You are on page 1of 50

Accepted Manuscript

Polysaccharide-based films and coatings for food packaging: A review

Patricia Cazón, Gonzalo Velazquez, José A. Ramírez, Manuel Vázquez

PII: S0268-005X(16)30415-5
DOI: 10.1016/j.foodhyd.2016.09.009
Reference: FOOHYD 3584

To appear in: Food Hydrocolloids

Received Date: 3 July 2016


Revised Date: 5 September 2016
Accepted Date: 7 September 2016

Please cite this article as: Cazón, P., Velazquez, G., Ramírez, J.A., Vázquez, M., Polysaccharide-
based films and coatings for food packaging: A review, Food Hydrocolloids (2016), doi: 10.1016/
j.foodhyd.2016.09.009.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT
1 Polysaccharide-based films and coatings for food packaging: a review

3 Patricia Cazóna,c, Gonzalo Velazqueza, José A. Ramírezb, Manuel Vázquezc,*

4
a
5 Instituto Politécnico Nacional. CICATA unidad Querétaro. Cerro Blanco No. 141.

PT
6 Colinas del cimatario, Querétaro, Qro. 76090, México.
b

RI
7 Unidad Académica de Trabajo Social y Ciencias para el Desarrollo Humano.

8 Universidad Autónoma de Tamaulipas. Centro Universitario. Ciudad Victoria,

SC
9 Tamaulipas, 87149, México
c
10 Department of Analytical Chemistry, Faculty of Veterinary Science, University of

11
U
Santiago de Compostela, Campus Lugo, 27002-Lugo, Spain
AN
12
M

13 *Corresponding Author: E-mail: manuel.vazquez@usc.es. phone

14 +34982822420; fax +34982254592


D

15
TE

16

17 ABSTRACT
EP

18 The accumulation of synthetic plastics, mainly from food packaging, is causing

19 a serious environmental problem. It is driving research efforts to the development of


C
AC

20 biodegradable films and coatings. The biopolymers used as raw material to prepare

21 biodegradable films should be renewable, abundant and low-cost. In some cases, they

22 can be obtained from wastes. This review summarizes the advances in polysaccharide-

23 based films and coatings for food packaging. Among the materials studied to develop

24 biodegradable packaging films and coatings are polysaccharides such as cellulose,

25 chitosan, starch, pectin and alginate. These polysaccharides are able to form films and

1
ACCEPTED MANUSCRIPT
26 coatings with good barrier properties against the transport of gases such as oxygen and

27 carbon dioxide. On the other hand, tensile strength and percentage of elongation are

28 important mechanical properties. Desirable values of them are required to maintain the

29 integrity of the packed food. The tensile strength values showed by polysaccharide-

30 based films vary from each other, but some of them exhibit similar values to those

PT
31 observed in synthetic polymers values. For example, tensile strength values of films

RI
32 based on high amylose starch or chitosan are comparable to those values found in high-

33 density polyethylene films. The values of percentage of elongation are the main

SC
34 concern, which are far from the desirable values found for synthetic polymers.

35 Researchers are studying combinations of polysaccharides with other materials to

36
U
improve the barrier and mechanical properties in order to obtain biopolymers that could
AN
37 replace synthetic polymers. Functional polymers with antimicrobial properties, as that
M

38 the case of chitosan, are also being studied.

39
D

40 Keywords: Cellulose; chitosan; starch; water vapour permeability; tensile strength;


TE

41 percentage of elongation at break.

42
EP

43 1. Introduction

44
C
AC

45 The production and applications of synthetic films in food packaging have

46 grown quickly over the past few decades resulting in serious environmental concerns

47 because the synthetic plastics are resistant to degradation (Muscat, Adhikari, Adhikari,

48 & Chaudhary, 2012). Nowadays, consumers seek to reduce the environmental problems

49 associated with food packaging and demand biodegradable materials. Research studies

50 have been focused on biopolymeric materials as raw material for food packaging and

2
ACCEPTED MANUSCRIPT
51 preservation (Persin et al., 2011). Edible and biodegradable films could be an alternative

52 to synthetic packaging materials in several applications due to their capabilities to

53 prevent moisture loss, aromas loss, solute transport, water absorption in the food matrix

54 or oxygen penetration (Aider, 2010; Dutta, Tripathi, Mehrotra, & Dutta, 2009). For this

55 reason, food scientists and engineers are trying to develop new materials for edible and

PT
56 biodegradable films, mainly based on materials from renewable sources that are

RI
57 abundant in nature. In general, these materials are cheap and many of them are

58 considered as waste or by-products (Kim, Son, Kim, Weller, & Hanna, 2006). The

SC
59 applications of polysaccharide-based films in food products could offer new

60 opportunities to develop novel food packaging systems. Also, biodegradable films can

61
U
reduce environmental problems associated with food packaging.
AN
62 Currently, the applications of biodegradable polymers in the food industry
M

63 include disposable cutlery, drinking cups, lids, plates, overwrap and lamination films,

64 straws, stirrers, containers for food dispensed at gourmet food stores and fast-food
D

65 establishments (Siracusa, Rocculi, Romani, & Rosa, 2008). The materials obtained from
TE

66 renewable natural sources that have been used to produce edible film include mainly

67 proteins, lipids, polysaccharides and all possible combinations among them and
EP

68 sometimes, incorporation of additives aimed to improve the properties of the film.

69 The film-forming ability of several polysaccharides has been studied, including


C
AC

70 cellulose, chitosan, starch, pectin, alginate, carrageenan, pullulan and kefiran (Maher Z.

71 Elsabee & Entsar S. Abdou, 2013; Espitia, Du, Avena-Bustillos, Soares, & McHugh,

72 2014; Galus & Kadzińska, 2015; Jiménez, Fabra, Talens, & Chiralt, 2012; Motedayen,

73 Khodaiyan, & Salehi; Qiu & Hu, 2013; Xu, Chen, Rosswurm, Yao, & Janaswamy,

74 2016; Zolfi, Khodaiyan, Mousavi, & Hashemi, 2014).

3
ACCEPTED MANUSCRIPT
75 Lipid-based materials include waxes, acylglycerols, and fatty acids (Pérez-Gago

76 & Rhim, 2014). Protein materials include collagen, caseins, fish gelatin, quinoa protein,

77 whey protein, corn zein, wheat gluten, egg white protein, myofibrillar protein, soy

78 protein and keratin (Fakhreddin Hosseini, Rezaei, Zandi, & Ghavi, 2013; Galus &

79 Kadzińska, 2015; Hernandez‐Izquierdo & Krochta, 2008; Jia, Fang, & Yao, 2009;

PT
80 Valenzuela, Abugoch, & Tapia, 2013).

RI
81 In most of the cases, the addition of plasticizers is required in order to obtain

82 protein- and polysaccharides-based films. With no plasticizer added, the films obtained

SC
83 from several polysaccharides are brittle usually due to interactions between polymer

84 chains (Han, 2014). Plasticizers reduce cohesion within the film network by weakening

85
U
the intermolecular forces between adjacent polymer chains (Espitia et al., 2014; Marcos,
AN
86 Aymerich, Monfort, & Garriga, 2010). In this way, plasticizers modify or improve the
M

87 mechanical properties, reduce the tension of deformation, hardness, density and

88 viscosity and increase the polymer chain flexibility as well as the resistance to fracture
D

89 (Vieira, da Silva, dos Santos, & Beppu, 2011). In hydrophilic materials, plasticizers
TE

90 modify the barrier properties and usually increase the water vapour permeability. In

91 contrast, the addition of hydrophobic plasticizers could result in a decrease in water


EP

92 uptake (Espitia et al., 2014; Han, 2014; Vieira et al., 2011).

Some plasticizers commonly used are polyols (glycerol, sorbitol, and


C

93
AC

94 polyethylene glycol), sugars (glucose and sucrose), and lipids (monoglycerides,

95 phospholipids and surfactants) (Espitia et al., 2014). The use of natural plasticizers, like

96 triglycerides from vegetable oils or fatty acid esters, is increasing because they have low

97 toxicity and low migration (Vieira et al., 2011). This review summarizes the advances

98 in polysaccharide-based films and coatings for food packaging.

99

4
ACCEPTED MANUSCRIPT
100 2. Polysaccharides for biodegradable films

101 A biodegradable film could be defined as a primary packaging made from

102 biodegradable polymers and food-grade additives. A thin layer of biodegradable

103 material can be formed into a film and might be used as a food wrap without changing

104 the original ingredients or the processing method. Biodegradable films have been used

PT
105 for protection and to extend the shelf life of several food products (Galus & Kadzińska,

RI
106 2015; Han, 2014).

107 For film preparation, the raw materials must be first dissolved or dispersed using

SC
108 a solvent such as water, alcohol, mixture of water and alcohol or a mixture of other

109 solvents. Plasticizers, antimicrobial agents, colouring or flavouring agents can be added

110
U
in this process. Adjusting the pH and/or heating the solutions may be necessary to
AN
111 facilitate the solubility of some biopolymers. Then, the film-forming solution is cast and
M

112 dried at desired temperature and relative humidity conditions to obtain free-standing

113 films. As a packaging material, film-forming solutions could be applied to food as


D

114 coating by several methods including dipping, spraying, brushing and panning followed
TE

115 by drying (Bourtoom, 2008).

116 Polysaccharides such as cellulose derivatives, chitosan, starches and gums have
EP

117 been reported as raw material to prepare edible films and coatings that can be used as

118 packaging material for food preservation (Aider, 2010; M. Z. Elsabee & E. S. Abdou,
C
AC

119 2013; Espitia et al., 2014; Han, 2014; Jiménez et al., 2012; Klemm, Heublein, Fink, &

120 Bohn, 2005; Psomiadou, Arvanitoyannis, & Yamamoto, 1996; Vargas & Gonzalez-

121 Martinez, 2010; Zhang, Han, & Liu, 2008). The polysaccharides-based packaging

122 materials must have low cost, availability and some functional or specific properties.

123 The water barrier efficiency of films is important to retard the dehydration of

124 fresh products, or on the other hand, the loss of crispness in dry products when the

5
ACCEPTED MANUSCRIPT
125 packaging material is used to avoid moisture absorption. The permeability to other

126 gases like oxygen is also important because it allows controlling the ripening of fruits or

127 reducing the oxidation of some foods components like polyunsaturated fats (Dutta et al.,

128 2009; Vieira et al., 2011).

129 The biodegradable films based on polysaccharides are known to be an effective

PT
130 barrier to gas transference like O2 and CO2 (Bertuzzi, Castro Vidaurre, Armada, &

RI
131 Gottifredi, 2007) although these materials are generally very hydrophilic resulting in

132 poor water vapour barrier properties.

SC
133 Although some polysaccharides have poor water vapour barrier properties i.e.

134 alginate and carrageenan, they are highly hygroscopic and can be applied as relatively

135
U
thick films on the surface of food in order to absorb water providing temporary
AN
136 protections to further moisture loss from the food. In this way, the product does not lose
M

137 significant moisture until the coating is dehydrated (Bourtoom, 2008; Embuscado &

138 Huber, 2009).


D

139 Moreover, the gas barrier properties of the polysaccharides-based films are
TE

140 important to retard loss of organic vapours (aroma compounds) during storage or to

141 prevent solvent penetration in foods, which could result in toxicity or quality loss (Dutta
EP

142 et al., 2009; Vieira et al., 2011)

143 The types of polysaccharide and the wide range of additives that could be
C
AC

144 incorporated during the preparation of biodegradable films could allow enhancing the

145 shelf-life of ready-to-eat foods.

146

147 2.1. Cellulose

148 Cellulose constitutes the most abundant renewable polymer resource available in

149 nature and it has been widely reported as a raw material for biodegradable films mainly

6
ACCEPTED MANUSCRIPT
150 because of its renewability, low cost, non-toxicity, biocompatibility, biodegradability

151 and chemical stability (Wang, Lu, & Zhang, 2016). Cellulose can be isolated from

152 wood, cotton, hemp and plant-based materials as well as synthesized by tunicates and

153 microorganisms (Xu et al., 2016). These characteristics make cellulose fibres the most

154 common choice for natural fillers in plastic materials (Kolybaba et al., 2006).

PT
155 Cellulose is a linear chain with two anhydroglucose rings ((C6H10O5)n),

covalently linked through an oxygen in a β1-4 glycosidic bond. The number of repeated

RI
156

157 units per chain depends on the source (Wang et al., 2016).

SC
158 Cellulose is insoluble in polar solvents, but it is soluble in a few solvents which

159 have no similar chemical properties. The insolubility in water has been the main focus

160
U
of several investigations and so far there are two main hypotheses, one of them states
AN
161 that the insolubility of cellulose is due to the relatively high length of the cellulose
M

162 molecular chain and its close packing through numerous hydrogen bonds (Wang et al.,

163 2016). Other authors explain the insolubility taking into account the crystallinity of
D

164 cellulose (Lindman, Karlström, & Stigsson, 2010).


TE

165 As cellulose is a potential raw material for biodegradable films, several methods

166 have been developed to dissolve it as a previous step in the process of film production.
EP

167 Some of these methods use solvents like N-methylmorpholine N-oxide, ionic liquids,

168 LiCl/N,N-dimethylacetamide, NaOH aqueous solution, alkali/urea, NaOH/thiourea


C
AC

169 aqueous solution, tetra butyl ammonium fluoride/dimethyl sulfoxide system, molten

170 inorganic salt hydrates (e.g.: ZnCl2·4H2O, LiI·2H2O, LiBr) and other more complex

171 solutions like aqueous metal solutions composed of transition metal ions and nitrous

172 ligands including cuprammonium hydroxide (Cuam: [Cu (NH3)4](OH)2) and

173 cupriethylenediamine hydroxide (Cuen:[Cu(H2N(CH2 )2NH2)2](OH)2) (Isik, Sardon, &

174 Mecerreyes, 2014; Lindman et al., 2010; Wang et al., 2016).

7
ACCEPTED MANUSCRIPT
175 Water solubility can be increased by soaking cellulose with alkali to swell the

176 structure, followed by reaction with chloroacetic acid, methyl chloride or propylene

177 oxide to yield carboxymethyl cellulose (CMC), methylcellulose (MC),

178 hydroxypropylmethyl cellulose (HPMC) or hydroxypropyl cellulose (HPC) (Bourtoom,

179 2008). As the solubility increases after the chemical treatments, these cellulose

PT
180 derivatives are used as raw material to prepare biodegradable films. The films obtained

RI
181 from these cellulose derivatives are generally transparent, water soluble, odourless,

182 tasteless, flexible, with moderate strength and resistance to lipid compounds (Dhall,

SC
183 2013).

184 Methylcellulose is the most hydrophilic water-soluble cellulose derivative. It is

185
U
also more economical and readily available (Erdohan & Turhan, 2005). Films from
AN
186 Hydroxypropyl cellulose and methylcellulose are very efficient oxygen, carbon dioxide,
M

187 and lipid barriers, but with poor resistance to water vapour transport. Although, the

188 water vapour barrier properties can be improved by adding hydrophobic materials such
D

189 as lipids into the film-forming solution (Villalobos, Chanona, Hernández, Gutiérrez, &
TE

190 Chiralt, 2005). On the other hand, methylcellulose and hydroxypropylmethyl cellulose

191 have the ability to form a thermally induced coating that has been used to cover fried
EP

192 products (Varela & Fiszman, 2011). Moreover methylcellulose could be used as a

193 barrier to lipid migration. Hydroxypropyl cellulose can be extruded into films because
C
AC

194 of its thermoplastic characteristics (Bourtoom, 2008).

195 Other cellulose derivative like cellophane, a regenerated cellulose film, has not

196 been commonly used for food packaging because has low barrier properties (high water

197 vapour permeability) compared to synthetic polymers (Bedane, Eić, Farmahini-

198 Farahani, & Xiao, 2015).

8
ACCEPTED MANUSCRIPT
199 Cellulose obtained from vegetal sources contains impurities in the form of other

200 natural compounds like lignin and hemicellulose. The bacterial cellulose is cellulose

201 synthesized by microorganisms (Esa, Tasirin, & Rahman, 2014; Klemm et al., 2005;

202 Nguyen, Gidley, & Dykes, 2008), It is pure cellulose, free of other compounds.

203 Bacterial cellulose can be extracellularly synthesized into fibrils by several bacteria

PT
204 using mainly glucose as a carbon source, although the use of other sugars like fructose

RI
205 as substrate have been reported (Shoda & Sugano, 2005). Cellulose from plants and

206 bacterial cellulose have the same chemical structure, but because of the characteristics

SC
207 of fibres and its purity, bacterial cellulose has several advantages over plant cellulose

208 such as higher mechanical strength, crystallinity and hydrophilicity (Rozenberga et al.,

209 2016).
U
AN
210
M

211 2.2 Chitosan

212 The second most abundant polysaccharide found in nature after cellulose is
D

213 chitin. Chitosan is obtained by alkaline N-deacetylation of chitin. Chitin is found in the
TE

214 exoskeleton of crustaceans and several insects. For this reason, chitosan is commercially

215 available from plentiful renewable sources, primarily waste from the shellfish industry
EP

216 (Kim et al., 2006).

217 Chitosan is a linear polysaccharide consisting of (1,4)-linked 2-amino-deoxy-β-


C

Ɗ-glucan. It is non-toxic, biodegradable, biofunctional and biocompatible (Dutta et al.,


AC

218

219 2009). Additionally, chitosan has attracted attention as a potential food preservative

220 due to its antimicrobial activity against a wide range of fungi, yeasts and bacteria

221 (Aider, 2010; Darmadji & Izumimoto, 1994; do Amaral et al., 2015; Dutta et al., 2009;

222 Jo, Lee, Lee, & Byun, 2001; Lekjing, 2016). Although the mechanism of the

223 antimicrobial activity of chitosan is not discerned yet, there are some hypotheses. The

9
ACCEPTED MANUSCRIPT
224 most accepted hypothesis attributes the antimicrobial activity to a change in cell

225 permeability due to interactions between the positively charged chitosan molecules and

226 the negatively charged microbial cell membrane. The other hypothesis is the interaction

227 of diffused hydrolysis products with microbial DNA inhibiting the mRNA and protein

228 synthesis and chelating metals, spore elements and essential nutrients (No, Meyers,

PT
229 Prinyawiwatkul, & Xu, 2007).

RI
230 Chitosan is insoluble in water and in common organic solvents (van den Broek,

231 Knoop, Kappen, & Boeriu, 2015). However, the solubility depends on the degree of N-

SC
232 acetylation (DA) and molecular weight (Mw). Chitosan can be easily dissolved in acid

233 solutions below pH 6.3, although at concentrations above >2 % w/w the resulting

234
U
solution becomes very viscous (Kaur & Dhillon, 2014; Yeul & Rayalu, 2012).
AN
235 Several methods to obtain films based on chitosan have been reported. In the
M

236 casting method, chitosan is dissolved in suitable solvents or slightly acidified water and,

237 if necessary, a plasticizer is added. The solution is poured on a flat surface and the
D

238 solvent is allowed to evaporate.


TE

239 When obtaining foils, extrusion is the most economical option (van den Broek et

240 al., 2015). In extrusion, one or two rotating screws fitted in a barrel are used to
EP

241 progressively increase the pressure and mix the ingredients during the manufacture of

242 the films. At the end of the screws, the mixture is expanded after passing through a die
C
AC

243 (Espitia et al., 2014). But in this case, the chitosan has a disadvantage, because chitosan

244 is not thermoplastic as it degrades before the melting point. Therefore, unlike

245 conventional thermoplastic polymers, chitosan cannot be extruded or molded and the

246 films cannot be heat-sealed. This behaviour limits the production of chitosan films at a

247 commercial level and narrows the applications (Pelissari, Yamashita, & Grossmann,

248 2011). The blending of chitosan with thermoplastic polymers, like poly(butylene

10
ACCEPTED MANUSCRIPT
249 succinate), poly(butyleneterephthalate adipate), poly(butylene succinate adipate),

250 represents an alternative to improve thermal properties of this material (van den Broek

251 et al., 2015).

252 The good film-forming properties of chitosan allow the production of films and

253 coating materials with good mechanical properties and a selective permeability to CO2

PT
254 and O2. However, chitosan films are highly permeable to water vapour limiting their use

RI
255 in food products because usually, an effective control of moisture transfer is an

256 important property for the preservation of the food quality, especially in high relative

SC
257 humidity environments (Aider, 2010).

258 For this reason, several strategies have been proposed to improve the functional

259
U
properties of chitosan films. For example, modifications of the deacetylation degree,
AN
260 pH, solvent type or plasticizers. Mixing with other components (proteins or
M

261 polysaccharides) can improve the functional properties of the chitosan films (Maher Z.

262 Elsabee & Entsar S. Abdou, 2013). Unfortunately, the high sensibility of chitosan to
D

263 moisture limits the application for food packaging as chitosan films can remain water
TE

264 sensitive or even water soluble. Cross-linking of chitosan with several reagents like

265 genipin, glutaraldehyde or formaldehyde has been proposed to prevent dissolving and/or
EP

266 swelling of chitosan-based films (Dutta et al., 2009).

267
C
AC

268 2.3. Starch

269 Starch is a natural polymer that could be an alternative to produce food-

270 packaging materials. It has been extensively studied because it is abundant, cheap,

271 biodegradable and edible. Starch is an agricultural biopolymer found in a variety of

272 plants including wheat, corn, rice, beans, and potatoes (Kolybaba et al., 2006). This

273 polymer constitutes more than 60 % of cereal kernels and it is relatively easy to separate

11
ACCEPTED MANUSCRIPT
274 from the other components (Jiménez et al., 2012). Depending on the botanical source,

275 starch granules vary in shape, size, structure, and chemical composition (Molavi,

276 Behfar, Shariati, Kaviani, & Atarod, 2015).

277 The starch granules are essentially composed of two main polysaccharides:

278 amylose and amylopectin. Starch granules contain also trace amounts of other

PT
279 components such as lipids and proteins. Amylose is a linear chain polymer of α-1,4

RI
280 anhydroglucose units with a molecular size ranging from 20 to 800 kg/mol. It accounts

281 for about 20-25% of most granular starches. The amylose is responsible for the film-

SC
282 forming properties of the starch (Bertuzzi et al., 2007; Bonilla, Atarés, Vargas, &

283 Chiralt, 2013; Jiménez et al., 2012; Peressini, Bravin, Lapasin, Rizzotti, & Sensidoni,

284
U
2003). Amylopectin is a highly branched polymer of short α-1,4 chains linked by α-1,6
AN
285 glycosidic branching points occurring every 25–30 glucose units and with a very high
M

286 molecular weight (5000–30,000 kg/mol) (Jiménez et al., 2012; Peressini et al., 2003).

287 The differences of structure and molecular weight between amylose and amylopectin
D

288 lead to differences in their molecular properties and film-forming properties.


TE

289 The majority of starches are semi-crystalline materials ranging from 15 to 45%

290 of crystallinity depending on the ratio of amylose/amylopectin (usually from 20-25/75-


EP

291 80%). The crystalline regions are formed by the short-branched chains in the

292 amylopectin meanwhile the amylose and the branching points of amylopectin form the
C
AC

293 amorphous regions.

294 Starch granules are not soluble in cold water as the hydrogen bonds hold the

295 starch chains together. However, when starch is heated in water, the crystalline structure

296 is disrupted and water molecules interact with the hydroxyl groups of amylose and

297 amylopectin resulting in a partial solubilisation. Heating starch suspensions in excess of

298 water allows forming hydrogen bonding and at temperatures between 65 and 90 °C,

12
ACCEPTED MANUSCRIPT
299 depending on the type of starch, an irreversible gelatinization process takes place

300 (Jiménez et al., 2012).

301 To obtain a homogeneous film-forming solution of starch, it is necessary to

302 gelatinize the granules in an excess of water (>90 % w/w). This process breaks the

303 amylopectin matrix and releases the amylose. These changes happen because the water

PT
304 diffuses through the granules promoting the melting of the starch crystallites. The

RI
305 gelatinization process is very complex (Carvalho, 2008) and it is needed to obtain a

306 homogeneous solution during film preparation.

SC
307 To obtain films from starch, two techniques are mainly employed: dry process

308 and wet process. The dry process is based on the thermoplastic properties of starch by

309
U
extrusion. In this method, the starch is plasticized and heated above its glass transition
AN
310 temperature in conditions of low water content. In the wet process, the polymers are
M

311 solubilized and then the film-forming solution is dried. The wet process is generally

312 preferred to form edible pre-formed films, or to apply coatings by dipping, brushing or
D

313 spraying onto food products (Peressini et al., 2003). Nevertheless, dry methods are more
TE

314 feasible for film manufacturing at industrial level (Jiménez et al., 2012).

315 Starch films have excellent oxygen barrier properties due to their high-ordered
EP

316 hydrogen-bonded network structure in which the amylose and amylopectin form

317 crystalline and non-crystalline regions in alternating layers. Therefore the barrier
C
AC

318 properties are improved by increased crystallinity or higher content of amylopectin in

319 the material. Starch-based films have some drawbacks compared to conventional

320 synthetic polymers. The tensile strength is relatively high but the elongation percentage

321 is low resulting in poor mechanical properties. Although these films are hydrophilic

322 materials, starch films with higher crystalline structure are less sensitive to moisture and

13
ACCEPTED MANUSCRIPT
323 to the environmental relative humidity (Jonhed, Andersson, & Järnström, 2008; Mali,

324 Grossmann, Garcı́a, Martino, & Zaritzky, 2004; Molavi et al., 2015).

325 The poor mechanical properties, mainly the brittleness, of starch films are due to

326 the amorphous regions formed by amylose. This is why it is necessary to add a

327 plasticizer to overcome film brittleness caused by extensive intermolecular forces,

PT
328 thereby improving flexibility and extensibility. Another alternative to improve these

RI
329 properties is blending the starch with other biopolymers or synthetic polymers, like

330 polyvinyl alcohol (PVA), which allows maintaining its biodegradability while

SC
331 improving the mechanical properties (Bertuzzi et al., 2007; Xiong, Tang, Tang, & Zou,

332 2008).

333
U
AN
334 2.4. Pectin
M

335 Pectin is one of the main components of the plant cell and constitutes

336 approximately the third part, on dry basis, of the cell wall of peels in several fruits. Only
D

337 a few plants are utilized as raw material for commercial production of pectins, mainly
TE

338 apple and citrus peels, with 10–15 and 20–30 wt % in dry basis, respectively. The

339 selection of these sources depends on the yield, time, cost of the extraction procedures,
EP

340 the desired properties of the extracted pectins and the availability of the raw materials

341 (Casas-Orozco, Villa, Bustamante, & González, 2015; Munarin, Tanzi, & Petrini,
C
AC

342 2012).

343 Pectin properties are strictly related to the microstructure, even though it is not

344 yet completely understood. There is a traditional hypothesis about pectin structure,

345 which assumes that the polysaccharide is constituted by poly α-1-4-galacturonic acids,

346 known as homogalacturonan (De Cindio, Gabriele, & Lupi, 2016; Espitia et al., 2014).

347 Pectin is composed of at least three polysaccharide domains: homogalacturonan,

14
ACCEPTED MANUSCRIPT
348 rhamnogalacturonan-I and rhamnogalacturonan-II, but homogalacturonan is the major

349 component of pectin polysaccharides (Munarin et al., 2012).

350 The carboxyl groups of the galacturonic acid units are esterified with methanol

351 and sometimes, partially acetyl-esterified. Depending on its degree of esterification with

352 methanol, pectin can be classified as high methoxyl pectin (containing more than 50%

PT
353 esterified carboxyl groups), or low methoxyl pectin (pectins with < 50% esterified

RI
354 carboxyl groups) (De Cindio et al., 2016; Espitia et al., 2014).

355 The molecular weight, degree of esterification and acetyl-esterification depend

SC
356 on the source and the extraction conditions. These parameters determine the properties

357 of the pectin including gelling, textural properties and stability (Munarin et al., 2012).

358
U
Pectins are acid and water soluble, and they are used mainly as a gelling agent in
AN
359 the production of jams, fruit juices and bakery fillings, and as stabilizing agents in milk
M

360 drinks and yogurt (Gutierrez-Pacheco et al., 2016; Willats, Knox, & Mikkelsen, 2006).

361 Furthermore, because of its gelling properties, the use of pectin as raw material to
D

362 prepare edible films has been investigated in recent years. Pectin-based films have good
TE

363 gas permeability properties, but poor water barrier properties although these films have

364 been used to retard moisture loss and migration of lipids (Gutierrez-Pacheco et al.,
EP

365 2016). Moreover pectin does not have antimicrobial properties. Films elaborated with

366 pure pectin promoted microbial growth because pectin is used as a carbon source by
C
AC

367 fungi and bacteria (Gutierrez-Pacheco et al., 2016).

368

369 2.5. Alginate

370 Alginates are extracted from brown seaweeds called Phaephyceae. Alginates are

371 the salts of alginic acid, a linear co-polymer composed of 1-4β-D-mannuronic acid (M)

15
ACCEPTED MANUSCRIPT
372 and α-L-guluronic acid (G) (Gennadios, Hanna, & Kurth, 1997). In polymer chains,

373 monomers are arranged alternately in GG and MM blocks, together with MG blocks.

374 The most important property of alginates is their ability to react with di- and tri-

375 valent cations to form films. Calcium ions are more effective than magnesium,

376 manganese, aluminum, ferrous, and ferric ions as gelling agents. The ions establish an

PT
377 association between M and G blocks, resulting in a stable and ordered three-

RI
378 dimensional network pictured as the ‘‘eggbox’’ model (Benavides, Villalobos-Carvajal,

379 & Reyes, 2012; Gennadios et al., 1997). The crosslinking process with polyvalent

SC
380 cations has been used to improve the water barrier properties, mechanical resistance,

381 cohesiveness and stiffness and to delay the release of some drugs. Due to the fast

382
U
crosslinking process with calcium ions, localized gelling areas are produced,
AN
383 compromising the uniformity of the films. A technique to form homogeneous matrices
M

384 through a slow release of calcium in an acid medium have been proposed (Benavides et

385 al., 2012).


D

386 Alginate forms films after evaporation of the solvent and the crosslinking
TE

387 promoted by added calcium can improve the mechanical and barrier properties. As

388 mentioned earlier, although alginate films or coatings have poor moisture barriers, their
EP

389 hygroscopicity slows food dehydration on which is applied. Edible films made from

390 pectins possess similar characteristics to the alginate ones retarding moisture loss
C
AC

391 (Benavides et al., 2012; Olivas & Barbosa-Cánovas, 2008).

392

393 3. Film properties

394 Controlling water vapour and gas transmission rates are critical factors in order

395 to achieve the required quality, safety and extended shelf-life for moisture sensitive

396 foods. High-performance films with high flexibility, optical transparency, thermal

16
ACCEPTED MANUSCRIPT
397 stability, mechanical strength, biodegradability, and gas barrier properties are highly

398 demanded for several applications beyond packaging (Bedane et al., 2015).

399 Polysaccharides form a network responsible for good mechanical properties but

400 it has poor barrier for water vapour transfer. On the other hand, adding lipids allows

401 obtaining excellent water vapour barriers properties but also the obtained films are

PT
402 usually opaque, unstable, brittle, and they have a waxy tasting. Several studies have

RI
403 reported films prepared from polysaccharides combined with lipids or proteins to take

404 advantage of the properties of each component (Rubilar, Zúñiga, Osorio, & Pedreschi,

SC
405 2015).

406

407
U
3.1 Barrier properties. Water vapour permeability
AN
408 The barrier properties of a polymeric film are crucial for estimating or predicting
M

409 the product shelf-life. Commonly, plastics are relatively permeable to small molecules

410 such as gases, water vapour, organic vapours or liquids. Water vapour and oxygen are
D

411 two of the main permeates studied in packaging applications. These compounds may
TE

412 transfer from the internal or external environment through the polymer package wall,

413 resulting in a continuous change in product quality decreasing the shelf-life (Siracusa et
EP

414 al., 2008).

415 The water vapour barrier properties can be quantified by the water vapour
C
AC

416 permeability (WVP), which indicates the amount of water that permeates per unit of

417 area and time (kg/m s Pa). It takes into account the differential pressure and thickness of

418 the packaging material. Water vapour barrier properties are important either for fresh

419 food products, i.e. vegetables, where it is important to avoid dehydration or in some

420 other foods like bread or dry foods where it is important to avoid moisture uptake from

421 the environment.

17
ACCEPTED MANUSCRIPT
422 Most of data on WVP of biodegradable films available in the literature have

423 been obtained gravimetrically following the ASTM Standard Test Method E96/E96M

424 (ASTM), known as the “cup method”, or variations of it. According to this method, a

425 cup with an open mouth of known area is filled with distilled water or desiccant. In a

426 modification of the method, saturated salt solutions are also used. The film is sealed on

PT
427 the open mouth of the cup, the assembly is weighed, and placed under controlled

RI
428 temperature and usually low relative humidity conditions. The driving force of water

429 vapour transport is the gradient of partial pressure between both sides of the film. The

SC
430 weight change of the cup is monitored (Gennadios, Weller, & Gooding, 1994). With the

431 obtained data, WVP is calculated according to the combined Fick and Henry laws for

432
U
gas diffusion through films, using the equation 1:
AN
433 WVP = ∆w·x/A ∆P (1)
M

434 where ∆w is the weight gain of the cell after 24 h, x is the film thickness, A is the area of

435 exposed film, and ∆P is the differential vapour pressure of water through the film.
D

436 (Khan et al., 2010).


TE

437 To express water transfer, the water vapour transmission rate (WVTR),

438 expressed in (g/m2 day) is also used. The WVP can be calculated from to the WVTR
EP

439 according to equation 2 (Siracusa et al., 2008):

440 WVP = WVTR·x/∆P (2)


C
AC

441 where x is the thickness of the film (m) and ∆P is the partial pressure difference across

442 the film (Pa).

443 Usually, polysaccharide-based films have higher values of WVP than

444 commercial synthetic materials. For comparative purposes, Table 1 shows water vapour

445 permeability of some polysaccharide-based films and Table 2 shows water vapour

446 permeability of several synthetic polymers.

18
ACCEPTED MANUSCRIPT
447 Biodegradable films based on cellulose-derivative are very efficient barriers to

448 oxygen and aroma compounds. Among the cellulose derivatives methylcellulose is one

449 of the most extensively studied (Nazan Turhan & Şahbaz, 2004). Studies about the

450 effect of methylcellulose concentration on the water vapour barrier properties reported

451 no significant variation in the WVP obtaining values from 0.525x10-10 to 0.598x10-10

PT
452 (g/s m Pa) for concentrations ranging from 1.5 to 6 %. For these films, the thickness

RI
453 ranged from 0.01 to 0.024 mm. These values could be explained by the hydrophilic

454 characteristic of the films resulting from the hydroxyl groups in the polymer chain.

SC
455 When ethanol was used as a solvent, WVP increased significantly as the ethanol

456 concentration increased. The same behaviour was observed with the addition of

457
U
plasticizers to the polymer matrix resulting in an increase in WVP (Nazan Turhan &
AN
458 Şahbaz, 2004).
M

459 In chitosan-based films, the WVP values were significantly affected by the

460 degree of deacetylation of chitosan, solvent pH and type of acid (Kim et al., 2006). In
D

461 the pH range 3-4, the WVP values increased with pH. Regarding the solvent effect,
TE

462 WVP values of chitosan-based films obtained using formic or lactic acid as solvents

463 were 3.04x10-5 or 2.34x10-5 g/m h Pa, respectively. These values were significantly
EP

464 higher than those obtained from chitosan-based films using acetic or propionic acid as

465 solvents. The addition of starch to chitosan decreased the values of WVP (Santacruz,
C
AC

466 Rivadeneira, & Castro, 2015). About the deacetylation degree (DA) effect, low DA

467 chitosan-based films showed lower values of WVP compared to high DA chitosan-

468 based films (Kim et al., 2006). The molecular weight of chitosan did not influence WVP

469 but it did affect the mechanical properties (Kerch & Korkhov, 2010; Park, Marsh, &

470 Rhim, 2002).

19
ACCEPTED MANUSCRIPT
471 In starch-based films, the values of WVP gave similar values using starches of

472 high (0.52 ± 0.03 g mm/m2 h kPa) or low (0.59±0.13 g mm/m2 h kPa) amylose content.

473 As in other materials, the effect of plasticizers and their interaction with the matrix

474 should be considered when working with starch films. The addition of glycerol as a

475 plasticizer increased the WVP of high amylose starch-based films. An opposite effect

PT
476 was observed using xylitol as a plasticizer. However, for low amylose starch-based

RI
477 films, adding glycerol increased considerably WVP values while no effect of xylitol

478 was observed (Muscat et al., 2012).

SC
479 When the starch is treated with acids, the type of acid should be considered.

480 Several authors reported that the addition of citric acid in starch films led to a decrease

481
U
in WVP because the hydrophilic OH groups are substituted with hydrophobic ester
AN
482 groups (Seligra, Medina Jaramillo, Famá, & Goyanes, 2016).

Chitosan in tapioca starch films decreased WVP values from 12.1x10-10 to


M

483

484 2.8x10-10 g/m s Pa (Vásconez, Flores, Campos, Alvarado, & Gerschenson, 2009). This
D

485 decreasing promoted by chitosan can be due to its higher hydrophobicity when
TE

486 compared to starch. Additionally, hydrogen bond interactions between tapioca starch

487 and chitosan reduce the availability of the hydrophilic groups, decreasing their
EP

488 interactions with water molecules and reducing the water vapour transmission rate (Park

489 et al., 2004; Pinotti, Garcia, Martino, & Zaritizky, 2007; Xu et al., 2005). The addition
C
AC

490 of potassium sorbate to chitosan-starch films increased WVP. In this case, electrostatic

491 and/or hydrogen bond interactions between potassium sorbate and chitosan might have

492 prevailed over starch-chitosan hydrogen bond interactions (Vásconez, Flores, Campos,

493 Alvarado, & Gerschenson, 2009).

494 In the case of alginate-based films, alginate composition significantly affected

495 the WVP properties of films. Films with a higher proportion of α-L-guluronic acid

20
ACCEPTED MANUSCRIPT
496 proved to be better moisture barriers. Films containing a higher concentration of

497 guluronate (G) than mannuronate (M) (M/G 0.45 films) showed WVP two or three

498 times lower than M/G 1.5 films when immersed in calcium solution (Olivas & Barbosa-

499 Cánovas, 2008). Alginate–Ca films containing fructose showed low values of WVP.

500 However, no statistically significant difference was found between WVP of films

PT
501 containing fructose and that of films containing sorbitol at 76% and 100% RH (Olivas

RI
502 & Barbosa-Cánovas, 2008).

503 The gelation process of the alginate films can affect the WVP values of the

SC
504 films. WVP of the films was not significantly affected by the internal gelation process,

505 except at high calcium concentration (0.03 g CaCO3/g alginate), which revealed a

506
U
significant decrease in WVP (Benavides et al., 2012).
AN
507 The changes in the film structure and properties when other components are
M

508 added to the chitosan matrix have been evaluated. An increase in WVP values in

509 chitosan films containing proteins (fish gelatin, quinoa protein, whey protein) has been
D

510 reported. Results showed that the chitosan-protein films obtained were more hydrophilic
TE

511 than chitosan films (Abugoch, Tapia, Villamán, Yazdani-Pedram, & Díaz-Dosque,

512 2011; Fakhreddin Hosseini et al., 2013).


EP

513 The addition of lipids to decrease the WVP values has been evaluated

514 (Zivanovic, Chi, & Draughon, 2005). When combining methylcellulose and
C
AC

515 hydroxypropylmethyl cellulose with stearic/palmitic fatty acids and beeswax for a

516 surface coat (Kester & Fennema, 1989), low WVP values were obtained (6.615x10-13

517 g/s m Pa). This is in the same order of magnitude as that of synthetic polymers. These

518 wax-laminated films are potentially useful in any food product where undesirable

519 intercomponent moisture migration takes place (Kester & Fennema, 1989). Other

520 examples are tara gum added with oleic acid (OA) (Ma, Hu, Wang, & Wang, 2016),

21
ACCEPTED MANUSCRIPT
521 brea gum added with beeswax (Spotti, Cecchini, Spotti, & Carrara, 2016), chitosan

522 added with citronella essential oil and cedarwood oil (Shen & Kamdem, 2015) and

523 alginate added with essential oregano oil (Benavides et al., 2012).

524 In pectin-based film, the WVP increased when the concentration of pectin and

525 sorbitol was increased meanwhile adding beeswax decreased WVP as expected taking

PT
526 into account its hydrophobicity (Maftoonazad, Ramaswamy, & Marcotte, 2007)

RI
527 Recently, a nanotechnological approach for improving barrier properties of films

528 has been the main focus of several studies. In polysaccharides-based films, nanoclays

SC
529 have been added to the polymeric matrix to improve the barrier properties (Chivrac,

530 Pollet, & Avérous, 2009). Moreover polysaccharides are used as raw material to prepare

531
U
nano-fillers and the most common polysaccharides used for nanoparticle production in
AN
532 biodegradable films are cellulose, starch and chitosan. However, chitosan is the most
M

533 attractive material to produce nano-fillers.

534 Chitosan nanoparticles can be produced based on ionic gelation as a result of


D

535 inter- and intra-molecular cross-linking of chitosan's protonated amino groups by


TE

536 multivalent polyanions. Sodium tripolyphosphate (TPP) is a very popular polyanion

537 because it is non-toxic and forms gels with desirable properties. Chitosan nanoparticles
EP

538 have been successfully used as fillers to improve mechanical and barrier properties as

539 well as the thermo-stability of films, decrease solubility and produce more compact and
C
AC

540 dense materials. Additionally, chitosan has the advantage of its antibacterial activity

541 when compared to other bio-based food packaging materials (Antoniou, Liu, Majeed, &

542 Zhong, 2015). The studies of chitosan nanoparticles and bulk chitosan with glycerol-

543 plasticized tara gum films (Antoniou et al., 2015) showed that the nanoparticles did not

544 affect the WVP values (Lorevice, Otoni, Moura, & Mattoso, 2016).

22
ACCEPTED MANUSCRIPT
545 Cellulose and derivatives nanoparticles were also studied. Hydroxypropylmethyl

546 cellulose with microcrystalline cellulose (MCC) (Dogan & McHugh, 2007),

547 methylcellulose with nanocellulose (Khan et al., 2010), tara gum film with cellulose

548 nanocrystals (Ma, Hu, & Wang, 2016), gelatin nanocomposite films containing bacterial

549 cellulose nanocrystals (George, 2012). All films did not show a significant change in the

PT
550 WVP with the addition of fillers.

RI
551 Other materials studied are hydroxypropylmethyl cellulose with silver

552 nanoparticles (de Moura, Mattoso, & Zucolotto, 2012), potato starch films with

SC
553 montmorillonite nanoparticles (Avella et al., 2005), agar films with silver nanoparticles

554 (Rhim, Wang, & Hong, 2013), potato starch plasticized with glycerol and further

555
U
reinforced with catechin and starch nanocrystals (Sessini, Arrieta, Kenny, & Peponi,
AN
556 2016), chitosan with Ag-nanoparticles and chitosan with ZnO nanoparticles (Youssef,
M

557 Abou-Yousef, El-Sayed, & Kamel, 2015) and chitosan with silver oxide encapsulated

558 nanocomposite (Tripathi, Mehrotra, & Dutta, 2011).


D

559 On the other hand, an important factor that must be taken into account when
TE

560 establishing the feasibility of these films as packaging material is the modification of

561 the properties over time and storage temperature. Experimental data showed WVP rate
EP

562 increases when the storage time increases and with the decrease of storage temperature.

563
C
AC

564 3.2. Mechanical properties

565 Mechanical properties including tensile strength (TS), percentage of elongation

566 at break (%E) and Young’s modulus, depend on the film composition and on the nature

567 of components (Vieira et al., 2011). These properties are important when the packaging

568 films are intended to withstand external stress while maintaining its integrity

569 (Fakhreddin Hosseini et al., 2013). The measurement of these properties allows

23
ACCEPTED MANUSCRIPT
570 predicting how the material will behave under different food processing conditions and

571 comparing it with commercial polymers.

572 Tensile strength is the maximum stress supported by the film before breaking

573 (Gennadios et al., 1997) and is calculated from the tensile test using the equation 4

574 (Antoniou et al., 2015):

PT
575 TS = F/(L·x) (4)

RI
576 where F is the tensile force (N), L the width of the film (mm) and x the thickness (mm).

577 Percentage of elongation at break is the maximum elongation of the film before

SC
578 rupture (Krochta and De Mulder-Johnston, 1997), calculated also from the tensile test

579 using the equation 5 (Antoniou et al., 2015).

580 %E = 100 · (l-l1)/l1


U (5)
AN
581 Where l1 is the initial length and l is the length of the film at the breaking point.
M

582 A materials testing machine (Instron) or a texture analyser in tensile mode

583 following the ASTM D882-12 can be used to determine the tensile strength, the
D

584 percentage of elongation at break (E%) and the elastic Young’s modulus of packaging
TE

585 material (Siracusa et al., 2008).

586 The puncture test is an alternative for measuring the mechanical properties when
EP

587 no probes or fixtures are available to perform tensile test. In the puncture test the film is

588 fixed to a support with a circular opening, and a cylindrical probe penetrated the film.
C
AC

589 Force-deformation curves are used to calculate the maximum strength and maximum

590 deformation at the rupture point (Spotti et al., 2016).

591 The TS values of the polysaccharides-based films are similar to the values of the

592 synthetic polymers. The most important differences are observed in the values of %E

593 (Table 3). For comparative purposes, Table 4 shows TS and %E values of some

594 synthetic polymers.

24
ACCEPTED MANUSCRIPT
595 In methylcellulose films, an increase in methylcellulose concentration decreased

596 the TS and %E values. Values of 33±3 and 14±1% were obtained using a concentration

597 of 1.5% methylcellulose, respectively. The values were 8±1 and 6± 2% using a

598 concentration of 6% methylcellulose. The decrease in mechanical properties with

599 increasing methylcellulose concentration may be explained by the partial insolubility at

PT
600 high concentrations. On the other hand, the addition of plasticizer (PEG400) decreased

the TS values and increased the %E values (Nazan Turhan & Şahbaz, 2004).

RI
601

602 In chitosan films, mechanical properties were affected by deacetylation degree,

SC
603 solvent, pH and molecular weight. Chitosan films with formic or lactic acid showed

604 lower TS values when compared to films prepared using acetic or propionic acid (Kim

605
U
et al., 2006; Park et al., 2002). The TS values significantly decreased when pH
AN
606 increased. High deacetylation degree chitosan films were very sensitive to pH changes
M

607 in comparison to low deacetylation degree chitosan film. Also, the degree of

608 dissociation of chitosan decreased when the pH increased (Kim et al., 2006).
D

609 An increase in molecular weight produced films with high values of TS y %E


TE

610 when the acetic acid was used as a solvent. On the other hand, lower values in

611 mechanical properties were observed when the citric acid was used as a solvent (Park et
EP

612 al., 2002; Zivanovic et al., 2005).

613 For chitosan films, temperature and relative humidity exerted greater influence
C
AC

614 on TS and %E, while storage period showed moderate influence (Kerch & Korkhov,

615 2010). Also, chitosan films had significantly higher %E values compared to films made

616 with other biopolymers such as wheat gluten, corn zein protein, and soy protein isolate

617 (Cunningham, Ogale, Dawson, & Acton, 2000).

618 Films obtained adding proteins to chitosan showed a decrease in TS values and

619 an increase in %E values. For example, the addition of quinoa protein decreased the

25
ACCEPTED MANUSCRIPT
620 values of TS from 22.2 to 2.3 MPa and increased the %E from 73.6 to 237% when

621 compared to control films with no quinoa added (Abugoch et al., 2011).

622 For starch-based films, when high amylose starch was used, films showed higher

623 tensile strength, higher Young´s modulus and lower elongation at break than those

624 obtained from low amylose starch. Both, the low and high amylose starch films, showed

PT
625 a decrease in the tensile strength and Young´s modulus and an increase in the

RI
626 elongation when increased plasticizer (i.e. glycerol or xylitol) concentrations above

627 15%. These plasticizers reduced the intra-molecular attraction forces between the starch

SC
628 chains promoting the hydrogen bonds formation between plasticizer and starch

629 molecules resulting in greater flexibility and subsequently decreasing the tensile

630 strength (Muscat et al., 2012).


U
AN
631 Adding starch from different sources to chitosan did not change the mechanical
M

632 properties as the obtained TS values were 10.61 and 9.27 MPa for potato and cassava

633 starch, respectively (Santacruz et al., 2015).


D

634 Lipids were used to improve barrier properties, however, they affect negatively
TE

635 the mechanical properties. For example, tara gum films added with oleic acid (OA)

636 significantly decreased TS value from 57.4 to 26.8 MPa when the OA concentration
EP

637 increased from 0 to 20 wt%. However, the %E increased from 2.7 to a maximum of 8.5

638 % at 10 % of OA and then decreased to 2.85 % when the OA concentration reached 20


C
AC

639 %. This behaviour was explained by a heterogeneous film structure with discontinuities

640 in the polymer network at higher concentration of OA resulting in a loss of flexibility

641 (Ma, Hu, Wang, et al., 2016).

642 The effect of the lipids depends on the polysaccharide source. The addition of

643 beeswax to brean gum films decreased values of both TS and %E (Spotti et al., 2016). A

26
ACCEPTED MANUSCRIPT
644 similar effect was observed in films obtained from chitosan added with citronella

645 essential oil and cedarwood oil (Shen & Kamdem, 2015).

646 The effect of adding cinnamon, allspice, and clove bud essential oils on the

647 mechanical properties of pectin films has been reported. The incorporation of essential

648 oils into the formulation resulted in a significant reduction of the tensile strength and the

PT
649 elastic modulus (Gutierrez-Pacheco et al., 2016).

RI
650 In films of HPMC containing MCC fillers, the tensile strength of the control film

651 prepared using a 3 wt% HPMC solution, increased from 29.7±1.6 to 70.1±7.9 MPa with

SC
652 the addition of MCC of 500-nm particles, while it increased only to 37.4±5.5 MPa with

653 the addition of 3-µm particles (Dogan & McHugh, 2007). The addition of chitosan

654
U
nanoparticles to HDM pectin and LDM pectin improved the mechanical properties. The
AN
655 tensile strength was the most affected mechanical attribute increasing from 30.81 to
M

656 46.95 MPa and from 26.07 to 58.51 MPa for HDM and LDM pectin-based films,

657 respectively.
D

658 The addition of nanocrystalline (CNC) cellulose to tara gum films was studied.
TE

659 The reduction in E% value with CNC addition may be related to a restriction in tara

660 gum matrix motion by rigid CNC (Ma, Hu, & Wang, 2016). In another study, the
EP

661 addition of chitosan nanoparticles increased considerably the TS values from 22.71 ±

662 2.98 to 58.44 ± 3.23 MPa. Using bulk chitosan, the increasing was not so important (49
C
AC

663 MPa). It is hypothesized that the incorporation of chitosan nanoparticles increased the

664 polymer's stiffness confirming the reinforcing effect of the nanoparticles on the

665 polymeric matrix. However, the percentage of elongation of the films decreased slightly

666 when CSNPs were incorporated, independently of the chitosan form (nanoparticles or

667 bulk) (Antoniou et al., 2015).

27
ACCEPTED MANUSCRIPT
668 In studies with alginate, mechanical properties were strongly affected by relative

669 humidity (Rhim, 2004). A higher relative humidity resulted in a decreasing of TS in

670 alginate films, while elongation increased, showing that water acts as a plasticizer in

671 hydrophilic films (Olivas & Barbosa-Cánovas, 2008).

672

PT
673 4. Applications

RI
674 Edibles films from polysaccharides or blend of polysaccharides with several

675 compounds including other polysaccharides, proteins, lipids and additives, have been

SC
676 applied to extend the shelf-life and preserve the quality of foods. Films and coatings

677 usually can be obtained from the same formulation. Coatings are applied in liquid form

678
U
before forming the coating meanwhile films are obtained as solid sheets and then
AN
679 applied to food products (Galus & Kadzińska, 2015). Films and coatings act as a barrier
M

680 and their mechanical properties depending on the specific requirements of food

681 preservation. Mechanical properties of coatings cannot be measured separately from the
D

682 surface coated (Gennadios et al., 1997).


TE

683 The development of polysaccharide films brought a significant increase in their

684 applications and in the number of products that can be treated for extending the shelf-
EP

685 life. For example, polysaccharide films have generally good gas barriers properties and

686 have also good adherence to cut surfaces of fruits or vegetables. However, their
C
AC

687 hydrophilic nature makes them poor barriers to moisture (E. Baldwin, Nisperos-

688 Carriedo, & Baker, 1995; Falguera, Quintero, Jiménez, Muñoz, & Ibarz, 2011).

689 Polysaccharide-based films exhibit excellent gas permeability properties,

690 resulting in desirable modified atmospheres that enhance the shelf life of the product

691 without creating anaerobic conditions unlike lipid-based films, which could promote

692 anaerobic conditions (Cutter, 2006).

28
ACCEPTED MANUSCRIPT
693 A cellulose-based edible film as a carrier of antioxidants, acidulants and

694 preservatives was used to prolong the storage life of cut apple and potato for 1 week at 4

695 °C (E. A. Baldwin, Nisperos, Chen, & Hagenmaier, 1996).

696 Chitosan films protect postharvest peaches from brown rot by decreasing the

697 incidence, prolonging the incubation period and reducing the rot area (Li & Yu, 2001).

PT
698 Chitosan coatings have been also used to increase the shelf life of peeled litchi

RI
699 fruit, which has a short shelf-life. The results showed that the application of chitosan

700 coatings effectively maintained quality attributes and extended shelf-life of the peeled

SC
701 fruit. Chitosan coatings retarded weight loss and the decline in sensory quality, keeping

702 higher contents of total soluble solids, titratable acid, and ascorbic acid, and suppressed

703
U
the activity of polyphenol oxidase and peroxidase (Dong, Cheng, Tan, Zheng, & Jiang,
AN
704 2004).
M

705 Chitosan edible films were used to treat mango pulp, which has a short shelf life.

706 Sliced mango was treated with different concentrations of chitosan aqueous solutions
D

707 (0%, 0.5%, 1% or 2% chitosan); placed into plastic trays and over-wrapped with a
TE

708 commercial synthetic film, then they stored at 6ºC. The results showed that a chitosan

709 coating retarded water loss and the drop in sensory quality, increasing the soluble solid
EP

710 content, titratable acidity and ascorbic acid content. It inhibited the growth of

711 microorganisms, concluding that applying a chitosan coating effectively prolongs the
C
AC

712 quality attributes and extends the shelf life of sliced mango fruit (Chien, Sheu, & Yang,

713 2007).

714 The effect of chitosan as antimicrobial coating was also studied in refrigerated

715 broccoli. The antimicrobial effect of chitosan on the native microflora and Escherichia

716 coli O157:H7 inoculated in broccoli was evaluated. Chitosan treatments resulted in a

717 significant reduction in total mesophilic and psychotropic bacteria counts with respect

29
ACCEPTED MANUSCRIPT
718 to the control samples during the entire storage period. There was an immediate

719 decontaminating activity of chitosan. This work showed that the use of chitosan coating

720 is a viable alternative to control the microbiota present in food products minimally

721 processed like broccoli (Moreira, Roura, & Ponce, 2011).

722 Two types of oxidized banana starch films (betalains and no betalains) were

PT
723 prepared and used as covering for sausages. The results indicated that films did not

RI
724 significantly alter the color or moisture loss during storage while the films with

725 betalains maintained the antoxidative properties in the sausages during storage

SC
726 (Zamudio-Flores et al., 2015).

727 Starch-based coatings were applied to extend the storage life of strawberries

728
U
(Fragaria × ananassa). The effect of amylose content, plasticizer (glycerol or sorbitol),
AN
729 and the antimicrobial agent (potassium sorbate or citric acid) on the strawberry quality
M

730 attributes was studied. The inclusion of starch with higher amylose content into the

731 formulation decreased WVP and weight losses while retaining fruit firmness for longer
D

732 periods. The plasticizer used is also important. Coatings containing sorbitol showed
TE

733 lower WVPs than those containing glycerol, although both plasticizers allowed reducing

734 weight losses and maintained texture and surface colour of strawberries. Adding
EP

735 potassium sorbate to the formulation extended the storage life up to 28 days. Also, the

736 antimicrobial action of potassium sorbate was enhanced with the addition of citric acid
C
AC

737 (García, Martino, & Zaritzky, 1998).

738 An edible coating based on starch-gelatin blend was developed to keep quality

739 of refrigerated Red Crimson grapes during 21 days, concluding that the use of this

740 coatings in refrigerated grapes was effective as a natural postharvest treatment with the

741 aim to extend the shelf-life of grapes and maintain fruit quality, particularly for

30
ACCEPTED MANUSCRIPT
742 reducing the total weight loss. About the taste, the coating did not influence consumer

743 acceptance (Fakhouri, Martelli, Caon, Velasco, & Mei, 2015).

744 An edible antimicrobial coating based on a starch-chitosan blend was developed.

745 The presence of 1.5% chitosan in the coating inhibited the growth of total coliforms and

746 lactic acid bacteria throughout the storage period, thus demonstrating potential

PT
747 antimicrobial activity in the coating (Durango, Soares, & Andrade, 2006).

RI
748 The use of pectin films increased the water vapour resistance of coated fresh-cut

749 melon, previously dehydrated and immersed in calcium lactate, in comparison with

SC
750 uncoated fruit (Ferrari, Sarantópoulos, Carmello-Guerreiro, & Hubinger, 2013).

751 The effect of edible coatings made from alginate, pectin or gellan on the shelf-

752
U
life of fresh-cut melon has been reported. Shelf-life was assessed evaluating the gas
AN
753 exchange, firmness, sensory quality and microbial growth of fresh-cut melon for 15
M

754 days at 4 ºC. Coating from gellan, pectin or alginate decreased the water vapour

755 transmission, preventing desiccation and maintaining fruit firmness during storage, but
D

756 the coatings did not provide sufficient barrier to gas diffusion and did not exhibit
TE

757 antimicrobial effects (Oms-Oliu, Soliva-Fortuny, & Martín-Belloso, 2008).

758 In another study about pear wedges coated with pectin films containing
EP

759 antioxidants reported an increase in vitamin C and total phenolic content which allowed

760 maintaining their antioxidant potential (Gutierrez-Pacheco et al., 2016). Pectin coatings
C
AC

761 were applied on Ataulfo mangoes with positive results, reducing the postharvest

762 physiological changes and extending the storage life (Moalemiyan, Ramaswamy, &

763 Maftoonazad, 2012). The preservation of mango fruits was studied also with chitosan

764 films. The fruits were kept in a box covered with a chitosan film and using low-density

765 polyethylene film as a positive control. Several quality parameters including colour,

766 chlorophyll, acidity, vitamin C, carotenoid and sugar contents, were studied during

31
ACCEPTED MANUSCRIPT
767 storage for 9 days. Results showed the feasibility of using chitosan films for better

768 storage quality and shelf-life extension of fruits such as mangoes (Srinivasa, Baskaran,

769 Ramesh, Harish Prashanth, & Tharanathan, 2002).

770 The effect of alginate and gellan-based edible coatings on the shelf-life of fresh-

771 cut Fuji apples was evaluated by measuring changes in headspace atmosphere, colour,

PT
772 firmness and microbial growth during 23 days of storage at 4 ºC. Results demonstrated

RI
773 that the shelf-life of the apples with alginate and gellan coatings was extended

774 approximately three times as compared to the control. A significant decrease of ethylene

SC
775 production was observed and the initial firmness and colour of the apple wedges were

776 maintained during the refrigerated storage (Rojas-Graü, Tapia, & Martín-Belloso,

777 2008).
U
AN
778 The effect of alginate coating was studied for sweet cherry fruits. These fruits
M

779 were treated with an edible coating based on sodium alginate at several concentrations.

780 The study showed that the alginate coating was effective to delay the postharvest
D

781 ripening process of sweet cherry (reduced colour changes, acidity and firmness losses,
TE

782 higher concentration of total phenolics and total antioxidant activity). About shelf life of

783 the sweet cherry, results suggest that the maximum storability period of control fruits
EP

784 could be established in 8 days at 2 °C plus 2 days at 20 °C, while cherries with alginate

785 coating could be stored with optimal quality and enhanced antioxidant activity up to 16
C
AC

786 days at 2°C plus 2 days at 20°C (Díaz-Mula, Serrano, & Valero, 2012).

787 Alginate and gellan-based edible coatings with sunflower oil were also applied

788 to fresh-cut papaya. This study demonstrated that gellan-based formulations, with lipid

789 addition, had a better performance in limiting water loss compared to the coatings

790 obtained from alginate (Tapia et al., 2008).

32
ACCEPTED MANUSCRIPT
791 Polysaccharides-based films or coatings intended for meat preservation have

792 been available commercially for the Japanese meat industry for several years. The meat

793 products are wrapped or coated and then smoked or steam cooked. During the

794 processing, the polysaccharide-based coating is dissolved and integrated into the meat

795 surface resulting in higher yields, improved texture and reduced moisture loss (Cutter,

PT
796 2006).

RI
797 Cellulose derivatives are incorporated in most commercial products such as

798 Semperfresh (AgriCoat Industries Ltd., Berkshire, UK), Pro-long (Courtaulds Group,

SC
799 London), Nature-Seal (Ecoscience Product System Divison, Orlando, FL), and Natural

800 Shine 9000 (Pace International, Seattle, USA) among others (Dhall, 2013).

801
U
Some examples of starch-based commercial products derived are Materbi
AN
802 (Novamont-Italy) and Biopar (Biop. Germany) (Avella et al., 2005).
M

803

804 5. Conclusions
D

805 The applications of hydrocolloids as polysaccharide-based films in food


TE

806 technology offer new opportunities to develop novel food biodegradable packaging and

807 the possibility of being an alternative to synthetic plastics. This can solve the problem
EP

808 of the accumulation of waste due to the use of non-biodegradable petroleum-based

809 plastics.
C
AC

810 The barrier and mechanical properties reported for polysaccharides-based films

811 are away from properties that have the most commonly synthetic plastics used as

812 packaging in the food industry. However, they have been improved by combining

813 polysaccharides with other biopolymers.

814 A great discussion exists about the potential applications of edible materials on

815 food products. The main efforts are focused on searching for the correct combination

33
ACCEPTED MANUSCRIPT
816 between materials mixed and food products applied, because of the efficiency and

817 functional properties of polysaccharide films depend on biopolymers-based and other

818 additives used to improve their properties and, the final food products with which the

819 edible films are used. Further research to enhance the properties of these biopolymers to

820 achieve similar properties of commercial plastics are needed. Probably in a near future,

PT
821 polysaccharide-based films will be a good alternative to replace synthetic plastics.

RI
822

823 Acknowledgements

SC
824 A grant from the CONACYT-México to author Cazón is gratefully acknowledged.

825

826 References
U
AN
827 Abugoch, L. E., Tapia, C., Villamán, M. C., Yazdani-Pedram, M., & Díaz-Dosque, M. (2011).
828 Characterization of quinoa protein–chitosan blend edible films. Food Hydrocolloids,
829 25(5), 879-886. doi: http://dx.doi.org/10.1016/j.foodhyd.2010.08.008
M

830 Aider, M. (2010). Chitosan application for active bio-based films production and potential in
831 the food industry: Review. LWT - Food Science and Technology, 43(6), 837-842. doi:
832 http://dx.doi.org/10.1016/j.lwt.2010.01.021
D

833 AlMaadeed, M. A., Nógellová, Z., Mičušík, M., Novák, I., & Krupa, I. (2014). Mechanical,
834 sorption and adhesive properties of composites based on low density polyethylene
TE

835 filled with date palm wood powder. Materials & Design, 53, 29-37. doi:
836 http://dx.doi.org/10.1016/j.matdes.2013.05.093
837 Antoniou, J., Liu, F., Majeed, H., & Zhong, F. (2015). Characterization of tara gum edible films
EP

838 incorporated with bulk chitosan and chitosan nanoparticles: A comparative study.
839 Food Hydrocolloids, 44, 309-319. doi:
840 http://dx.doi.org/10.1016/j.foodhyd.2014.09.023
841 ASTM, E. Test Methods for Water Vapor Transmission of Materials ASTM E (Vol. 283): ASTM
C

842 International.
843 Avella, M., De Vlieger, J. J., Errico, M. E., Fischer, S., Vacca, P., & Volpe, M. G. (2005).
AC

844 Biodegradable starch/clay nanocomposite films for food packaging applications. Food
845 Chemistry, 93(3), 467-474.
846 Baldwin, E., Nisperos-Carriedo, M., & Baker, R. (1995). Edible coatings for lightly processed
847 fruits and vegetables. HortScience, 30(1), 35-38.
848 Baldwin, E. A., Nisperos, M. O., Chen, X., & Hagenmaier, R. D. (1996). Special Issue: Lightly-
849 Processed Horticultural ProductsImproving storage life of cut apple and potato with
850 edible coating. Postharvest Biology and Technology, 9(2), 151-163. doi:
851 http://dx.doi.org/10.1016/S0925-5214(96)00044-0
852 Bedane, A. H., Eić, M., Farmahini-Farahani, M., & Xiao, H. (2015). Water vapor transport
853 properties of regenerated cellulose and nanofibrillated cellulose films. Journal of
854 Membrane Science, 493, 46-57. doi: http://dx.doi.org/10.1016/j.memsci.2015.06.009

34
ACCEPTED MANUSCRIPT
855 Benavides, S., Villalobos-Carvajal, R., & Reyes, J. E. (2012). Physical, mechanical and
856 antibacterial properties of alginate film: Effect of the crosslinking degree and oregano
857 essential oil concentration. Journal of Food Engineering, 110(2), 232-239. doi:
858 http://dx.doi.org/10.1016/j.jfoodeng.2011.05.023
859 Bertuzzi, M. A., Castro Vidaurre, E. F., Armada, M., & Gottifredi, J. C. (2007). Water vapor
860 permeability of edible starch based films. Journal of Food Engineering, 80(3), 972-978.
861 doi: http://dx.doi.org/10.1016/j.jfoodeng.2006.07.016
862 Boldt, R., Gohs, U., Wagenknecht, U., & Stamm, M. (2016). Effect of electron-induced reactive
863 processing on morphology and structural properties of high-density polyethylene.

PT
864 Polymer, 95, 1-8. doi: http://dx.doi.org/10.1016/j.polymer.2016.04.044
865 Bonilla, J., Atarés, L., Vargas, M., & Chiralt, A. (2013). Properties of wheat starch film-forming
866 dispersions and films as affected by chitosan addition. Journal of Food Engineering,
867 114(3), 303-312. doi: http://dx.doi.org/10.1016/j.jfoodeng.2012.08.005

RI
868 Bourtoom, T. (2008). Edible films and coatings: characteristics and properties.
869 Carvalho, A. J. (2008). Starch: major sources, properties and applications as thermoplastic
870 materials: Elsevier, Amsterdam.

SC
871 Casas-Orozco, D., Villa, A. L., Bustamante, F., & González, L.-M. (2015). Process development
872 and simulation of pectin extraction from orange peels. Food and Bioproducts
873 Processing, 96, 86-98. doi: http://dx.doi.org/10.1016/j.fbp.2015.06.006

U
874 Cunningham, P., Ogale, A. A., Dawson, P., & Acton, J. (2000). Tensile properties of soy protein
875 isolate films produced by a thermal compaction technique. J Food Sci, 65(4), 668-671.
AN
876 Cutter, C. N. (2006). Opportunities for bio-based packaging technologies to improve the quality
877 and safety of fresh and further processed muscle foods. Meat Sci, 74(1), 131-142. doi:
878 http://dx.doi.org/10.1016/j.meatsci.2006.04.023
879 Chien, P.-J., Sheu, F., & Yang, F.-H. (2007). Effects of edible chitosan coating on quality and
M

880 shelf life of sliced mango fruit. Journal of Food Engineering, 78(1), 225-229. doi:
881 http://dx.doi.org/10.1016/j.jfoodeng.2005.09.022
882 Chivrac, F., Pollet, E., & Avérous, L. (2009). Progress in nano-biocomposites based on
D

883 polysaccharides and nanoclays. Materials Science and Engineering: R: Reports, 67(1),
884 1-17. doi: http://dx.doi.org/10.1016/j.mser.2009.09.002
TE

885 Darmadji, P., & Izumimoto, M. (1994). Effect of chitosan in meat preservation. Meat Sci, 38(2),
886 243-254. doi: 10.1016/0309-1740(94)90114-7
887 De Cindio, B., Gabriele, D., & Lupi, F. R. (2016). Pectin: Properties Determination and Uses
888 Encyclopedia of Food and Health (pp. 294-300). Oxford: Academic Press.
EP

889 de Moura, M. R., Mattoso, L. H. C., & Zucolotto, V. (2012). Development of cellulose-based
890 bactericidal nanocomposites containing silver nanoparticles and their use as active
891 food packaging. Journal of Food Engineering, 109(3), 520-524. doi:
892 http://dx.doi.org/10.1016/j.jfoodeng.2011.10.030
C

893 Dhall, R. K. (2013). Advances in Edible Coatings for Fresh Fruits and Vegetables: A Review. Crit
894 Rev Food Sci Nutr, 53(5), 435-450. doi: 10.1080/10408398.2010.541568
AC

895 Díaz-Mula, H. M., Serrano, M., & Valero, D. (2012). Alginate Coatings Preserve Fruit Quality and
896 Bioactive Compounds during Storage of Sweet Cherry Fruit. Food and Bioprocess
897 Technology, 5(8), 2990-2997. doi: 10.1007/s11947-011-0599-2
898 do Amaral, D. S., Cardelle-Cobas, A., do Nascimento, B. M., Monteiro, M. J., Madruga, M. S., &
899 Pintado, M. M. (2015). Development of a low fat fresh pork sausage based on chitosan
900 with health claims: impact on the quality, functionality and shelf-life. Food Funct, 6(8),
901 2768-2778. doi: 10.1039/c5fo00303b
902 Dogan, N., & McHugh, T. H. (2007). Effects of Microcrystalline Cellulose on Functional
903 Properties of Hydroxy Propyl Methyl Cellulose Microcomposite Films. J Food Sci, 72(1),
904 E016-E022. doi: 10.1111/j.1750-3841.2006.00237.x

35
ACCEPTED MANUSCRIPT
905 Dong, H., Cheng, L., Tan, J., Zheng, K., & Jiang, Y. (2004). Effects of chitosan coating on quality
906 and shelf life of peeled litchi fruit. Journal of Food Engineering, 64(3), 355-358. doi:
907 http://dx.doi.org/10.1016/j.jfoodeng.2003.11.003
908 Durango, A. M., Soares, N. F. F., & Andrade, N. J. (2006). Microbiological evaluation of an
909 edible antimicrobial coating on minimally processed carrots. Food Control, 17(5), 336-
910 341. doi: http://dx.doi.org/10.1016/j.foodcont.2004.10.024
911 Dutta, P. K., Tripathi, S., Mehrotra, G. K., & Dutta, J. (2009). Perspectives for chitosan based
912 antimicrobial films in food applications. Food Chemistry, 114(4), 1173-1182. doi:
913 http://dx.doi.org/10.1016/j.foodchem.2008.11.047

PT
914 Elsabee, M. Z., & Abdou, E. S. (2013). Chitosan based edible films and coatings: A review.
915 Materials Science and Engineering: C, 33(4), 1819-1841. doi:
916 http://dx.doi.org/10.1016/j.msec.2013.01.010
917 Embuscado, M. E., & Huber, K. C. (2009). Edible films and coatings for food applications:

RI
918 Springer.
919 Erdohan, Z. Ö., & Turhan, K. N. (2005). Barrier and mechanical properties of methylcellulose–
920 whey protein films. Packaging Technology and Science, 18(6), 295-302. doi:

SC
921 10.1002/pts.700
922 Esa, F., Tasirin, S. M., & Rahman, N. A. (2014). Overview of Bacterial Cellulose Production and
923 Application. Agriculture and Agricultural Science Procedia, 2, 113-119. doi:

U
924 http://dx.doi.org/10.1016/j.aaspro.2014.11.017
925 Espitia, P. J. P., Du, W.-X., Avena-Bustillos, R. d. J., Soares, N. d. F. F., & McHugh, T. H. (2014).
AN
926 Edible films from pectin: Physical-mechanical and antimicrobial properties - A review.
927 Food Hydrocolloids, 35, 287-296. doi:
928 http://dx.doi.org/10.1016/j.foodhyd.2013.06.005
929 Fakhouri, F. M., Martelli, S. M., Caon, T., Velasco, J. I., & Mei, L. H. I. (2015). Edible films and
M

930 coatings based on starch/gelatin: Film properties and effect of coatings on quality of
931 refrigerated Red Crimson grapes. Postharvest Biology and Technology, 109, 57-64. doi:
932 http://dx.doi.org/10.1016/j.postharvbio.2015.05.015
D

933 Fakhreddin Hosseini, S., Rezaei, M., Zandi, M., & Ghavi, F. F. (2013). Preparation and functional
934 properties of fish gelatin-chitosan blend edible films. Food Chem, 136(3-4), 1490-1495.
TE

935 doi: 10.1016/j.foodchem.2012.09.081


936 Falguera, V., Quintero, J. P., Jiménez, A., Muñoz, J. A., & Ibarz, A. (2011). Edible films and
937 coatings: Structures, active functions and trends in their use. Trends in Food Science &
938 Technology, 22(6), 292-303. doi: http://dx.doi.org/10.1016/j.tifs.2011.02.004
EP

939 Ferrari, C. C., Sarantópoulos, C. I. G. L., Carmello-Guerreiro, S. M., & Hubinger, M. D. (2013).
940 Effect of Osmotic Dehydration and Pectin Edible Coatings on Quality and Shelf Life of
941 Fresh-Cut Melon. Food and Bioprocess Technology, 6(1), 80-91. doi: 10.1007/s11947-
942 011-0704-6
C

943 Galus, S., & Kadzińska, J. (2015). Food applications of emulsion-based edible films and
944 coatings. Trends in Food Science & Technology, 45(2), 273-283. doi:
AC

945 http://dx.doi.org/10.1016/j.tifs.2015.07.011
946 García, M. A., Martino, M. N., & Zaritzky, N. E. (1998). Plasticized Starch-Based Coatings To
947 Improve Strawberry (Fragaria × Ananassa) Quality and Stability. J Agric Food Chem,
948 46(9), 3758-3767. doi: 10.1021/jf980014c
949 Gennadios, A., Brandenburg, A. H., Park, J. W., Weller, C. L., & Testin, R. F. (1994). Water vapor
950 permeability of wheat gluten and soy protein isolate films. Industrial Crops and
951 Products, 2(3), 189-195. doi: http://dx.doi.org/10.1016/0926-6690(94)90035-3
952 Gennadios, A., Hanna, M. A., & Kurth, L. B. (1997). Application of edible coatings on meats,
953 poultry and seafoods: a review. LWT-Food Science and Technology, 30(4), 337-350.
954 Gennadios, A., Weller, C. L., & Gooding, C. H. (1994). Measurement errors in water vapor
955 permeability of highly permeable, hydrophilic edible films. Journal of Food
956 Engineering, 21(4), 395-409. doi: http://dx.doi.org/10.1016/0260-8774(94)90062-0

36
ACCEPTED MANUSCRIPT
957 George, J. (2012). High performance edible nanocomposite films containing bacterial cellulose
958 nanocrystals. Carbohydrate Polymers, 87(3), 2031-2037.
959 Goetz, L. A., Jalvo, B., Rosal, R., & Mathew, A. P. (2016). Superhydrophilic anti-fouling
960 electrospun cellulose acetate membranes coated with chitin nanocrystals for water
961 filtration. Journal of Membrane Science, 510, 238-248. doi:
962 http://dx.doi.org/10.1016/j.memsci.2016.02.069
963 Gutierrez-Pacheco, M. M., Ortega-Ramirez, L. A., Cruz-Valenzuela, M. R., Silva-Espinoza, B. A.,
964 Gonzalez-Aguilar, G. A., & Ayala-Zavala, J. F. (2016). Chapter 50 - Combinational
965 Approaches for Antimicrobial Packaging: Pectin and Cinnamon Leaf Oil A2 - Barros-

PT
966 Velázquez, Jorge Antimicrobial Food Packaging (pp. 609-617). San Diego: Academic
967 Press.
968 Han, J. H. (2014). Chapter 9 - Edible Films and Coatings: A Review Innovations in Food
969 Packaging (Second Edition) (pp. 213-255). San Diego: Academic Press.

RI
970 Hernandez-Izquierdo, V., & Krochta, J. (2008). Thermoplastic processing of proteins for film
971 formation—a review. J Food Sci, 73(2), R30-R39.
972 Isik, M., Sardon, H., & Mecerreyes, D. (2014). Ionic Liquids and Cellulose: Dissolution, Chemical

SC
973 Modification and Preparation of New Cellulosic Materials. Int J Mol Sci, 15(7), 11922.
974 Jia, D., Fang, Y., & Yao, K. (2009). Water vapor barrier and mechanical properties of konjac
975 glucomannan–chitosan–soy protein isolate edible films. Food and Bioproducts

U
976 Processing, 87(1), 7-10. doi: http://dx.doi.org/10.1016/j.fbp.2008.06.002
977 Jiménez, A., Fabra, M. J., Talens, P., & Chiralt, A. (2012). Edible and Biodegradable Starch Films:
AN
978 A Review. Food and Bioprocess Technology, 5(6), 2058-2076. doi: 10.1007/s11947-012-
979 0835-4
980 Jo, C., Lee, J. W., Lee, K. H., & Byun, M. W. (2001). Quality properties of pork sausage prepared
981 with water-soluble chitosan oligomer. Meat Sci, 59(4), 369-375.
M

982 Jonhed, A., Andersson, C., & Järnström, L. (2008). Effects of film forming and hydrophobic
983 properties of starches on surface sized packaging paper. Packaging Technology and
984 Science, 21(3), 123-135.
D

985 Kaur, S., & Dhillon, G. S. (2014). The versatile biopolymer chitosan: potential sources,
986 evaluation of extraction methods and applications. Crit Rev Microbiol, 40(2), 155-175.
TE

987 doi: 10.3109/1040841X.2013.770385


988 Kerch, G., & Korkhov, V. (2010). Effect of storage time and temperature on structure,
989 mechanical and barrier properties of chitosan-based films. European Food Research
990 and Technology, 232(1), 17-22. doi: 10.1007/s00217-010-1356-x
EP

991 Kester, J. J., & Fennema, O. (1989). An Edible Film of Lipids and Cellulose Ethers: Barrier
992 Properties to Moisture Vapor Transmission and Structural Evaluation. J Food Sci, 54(6),
993 1383-1389. doi: 10.1111/j.1365-2621.1989.tb05118.x
994 Khan, R. A., Salmieri, S., Dussault, D., Uribe-Calderon, J., Kamal, M. R., Safrany, A., & Lacroix, M.
C

995 (2010). Production and properties of nanocellulose-reinforced methylcellulose-based


996 biodegradable films. J Agric Food Chem, 58(13), 7878-7885.
AC

997 Kim, K. M., Son, J. H., Kim, S.-K., Weller, C. L., & Hanna, M. A. (2006). Properties of Chitosan
998 Films as a Function of pH and Solvent Type. J Food Sci, 71(3), E119-E124. doi:
999 10.1111/j.1365-2621.2006.tb15624.x
1000 Klemm, D., Heublein, B., Fink, H. P., & Bohn, A. (2005). Cellulose: Fascinating Biopolymer and
1001 Sustainable Raw Material. Angewandte Chemie International Edition, 44(22), 3358-
1002 3393. doi: 10.1002/anie.200460587
1003 Kolybaba, M., Tabil, L., Panigrahi, S., Crerar, W., Powell, T., & Wang, B. (2006). Biodegradable
1004 polymers: past, present, and future. Paper presented at the ASABE/CSBE North Central
1005 Intersectional Meeting.
1006 Lekjing, S. (2016). A chitosan-based coating with or without clove oil extends the shelf life of
1007 cooked pork sausages in refrigerated storage. Meat Sci, 111, 192-197. doi:
1008 10.1016/j.meatsci.2015.10.003

37
ACCEPTED MANUSCRIPT
1009 Li, H., & Yu, T. (2001). Effect of chitosan on incidence of brown rot, quality and physiological
1010 attributes of postharvest peach fruit. Journal of the Science of Food and Agriculture,
1011 81(2), 269-274. doi: 10.1002/1097-0010(20010115)81:2<269::AID-JSFA806>3.0.CO;2-F
1012 Lindman, B., Karlström, G., & Stigsson, L. (2010). On the mechanism of dissolution of cellulose.
1013 Journal of Molecular Liquids, 156(1), 76-81. doi:
1014 http://dx.doi.org/10.1016/j.molliq.2010.04.016
1015 Lorevice, M. V., Otoni, C. G., Moura, M. R. d., & Mattoso, L. H. C. (2016). Chitosan
1016 nanoparticles on the improvement of thermal, barrier, and mechanical properties of
1017 high- and low-methyl pectin films. Food Hydrocolloids, 52, 732-740. doi:

PT
1018 http://dx.doi.org/10.1016/j.foodhyd.2015.08.003
1019 Ma, Q., Hu, D., Wang, H., & Wang, L. (2016). Tara gum edible film incorporated with oleic acid.
1020 Food Hydrocolloids, 56, 127-133. doi:
1021 http://dx.doi.org/10.1016/j.foodhyd.2015.11.033

RI
1022 Ma, Q., Hu, D., & Wang, L. (2016). Preparation and physical properties of tara gum film
1023 reinforced with cellulose nanocrystals. Int J Biol Macromol, 86, 606-612. doi:
1024 http://dx.doi.org/10.1016/j.ijbiomac.2016.01.104

SC
1025 Maftoonazad, N., Ramaswamy, H. S., & Marcotte, M. (2007). EVALUATION OF FACTORS
1026 AFFECTING BARRIER, MECHANICAL AND OPTICAL PROPERTIES OF PECTIN-BASED FILMS
1027 USING RESPONSE SURFACE METHODOLOGY. Journal of Food Process Engineering,

U
1028 30(5), 539-563.
1029 Mali, S., Grossmann, M. V. E., Garcı ́a, M. A., Martino, M. N., & Zaritzky, N. E. (2004). Barrier,
AN
1030 mechanical and optical properties of plasticized yam starch films. Carbohydrate
1031 Polymers, 56(2), 129-135.
1032 Marcos, B., Aymerich, T., Monfort, J. M., & Garriga, M. (2010). Physical performance of
1033 biodegradable films intended for antimicrobial food packaging. J Food Sci, 75(8), E502-
M

1034 E507.
1035 Moalemiyan, M., Ramaswamy, H. S., & Maftoonazad, N. (2012). PECTIN-BASED EDIBLE
1036 COATING FOR SHELF-LIFE EXTENSION OF ATAULFO MANGO. Journal of Food Process
D

1037 Engineering, 35(4), 572-600. doi: 10.1111/j.1745-4530.2010.00609.x


1038 Molavi, H., Behfar, S., Shariati, M. A., Kaviani, M., & Atarod, S. (2015). A REVIEW ON
TE

1039 BIODEGRADABLE STARCH BASED FILM. The Journal of Microbiology, Biotechnology and
1040 Food Sciences, 4(5), 456.
1041 Moreira, M. d. R., Roura, S. I., & Ponce, A. (2011). Effectiveness of chitosan edible coatings to
1042 improve microbiological and sensory quality of fresh cut broccoli. LWT - Food Science
EP

1043 and Technology, 44(10), 2335-2341. doi: http://dx.doi.org/10.1016/j.lwt.2011.04.009


1044 Motedayen, A. A., Khodaiyan, F., & Salehi, E. A. Development and characterisation of
1045 composite films made of kefiran and starch. Food Chemistry, 136(3–4), 1231-1238. doi:
1046 http://dx.doi.org/10.1016/j.foodchem.2012.08.073
C

1047 Munarin, F., Tanzi, M. C., & Petrini, P. (2012). Advances in biomedical applications of pectin
1048 gels. International Journal of Biological Macromolecules, 51(4), 681-689. doi:
AC

1049 http://dx.doi.org/10.1016/j.ijbiomac.2012.07.002
1050 Muscat, D., Adhikari, B., Adhikari, R., & Chaudhary, D. S. (2012). Comparative study of film
1051 forming behaviour of low and high amylose starches using glycerol and xylitol as
1052 plasticizers. Journal of Food Engineering, 109(2), 189-201. doi:
1053 http://dx.doi.org/10.1016/j.jfoodeng.2011.10.019
1054 Nazan Turhan, K., & Şahbaz, F. (2004). Water vapor permeability, tensile properties and
1055 solubility of methylcellulose-based edible films. Journal of Food Engineering, 61(3),
1056 459-466. doi: http://dx.doi.org/10.1016/S0260-8774(03)00155-9
1057 Nguyen, V., Gidley, M., & Dykes, G. (2008). Potential of a nisin-containing bacterial cellulose
1058 film to inhibit Listeria monocytogenes on processed meats. Food Microbiol, 25(3), 471-
1059 478.

38
ACCEPTED MANUSCRIPT
1060 No, H., Meyers, S., Prinyawiwatkul, W., & Xu, Z. (2007). Applications of chitosan for
1061 improvement of quality and shelf life of foods: a review. J Food Sci, 72(5), R87-R100.
1062 Olivas, G. I., & Barbosa-Cánovas, G. V. (2008). Alginate–calcium films: Water vapor
1063 permeability and mechanical properties as affected by plasticizer and relative
1064 humidity. LWT - Food Science and Technology, 41(2), 359-366. doi:
1065 http://dx.doi.org/10.1016/j.lwt.2007.02.015
1066 Oms-Oliu, G., Soliva-Fortuny, R., & Martín-Belloso, O. (2008). Using polysaccharide-based
1067 edible coatings to enhance quality and antioxidant properties of fresh-cut melon. LWT-
1068 Food Science and Technology, 41(10), 1862-1870.

PT
1069 Park, S. Y., Marsh, K. S., & Rhim, J. W. (2002). Characteristics of Different Molecular Weight
1070 Chitosan Films Affected by the Type of Organic Solvents. J Food Sci, 67(1), 194-197.
1071 doi: 10.1111/j.1365-2621.2002.tb11382.x
1072 Pelissari, F. M., Yamashita, F., & Grossmann, M. V. E. (2011). Extrusion parameters related to

RI
1073 starch/chitosan active films properties. International Journal of Food Science &
1074 Technology, 46(4), 702-710. doi: 10.1111/j.1365-2621.2010.02533.x
1075 Peressini, D., Bravin, B., Lapasin, R., Rizzotti, C., & Sensidoni, A. (2003). Starch–methylcellulose

SC
1076 based edible films: rheological properties of film-forming dispersions. Journal of Food
1077 Engineering, 59(1), 25-32. doi: http://dx.doi.org/10.1016/S0260-8774(02)00426-0
1078 Pérez-Gago, M. B., & Rhim, J.-W. (2014). Chapter 13 - Edible Coating and Film Materials: Lipid

U
1079 Bilayers and Lipid Emulsions Innovations in Food Packaging (Second Edition) (pp. 325-
1080 350). San Diego: Academic Press.
AN
1081 Persin, Z., Stana-Kleinschek, K., Foster, T. J., van Dam, J. E. G., Boeriu, C. G., & Navard, P.
1082 (2011). Challenges and opportunities in polysaccharides research and technology: The
1083 EPNOE views for the next decade in the areas of materials, food and health care.
1084 Carbohydrate Polymers, 84(1), 22-32. doi:
M

1085 http://dx.doi.org/10.1016/j.carbpol.2010.11.044
1086 Psomiadou, E., Arvanitoyannis, I., & Yamamoto, N. (1996). Edible films made from natural
1087 resources; microcrystalline cellulose (MCC), methylcellulose (MC) and corn starch and
D

1088 polyols—Part 2. Carbohydrate Polymers, 31(4), 193-204. doi:


1089 http://dx.doi.org/10.1016/S0144-8617(96)00077-X
TE

1090 Qiu, X., & Hu, S. (2013). “Smart” materials based on cellulose: a review of the preparations,
1091 properties, and applications. Materials, 6(3), 738-781.
1092 Rhim. (2004). Physical and mechanical properties of water resistant sodium alginate films. LWT
1093 - Food Science and Technology, 37(3), 323-330. doi:
EP

1094 http://dx.doi.org/10.1016/j.lwt.2003.09.008
1095 Rhim, Wang, L. F., & Hong, S. I. (2013). Preparation and characterization of agar/silver
1096 nanoparticles composite films with antimicrobial activity. Food Hydrocolloids, 33(2),
1097 327-335. doi: 10.1016/j.foodhyd.2013.04.002
C

1098 Rojas-Graü, M. A., Tapia, M. S., & Martín-Belloso, O. (2008). Using polysaccharide-based edible
1099 coatings to maintain quality of fresh-cut Fuji apples. LWT - Food Science and
AC

1100 Technology, 41(1), 139-147. doi: http://dx.doi.org/10.1016/j.lwt.2007.01.009


1101 Rozenberga, L., Skute, M., Belkova, L., Sable, I., Vikele, L., Semjonovs, P., . . . Paegle, L. (2016).
1102 Characterisation of films and nanopaper obtained from cellulose synthesised by acetic
1103 acid bacteria. Carbohydrate Polymers, 144, 33-40. doi:
1104 http://dx.doi.org/10.1016/j.carbpol.2016.02.025
1105 Rubilar, J. F., Zúñiga, R. N., Osorio, F., & Pedreschi, F. (2015). Physical properties of emulsion-
1106 based hydroxypropyl methylcellulose/whey protein isolate (HPMC/WPI) edible films.
1107 Carbohydrate Polymers, 123, 27-38. doi:
1108 http://dx.doi.org/10.1016/j.carbpol.2015.01.010
1109 Santacruz, S., Rivadeneira, C., & Castro, M. (2015). Edible films based on starch and chitosan.
1110 Effect of starch source and concentration, plasticizer, surfactant's hydrophobic tail

39
ACCEPTED MANUSCRIPT
1111 and mechanical treatment. Food Hydrocolloids, 49, 89-94. doi:
1112 http://dx.doi.org/10.1016/j.foodhyd.2015.03.019
1113 Seligra, P. G., Medina Jaramillo, C., Famá, L., & Goyanes, S. (2016). Biodegradable and non-
1114 retrogradable eco-films based on starch–glycerol with citric acid as crosslinking agent.
1115 Carbohydrate Polymers, 138, 66-74. doi:
1116 http://dx.doi.org/10.1016/j.carbpol.2015.11.041
1117 Sessini, V., Arrieta, M. P., Kenny, J. M., & Peponi, L. (2016). Processing of edible films based on
1118 nanoreinforced gelatinized starch. Polymer Degradation and Stability. doi:
1119 10.1016/j.polymdegradstab.2016.02.026

PT
1120 Shen, Z., & Kamdem, D. P. (2015). Development and characterization of biodegradable
1121 chitosan films containing two essential oils. Int J Biol Macromol, 74, 289-296. doi:
1122 10.1016/j.ijbiomac.2014.11.046
1123 Shoda, M., & Sugano, Y. (2005). Recent advances in bacterial cellulose production.

RI
1124 Biotechnology and Bioprocess Engineering, 10(1), 1-8.
1125 Siracusa, V., Rocculi, P., Romani, S., & Rosa, M. D. (2008). Biodegradable polymers for food
1126 packaging: a review. Trends in Food Science & Technology, 19(12), 634-643. doi:

SC
1127 http://dx.doi.org/10.1016/j.tifs.2008.07.003
1128 Spotti, M. L., Cecchini, J. P., Spotti, M. J., & Carrara, C. R. (2016). Brea Gum (from Cercidium
1129 praecox) as a structural support for emulsion-based edible films. LWT - Food Science

U
1130 and Technology, 68, 127-134. doi: http://dx.doi.org/10.1016/j.lwt.2015.12.018
1131 Srinivasa, P., Baskaran, R., Ramesh, M., Harish Prashanth, K., & Tharanathan, R. (2002). Storage
AN
1132 studies of mango packed using biodegradable chitosan film. European Food Research
1133 and Technology, 215(6), 504-508. doi: 10.1007/s00217-002-0591-1
1134 Su, J., & Zhang, J. (2015). Comparison of rheological, mechanical, electrical properties of HDPE
1135 filled with BaTiO3 with different polar surface tension. Applied Surface Science. doi:
M

1136 10.1016/j.apsusc.2015.10.156
1137 Tapia, M. S., Rojas-Graü, M. A., Carmona, A., Rodríguez, F. J., Soliva-Fortuny, R., & Martin-
1138 Belloso, O. (2008). Use of alginate- and gellan-based coatings for improving barrier,
D

1139 texture and nutritional properties of fresh-cut papaya. Food Hydrocolloids, 22(8),
1140 1493-1503. doi: http://dx.doi.org/10.1016/j.foodhyd.2007.10.004
TE

1141 Tripathi, S., Mehrotra, G. K., & Dutta, P. K. (2011). Chitosan–silver oxide nanocomposite film:
1142 Preparation and antimicrobial activity. Bulletin of Materials Science, 34(1), 29-35. doi:
1143 10.1007/s12034-011-0032-5
1144 Valenzuela, C., Abugoch, L., & Tapia, C. (2013). Quinoa protein–chitosan–sunflower oil edible
EP

1145 film: Mechanical, barrier and structural properties. LWT - Food Science and
1146 Technology, 50(2), 531-537. doi: http://dx.doi.org/10.1016/j.lwt.2012.08.010
1147 van den Broek, L. A. M., Knoop, R. J. I., Kappen, F. H. J., & Boeriu, C. G. (2015). Chitosan films
1148 and blends for packaging material. Carbohydrate Polymers, 116, 237-242. doi:
C

1149 http://dx.doi.org/10.1016/j.carbpol.2014.07.039
1150 Varela, P., & Fiszman, S. M. (2011). Hydrocolloids in fried foods. A review. Food Hydrocolloids,
AC

1151 25(8), 1801-1812. doi: http://dx.doi.org/10.1016/j.foodhyd.2011.01.016


1152 Vargas, M., & Gonzalez-Martinez, C. (2010). Recent patents on food applications of chitosan.
1153 Recent Pat Food Nutr Agric, 2(2), 121-128.
1154 Vásconez, M. B., Flores, S. K., Campos, C. A., Alvarado, J., & Gerschenson, L. N. (2009).
1155 Antimicrobial activity and physical properties of chitosan–tapioca starch based edible
1156 films and coatings. Food Research International, 42(7), 762-769. doi:
1157 http://dx.doi.org/10.1016/j.foodres.2009.02.026
1158 Vieira, M. G. A., da Silva, M. A., dos Santos, L. O., & Beppu, M. M. (2011). Natural-based
1159 plasticizers and biopolymer films: A review. European Polymer Journal, 47(3), 254-263.
1160 doi: http://dx.doi.org/10.1016/j.eurpolymj.2010.12.011
1161 Villalobos, R., Chanona, J., Hernández, P., Gutiérrez, G., & Chiralt, A. (2005). Gloss and
1162 transparency of hydroxypropyl methylcellulose films containing surfactants as affected

40
ACCEPTED MANUSCRIPT
1163 by their microstructure. Food Hydrocolloids, 19(1), 53-61. doi:
1164 http://dx.doi.org/10.1016/j.foodhyd.2004.04.014
1165 Wang, S., Lu, A., & Zhang, L. (2016). Recent advances in regenerated cellulose materials.
1166 Progress in Polymer Science, 53, 169-206. doi:
1167 http://dx.doi.org/10.1016/j.progpolymsci.2015.07.003
1168 Willats, W. G. T., Knox, J. P., & Mikkelsen, J. D. (2006). Pectin: new insights into an old polymer
1169 are starting to gel. Trends in Food Science & Technology, 17(3), 97-104. doi:
1170 http://dx.doi.org/10.1016/j.tifs.2005.10.008
1171 Xiong, H., Tang, S., Tang, H., & Zou, P. (2008). The structure and properties of a starch-based

PT
1172 biodegradable film. Carbohydrate Polymers, 71(2), 263-268.
1173 Xu, Q., Chen, C., Rosswurm, K., Yao, T., & Janaswamy, S. (2016). A facile route to prepare
1174 cellulose-based films. Carbohydrate Polymers, 149, 274-281. doi:
1175 http://dx.doi.org/10.1016/j.carbpol.2016.04.114

RI
1176 Yeul, V. S., & Rayalu, S. S. (2012). Unprecedented Chitin and Chitosan: A Chemical Overview.
1177 Journal of Polymers and the Environment, 21(2), 606-614. doi: 10.1007/s10924-012-
1178 0458-x

SC
1179 Youssef, A. M., Abou-Yousef, H., El-Sayed, S. M., & Kamel, S. (2015). Mechanical and
1180 antibacterial properties of novel high performance chitosan/nanocomposite films. Int J
1181 Biol Macromol, 76, 25-32. doi: 10.1016/j.ijbiomac.2015.02.016

U
1182 Zamudio-Flores, P.B., Ochoa-Reyes, E., Ornelas-Paz, J. De J., Aparicio-Saguilán, A., Vargas-
1183 Torres, A., Bello-Pérez, L. A., Rubio-Ríos, A., & Cárdenas-Félix, R.G. (2015). Effect of
AN
1184 storage time on physicochemical and textural properties of sausages covered with
1185 oxidized banana starch film with and without betalains. CyTA - Journal of Food, 13(3),
1186 456-463. doi: 10.1080/19476337.2014.998713
1187 Zhang, Y., Han, J., & Liu, Z. (2008). 5 - Starch-based edible films A2 - Chiellini, Emo
M

1188 Environmentally Compatible Food Packaging (pp. 108-136): Woodhead Publishing.


1189 Zivanovic, S., Chi, S., & Draughon, A. F. (2005). Antimicrobial activity of chitosan films enriched
1190 with essential oils. J Food Sci, 70(1), M45-M51.
D

1191 Zolfi, M., Khodaiyan, F., Mousavi, M., & Hashemi, M. (2014). The improvement of
1192 characteristics of biodegradable films made from kefiran–whey protein by
TE

1193 nanoparticle incorporation. Carbohydrate Polymers, 109, 118-125. doi:


1194 http://dx.doi.org/10.1016/j.carbpol.2014.03.018
1195
EP

1196
C
AC

41
ACCEPTED MANUSCRIPT
1197 Table captions

1198
1199 Table 1. Water vapor permeability of several polysaccharide-based films.

1200

1201 Table 2. Water vapour permeability of several synthetic polymers. Modified from

PT
1202 (Gennadios, Brandenburg, Park, Weller, & Testin, 1994)

1203

RI
1204 Table 3. Tensile properties of several polysaccharide-based films.

SC
1205

1206 Table 4. Tensile strength and elongation at break of several synthetic polymers.

U
AN
M
D
TE
C EP
AC

42
ACCEPTED MANUSCRIPT

1207 Table 1. Water vapor permeability (WVP) of several polysaccharide-based films. *In some data, the units were normalised.

Film composition WVP Temperature Relative

PT
References
(g/m s Pa)* (°C) humidity (%)
Cellulose derivatives (methylcellulose and

RI
hydroxypropyl methylcellulose mixture) with
6.615x10-13 - 1.632x10-11 25 65 - 97 Kester and Fennema (1989)

SC
polyethylene glycol and stearic/palmitic fatty with a
beeswax surface coat
Nazan Turhan and Şahbaz

U
Methylcellulose mixtures with ethanol 0.51x10-10 - 1.08x10-10 25 52
(2004)

AN
-11 -11
Methylcellulose mixtures with/without nanocellulose 5.44x10 - 7.29x10 25 60 Khan et al. (2010)
Hydroxypopyl methylcellulose mixtures with/without
1.54x10-10 - 2.22x10-10

M
25 70.5 - 71.9 de Moura et al. (2012)
silver nanoparticles
Agar mixtures with/without silver nanoparticles 1.67x10-9 - 1.97x10-9 25 50 Rhim et al. (2013)

D
-10
Chitosan 1.05x10 22 58 Abugoch et al. (2011)

TE
Chitosan and starch mixtures 1.18x10-10 - 1.55x10-10 25 100 Santacruz et al. (2015)

Chitosan mixtures with different molecular weight and


EP
0.32x10-9 - 0.51x10-9 25 50 Park et al. (2002)
different solvents

1.77x10-9 - 2.26x10-8
C

Chitosan mixtures with different solvents and pH 25 50 Kim et al. (2006)


AC

Chitosan-tapioca starch mixtures 2.8x10-10 - 6.7x10-10 32 65 Vásconez et al. (2009)

Tapioca starch mixtures 12.1x10-10 32 65 Vásconez et al. (2009)

43
ACCEPTED MANUSCRIPT

Cassava starch mixtures with glycerol 1.8x10-10 - 2.8x10-10 25 70 Seligra et al. (2016)
-10 -10
Tara gum mixtures with sorbitol and glycerol 0.52x10 - 0.69x10 25 75 Ma, Hu, Wang, et al. (2016)

PT
Tara gum mixtures with glycerol and with/without
9.91x10-11-1.28x10-10 25 53 Antoniou et al. (2015)
chitosan
2.5x10-11 - 4.75x10-11

RI
Brean gum with glycerol and beeswax mixtures 32 75.1 Spotti et al. (2016)
Starch mixtures with glycerol and xylitol as
1.94x10-11 - 1.77x10-10 20 52.9 Muscat et al. (2012)

SC
plasticizers
Olivas and Barbosa-Cánovas
-10 -9
Alginate mixtures with/without polyethylene glycol 6.5x10 - 0.93x10 29 - 25 76 - 50 (2008)

U
Rhim (2004)

AN
-9 -9
Agar with/without silver nanoparticules 1.67x10 - 1.97x10 25 50 Rhim et al. (2013)
-10
Quinoa and chitosan mixtures 2.61x10 22 58 Abugoch et al. (2011)

M
1208

D
TE
C EP
AC

44
ACCEPTED MANUSCRIPT

1209 Table 2. Water vapour permeability (WVP) of several synthetic polymers. Modified from Gennadios, Brandenburg, Park, Weller, and Testin (1994)

Film WVP (g/m s Pa)

PT
Polyvinylidene chloride 0.7x10-13 - 2.4x10-13
High-density polyethylene 2.4x10-13

RI
Polypropylene 4.9x10-13
Low density polyethylene 7.3x10-13 - 9.7x10-13

SC
Ethylene-vinylacetate 2.4x10-12 - 4.9x10-12
Polyester 1.2x10-12 - 1.5x10-12

U
Cellulose acetate 0.5x10-11 - 1.6x10-11

AN
1210

1211

M
1212

D
1213

1214
TE
EP
1215
C
AC

45
ACCEPTED MANUSCRIPT

1216 Table 3. Tensile properties and elongation at break of several polysaccharide-based films. NR = Data not reported

Tensile strength Elongation Temperature Relative

PT
Film composition References
(N/mm2) (%) (°C) humidity (%)
Methylcellulose and ethanol mixtures
25 - 33 29 - 14 25 52 Nazan Turhan and Şahbaz (2004)

RI
with/without polyethylene glycol
Hydroxypopyl methylcellulose mixtures
51.0 - 28.3 NR 24 NR de Moura et al. (2012)
with/without silver nanoparticles

SC
Abugoch et al. (2011)
Chitosan 22.2 - 39.6 13 - 73.6 22 30-60
Shen and Kamdem (2015)

U
Chitosan mixtures with different molecular
6.7 - 150.2 4.1 - 117.8 25 50 Park et al. (2002)
weight and different solvents

AN
Chitosan mixtures with different solvents and
0.56 - 19.2 22 - 494.8 25 50 Kim et al. (2006)
pH
Quinoa-chitosan mixtures 2.3 - 8.3 273 - 117.4 22 60 Abugoch et al. (2011)

M
Starch mixtures with glycerol and xylitol as
5.06 - 44.3 2.4 - 70.7 25 53 Muscat et al. (2012)
plasticizers
Potato starch with/without glycerol 30 - 68 3-5 NR NR (Sessini et al., 2016)

D
Tara gum mixtures with sorbitol and glycerol
26.8 - 57.4 8.5 - 2.7 NR NR Ma, Hu, Wang, et al. (2016)

TE
as plasticizers and with/without oleic acid
Tara gum mixtures with glycerol and
58.44 - 22.71 44 - 46 25 53 Antoniou et al. (2015)
with/without chitosan
Brean gum with glycerol and with/without
EP
1.64 - 7.58 8.22 - 4.85 NR NR Spotti et al. (2016)
beeswax mixtures
Pectin mixtures with/without chitosan
58.5 - 26.07 2.91 - 0.94 27 50 Lorevice et al. (2016)
C

nanoparticles
Alginate mixtures with sorbitol or glycerol as
AC

65.9 - 64.7 2.5 - 2.8 25 56 Olivas and Barbosa-Cánovas (2008)


plasticizers
Alginate mixtures with sorbitol or glycerol as
18.4 - 24.1 6.6 - 7.9 25 98 Olivas and Barbosa-Cánovas (2008)
plasticizers
Sodium alginate mixture 33.6 - 75.8 3.4 - 14.0 25 50 Rhim (2004)
Agar with/without silver nanoparticles 51.5 - 46.38 33.02-33.64 25 50 Rhim et al. (2013)

46
ACCEPTED MANUSCRIPT

1217 Table 4. Tensile strength and elongation at break of several synthetic polymers.

Tensile strength Elongation at break Reference

PT
Film
(MPa) (%)

RI
High density polyethylene 29.3 859 Su and Zhang (2015)
Polypropylene 31 - 38 Not reported Rhim et al. (2013)

SC
Boldt, Gohs, Wagenknecht, and Stamm (2016)
Low density polyethylene 20 633 AlMaadeed, Nógellová, Mičušík, Novák, and

U
Krupa (2014)

AN
Cellulose acetate 1.43 Not reported Goetz, Jalvo, Rosal, and Mathew (2016)
1218

M
1219

D
1220

TE
C EP
AC

47
ACCEPTED MANUSCRIPT
Highlights

- Edible and biodegradable films could be an alternative to synthetic packaging


films
- Polysaccharide-based films allow enhancing the shelf-life of ready-to-eat foods
- Polysaccharide-based films are an effective barrier to gas transference
- Polysaccharide-based films have poor water vapour permeability
- Tensile strength is similar to that of synthetic polymers

PT
RI
U SC
AN
M
D
TE
C EP
AC

You might also like