefreview. Drying Technology 19, 113.

    Rahman, M.S., 2005. Mass-volume-area-related properties of foods. In: Rao, M.A.,Rizvi, S.S.H., Datta, A.K. (Eds.), Engineering Properties of Foods. Taylor andFrancis Group, Boca Raton, FL.

    Riva, M., Campolongo, S., Leva, A.A., Maestrelli, A., Torreggiani, D., 2005. Structureproperty relationships in osmo-air-dehydrated apricot cubes. Food ResearchInternational 38, 533542.

    Sereno, A.M., Silva, M.A., Mayor, L., 2007. Determination of particle density andporosity in foods and porous materials with high moisture content.International Journal of Food Properties 10, 455469.

    Shimada, Y., Roos, Y., Karel, M., 1991. Oxidation of methyl linoleate encapsulated in

    amorphous lactose based food model. Journal of Agricultural and FoodChemistry 39, 637641.

    Simal, S., Rossello, C., Berna, A., Mulet, A., 1998. Drying of shrinking cylinder-shapedbodies. Journal of Food Engineering 37, 423435.

    Souza, J.S., Medeiros, M.F.D., Magalhaes, M.M.A., Rodrigues, S., Fernandes, F.A.N.,2007. Optimization of osmotic dehydration of tomatoes in a ternary systemfollowed by air-drying. Journal of Food Engineering 83, 501509.

    Trujillo,F.J., Wiangkaew, C., Pham, Q.T., 2007. Drying modelingand water diffusivityin beef meat. Journal of Food Engineering 78, 7485.

    Viberg, U., Freuler, S., Gekas, V., Sjholm, I., 1998. Osmotic pretreatment ofstrawberries and shrinkage effects. Journal of Food Engineering 35, 135145.

    White, K.L., Bell, L.N., 1999. Glucose loss and Maillard browning in solids as affectedby porosity and collapse. Journal of Food Science 64 (6), 10101014.

    Zogzas, N.P., Maroulis, Z.B., Marinos-Kouris, D., 1994. Densities, shrinkage andporosity of some vegetables during air drying. Drying Technology 12, 16531666.

    L. Mayor et al. / Journal of Food Engineering 103 (2011) 2937 37