Sie sind auf Seite 1von 11

Journal of Food Engineering 76 (2006) 280290

www.elsevier.com/locate/jfoodeng

Development and application of polysaccharidelipid edible


coating to extend shelf-life of dry bakery products
Barbara Bravin, Donatella Peressini *, Alessandro Sensidoni
Department of Food Science, University of Udine, Via Marangoni, 97-33100 Udine, Italy
Received 5 November 2004; accepted 19 May 2005
Available online 11 July 2005

Abstract
This study investigated the eect of the deposition process used for lm-forming dispersion (spreading and spraying), relative
humidity gradient across the lm (from 2265% to 2285%) and lm thickness (1590 lm) on water vapor permeability (WVP),
tensile strength (TS), percentage elongation at breaking (E) and structure of an emulsied edible lm composed of corn starch, methylcellulose (MC) and soybean oil. The eectiveness of edible coating in controlling moisture transfer in moisture-sensitive products
was evaluated by coating crackers, a low aw-type cereal food. Spread lm gave better water vapor barrier and mechanical properties
than sprayed lm. High atomization pressure and thickness increased lm WVP. Atomization pressure of 2 bar and lm thickness of
30 lm were identied as optimum for the application of edible coating to bakery products. Coated and uncoated (reference) crackers
were stored at 65%, 75% and 85% relative humidity. Moisture uptake and resistance to water vapor transmission (r) were then
calculated. Coated crackers had longer shelf-life and higher r than reference at all storage conditions.
 2005 Elsevier Ltd. All rights reserved.
Keywords: Edible lm; Water vapor permeability; Mechanical properties; Bakery products; Shelf-life

1. Introduction
Low-moisture bakery and extruded products such as
biscuits, snacks, and breakfast cereals have a crispy texture, which contributes to their consumer appeal. Loss
of crispness and softening during storage under high relative humidity conditions are due to increased water
content (Katz & Labuza, 1981; Roudaut, Dacremont,
Valle`s Pa`mies, Colas, & Le Meste, 2002; Sauvageot &
Blond, 1991). The critical eect of hydration has been
associated with plasticization, which causes a decrease
in glass transition temperature (Tg) below ambient
temperature (Champion, Le Meste, & Simatos, 2000;
Nikolaidis & Labuza, 1996). Moisture adsorption is
inuenced by thermodynamics (water activity equilibrium) and the dynamics of mass transfer process. Any
*

Corresponding author. Tel.: +39 0432590726.


E-mail address: donatella.peressini@uniud.it (D. Peressini).

0260-8774/$ - see front matter  2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jfoodeng.2005.05.021

dierence in the water activity of food and the surrounding environment, or in the components of multidomain
foods, induces a driving force for water transfer (Labuza
& Hyman, 1998).
In recent years, the potential of edible lms to control
water transfer, and to improve food quality and shelflife, has received increasing attention from researchers
and industry (Garca, Martino, & Zaritzky, 1998;
Kamper & Fennema, 1985; Rico-Pena & Torres, 1990;
Yaman & Bayoindirli, 2002). An edible coating or lm
has been dened as a thin, continuous layer of edible
material formed or placed on or between foods or
food components. The aim is to produce natural biopolymer-based coating materials with specic properties, which may be eaten together with the food.
Materials that can be used to make edible lms include
polysaccharides (Nisperos-Carriedo, 1994), proteins
(Gennadios, McHugh, Weller, & Krochta, 1994) and
lipids (Hernandez, 1994), or a combination of these.

B. Bravin et al. / Journal of Food Engineering 76 (2006) 280290

Certain additives, such as plasticizers and emulsiers,


may also be used (Debeaufort & Voilley, 1995, 1997).
The functional properties of edible lms are greatly
inuenced by parameters such as formulation, lmforming technology, solvent characteristics, and additives (Gontard, Guilbert, & Cuq, 1992, 1993).
Cellulose and its derivatives such as ethers and esters
are, with starch, the most important raw materials for
the preparation of lms. The usefulness of cellulose as
a material for edible lms may be extended by chemical
modication. One such modied version is methylcellulose (MC), a water-soluble ether with good lm-forming
properties. Glycerol and polyethylene glycols have
been shown to be the most eective plasticizers for
MC (Donhowe & Fennema, 1993; Park, Weller, Vergano, & Testin, 1993). The water vapor barrier properties
of cellulose-based lms have been improved by the
addition of lipids (Kester & Fennema, 1989; Koelsch
& Labuza, 1992).
Starch is a natural polymer that can readily be cast
into lms. Starch lms have poor physical properties,
but these can be improved by blending the starch with
cellulose derivatives and proteins (Arvanitoyannis,
Biliaderis, Ogawa, & Kawasaki, 1998; Arvanitoyannis,
Psomiadou, & Nakayama, 1996; Peressini, Bravin, &
Sensidoni, 2004; Psomiadou, Arvanitoyannis, &
Yamamoto, 1996).
Several starch lms have been prepared using dryprocess approaches, such as thermoplastic extrusion,
based on the thermoplastic properties of polymers when
plasticized and heated above their glass-transition temperature in low water-content conditions (Arvanitoyannis & Biliaderis, 1998; Arvanitoyannis et al., 1996;
Psomiadou et al., 1996; Warburton, Donald, & Smith,
1993). The disadvantage of extruded lms is that they
cannot be used to cover irregular surfaces. The wet-process mechanism is based on a lm-forming dispersion in
which polymers are rst dispersed into a liquid phase,
and then dried at the end of the application. The wet
process permits the application of coatings in liquid
form directly onto food products by dipping, brushing
or spraying.
Water permeability of edible lms is aected by many
factors that depend on the nature of the components,
lm structure, and thermodynamics (Morillon, Debeaufort, Blond, Capelle, & Voilley, 2002). Hydrophilic lms
exhibit water vapor pressure-dependent permeability
(Cuq, Gontard, Aymard, & Guilbert, 1997; Fennema,
Donhowe, & Kester, 1994; Gontard, Thibault, Cuq, &
Guilbert, 1996; McHugh, Aujard, & Krochta, 1994;
McHugh & Krochta, 1994; Roy, Gennadios, Weller, &
Testin, 2000). Film thickness inuences water vapor barrier performance (Cuq, Gontard, Cuq, & Guilbert, 1996;
McHugh, Avena-Bustillos, & Krochta, 1993).
In our previous works (Bravin, Peressini, & Sensidoni, 2004; Peressini et al., 2004; Peressini, Bravin,

281

Lapasin, Rizzotti, & Sensidoni, 2003) a starchMC


lipid edible lm was developed by evaluating and
optimizing the combined eects on lm properties of
plasticizer (glycerol) content, blending levels of methylcellulose (MC) with starch and lipid type and content.
The objective of the present study is to examine the
inuence of the deposition process of lm-forming dispersion, relative humidity gradient across the lm, and
lm thickness on the functional properties (water vapor
permeability and mechanical properties) and structure
of optimized starchMClipid edible lm. The study
also investigated the ability of this coating to retard
moisture transfer between dry bakery food and the
surrounding atmosphere, so prolonging the shelf-life.

2. Materials and methods


2.1. Materials
Corn starch (27% amylose, Sigma) and methylcellulose (MC, medium viscosity, 27.532% methoxyl content, Fluka) were used as lm-forming components of
the hydrophilic continuous phase for emulsion-based
edible lms; commercial soybean oil was used as the
hydrophobic disperse phase; and glycerol (Baker) was
added as a plasticizer. Ethyl alcohol and magnesium
nitrate were purchased from Fluka, while sodium nitrite,
potassium chloride, magnesium chloride, sodium iodide,
potassium carbonate, lithium chloride and potassium
acetate were supplied by Sigma.
2.2. Preparation of emulsied lms
Edible lm-forming dispersions were obtained by dispersion of MC (1.44 g) in 75 mL of distilled water-ethyl
alcohol mixture (2/1, v/v) at 75 C for 10 min, and dispersion and gelatinization of corn starch (3.19 g) in
75 mL of water at 95 C for 30 min. Gelatinized starch
was homogenized at 4000 rpm for 1 min (Polytron PT
3000, Kinematica AG). Glycerol (1.16 g) was then
added to the MC and the dispersion was homogenized
at 6000 rpm for 1 min. The components were mixed with
a magnetic stirrer (800 rpm). MC and starch preparations were mixed together, homogenized at 6000 rpm
for 90 s, and maintained at 75 C for 10 min under stirring. Soybean oil (1.45 g) was added to the starchMCglycerol dispersion, and the mixture was predispersed
under magnetic stirring for 2 min at 800 rpm and
75 C before being homogenized at 6000 rpm for
2 min. The emulsion was then maintained under magnetic stirring for 10 min at 800 rpm and 75 C before
being spread onto glass plates (25 50 1 cm) with a
thin layer chromatography spreader, or sprayed with a
spray system (ASTURO III EC mod. IM, Walmec) at
two atomization pressures (2 and 3.5 bar). Films were

282

B. Bravin et al. / Journal of Food Engineering 76 (2006) 280290

dried at 25 C and at relative humidity of about 40% for


15 h. Film thickness was 30 3 lm (Elcometer 345 digital micrometer, Elcometer Instruments). Films with
thicknesses ranging from 30 to 90 lm were prepared
by spraying at 2 bar to investigate the thickness eect.
Before measurements lms were maintained at a constant relative humidity of 53% and a temperature of
25 C for 10 days using saturated magnesium nitrate
solution.
2.3. Coating application
Commercial cracker was coated by spraying the
emulsion at a pressure of 2 bar and a temperature of
70 C. Coating application was performed in two steps:
the upper surface of the cracker was coated with the liquid lm and dried at 60 C for 2 h, then the same procedure was applied to the lower surface. Spray
conditions (liquid lm volume, distance between the
pneumatic nozzle and product, and spray surface and
time) were standardized to obtain a lm thickness on a
glass plate of 30 lm. Cracker was cooled to room temperature and placed in a desiccator containing P2O5 to
reach water activity of commercial cracker (aw = 0.05).
Untreated cracker was used as reference.
2.4. Water vapor permeability (WVP) of edible lm
Water vapor transfer rate (WVTR) was measured
gravimetrically at a constant dierence of relative
humidity (RH) and at 25 C, using a modication of
ASTM standard method E 96-80 (1990). The test lm
was sealed to a glass permeation cell containing distilled
water, or saturated salt solutions of potassium chloride
(85% RH) and sodium chloride (75% RH); the cell
was placed in a climatic room maintained at 22% RH,
with saturated salt solution of potassium acetate. The
climatic room was equipped with fans to eliminate stagnant air above the test cell. Cell weight was recorded
periodically on a computer to evaluate stationary-state
water vapor transfer. WVTR (g m2 s1) and WVP
(g m1 s1 Pa1) were calculated using the following
equations:
Dm
A  Dt
WVP WVTR  x
p1  p2

WVTR

1
2

where Dm is weight loss of the permeation cell, with


4.9 104 m2 exposed area (A), over time (Dt), x is lm
thickness and p1  p2 is real vapor partial pressure
dierence (Pa) across the lm. The real vapor partial
pressure at the lm inner surface (p1) was corrected for
the stagnant air gap inside the test cell, according to
Gennadios, Weller, and Gooding (1994). After this cor-

rection, the dierences in relative humidity were approximately 2286% instead of theoretical 22100%, 2277%
and 2267% instead of 2285% and 2275%, respectively. WVTR and WVP values are reported as the mean
of eight measurements from two dierent preparations.
2.5. Mechanical properties of edible lms
Tensile strength (TS, MPa) and elongation percentage (E, %) at breakpoint were measured uniaxially by
stretching the specimen (10 2.5 cm) in one direction
at 50 mm/min using an Instron Universal Testing
Instrument (Model 4301). The lms were analyzed in a
climatic room at 53% RH and 25 C. Initial grip separation was set at 4 cm. TS was calculated by dividing the
maximum load by the cross-sectional area of the lm,
E being expressed as a percentage of change in the original specimen length between grips (4 cm), according
to the ASTM standard method D 882-88 (1989). TS
and E values are reported as the mean of 14 measurements from two dierent preparations.
2.6. Moisture adsorption isotherm
Moisture adsorption isotherm at 25 C of lms,
coated crackers and uncoated crackers (about 3 g) was
determined for aw varying from 0.11 to 0.84 using saturated salt solutions in desiccators [LiCl, aw 0.11;
CH3COOK, aw 0.22; MgCl2, aw 0.33; NaI, aw 0.39;
K2CO3, aw 0.43; Mg(NO3)2, aw 0.53; NaNO2, aw 0.65;
KCl, aw 0.84]. Samples were checked at seven-day intervals to ensue saturation. Equilibrium was judged to have
been attained when the dierence between samples
weighed consecutively twice a week was less than
1 mg g1 solids (30 days). Moisture content was determined in the equilibrated samples as the dierence in
weight before and after drying in an oven at 130 C
for 1 h (AACC, 1995). Water activity was evaluated at
25 C by means of an AquaLab CX-2 instrument (Decagon Devices, Inc.). Prior to moisture and aw determination the bakery samples were cut into small pieces.
Moisture contents are reported as the mean of six measurements from two dierent preparations.
2.7. Mechanical properties of crackers
Crackers were equilibrated to aw varying from 0.11 to
0.84, as described for moisture adsorption isotherm
determination. All equilibrated samples were subjected
to a wedge test at 25 C using an Instron Universal Testing Instrument (Model 4301) (Katz & Labuza, 1981;
Vincent, 1998). Crackers were cleaved with a metal
wedge at a cross-head speed of 20 mm/min. Initial slope
(N/mm) was obtained from the forcedeformation
curve. Data are reported as the mean of 12 measurements from two dierent coating applications.

B. Bravin et al. / Journal of Food Engineering 76 (2006) 280290

2.8. Water vapor resistance (r)


Coated and uncoated crackers were held on individual weighing trays sitting on shelves inside a desiccator
at 25 C and dierent RH (65%, 75% and 85%) obtained
using saturated salt solutions. The samples were weighed
periodically. Moisture uptake was estimated by subtracting the initial weight from the weight measurements
taken at dierent times. Water vapor resistance (r,
s cm1) was calculated using a modied Ficks equation
(Avena-Bustillos, Krochta, Saltveit, Rojas-Villegas, &
Sauceda-Perez, 1994; Ben-Yehoshua, Burg, & Young,
1985):


P wv
RH A
3
r
 aw 

100
J
RT
where aw is the water activity of the cracker; RH (%) is
the storage relative humidity (65%, 75% and 85%); Pwv
is saturated water vapor pressure at 25 C (mm Hg); R
is the universal gas constant; T is the storage temperature (K); A is the surface area of the cracker (cm2); J
is initial slope of water uptake in the cracker versus storage time (g s1). r values are reported as the mean of 12
measurements from two dierent coating applications.
2.9. Environmental scanning electron microscopy
(ESEM)
Film cross-section was observed using environmental
scanning electron microscopy (ESEM, model XL 30,
Philips). Small lm strips (5 3 mm) were xed on the
support using double-sided adhesive tape. An accelerating voltage of 15 kV was used.
2.10. Statistical analysis of data
Statistical dierences in the lm functional properties
and cracker mechanical properties of samples were
determined by one-way analysis of variance (ANOVA)
and Ducans multiple range test (p = 0.05) (Statistica
software version 5, 1997). Two-way analysis of variance
was conducted to determine the eects of coating and
relative humidity on water resistance. A Students t-test
was used to identify signicant dierences in slope WVP
versus thickness obtained from linear regression analysis
(p = 0.05).

3. Results and discussion


3.1. Film formation
In the coating process, one or more liquid layers are
deposited on a solid substrate and subsequently dried
to form solid lms, which possess specic functions.
Spreading is widely used to prepare self-supporting edi-

283

ble lms in laboratory scale. Our previous works


(Bravin et al., 2004; Peressini et al., 2003, 2004) reported
the development and optimization of starchMClipid
edible lm, which was prepared by spreading. Application of the edible coating to bakery products such as
crackers requires the formation of lms directly on food
surfaces. Spraying was chosen as the most suitable technique for our purpose because it produces uniform coating. The two deposition processes imply shear strains
and strain rates of dierent magnitude, which may inuence coating performance (Barnes, Hutton, & Walters,
1989). The eect of lm formation on the functional
properties of edible lms was investigated before
proceeding to food applications.
Table 1 shows the inuence of deposition process on
WVP and mechanical properties of starchMC lm containing 20% oil or without oil. Adding oil determined a
signicant reduction in WVP with respect to the lipidfree reference, as observed by other Authors (Garca,
Martino, & Zaritzky, 2000). Water vapor permeability
ranged from 11.68 0.38 ( 1011 g m1 s1 Pa1) to
20.50 1.19 ( 1011 g m1 s1 Pa1) in dierent samples. These results fall within the range for other emulsied edible lms. WVP values of 13.20 ( 1011
g m1 s1 Pa1) have been reported for MC-hydrogenated palm oil (Morillon et al., 2002). Synthetic lms,
such as cellophane and LDPE, respectively, gave WVP
values of 8.4 and 0.02 ( 1011 g m1 s1 Pa1) (Shellhammer & Krochta, 1997).
No statistically signicant dierences (p > 0.05) in
WVP were observed for 20% oil lms prepared by
spreading and spraying at the pressure of 2 bar (Table
1). Water vapor barrier properties of 20% oil lm were
signicantly (p < 0.05) lower at the atomization pressure
of 3.5 bar than at 2 bar. For starchMC lm, spraying
at 2 bar gave a signicantly lower (p < 0.05) WVP value
than spreading.
Tensile strength ranged from 13.22 0.22 MPa to
20.58 2.69 MPa, and elongation from 8.95 1.36%
to 25.49 5.82% (Table 1). TS values are in the same
range as for low-density polyethylene lm (8.3
27.6 MPa) (Arvanitoyannis et al., 1998; Park et al.,
1993).
In general, E showed constant values and TS was signicantly (p < 0.05) higher for spread samples than
sprayed samples (Table 1). High atomization pressure
was detrimental for TS of 20% oil lm, whereas it gave
the highest TS and lowest E in starchMC formulation.
From the WVP and tensile properties, it is clear that the
polysaccharidelipid lm was sensitive to spraying conditions. Preparation of lms by spreading did not involve mechanical stresses or shear rates as high as
those entailed by spraying (Barnes et al., 1989), which
probably induce irreversible structural changes of the
polymeric-lipid dispersion and deterioration of lm performances. Atomization pressure inuences the size of

284

B. Bravin et al. / Journal of Food Engineering 76 (2006) 280290

Table 1
Eect of deposition process on WVP and mechanical properties of
starchMC lm and starchMCoil lm
Film properties

No oil

With oil

WVP (1011 g m1 s1 g1)


Spreading
20.50 1.19a
Spraying (2 bar)
16.52 0.99b
Spraying (3.5 bar)
18.94 0.17a

11.68 0.38a
12.25 0.53ab
13.78 0.09b

Tensile strength (MPa)


Spreading
Spraying (2 bar)
Spraying (3.5 bar)

17.55 0.17a
14.24 0.91b
20.58 2.69c

17.99 0.85a
15.44 0.69b
13.22 0.22c

Elongation (%)
Spreading
Spraying (2 bar)
Spraying (3.5 bar)

20.13 0.46a
25.49 5.82a
11.88 2.55b

12.66 0.08a
8.95 1.36a
11.08 3.16a

Means standard deviations. For each lm properties mean values in


the same column with dierent letters are dierent (p < 0.05).

liquid droplets, which become smaller at higher atomization pressures (Burns & Fast, 1991; Dewettinck &
Huyghebaert, 1998). Deposition by spraying requires
dropletdroplet contact and aggregation, which is a critical step for homogeneous lm structure. ESEM observations of lm cross-sections were used to obtain more
information on the eect of deposition technique on lm
structure (Fig. 1). Spread samples exist as multilaminar
packed layers of starch and MC, probably covered with
oil. These layers are arranged perpendicular to the direction of vapor ow (Fig. 1c). Less compact structures and
large cavities within the matrix were observed in sprayed
oil lms at 3.5 bar (Fig. 1d). Aggregation was probably

more dicult for oil formulation because of the higher


surface tension of liquid droplets. These cavities could
explain the increase in WVP of oil lm due to preferential paths for water vapor diusion and the low TS for
structure discontinuities (Table 1).
Changes in the functional properties of lm with oil
obtained by atomization at 2 bar did not jeopardize its
performance with respect to layered lm. The application spraying to real products could be an attractive option as the technique oers the major advantage over
dipping of a more uniform coating, especially on nonregular surfaces.
3.2. Eect of relative humidity
Fig. 2 shows WVP as a function of relative humidity
inside the permeation cell for edible lms obtained by
spraying at the pressure of 2 bar. Water vapor barrier
properties decreased signicantly (p < 0.05) with the increase of RH, in agreement with previous studies (Cuq
et al., 1997; Fennema et al., 1994; McHugh et al.,
1994; McHugh & Krochta, 1994; Roy et al., 2000). In
contrast, WVP of apolar synthetic polymers such as
polyehtylene does not depend on RH gradient (Morillon
et al., 2002). Permeability is dened as the product of
permeate solubility coecient (S) in the lm and permeate diusion coecient (D) across the lm (Donhowe &
Fennema, 1994). Ideally, if no interaction occurs between a polymer lm and the permeating water vapor,
WVP is independent of RH gradient. Hydrophilic lms
deviate from ideal behavior because of permeatelm

Fig. 1. ESEM cross-section micrographs of starchMC lms (upper) and starchMCoil lms (lower) prepared by spreading (a, c) and spraying at
3.5 bar (b, d).

B. Bravin et al. / Journal of Food Engineering 76 (2006) 280290

285

our edible lm containing oil gave WVP slightly higher


than cellophane (5.6 1011 g m1 s Pa1) (Martin
Polo, Mauguin, & Voilley, 1992).

2E-10
1.8E-10

WVP (g m-1 s-1 Pa-1)

1.6E-10

3.3. Eect of thickness

1.4E-10
1.2E-10
1E-10
8E-11
no oil

6E-11

with oil

4E-11
60

65

70

75

80

85

90

95

Relative humidity (%)


Fig. 2. Water vapor permeability (WVP) as a function of relative
humidity inside the permeation cell for starchMC lm and starch
MCoil lm, prepared by spraying at 2 bar (thickness 30 lm). Values
are means and standard deviations are represented by bars.

interaction. The water adsorption isotherm of edible


lms showed that water solubility increased for high
aw values (Fig. 3). Kamper and Fennema (1984) and
Roy et al. (2000) reported that D also increased for
hydrophilic lms as the RH gradient applied across
the lm moved upwards in the RH spectrum. Hydration
causes matrix swelling, increased water mobility, and a
rise in the diusion coecient.
Although oil lm was less hygroscopic (Fig. 3),
starchMC and starchMCoil lms exhibited the same
trend of WVP versus RH (Fig. 2). Stading, RindlavWestling, and Gatenholm (2001) observed that amylose
and amylopectin lms conditioned to dierent aw gave
nonhomogeneous swelling of polymer network with
the formation of high porosity regions.
One interesting result is that, under the same conditions of RH gradient (2271%) and thickness (30 lm),

Edible lms with thickness varying from 30 to 90 lm


were prepared by spraying at 2 bar to investigate the effect of this variable on their functional properties. Figs.
4 and 5 show changes in WVTR and WVP as a function
of thickness. No eect on thickness was observed for
WVTR (Fig. 4). WVP values rose as lm thickness increased, according to other studies (Cuq et al., 1996;
McHugh et al., 1993) (Fig. 5). Ideal polymeric lms exhibit no thickness eect on water vapor barrier properties. Experimental WVP data and thickness for
polysaccharide and polysaccharide-oil lms tted well
with rst-order regression Eqs. (4) and (5), respectively
(r2 > 0.93; p < 105) (Fig. 5):
WVP 5.84  106  thickness 5.58  1012
WVP 3.77  10

6

 thickness 1.88  10

11

WVP of starchMC lm showed signicantly (p < 0.05)


higher dependence on thickness than lm containing oil.
Various hypotheses have been proposed to explain
the increase in WVP with increasing lm thickness.
Barrer (1951) and Banker, Gore, and Swarbrick (1966)
attributed the thickness eect to lm swelling as a result
of waterlm interaction. Recently, other Authors (Cuq
et al., 1996; McHugh et al., 1993) observed that as lm
thickness increased, the lm oered enhanced resistance
to water vapor transfer across it. In consequence, a stagnant air layer formed, characterized by high water vapor
partial pressure at the inner lm surface. Macromolecular swelling at the stagnant air gaplm interface would
determine higher water diusion because of increasing
3000

0.028

2700

0.021

2400

0.014

2100

0.007

50

p1 (Pa)

no oil
with oil

40
30

WVTR (g m-2 s-1)

Moisture content (g / 100 g dry basis)

60

20
10

1800
0

0
0.1

0.00002

0.00004

0
0.00006

Thickness (m)
0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

Water activity
Fig. 3. Water adsorption isotherms of starchMC lm and starch
MCoil lm prepared by spraying at 2 bar (thickness 30 lm). Values
are means and standard deviations are represented by bars.

Fig. 4. Water vapor transfer rate (WVTR) and water vapor partial
pressure (p1) as a function of thickness for lms prepared by spraying
at 2 bar. WVTR: starchMC lm () and starchMCoil lm (); p1:
starchMC lm (D) and starchMCoil lm (d) (theoretical RH
gradient 22100%).

286

B. Bravin et al. / Journal of Food Engineering 76 (2006) 280290

Film thickness of 30 lm was identied as the most


appropriate to ensure low WVP and desirable mechanical properties in our edible lms.

3.2E-10

WVP (g m-1 s-1 Pa-1)

2.8E-10
2.4E-10

3.4. Eect of edible coating on crackers


2E-10
1.6E-10
1.2E-10
8E-11

with oil

4E-11

no oil

0.00002

0.00004

0.00006

Thickness (m)
Fig. 5. Water vapor permeability (WVP) as a function of thickness for
starchMC lm and starchMCoil lm prepared by spraying at 2 bar
(theoretical RH gradient 22100%).

mobility of polymeric chains (Cuq et al., 1996; McHugh


et al., 1993). Fig. 4 shows that the actual water vapor
partial pressure (p1) inside the permeation cell was not
inuenced by thickness in our experimental work.
Structural modications in the lm related to changes
in thickness could explain our results (Cuq et al.,
1996; Hauser & McLaren, 1948). Thickness aects the
drying kinetic of the liquid lm-forming dispersion,
which may cause dierences in lm structure (Debeaufort & Voilley, 1995). TS decreased with the increase
in thickness, conrming the hypothesis of the presence
of dierent structures (Fig. 6). Higher TS values were
observed for low thickness associated with high drying
rates. For polysaccharide-oil lm, high drying times
could favor coalescence of oil droplets, which generate
discontinuities inside the lm (Debeaufort & Voilley,
1995).

The eciency of polysaccharidelipid-based edible


coating in controlling moisture transfer in moisture-sensitive products was evaluated by coating crackers, a low
aw-type cereal food. Texture is a primary sensory attribute for these products, and loss of the desired texture
(crispness) results in shorter shelf-life.
Taking the above results into account, atomization
pressure of 2 bar and lm thickness of 30 lm were identied as optimum for the application of edible coating to
crackers. The choice of this product was suggested by its
regular, constant geometry, which enables uniform coating and control over exchange surfaces. Coating did not
modify cracker appearance.
Fig. 7 shows the relationship between the aw and
equilibrium moisture content of crackers at 25 C. The
BET model was used to t experimental data of moisture adsorption isotherm in the 0.050.44 aw range:
aw
1
c  1  aw

1  aw  m c  m0
c  m0

where m is the moisture content (g/100 g dry basis); m0 is


the monolayer moisture content (g/100 g dry basis); and
k and c are constants. BET model parameters and the
coecient of correlation for coated product and reference are given in Table 2. Untreated cracker had slightly
lower equilibrium moisture content until 0.44 aw, indicating the less hygroscopic nature of the reference.
Monolayer values were found to be 5.58 and 6.20
g/100 g dry basis for reference and coated product,
respectively (Table 2). Katz and Labuza (1981) and Tubert and Iglesias (1986) reported m0 values of 3.78
5.53 g/100 g dry basis at 20 C for dierent types of

25

Tensile strength (MPa)

Moisture content (g / 100 g dry basis)

24
no oil

20

with oil
15

10

0.00003

0.00006

0.00009

0.00012

Thickness (m)
Fig. 6. Tensile strength as a function of thickness for starchMC lm
and starchMCoil lm prepared by spraying at 2 bar.

20
untreated
16

coated

12
8
4
0
0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

Water activity
Fig. 7. Water adsorption isotherms for coated and untreated crackers.
Values are means and standard deviations are represented by bars.

B. Bravin et al. / Journal of Food Engineering 76 (2006) 280290

Model parameters
and R2

Untreated
cracker

Coated
cracker

BET (aw < 0.44)

c
m0
R2

1.119
5.58
0.97

1.106
6.20
0.99

cereal crackers, in line with to our results. As expected,


the initial moisture content of crackers was below the
monolayer value (Fig. 7), which may be considered as
a critical moisture content because it is related to the
physical deterioration of dehydrated foods (Bell & Labuza, 2000).
Fig. 8 shows the initial slope of the forcedeformation curve for coated and untreated crackers. Although
sensory analysis gives a more complete description of
the texture, this parameter can be used as indicator of
cracker crispness, and is validated by sensory data (Katz
& Labuza, 1981). Coating did not show any eect on initial slope (Fig. 8). No signicant changes in initial slope
were observed in the aw range 0.050.33, according to
Katz and Labuza (1981) (p > 0.05) (Fig. 8). For higher
aw values, the mechanical parameter decreased. The
drop is associated with a loss of crispness due to hydration, which induces glass transition in polymer amorphous regions (Roudaut et al., 2002). The Tg of
crackers was found to range from 115 C at 3.14
g/100 g dry-basis moisture content to 12 C at 14.02
g/100 g dry basis (Nikolaidis & Labuza, 1996). Water
plasticization transfers the material from the glassy state
(crisp texture) to the rubbery state, resulting in sogginess. Crispness was lost as the Tg was depressed below
ambient temperature because of water plasticization.
The loss in this texture attribute is the major cause of
dry food rejection by consumers.

18
16
14
12

14

Moisture (g / 100 g dry basis)

Model

A critical value for aw (awc) was estimated on the


basis of a statistical test involving the determination of
the rst aw related to crispness intensity signicantly
(p < 0.05) lower than the crispness of the sample at aw
0.050.11 (Fig. 8), according to Sauvageot and Blond
(1991). The awc value is the water activity which becomes before the aw previously determined. Coated
cracker and the reference showed awc of 0.33, corresponding to critical moisture contents (mc) respectively,
of 8.25 and 6.64 g/100 g dry basis (Fig. 7). Prior sensory
assessment of the saltine cracker reported awc of 0.39
and mc of 7.00 g/100 g dry basis (Katz & Labuza,
1981). Values of mc were higher than BET monolayer
values (Table 2). Identication of awc and mc are very
useful in determining product shelf-life.
Figs. 9 and 10 show the evolution of moisture content
with storage time of coated and untreated crackers at

Moisture content ( g / 100 g dry basis)

Table 2
BET model parameters and coecient of determination for untreated
and coated crackers

287

10
8
6
4

coated

2
0

12
10
mc coated
mc untreated

8
6
4

2
0

10

untreated
0

96

192

288

20
30
40
Storage time (hours)

384

50

60

480

576

Storage time (hours)


Fig. 9. Moisture content evolution with storage time of coated and
untreated crackers stored at 75% RH. (a) The inset illustrates
estimation of shelf-life using mc values. Values are means and standard
deviations are represented by bars.

70

Initial slope (N/mm)

60
coated
untreated

50
40
30
20
10
0

20

15

10

coated
untreated

Moisture (g / 100 g dry basis)

Moisture content ( g/ 100 g dry basis)

25

18
16
14
12
10
8
6
4
2
0

mc coated
mc untreated

a
0

20
40
Storage time (hours)

60

0
0
0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

Water activity
Fig. 8. Initial slope of the forcedeformation curve as a function of
water activity for coated and untreated crackers. Values are means and
standard deviations are represented by bars.

96

192

288

384

480

576

Storage time (hours)


Fig. 10. Moisture content evolution with storage time of coated and
untreated crackers stored at 85% RH. (a) The inset illustrates
estimation of shelf-life using mc values. Values are means and standard
deviations are represented by bars.

288

B. Bravin et al. / Journal of Food Engineering 76 (2006) 280290

0.05 initial aw, stored at 75% and 85% RH, respectively.


The same trend was observed at 65% RH (data not
shown). Since RH is a factor that aects both crispness
loss and the performances of edible coating, it was introduced as a variable in this work. Moisture content
showed a sharp increase over the rst 50 h. Subsequently,
the trend was towards constant value (Figs. 9 and 10).
Reference needed about 288 h to equilibrate with the surroundings, whereas the coated product did not reach
equilibrium with the controlled ambient humidity in the
time scale of experimental work. The kinetic of moisture
uptake was lower for coated crackers. Coating delayed
water vapor transfer from the controlled environment
and the dry bakery product. However, this barrier is
not total. Previous studies (Avena-Bustillos et al., 1994;
Fennema, Donhowe, & Kester, 1993; Garca et al.,
1998; Park, Chinnan, & Shewfelt, 1994; Yaman &
Bayoindirli, 2002) reported that edible coatings are not
completely impervious to moisture.
Water vapor resistance for coated cracker was higher
than the value for the untreated sample at all RH
conditions (Fig. 11). Table 3 shows results from the
analysis of the eects of coating, RH, and their interaction (coatingRH) on water vapor resistance of coated

600

Table 4
Shelf-life of coated and untreated crackers stored in dierent relative
humidity conditions
Relative humidity (%)

65
75
85

Shelf-life (h)
Untreated cracker

Coated cracker

27
12
10

42
32
21

cracker. Both independent variables and their interaction had a signicant eect on r. The signicant eect
of RH is in accordance with the inuence of RH gradient on lm WVP (Fig. 2). The presence of the coating
RH interaction explains why r values of coated crackers
are less sensitive to RH changes than the untreated
product.
Table 4 shows the shelf-life (h) of crackers stored at
dierent RH, estimated from moisture content evolution
(Figs. 9a and 10a) as the time to reach mc value. The
coated crackers exhibited higher shelf-life than reference. At 85% RH, coating doubled the shelf-life of
bakery products. Results clearly suggest that polysaccharidelipid-based edible coating was successful in
delaying moisture uptake in a low aw food.

4. Conclusions
untreated

Resistance (s cm-1)

500

coated

400
300
200
100
0

65

75

85

Relative humidity (%)


Fig. 11. Water vapor resistance for coated and untreated crackers
stored at dierent relative humidity conditions. Values are means and
standard deviations are represented by bars.

Table 3
Variance (F-value) of coating, relative humidity (RH) and their
interaction (coatingRH) on water vapor resistance of crackers
Independent variable

Coating
RH
Interaction: coatingRH

253.1*
28.3**
5.6***

*
**
***

Signicant at p < 105.


Signicant at p < 103.
Signicant at p < 0.05.

The development and optimization of an edible lm


with water vapor barrier properties was undertaken by
investigating in detail factors associated with deposition
of uid-state dispersion, lm thickness and the relative
humidity gradient across the lm. Deposition techniques
for spreading and spraying of the lm-forming dispersion are characterized by very dierent strains and strain
rates, yet enable the acquisition of comparable water
vapor barriers through the control of operating conditions. Atomization pressure is a critical parameter,
which it is important to maintain at values below
3.5 bar to avoid excessive destructuring of the lm-forming system and negative repercussions on the barrier
properties of the lm. High lm thickness produces an
increase in WVP, so control of this parameter is crucial.
Application of an edible coating to crackers, a low aw
food, conrmed the potential of edible packaging to become an integral part of the food, and reduce the hydration kinetic in a high aw environment.
Acknowledgements
This work was supported by a grant from MURST
(Ministero dellUniversita` e della Ricerca Scientica e
Tecnologica): Piano Nazionale di Ricerca per il settore
Agroalimentare-tema 9 Innovazione di processo per
la produzione di alimenti disidratati.

B. Bravin et al. / Journal of Food Engineering 76 (2006) 280290

References
AACC (1995). Approved methods of the AACC (American Association
of Cereal Chemists, 9th ed.). S. Paul, Minnesota: The Association.
Arvanitoyannis, I., & Biliaderis, C. G. (1998). Physical properties of
polyol-plasticized edible lms made from sodium caseinate and
soluble starch blends. Food Chemistry, 62, 333342.
Arvanitoyannis, I., Biliaderis, C. G., Ogawa, H., & Kawasaki, N.
(1998). Biodegradable lms made from low-density polyethylene
(LDPE), rice starch and potato starch for food packaging
applications: Part 1. Carbohydrate Polymers, 36, 89104.
Arvanitoyannis, I., Psomiadou, E., & Nakayama, A. (1996). Edible
lms made from sodium caseinate, starches, sugars or glycerol.
Part 1. Carbohydrate Polymers, 31, 179192.
ASTM D 882-88 (1989). Standard test methods for tensile properties
of thin plastic sheeting. In Annual book of ASTM standards.
Philadelphia, Pennsylvania: American Society for Testing and
Materials.
ASTM E 96-80 (1990). Standard test method for water vapour
transmission of materials. In Annual book of ASTM standards.
Philadelphia, Pennsylvania: American Society for Testing and
Materials.
Avena-Bustillos, R. J., Krochta, M. J., Saltveit, M. E., Rojas-Villegas,
R. J., & Sauceda-Perez, J. A. (1994). Optimization of edible coating
formulations on zucchini to reduce water loss. Journal of Food
Engineering, 21, 197214.
Banker, G. S., Gore, A. Y., & Swarbrick, J. (1966). Water vapour
transmission properties of free polymer lms. Journal of Pharmacy
and Pharmacology, 18, 457466.
Barnes, H. A., Hutton, J. F., & Walters, K. F. R. S. (1989). An
introduction to rheology (1st ed., pp. 1213). Amsterdam, The
Netherlands: Elsevier Science Publishers B.V.
Barrer, R. M. (1951). Permeation of vapours through, and diusion
in, organic solid. In R. M. Barrer (Ed.), Diusion in and through
solids (pp. 430453). Cambridge, United Kingdom: University
Press.
Bell, L. N., & Labuza, T. P. (2000). Moisture sorption: Practical
aspects of isotherm measurement and use (2nd ed., pp. 5769). St.
Paul, Minnesota: Eagan Press, American Association of Cereal
Chemistry, Inc.
Ben-Yehoshua, S., Burg, S. P., & Young, R. (1985). Resistance of
citrus fruit to mass transport of water vapor and other gases. Plant
Physiology, 79, 10481053.
Bravin, B., Peressini, D., & Sensidoni, A. (2004). Inuence of
emulsier type and content on functional properties of polysaccharide lipid-based edible lms. Journal of Agricultural and Food
Chemistry, 52, 64486455.
Burns, R. E., & Fast, R. B. (1991). Design and operation of equipment
for coating breakfast cereals. Cereal Foods World, 36, 879881,
884887.
Champion, D., Le Meste, M., & Simatos, D. (2000). Towards an
improved understanding of glass transition and relaxations in
foods: molecular mobility in the glass transition range. Trends in
Food Science and Technology, 11, 4155.
Cuq, B., Gontard, N., Aymard, C., & Guilbert, S. (1997). Relative
humidity and temperature eects on mechanical and water vapor
barrier properties of myobrillar protein-based lms. Polymer Gels
and Networks, 5, 115.
Cuq, B., Gontard, N., Cuq, J. L., & Guilbert, S. (1996). Functional
properties of myobrillar protein-based biopackaging as aected
by lm thickness. Journal of Food Science, 61, 580584.
Debeaufort, F., & Voilley, A. (1995). Eect of surfactants and drying
rate on barrier properties of emulsied edible lms. International
Journal of Food Science and Technology, 30, 183190.
Debeaufort, F., & Voilley, A. (1997). Methylcellulose-based edible
lms and coatings: 2. Mechanical and thermal properties as a

289

function of plasticizer content. Journal of Agricultural and Food


Chemistry, 45, 685689.
Dewettinck, K., & Huyghebaert, A. (1998). Top-spray uidized bed
coating: eect of process variables on coating eciency. Lebensmittel-Wissenschaft und Technologie, 31, 568575.
Donhowe, I. G., & Fennema, O. (1993). The eects of plasticizers on
crystallinity, permeability, and mechanical properties of methylcellulose lms. Journal of Food Processing and Preservation, 17,
247257.
Donhowe, I. G., & Fennema, O. (1994). Edible lms and coatings: Characteristics, formation, denitions, and testing methods.
In J. M. Krochta, E. A. Baldwin, & M. O. Nisperos-Carriedo
(Eds.), Edible coatings and lms to improve food quality
(pp. 124). Lancaster, Pennsylvania: Technomic Publishing Co.,
Inc.
Fennema, O., Donhowe, I. G., & Kester, J. J. (1993). Edible lms:
barriers to moisture migration in frozen foods. Food Australia, 45,
521525.
Fennema, O., Donhowe, I. G., & Kester, J. J. (1994). Lipid type and
location of the relative humidity gradient inuence on the barrier
properties of lipids to water vapor. Journal of Food Engineering, 22,
225239.
Garca, M. A., Martino, M. N., & Zaritzky, N. E. (1998). Plasticized
starch-based coatings to improve strawberry (Fragaria x Ananassa)
quality and stability. Journal of Agricultural and Food Chemistry,
46, 37583767.
Garca, M. A., Martino, M. N., & Zaritzky, N. E. (2000). Lipid
addition to improve barrier properties of edible starch-based lms
and coatings. Journal of Food Science, 65, 941947.
Gennadios, A., McHugh, T. H., Weller, C. L., & Krochta, J. M.
(1994). Edible coatings and lms based on proteins. In J. M.
Krochta, E. A. Baldwin, & M. O. Nisperos-Carriedo (Eds.), Edible
coatings and lms to improve food quality (pp. 201277). Lancaster,
Pennsylvania: Technomic Publishing Co., Inc.
Gennadios, A., Weller, C. L., & Gooding, C. H. (1994). Measurement errors in water vapor permeability of highly permeable,
hydrophilic edible lms. Journal of Food Engineering, 21,
395409.
Gontard, N., Guilbert, S., & Cuq, J. L. (1992). Edible wheat gluten
lms: inuence of the main process variables on lm properties
using response surface methodology. Journal of Food Science, 57,
190195, 199.
Gontard, N., Guilbert, S., & Cuq, J. L. (1993). Water and glycerol as
plasticizers aect mechanical and water vapour barrier properties
of an edible wheat gluten lm. Journal of Food Science, 58,
206211.
Gontard, N., Thibault, R., Cuq, B., & Guilbert, S. (1996). Inuence of
the relative humidity and lm composition on oxygen and carbon
dioxide permeabilities of edible lms. Journal of Agricultural and
Food Chemistry, 44, 10641069.
Hauser, P. M., & McLaren, A. D. (1948). Permeation through and
sorption of water vapor by high polymers. Industrial and Engineering Chemistry Research, 40, 112117.
Hernandez, E. (1994). Edible coatings from lipids and resins. In J. M.
Krochta, E. A. Baldwin, & M. O. Nisperos-Carriedo (Eds.), Edible
coatings and lms to improve food quality (pp. 279303). Lancaster,
Pennsylvania: Technomic Publishing Co., Inc.
Kamper, S. L., & Fennema, O. (1984). Water vapor permeability of an
edible, fatty acid, bilayer lm. Journal of Food Science, 49,
14821485.
Kamper, S. L., & Fennema, O. (1985). Use of edible lm to maintain
water vapor gradients in foods. Journal Food Science, 50, 382
384.
Katz, E. E., & Labuza, T. P. (1981). Eect of water activity on the
sensory crispness and mechanical deformation of snack food
products. Journal of Food Science, 46, 403409.

290

B. Bravin et al. / Journal of Food Engineering 76 (2006) 280290

Kester, J. J., & Fennema, O. (1989). An edible lm of lipids and


cellulose ethers: performance in a model frozen-food system.
Journal of Food Science, 54, 13901392, 1406.
Koelsch, C. M., & Labuza, T. P. (1992). Functional, physical and
morphological properties of methylcellulose and fatty acid-based
edible barriers. Lebensmittel-Wissenschaft und Technologie, 25,
404411.
Labuza, T. P., & Hyman, C. R. (1998). Moisture migration and
control in multi-domain foods. Trends in Food Science and
Technology, 9, 4755.
Martin Polo, M., Mauguin, C., & Voilley, A. (1992). Hydrophobic
lms and their eciency against moisture transfer. 1. Inuence of
lm preparation technique. Journal of Agricultural and Food
Chemistry, 40, 407412.
McHugh, T. H., Aujard, J. F., & Krochta, J. M. (1994). Plasticized
whey protein edible lms: water vapor permeability properties.
Journal of Food Science, 59, 416419, 423.
McHugh, T. H., Avena-Bustillos, R., & Krochta, J. M. (1993).
Hydrophilic edible lms: modied procedure for water vapor
permeability and explanation of thickness eects. Journal of Food
Science, 58, 899903.
McHugh, T. H., & Krochta, J. M. (1994). Water vapor permeability
properties of edible whey proteinlipid emulsion lms. Journal of
the American Oil Chemists Society, 71, 307312.
Morillon, V., Debeaufort, F., Blond, G., Capelle, M., & Voilley, A.
(2002). Factors aecting the moisture permeability of lipid-based
edible lms: a review. Critical Reviews in Food Science and
Nutrition, 42(1), 6789.
Nikolaidis, A., & Labuza, T. P. (1996). Glass transition state diagram
of a baked cracker and its relationship to gluten. Journal of Food
Science, 61, 803806.
Nisperos-Carriedo, M. O. (1994). Edible coatings and lms based on
polysaccharides. In J. M. Krochta, E. A. Baldwin, & M. O.
Nisperos-Carriedo (Eds.), Edible coatings and lms to improve food
quality (pp. 305335). Lancaster, Pennsylvania: Technomic Publishing Co., Inc.
Park, H. J., Chinnan, M. S., & Shewfelt, R. L. (1994). Edible coating
eects on storage life and quality of tomatoes. Journal of Food
Science, 59, 568570.
Park, H. J., Weller, C. L., Vergano, P. J., & Testin, R. F. (1993).
Permeability and mechanical properties of cellulose-based edible
lms. Journal of Food Science, 58, 13611364, 1370.

Peressini, D., Bravin, B., Lapasin, R., Rizzotti, C., & Sensidoni, A.
(2003). Starchmethylcellulose based edible lms: rheological
properties of lm-forming dispersions. Journal of Food Engineering,
59, 2532.
Peressini, D., Bravin, B., & Sensidoni, A. (2004). Tensile properties,
water vapour permeabilities and solubilities of starchmethylcellulose-based edible lms. Italian Journal of Food Science, 16,
516.
Psomiadou, E., Arvanitoyannis, I., & Yamamoto, N. (1996). Edible
lms made from natural resources; microcrystalline cellulose
(MCC), methylcellulose (MC) and corn starch and polyols. Part
2. Carbohydrate Polymers, 31, 193204.
Rico-Pena, D. C., & Torres, J. A. (1990). Edible methylcellulose-based
lms as moisture-impermeable barriers in sundae ice cream cones.
Journal of Food Science, 55, 14681469.
Roudaut, G., Dacremont, C., Valle`s Pa`mies, B., Colas, B., & Le
Meste, M. (2002). Crispness: a critical review on sensory and
material science approaches. Trends in Food Science and Technology, 13, 217227.
Roy, S., Gennadios, A., Weller, C. L., & Testin, R. F. (2000). Water
vapor transport parameters of a cast wheat gluten lm. Industrial
Crops and Products, 11, 4350.
Sauvageot, F., & Blond, G. (1991). Eect of water activity on crispness
of breakfast cereals. Journal of Texture Studies, 22, 423442.
Shellhammer, T. H., & Krochta, J. M. (1997). Whey protein emulsion
lm performances as aected by lipid type and amount. Journal of
Food Science, 62, 390394.
Stading, M., Rindlav-Westling, A., & Gatenholm, P. (2001). Humidity-induced structural transitions in amylose and amylopectin lms.
Carbohydrate Polymers, 54, 209217.
Tubert, A. H., & Iglesias, H. A. (1986). Water sorption isotherms and
prediction of moisture gain during storage of packaged cereal
crackers. Lebensmittel-Wissenschaft und Technologie, 19, 365
368.
Vincent, J. F. V. (1998). The quantication of crispness. Journal of the
Science of Food and Agriculture, 78, 162168.
Warburton, S. C., Donald, A. M., & Smith, A. C. (1993). The
deformation of thin lms made from extruded starch. Carbohydrate
Polymers, 21, 1721.
., & Bayoindirli, L. (2002). Eects of an edible coating and
Yaman, O
cold storage on shelf-life and quality of cherries. LebensmittelWissenschaft und Technologie, 35, 146150.

Das könnte Ihnen auch gefallen