You are on page 1of 21

International Journal of Biological Macromolecules 104 (2017) 687–707

Contents lists available at ScienceDirect

International Journal of Biological Macromolecules


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

Review

Improving the integrity of natural biopolymer films used in food


packaging by crosslinking approach: A review
Farhad Garavand a , Milad Rouhi b , Seyed Hadi Razavi a,∗ , Ilaria Cacciotti c ,
Reza Mohammadi b
a
Bioprocess Engineering Laboratory (BPEL), Department of Food Science and Engineering, Faculty of Agricultural Engineering and Technology, University of
Tehran, Karaj, Iran
b
Department of Food Science and Technology, School of Nutrition Sciences and Food Technology, Kermanshah University of Medical Sciences, Kermanshah,
Iran
c
Department of Engineering, University of Rome “Niccolo Cusano”, INSTM RU, Via Don Carlo Gnocchi, 3, 00166 Rome, Italy

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

Article history: Currently used approaches for biopolymer modification are either expensive, poisonous or do not lead
Received 14 March 2017 into the well-desired characteristics to the final film materials. Development of crosslinking procedure is
Received in revised form 17 June 2017 an innovative strategy to improve mechanical, physical and thermal properties of biopolymer films. This
Accepted 21 June 2017
review provides a brief description of film-forming biopolymers (e.g. chitosan, whey protein, alginate
Available online 23 June 2017
and starch) followed by a detailed introduction to definition and classification of various crosslinkers,
the effect of crosslinking on emerging attributes of biopolymer films including physical, mechanical and
Keywords:
thermal properties, identification of crosslinking occurrence, and cytotoxicity status of commonly used
Biopolymer
Crosslinking
crosslinkers in the field of food and food-related packaging materials.
Film © 2017 Elsevier B.V. All rights reserved.
Food packaging
Mechanical properties

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 688
2. Biopolymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 688
3. Crosslinking agents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 689
3.1. Crosslinking at ambient temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 689
3.2. Crosslinking at intermediate temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 689
4. Crosslinking agents in biopolymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 690
5. Influence of crosslinking on biopolymers characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 691
5.1. Physical properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 691
5.1.1. Water absorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 691
5.1.2. Water vapor permeability (WVP) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 691
5.1.3. Solubility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 691
5.1.4. Film thickness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 698
5.1.5. Light barrier properties and transparency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 698
5.1.6. Color changes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 698
5.2. Mechanical properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 698
5.3. Thermal properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 699
5.3.1. Differential scanning calorimetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 700
5.3.2. Thermogravimetric analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 701

∗ Corresponding author.
E-mail address: srazavi@ut.ac.ir (S.H. Razavi).

http://dx.doi.org/10.1016/j.ijbiomac.2017.06.093
0141-8130/© 2017 Elsevier B.V. All rights reserved.
688 F. Garavand et al. / International Journal of Biological Macromolecules 104 (2017) 687–707

6. Complementary techniques for identification of crosslinking occurrence in biopolymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 701


6.1. X-ray diffraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 703
6.2. Sodium dodecyl sulfate–polyacrylamide gel electrophoresis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 703
6.3. Fourier transform infrared spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 703
6.4. Nuclear magnetic resonance (NMR) spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 704
7. Cytotoxicity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 704
8. Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 704
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 705

1. Introduction Crosslinking agents are suitable for biopolymer materials partic-


ularly those originated from carbohydrates or proteins, providing
The main mission of food packaging is to maintain the quality a reduced water vapor and gases permeability in food packaging
and safety of food products during storage and transportation, and materials. Crosslinkers can overcome the inherent deficiencies in
to broaden their shelf-life by avoiding unlikeable circumstances, mechanical and barrier properties of biopolymers, making them
such as hazardous microorganisms and their corresponding toxins, more applicable compared to petroleum-based counterparts for
external physical forces, chemicals compounds, sunlight, perme- food packaging purposes. A heat-light resistance biopolymer along
able volatile compounds, oxygen and moisture [1]. In order to with a reduced water solubility and swelling can be expected by
accomplish such a mission, packaging materials should arrange introducing crosslinking agents. Chemical crosslinking is the most
for physical protection and establish appropriate physiochemical accredited procedure than the others. Protein-containing biopoly-
and microbiological situations to products that are indispensable mers are inclined to participate more in the reactions induced by
to achieve an acceptable endurance, as well as uphold the quality- crosslinkers, because of their various reactive and functional groups
safety issues of food items [2,3]. [13]. The typical crosslinking process involves the use of a wide
Currently, a large number of petroleum based polymers are range of crosslinking agents (e.g. glutaraldehyde, epichlorohydrin,
used in food, beverage and foodstuff packaging industries [4], with Ca2+ ions, sodium trimetaphosphate, sodium benzoate, phosphoryl
consequent accumulation of non-degradable and non-recyclable chloride, citric acid, boric acid and Maillard reaction) [11,14–17].
waste materials and concerns over environmental issues. Thus, This review will present recent developments of crosslinked
there is an increasing demand for bio-based raw materials in order natural biopolymeric films for food packaging applications. It will
to unravel the waste disposal problems to an assured magni- start with an introduction to definition and classification of biopoly-
tude [5]. In this framework, biopolymers, particularly those from mers and crosslinkers, followed by a comprehensive report about
renewable organic resources, have been regarded as promising film-forming natural biopolymers. Then the effect of crosslink-
green substitutes of non-biodegradable plastic materials [6,7]. ing on emerging properties of film-forming biopolymers such
Numerous subjects must be addressed before commercial use of as chitosan, gelatin, whey protein and starch, will be discussed.
bio-based packaging raw materials. These subjects comprise the Finally, the different characterization methods to demonstrate the
rates of degradation in different situations, possible alterations of crosslinking occurrence and the environmental concerns regarding
some physical, mechanical and thermal properties during stor- crosslinkers will be presented and discussed.
age, their resistance against the growth of microorganisms, and
the risk of harmful compounds release into wrapped foodstuffs
[8]. Among biopolymer materials that can be used for construc- 2. Biopolymers
tion of biocomposite films, a lot of attention has been dedicated to
polysaccharides, proteins, lipids and polyesters alone or in combi- The use of plastics over glass and metals for packaging means
nation as materials for the film production [9]. Most of the protein has continued to increase due to their good properties, accessi-
and polysaccharide-based films commonly present suitable barrier bility, easy processability, low weight, and low cost. Classically,
properties against oxygen at low and intermediate relative humid- plastic materials are prepared by condensation polymerization
ity, and have relatively acceptable mechanical strength, alongside process or polymerization of hydrocarbon or hydrocarbon-like
poor moisture and water vapor barrier features, due to their high monomers from different raw materials. Therefore, they find
hydrophilic nature [9]. Although many studies were focused on their origin in the petrochemical industry, making them non-
improving the film-forming aspects of biopolymers and designated biodegradable and non-renewable [18]. With particular measures
a noteworthy progress in this field, some physical, structural, ther- taken to discover substitutes to petroleum-based materials and
mal, and mechanical attributes are not pleasing so far and there with regard to environmental issues, several studies have been
are a number of complications for their industrial scale-up and pushed to propose new biodegradable packaging materials and
employment. In recent years various strategies have been followed polymers achieved starting from renewable resources. Employing
to overcome such drawbacks, including chemical modifications, biopolymer-based materials can overcome waste disposal prob-
plasticizer addition, blending with other biodegradable polymers, lems to a considerable magnitude. Nevertheless, in some cases,
inclusion of nanoparticles as reinforcing fillers, and addition of weakened mechanical and physical attributes, as well as inappro-
compatibilizers to improve miscibility of some incompatible poly- priate gas and water vapor barrier properties, in comparison with
mers [10,11]. frequently used petroleum derived plastic materials, restrict their
In details, chemical processes, such as etherification, esterifica- commercial use. Consequently, several researches have been done
tion, grafting and crosslinking, are able to prevent the excessive to boost new procedures aimed to improve the mechanical and
moisture uptake and macromolecular reformations, in addition barrier traits of biopolymer-based packaging raw materials, along
to the accessibility and affordable costs compared to the novel with incorporation of bioactive compounds to provide antioxidant
approaches like the addition of nanoparticles as fillers [12]. and antimicrobial properties which are indispensible in the food
A polymer with an integrated network can be obtained by packaging sector [10,19].
crosslinking process, in which the polymer chains are connected The biodegradable biopolymers can be classified in three main
intra- or intermolecularly by covalent or non-covalent links. categories on the basis of their origin (Fig. 1) [10,20–23] as follows:
F. Garavand et al. / International Journal of Biological Macromolecules 104 (2017) 687–707 689

Fig. 1. Classification of different film-forming biopolymers.

1. Biopolymers originated from agricultural resources, including ing and post-crosslinking reactions can occur either at ambient
polysaccharides (e.g. starch), ligno-cellulosic products (e.g. cel- temperature or at intermediate temperature (<150 ◦ C) [25].
lulose and its derivatives), proteins (e.g. whey and collagen),
lipids (e.g. bee wax) and free fatty acids; 3.1. Crosslinking at ambient temperature
2. Biopolymers achieved by means of microbial fermentation, such
as pullulan and polyhydroxyalkanoates; Development of integrated networks in heat-sensitive sys-
3. Chemically synthesized biopolymers using monomers attained tems and improving energy efficiency are interesting subjects in
by natural raw materials such as poly(lactic acid). crosslinking reactions of polymers at ambient temperature. Sev-
eral examples of crosslinkers commonly used at room temperature
Recent advances in case of natural-based or natural-occurred are reported in Table 1, evidencing drawbacks, advantages and
biopolymer films are donating considerable progress in food applications [25]. Various fundamental functional groups includ-
packaging, pharmaceutical and surgery means. In particular, the ing carboxylic acid reactions, reactions which activated amines
present review will be mainly focused on the modification of agro- with carbonyl functions, acetoacetyl groups, acetals, acrylamide
resources derived biopolymers and on the influence of the added occupations, crosslinkers which capable to react with water
crosslinkers on their properties. and derivatives containing isocyanate functions, enzymatically
occurred reactions and physical crosslinking can be used in room
(ambient) temperature crosslinking. It is well understood that the
3. Crosslinking agents reactivity is a crucial limitation in these reactions and it should
be considered to avoid processes with over-reactive components
The crosslinking method involves the formation of chemical prior to introducing crosslinkers. This type of crosslinking is exten-
bonds between different molecular chains to generate a stronger sively used in biology, coatings and paints, particularly to develop
three-dimensional network (Fig. 2). There are various classifica- chemical resistance and improved mechanical properties.
tions for the employed crosslinking agents. They can be classified
on the basis of the occurred bond kind, leading to well-stabilized 3.2. Crosslinking at intermediate temperature
covalent crosslinking, formation of ionic bonds, and physical
crosslinking created by hydrogen and Van der Waals bonds. In other Table 2 shows main crosslinking agents involved in inter-
classifications, crosslinkers are divided into three main types: phys- mediate temperature approach [25]. Intermediate temperatures
ical, chemical, and enzymatical [24]. In the case of polymers, it are referred to temperatures in the range of 40–150 ◦ C, while
is possible to follow two different approaches: crosslinking and high temperatures are 150–300 ◦ C. According to Table 2, car-
post-crosslinking. In the former case, the crosslinker is directly boxylic acid reacts with oxazoline-epoxide functions, azetidines,
added during the polymer preparation in order to obtain the alcohols reacting with sheltered urethanes, azlactones and methy-
ultimate material at one stage, whereas in the latter case the lol amide, reversible Diels–Alder reaction, and Huisgen reactions
crosslinking occurs after polymer formation. Moreover, if the between azido and triple bonds (click chemistry) are key functions
desired efficiency was not achieved, a complementary stage (post- involved in intermediate crosslinking. The above-mentioned reac-
crosslinking) should be considered, although pre-crosslinking may tions require particular conditions and in certain circumstances
occur as a result of the first-stage crosslinking. Moreover, crosslink- (e.g. bisfuran/bismaleimide) they may be reversible as the temper-
690 F. Garavand et al. / International Journal of Biological Macromolecules 104 (2017) 687–707

Fig. 2. A schematic of the crosslinker’s mode of action in biopolymer chains.

Table 1
Main crosslinking agents applied at ambient (room) temperature.

Crosslinking agent Examples Drawbacks Benefits Applications

Carboxylic acid Aziridine and carbodiimide Poor water stability, and some Capability to use in in vivo Emulsions preparation and
irritant, toxic and mutagen conditions fabrication of composite films
properties and coatings
Acetoacetyl groups Acetylacetonate, pH- dependency Reaction with various active Polymer synthesis, acrylic
acetoacetoxyethyl groups, reduction of the latexes production
methacrylate viscosity and the glass
transition temperature of
polymers, less sensitivity to
hydrolysis
Amines Hydrazine derivatives such Time consuming of some Occurring several reactions at Use in waterborne acrylic
as piperazine and tertiary reactions room temperature because of dispersions
amines high nucleophily
Acetal function N-(2,2-dimethoxyethyl) Self- reaction under acidic pH responsive Production of acrylic coatings
methacrylamide conditions crosslinking agent and chemically stable latex for
paints
Acrylamide derivative N-methylolacrylamide and Formation of toxic Copolymerization with Self-crosslinkable latexes
isobutoxymethylacry- formaldehyde monomers containing
lamide by-products due to undesirable hydroxyl groups and
side reactions self-crosslinkability
Enzymes Transglutaminase Heat sensitivity, specificity Bio-based environment friend Blend film especially
crosslinking agent protein-polysaccharide ones
Physical crosslinkers Ultraviolet beams, ␥-rays, Inherent dangers due to Successful crosslinking Genetic aims, hydrogel
electron beams, ozone, working with radiation between proteins and nucleic formation and etc.
X-rays etc. acids
Other crosslinking moieties Water (isocyanates and Shrinkage and fracture Production of homogenous Generation of silane
sol–gel reactions) during the curing process networks with high purity at latex and emulsions
low temperature, coexisting of
organic and inorganic phases
to form polymers with hybrid
materials, green process
Oxygen (reaction Possibility of some unwanted Induction of the inter-chain Alkyde resins, mummification
generating peroxide and plasticizing effect due to long crosslinking though various and dental printings
acetylenic coupling) side chains side reactions

ature reaches to about 150 ◦ C. Well-known uses of such reactions divided into two main classes depending on their molecular weight,
are strengthening some properties in coatings, composite films, namely high molecular weight and low molecular weight with
self-healing polymers and textile finishing. molecular weights more than 105 g/mol and lower than 104 g/mol
(frequently in the range of 2–3 × 103 g/mol), respectively. Defi-
nitely, low molecular-weight biopolymers often need a consecutive
4. Crosslinking agents in biopolymers
crosslinking approach to achieve acceptable mechanical, thermal
and structural properties [25]. Some examples of the biopolymers
The possibility to introduce crosslinking agents into the
modified by such strategy are summarized in Table 3. As it is
structure of biopolymers extremely depends on the chemical
observed carboxylic acids and calcium ions improved the physio-
arrangement, presence of active groups and molecular weight of
chemical, thermal and mechanical properties of most biopolymers
biopolymers as well as compatibility of crosslinker and polymer
such as alginate, pectin, whey proteins, chitosan, starch and gelatin.
to form appropriate interactions. Accordingly, biopolymers can be
F. Garavand et al. / International Journal of Biological Macromolecules 104 (2017) 687–707 691

Table 2
Main crosslinking agents applied at intermediate temperatures.

Crosslinking agent Examples Drawbacks Benefits Applications

Carboxylic acid Reaction with epoxides and Not reported Production of durable coatings, Powder coatings and
reaction involving oxazolines easily introduction into the crosslinking of polyesters
polymers and
multi-functionality
Alcohol function Protected urethanes, azlactone Evolution of blocking agents Exhibiting UV-proofing Use in glass or glossy
function and methylol amide properties coatings and
function thermo-crosslinkable gels
Azetidine functions Acrylate and methacrylate Not reported Possessing double reactivity, Copolymers
monomers self- crosslinkability, and avoid
wool-felting during washing
Diels–Alder reactions Furane/bismaleimide reaction Sometimes reversibility is not Providing wide variety of Polyacrylic-type coatings
quantitative functional groups for both
polymerization and
crosslinking, obtaining
thermally reversible
crosslinked polymers
1,3-Dipolar cycloaddition Huisgen reactions between Explosiveness and toxicity of azides Reactions occur at low Synthesize some
and “click chemistry” azido and triple bond temperatures (<100 ◦ C), high organogels
reaction yields and tolerant to the
media condition

5. Influence of crosslinking on biopolymers characteristics film structure. On the other hand, crosslinker addition may intro-
duce a tortuous pathway for water molecules to transfer through
5.1. Physical properties the matrix. It is also observed that higher concentrations of citric
acid, from 10% to 20% (w/w), cause considerable enhancement in
5.1.1. Water absorption WVP, probably due to the plasticizing role of extra-added citric acid.
Kinetic diagrams of water absorption are imperative means to Free excessive citric acid molecules may be included between the
anticipate the stability and quality attributes of packaging materi- polymer chains, increasing the mobility of the inter-chain and chain
als during loading, transport and storage of various foodstuffs [54]. spaces and, thus, encouraging the film diffusivity and accelerating
In particular, crosslinking of polysaccharide-based films strength- the rate of water vapor transmission. In opposition with the results
ens the intermolecular interactions by generating covalent bonds reported the remarkable reduction in WVP as affected by incor-
that enhance natural intermolecular hydrogen bonds, resulting poration of crosslinkers, Cao et al. [28] pointed out that different
in improved water resistibility [43]. For example, crosslinking of concentrations of crosslinking agents (i.e. ferulic acid and tannin
O-acetyl galactoglucomannan, a softwood-derived polysaccharide acid) played negligible effects on WVP of gelatin biofilms. It was
from wood and pulping industry, by glyoxal caused a significant also stated that addition of ferulic acid did not influence the WVP of
decrease in moisture uptake of sorbitol containing films. Simi- the films obtained by soya protein isolate [56]. The supposed reason
larly, the moisture resistance of hemicellulose-based films can be may be due to the presence of numerous hydroxyl groups in fer-
strongly reinforced by crosslinkers such as glyoxal or citric acid ulic acid and/or tannin acid which tend to interact with water. The
[52]. In a research performed by Goetz et al. [44] about crosslinking introduction of transglutaminase to the casein, gelatin and their
of cellulose whiskers with poly(ethylene glycol) and poly(methyl blends, caused WVP enhancement, mainly because of improved
vinyl ether-co-maleic acid) and the resulting films, it was found chain mobility which resulted in an increase in the water diffusion
that the crosslinked films containing 25% whiskers absorb signif- coefficient [31].
icantly less amount of water compared to the both 50% and 75% In order to reduce WVP, it is also possible to use physical
non-crosslinked whisker films. crosslinking by ␥-irradiation as an operative technique to crosslink
proteins and to develop barrier and mechanical characteristics
of protein-based films. Lacroix et al. [57] demonstrated that ␥-
5.1.2. Water vapor permeability (WVP)
irradiation induces the formation of crosslinks in casein, soy, and
The solubility and diffusivity of water molecules are important
whey protein-based biofilms and, hence, improves the WVP, while
factors to control permeability in polymeric matrix [55]. Biopoly-
Ouattara et al. [58] affirmed that ␥-irradiation crosslinking causes
mers based films generally present relatively high inclination to
a decline in WVP of milk protein-based films and expands their
WVP. Therefore, crosslinking approach allows to deliver novel
microbial resistance and enzymatic biodegradation. Lee et al. [59]
arrangements of tightly linked three dimensional networks which
also revealed a decreased WVP in ␥- irradiated soy protein films,
can be established to obstruct migration of water within the food
mainly because of the generation of crosslinking and aggregation
packaging materials. As stated by Reddy et al., crosslinking of
of the polypeptide chains. On the basis of all the collected data, it
peanut protein films using citric acid could slightly reduce the WVP
can be concluded that incorporation of crosslinking agents in the
(about 12% reduction for 1% (w/w) citric acid) and no significant
desired range can modify the WVP of biopolymer films, otherwise
difference was revealed between the films crosslinked with dif-
the excess amount acts as a plasticizer agent.
ferent amounts of citric acid (1–3% w/w) [37]. In another work,
Ghanbarzadeh et al. demonstrated that the addition of different
concentrations of citric acid (0–20%, w/w) as crosslinker to the plas- 5.1.3. Solubility
ticized starch film caused a significant reduction in the WVP with Solubility is one of the main features of packaging because of
respect that of the untreated film (i.e. 4.64 × 10−7 g Pa−1 h−1 m−1 ) its protective role surroundings the food products. Water insolu-
[43]. As an example, the starch films containing 10% (w/w) of citric bility of biopolymers supports their integrity as well as guarantees
acid presented the lowest WVP value (2.62 × 10−7 g Pa−1 h−1 m−1 ). the quality of foods [60]. The film properties of wheat straw hemi-
The reason of such occurrence can be related to the substitution of cellulose, crosslinked by addition of citric acid, revealed that the
hydrophilic OH groups with hydrophobic ester groups inside the most intense influence of citric acid was on insoluble matter of
692
Table 3
Common crosslinking agents used for some biofilms.

Biopolymer(s) Crosslinking agent Concentration Crosslinking method Improved properties Reference

Whey proteins Formaldehyde and calcium 5.5 and 9% (w/w) Mixing crosslinking agents Physical, thermal and [26]
chloride formaldehyde and 0.05% (w/w) with whey protein solution mechanical properties
CaCl2 on the basis of film (10% (w/w)) at pH 7, drying at
solution dry matter 25 ◦ C for 18–24 h
Corriedale sheep wool keratin Ethylene glycol diglycidyl 75–300 mg/g of protein Mixing with keratin solution Physical and mechanical [27]
ether (EGDE) or glycerol (6.5% w/v protein), drying at properties
diglycidyl ether (GDE) 50 ◦ C overnight
Bovine bone gelatin Ferulic acid 5–40 mg/g of gelatin Mixing with gelatin solution Physical and mechanical [28]
(12% w/w) with pH 7 at 50 ◦ C properties

F. Garavand et al. / International Journal of Biological Macromolecules 104 (2017) 687–707


for 20 min, drying at ambient
temperature
Bovine bone gelatin Tannin acid 5–40 mg/g of gelatin Mixing with gelatin solution Physical and mechanical [28]
(12% w/w) with pH 9 at 50 ◦ C properties
for 20 min, drying at ambient
temperature
Bovine hide gelatin Dialdehyde carboxymethyl 1–10% (w/w) of gelatin Dissolving in distilled water Optical, physical, thermal and [29]
cellulose (DCMC) under stirring for 12 h at room mechanical properties
temperature to produce a 1%
(w/v) solution, mixing with
gelatin solution (10% w/v) at
60 ◦ C for 1 h, drying at 40 ◦ C for
24 h
Bovine hide gelatin Microbial transglutaminase 10 U/g of protein Mixing with gelatin solution Physical properties [30]
(10% w/w) at 50 ◦ C for 15 min
and subsequently enzyme
inactivation at 85 ◦ C for 10 min,
drying at room temperature for
24–48 h
Bovine hide gelatin Formaldehyde 8.8 mmol/100 ml of film Mixing with gelatin solution Physical, thermal and [30]
solution (10% w/w) at 50 ◦ C for 15 min, mechanical properties
drying at room temperature for
24–48 h
Bovine hide gelatin Glyoxal 26.5 mmol/100 ml of film Mixing with gelatin solution Physical and thermal [30]
solution (10% w/w) at 50 ◦ C for 15 min, properties
drying at room temperature for
24–48 h
Bovine hide gelatin/casein Microbial transglutaminase 10 U/g of protein Mixing with gelatin/casein Mechanical properties [31]
solution (7% w/w) at 50 ◦ C and
pH 7.0 for 15 min and
subsequently enzyme
inactivation at 85 ◦ C for 10 min,
drying at room temperature for
24 h
Pig skin gelatin Glutaraldehyde Glutaraldehyde solution Immersing films into solution Physical, thermal and [32]
0.125–2.5% (w/w) in phosphate at pH 7.4 for 24 h at room mechanical properties
buffer temperature, washing and
drying at room temperature
Pig skin gelatin Genipin Genipin solution 0.07–2.0% Immersing films into solution Physical, thermal and [33]
(w/w) in phosphate buffer at room temperature and pH mechanical properties
7.4 for 24 h, washing and
drying at room temperature
Fish gelatin Ultraviolet (UV)-crosslinking ␭ = 253.7 nm, 30 W Irradiating gelatin solution (5% Physical and mechanical [34]
w/v) by UV at a distance of properties
30 cm from the surface for
60 min (the film solution
received a UV radiation dose of
2.158 J/m2 ), drying at 40 ◦ C for
24 h

Fish gelatin Maillard reaction using ribose 1–2% w/w of gelatin granule Mixing with gelatin solution Physical, optical and [34]
(with 10.93% moisture) (5% w/v) at room temperature mechanical properties

F. Garavand et al. / International Journal of Biological Macromolecules 104 (2017) 687–707


(25 ◦ C), drying at 40 ◦ C for 24 h
Fish gelatin UV-crosslinking and Maillard 1–2% ribose (w/w of gelatin Mixing with gelatin solution Physical, optical and [34]
reaction using ribose granule (with 10.93% (5% w/v) at room temperature mechanical properties
moisture)) (25 ◦ C), irradiating gelatin
UV (␭= 253.7 nm, 30 W) solution (5% w/v) by UV at a
distance of 30 cm from the
surface for 60 min (the film
solution received a UV
radiation dose of 2.158 J/m2 ),
drying at 40 ◦ C for 24 h
Fish gelatin Lactose 1–2% w/w of gelatin granule Mixing with gelatin solution Physical properties [34]
(with 10.93% moisture) (5% w/v) at room temperature
(25 ◦ C), drying at 40 ◦ C for 24 h
Fish gelatin UV-crosslinking and lactose 1–2% lactose (w/w of gelatin Mixing with gelatin solution Physical and mechanical [34]
granule (with 10.93% (5% w/v) at room temperature properties
moisture)) (25 ◦ C), irradiating gelatin
UV (␭= 253.7 nm, 30 W) solution (5% w/v) by UV at a
distance of 30 cm from the
surface for 60 min (the film
solution received a UV
radiation dose of 2.158 J/m2 ),
drying at 40 ◦ C for 24 h
Fish gelatin Microbial transglutaminase 20–80 mg/ml enzyme solution Mixing 5 ml enzyme solution Thermal and mechanical [35]
with 200 ml fish gelatin properties
solution 10% (w/v) at 50 ◦ C and
pH 7 for 50 min and
subsequently enzyme
inactivation at 100 ◦ C for
15 min, degassing by filter
layers, drying at 25 ◦ C for 2 days
Fish gelatin/nanoclay Microbial transglutaminase 800 mg transglutaminase Mixing 5 ml enzyme solution None [36]
powder (1% purity and 100 with 100 ml fish gelatin
units of enzyme activity per solution 40% (w/v) at 50 ◦ C and
gram of powder)/5 ml enzyme subsequently enzyme
solution inactivation at 100 ◦ C for
15 min, mixing with nanoclay
solution (containing 5%
nanoclay and 20% sorbitol
(w/w of gelatin) at 35 ◦ C for
24 h, drying

693
Table 3 (Continued)

694
Biopolymer(s) Crosslinking agent Concentration Crosslinking method Improved properties Reference

Peanut protein Citric acid 1–3% w/w of film solution Mixing with sodium Physical, thermal and [37]
hypophosphite (as the catalyst, mechanical properties
50% w/w of citric acid), adding
to the peanut protein alkali
solution (containing 1% (w/w)
sodium hydroxide), heating to
90 ◦ C during 20 min with
vigorous stirring, holding at
90 ◦ C for 5 min, drying under
ambient conditions, curing
films at 150 ◦ C for 30 min
Soy protein Maillard reaction – Mixing at pH 10.0 at 80 ◦ C, Physical, thermal and [38]

F. Garavand et al. / International Journal of Biological Macromolecules 104 (2017) 687–707


isolate/carboxymethyl vacuum defoaming, drying at mechanical properties
cellulose (CMC) 50 ◦ C for 6 h
Soy protein isolate/CMC Maillard reaction – Mixing at pH 10.0 at 80 ◦ C, Optical, thermal and [17]
vacuum defoaming, drying at mechanical properties
50 ◦ C for 6 h

Sodium alginate Calcium chloride 1st stage: 0.03% calcium 1 st stage: mixing with sodium Physical and mechanical [39]
chloride (w/v) in film solution alginate solution (1.5% w/v) at properties
2nd stage: 2–7% (w/v) of 70 ◦ C for 30 min, drying at 40 ◦ C
calcium chloride solution for 18 h
2nd stage: spraying aqueous
solution of calcium chloride
over both sides of the film,
maintaining for 3 h at room
temperature (25 ◦ C), removing
the surface excess solution by
flushing water, drying at 30 ◦ C
for 15 h
Sodium alginate Calcium carbonate 0.01–0.03 g/g of alginate Mixing with glucono-␦-lactone Physical and mechanical [40]
(5.4 g/g CaCO3 ), mixing with properties
alginate solution (to provide
1.5% w/v alginate) at
13500 rpm and room
temperature for 3 min, gently
stirring overnight, drying at
30 ◦ C for 24 h
Sodium alginate Calcium chloride 1 st stage: 0.27% w/v of film 1 st stage: adding to the Not compared with blank [41]
solution alginate solution (1.5% w/v), treatments
2nd stage: 1.2% (w/v) of drying at 40 ◦ C for 20 h
calcium chloride solution 2nd stage: immersing films in
50 ml of calcium chloride
solution (containing 3% v/v
glycerol) for 10 min, removing
the surface excess solution,
drying at room temperature
(25 ◦ C) for 5 h
Sodium alginate Calcium chloride and barium 1 st stage: 0.27% calcium 1 st stage: adding to the Not compared with blank [41]
chloride chloride (w/v of film solution) alginate solution (1.5% w/v), treatments
2nd stage: 0.6% (w/v) of drying at 40 ◦ C for 20 h
barium chloride solution 2nd stage: immersing films in
50 ml of calcium chloride
solution (containing 3% v/v
glycerol) for 10 min, removing
the surface excess solution,
drying at room temperature
(25 ◦ C) for 5 h
Sodium alginate Barium chloride and calcium 1 st stage: 0.12% barium 1 st stage: adding to the Not compared with blank [41]
chloride chloride (w/v of film solution) alginate solution (1.5% w/v), treatments
2nd stage: 1.2% (w/v) of drying at 40 ◦ C for 20 h
calcium chloride solution 2nd stage: immersing films in
50 ml of calcium chloride
solution (containing 3% v/v
glycerol) for 10 min, removing
the surface excess solution,
drying at room temperature
(25 ◦ C) for 5 h

F. Garavand et al. / International Journal of Biological Macromolecules 104 (2017) 687–707


Alginate/pectin Calcium chloride 1 st stage: 0.02–0.04 g calcium 1 st stage: adding a dilute Not compared with blank [14]
chloride dihydrate/g aqueous calcium chloride treatments
biopolymers 2nd stage: 3% solution slowly (1 ml/min) to
(w/v) of calcium chloride the sodium alginate/pectin
solution solution (1.5% w/v) at 70 ◦ C for
30 min, strong mixing, drying
at 40 ◦ C for 20 h
2nd stage: immersing films in
50 ml of calcium chloride
solution for 30 min, removing
the surface excess solution,
drying at room temperature
and RH > 60% for 6 h

Starch Citric acid 5% (w/w) of starch Mixing with sodium Physical, thermal, optical and [42]
hypophosphite (as a catalyst, mechanical properties
50% w/w of citric acid) and
starch solution 3% (w/w) at
90 ◦ C for 20 min, cooling
solution to 65 ◦ C, drying for
48 h, treating films at 165 ◦ C for
5 min
Corn starch Citric acid 5–20% (w/w) of starch Mixing with starch solution Physical and mechanical [43]
(5% w/v) at room temperature properties
(25 ◦ C) for 5 min, agitating at
500 rpm and 90 ◦ C for 30 min,
drying at 60 ◦ C
Corn starch/CMC Citric acid 10% (w/w) of starch Mixing with starch solution Not compared with blank [43]
(5% w/v) at room temperature treatments
(25 ◦ C) for 5 min, mixing with
CMC solution at 75 ◦ C for
10 min, degassing with stirring
at 40 ◦ C for 30 min, drying at
60 ◦ C
Maize starch/microcrystalline UV-crosslinking using sodium Sodium benzoate (0.15% (w/w) Mixing sodium benzoate with Physical and mechanical [12]
cellulose benzoate (as a of dry starch) starch/cellulose solution (6.66% properties
photo-sensitizer) UV (␭ > 290 nm) irradiated by w/v) at 100–105 ◦ C, cooling to
medium-pressure mercury 85–90 ◦ C and degassing with
vapor lamps (400 W) stirring by vacuum, drying at
50–60 ◦ C, irradiating films by
UV at 60 ◦ C for 2–60 min

695
696
Table 3 (Continued)

Biopolymer(s) Crosslinking agent Concentration Crosslinking method Improved properties Reference

Cellulose nanowhiskers Poly(methyl vinyl 8.375% (w/v) of crosslinking Mixing 40 ml crosslinking Physical properties [44]
ether-co-maleic acid) solution (containing 1.25% solution with cellulose
(w/v) poly(ethylene glycol)) nanowhiskers, drying
overnight, curing at 135 ◦ C for
6.5 min
Methyl cellulose Glutaraldehyde 0.5–12.0% (w/w) of film Mixing with methyl cellulose Physical, thermal, mechanical [45]

F. Garavand et al. / International Journal of Biological Macromolecules 104 (2017) 687–707


solution solution (2% w/w of film and biodegradation properties
solution) and 2 drops of
hydrochloric acid (HCl, as a
catalyst to yield a solution of
pH 3) at room temperature and
700 rpm for 1 h, washing and
neutralizing dried films by
distilled water for 3 h
Hydroxy propyl methyl Citric acid 5–15% w/w of HPMC Mixing with Sodium Physical properties [46]
cellulose (HPMC) dihydrogenophosphate
(NaH2 PO4 , as a catalyst, 0–50%
w/w of citric acid) and HPMC
solution (3% w/w) for 15 min
(pH was reached to 2.65),
drying at 60 ◦ C for 60 min,
Treating films at 190 for
15 ◦ C min
Bacterial cellulose nanocrystals Boric acid 3–15% w/w of PVA Mixing with BCNC/PVA Mechanical properties [11]
(BCNC)/Poly(vinyl alcohol) solution at 80 ◦ C and 1600 rpm
(PVA) for 30 min (to achieve film
solution with 6% (w/v) dry
matter), synchronically
ultrasonicating (35 kHz) and
agitating (13000 rpm) at 60 ◦ C
for 5 min, degassing at 60 ◦ C
and 500 mmHg for 15 min,
drying at 50 ◦ C for 12 h

Wheat straw hemicellulose Citric acid 5–30% w/w of hemicellulose Mixing with hemicellulose Physical properties [47]
solution (5% w/v), degassing
under vacuum, drying at room
temperature for 24 h,
conditioning (50% RH, 24 ◦ C) in
an environmental chamber for
24 h to reach 11.91–12.32%
w/w moisture contents, curing
dried films at 150 ◦ C for 10 min
Chitosan H2 SO4 H2 SO4 solution 2% (w/w) Immersing films into solution Not compared with blank [48]
for 1 h, washing and drying treatments
Crab shell chitosan Tannic acid 20–80 mg/g of chitosan Mixing with chitosan solution Physical and mechanical [49]
(1.5% w/w), drying at 37 ◦ C for properties
24 h
Shrimp chitosan Main-chain-type benzoxazine 25–400% (w/w) of chitosan Dissolving in acetic acid Thermal and Mechanical [50]
polymer, (MCBP) solution (1% w/w acetic acid) of Properties
at ambient conditions, adding
slowly to the chitosan solution

F. Garavand et al. / International Journal of Biological Macromolecules 104 (2017) 687–707


(containing 1% w/w acetic acid)
under vigorous agitation to
obtain a total 2% (w/w) solid
content, degassing under
vacuum at room temperature,
drying at room temperature for
2 days, drying under vacuum at
50 ◦ C for 24 h, heating in
sequence at 75, 100, 125, 150,
and 175 ◦ C for 2 h
Chitosan/PVA Glutaraldehyde Glutaraldehyde solution 0.01% Immersing films into Not compared with blank [48]
(w/w) glutaraldehyde solution treatments
containing 0.5% (w/w) sulfuric
acid (as a catalyst) for 1 h,
washing and drying
Chitosan/corn cob Glutaraldehyde Glutaraldehyde solution 1% Mixing with chitosan solution Thermal and mechanical [51]
(v/v) (1.5% w/v) and corn cob properties
powder, drying at room
temperature for 48 h
Galactomannan/bovine serosa Glutaraldehyde Glutaraldehyde solution Immersing films into solution Physical and thermal [16]
collagen 0.001–1.5% in phosphate buffer at room temperature and pH properties
7.4 for 24 h, treating in
glycine solution (0.025 M
glycine:0.05 M borate, pH 9.2)
for 10 min, washing with water
and drying
Spruce Glyoxal 5% w/w of GGM Mixing with GGM solution (1% Physical, thermal and [52]
O-acetyl-galactoglucomannan w/v) at 60 ◦ C for 4 min, mechanical properties
(GGM) degassing by ultrasonication
under vacuum for 5 min,
drying at 50% RH and 23 ◦ C
Spruce GGM Ammonium zirconium 10% (w/w) of GGM Mixing with GGM solution (1% Physical and mechanical [53]
carbonate w/v), treating at pH 8.5 and properties
80 ◦ C for 5 h, degassing by
ultrasonication under vacuum
for 5 min, drying at 50% RH and
23 ◦ C

697
698 F. Garavand et al. / International Journal of Biological Macromolecules 104 (2017) 687–707

hemicellulose, whose insolubility was amplified 5-fold by incor- xanthan gum film. This result revealed that the transparency of
poration of 20% citric acid [47]. Former studies corroborated the the films is less than 1 at 600 nm and slightly improved with
above-mentioned changes in solubility [46,61], ascribed to the the increase in oxidization level of oxidized xanthan gum. The
generated ester bonds between citric acid and the polysaccharide slight increase in transparency at 600 nm could be a consequence
[62], resulting in a denser structure and corresponding improved of darker color caused by crosslinking. The light transmission in
water insolubility. The reduced and/or unavailable hydroxyl groups the UV and visible light range evidenced that the oxidization
and their replacement by hydrophobic ester groups could be also of xanthan gum can significantly improve the UV light barrier
related to the effects of citric acid concentration on water solubility attributes and has little effect on the transmission of visible light.
[47,63]. Consequently, crosslinking agents can overcome the poor The results obtained by Dou et al. on feather keratin film crosslinked
moisture resistance of polysaccharide films, originated from their by dialdehyde starch validated the transparency enhancement by
hydrophilic nature, reducing the water absorption by formation of introducing crosslinking agents [71].
some interconnects between polysaccharide molecules.
5.1.6. Color changes
5.1.4. Film thickness The color of biopolymeric films is regarded as an important fac-
Thickness of films and coatings which usually range from 80 ␮m tor to evaluate their quality. It is stated that the color is greatly
to 200 ␮m plays a pivotal role in some barrier properties, including influenced by various factors, such as pH value, degree crosslinking,
WVP and gases transport. Since most of the studied biopolymer- plasticizer content, thermal process and fabrication procedure. In
based films are originally hydrophilic, their thickness presumably addition, the color of protein-based biopolymers is more altered by
influences WVP because of differences between the water vapor the type and concentration of amino acids and proteins, rather than
pressure and the accumulation of moisture at the film/air inter- other treatments and processes [38]. Most of the polysaccharide-
face. As well, occurrence of crosslinking outwardly toughens the based films usually appear colorless, even though there are some
generated internal bonds of the biopolymer films [64] and vari- exceptions such as chitosan which displayed a slightly yellow
ous denser composites will be achieved, depending on the amount color. Crosslinking of chitosan by tannic acid showed that addition
of crosslinking agent incorporated into film structure. Denser of tannic acid triggered the major color changes in the obtained
structural biopolymeric patterns owing to crosslinking addition, films due to its affinity to oxidize, in turn yielding brownish films
along with the improvement in some physical characteristics, were [49]. Similarly, the crosslinking of proteins by citric acid results
reported in numerous documents [40,46,65–67]. in yellow-brown films principally when exposed to high tempera-
tures and/or after extended periods of time [72]. Reddy and Yang
5.1.5. Light barrier properties and transparency reported that yellowness index (YI) of the starch films was not
Because of consumer’s preference to see the food items, higher affected by different amounts of citric acid before curing treatment
transparency is greatly recommended for food packaging films and [42]. Indeed, curing of starch films at 165 ◦ C for 5 min, followed
coatings [68]. Film transparency is extensively related to the func- by addition of citric acid (1–20 wt%), intensified the yellowness
tionality of film ingredients because of their great influence on of the prepared films. In details, films crosslinked with 1–5 wt%
the appearance of the food products. As well, light barrier prop- citric acid showed YI values of about 10% higher than their corre-
erties are considered another fundamental requirement for food sponding YI before curing, whereas YI values of the films treated
packaging materials. Since the emitted UV light ranging from 200 with 10% citric acid sharply increased after curing treatment and
to 280 nm is considered as one of the oxidation initiators in food those of films treated with 20% citric acid were approximately 13
products, how to promote the barrier properties of packaging mate- times higher compared to non-crosslinked films after curing. The
rials against UV light has received a great deal of attention [69]. In revealed remarkably higher yellowness of films can be ascribed
particular, lipid oxidation represents a great concern for the food to the dehydration of citric acid owing to curing at high tempera-
manufacturers due to the development of undesirable off-flavors tures, with consequent development of an unsaturated acid [42,73].
and generated hazardous products, especially occurring free radical Comparable intensify in yellowness-pattern was also observed for
reactions. As a result, food manufacturers must improve procedures glutenin [74], soy protein isolate [75], and castor bean proteins [76]
to avoid or slow down lipid oxidation in food products. Leera- crosslinked by different aldehydes. On the other hand, Benavides
hawong et al. [69] reported that protein-based biofilms are good et al. noted that incorporation of CaCO3 did not significantly alter
UV protectors, mainly because of the presence of aromatic amino the total color difference of the alginate films so that calcium ions
acids which can captivate the UV light. In this regard Mu et al. [29] at low concentration (0.01–0.03 g CaCO3 /g alginate) was not suffi-
investigated the light barrier and transparency features of edible cient to change the light scattering emitted by alginate background
gelatin films crosslinked with dialdehyde carboxymethyl cellulose [40].
(DCMC). Crosslinked gelatin films prohibited lipid oxidation by
constructing proper barrier against UV light in the food system. 5.2. Mechanical properties
Furthermore, incorporating more DCMC content led to an increase
in transparency of gelatin–DCMC films at 280 nm, although it was Maintaining the integrity of packaging films is important during
insignificant at 600 nm. It was hypothesized that DCMC improves processing, transport, and handling [5]. Generally, the incorpo-
the UV barrier characters of gelatin films by generating abundant ration of more amounts and/or dosages of crosslinking agents
C N groups through formation of Schiff’s base between ␧-amino improve the tensile strength (TS) of biopolymer films and decrease
groups of lysine or hydroxylysine side groups of gelatin and the their elongation at break (EAB) [43]. Mechanical properties of
aldehyde groups in DCMC. some biofilms are compared in Table 4. All considered film making
In agreement with the aforementioned declarations, Guo et al. biopolymers display almost identical trend in case of incorporat-
expressed that the transmission of UV light was so slight at 200 nm ing crosslinkers so that significant and progressive increment of
and 280 nm for all gelatin-oxidized xanthan gum films, even though TS is detected once the amount/dosage of crosslinkers is amplified
transparency is high at 280 nm [70]. Therefore, the all films exhib- (commonly less than 10% for enzymatic and chemical crosslinkers
ited appropriate barrier properties against UV light, associated and less than 20 kGy for irradiation). Gou et al. reported that aver-
again to their high content of aromatic amino acids which can age TS value of gelatin-xanthan gum film is about 24 MPa, which is
absorb UV rays. Furthermore, the transparency of gelatin-oxidized considerably higher than that of gelatin film (about 2 MPa) [70].
xanthan gum films at 280 nm is about 3 times as that of gelatin- The significant increase of TS value by the addition of xanthan
F. Garavand et al. / International Journal of Biological Macromolecules 104 (2017) 687–707 699

Table 4
Mechanical properties of various biopolymer films as affected by crosslinkers.

Biopolymer(s) Crosslinker Mechanical properties Reference

Type Dosage Elongation at break Tensile strength

Pectin/gelatin Gamma irradiation 10 kGy 15% ↓ 34% ↑ [77]


Soy protein isolate Gamma irradiation 16 kGy 21% ↓ 25% ↑ [59]
Gelatin Ferulic acid 4% 35% ↓ 11% ↑ [28]
Gelatin Tannin acid 4% 4% ↓ 4% ↑ [28]

Starch Citric acid 5% – 150% ↑ [42]


Un-plasticized chitosan Tannic acid 4% 40% ↑ 15% ↑ [49]
Plasticized chitosan Tannic acid 4% 53% ↓ 25% ↑ [49]

Soy protein isolate Genipin 1% 103% ↑ 29% ↑ [78]


Fish gelatin/oxidized Sodium periodate 2.4% 50% ↓ 25% ↑ [70]
xanthan gum
Bovine gelatin/calcium Transglutaminase 8% 33% ↑ 58% ↑ [79]
carbonate
Soy protein isolate/fish Transglutaminase 3% 69% ↓ 27% ↑ [80]
gelatin
Fish gelatin Transglutaminase 3% 42% ↑ 14% ↑ [80]
Bacterial cellulose Boric acid 1.96% 13% ↓ 10% ↓ [11]
nanocrystals/Poly(vinyl
alcohol)

gum could be related to the created interactions between both addition. On the other hand, an increase in genipin concentration
biopolymers and the entanglement of gelatin chains within the (more than 1%) led to a reduction in EAB due to the formation of
film matrix. After introducing oxidized xanthan gum into gelatin new crosslinks, obtaining a stiffer structure in which the movement
as crosslinker, the obtained films presented substantially higher TS between the chains is limited. In another study performed by Wang
(average around 33 MPa) and reduced EAB values (average 11.3%), et al. [79] the EAB percent of gelatin-calcium carbonate compos-
suggesting the occurrence of crosslinking between gelatin and oxi- ite films was considerably enhanced with increase in the amount
dized xanthan gum by addition of sodium periodate. Increase in of transglutaminase enzyme. These experimental evidences were
TS value of protein films due to crosslinking often goes along with in agreement with those related to food proteins derived films
a reduction in EAB (less extensible films), because the crosslinks prepared using gelatin, soy protein isolates, sodium caseinate and
allow to obtain a tougher structure. whey protein concentrate which crosslinked by transglutaminase
It is important to highlight that some crosslinking agents may [84]. Different kinds and means of the interactions between plasti-
act both as a crosslinking agent and a plasticizer in the biopolymer cizers and protein chains and combination of plasticizers into some
films. For example, Ghanbarzadeh et al. [43] showed such bilat- particular areas in the polymeric matrix result in boosting their
eral effect by addition of wide range of citric acid in the starch distance, and thus, affect their mobility. Protein chain composi-
films. By increasing citric acid content up to 10 wt%, the TS of the tions, molecular weight, shape, and amino acid content of proteins
films increased and EAB decreased. Indeed, for low amounts, cit- derived from various origins could responsible for the aforesaid
ric acid probably induces the formation of highly linear structure results. The mechanical properties of the most of the crosslinked
by hydrolyzing branched chains of starch molecules, in turn gen- film-forming biopolymers were close to that of traditional syn-
erating more hydrogen connections between the starch chains and thetic films including polyvinylidene chloride (PVDC), polylactic
increasing the TS of the obtained films. For high contents, the extra acid (PLA), polyethylene and so on [85].
(not-reacted) portion of citric acid probably acts as a plasticizer and Rouhi et al. reported that increases in the concentrations of boric
reduces the interactions between the macromolecules, with con- acid improve the TS of the poly(vinyl alcohol) (PVA)/bacterial cel-
sequent TS decrement and EAB increase. Comparable results were lulose nanocrystals (BCNC) films due to the crosslinking reactions.
reported by Ning et al. about the effect of citric acid on the mechan- However, the TS of films decreased with increasing the content
ical behavior of the blend films based on thermoplastic starch of boric acid at high concentrations (5% (w/w)) of bacterial cellu-
and linear low-density polyethylene [81]. Contrariwise, Shi et al. lose nanocrystals. The reason was BCNC coalescence and decrease
remarked that increasing citric acid content (from 5 wt% to 30 wt%) in their aspect ratio due to intra and inter-molecular crosslinking
in polyvinyl alcohol-starch films did not significantly reduce the TS; [11].
however, the EAB remarkably increased [82]. On the other hand, Ma
et al. [83] showed an extraordinary crosslinking effect of citric acid
at 9 wt% concentration on pea starch biocomposite film, in such a
manner that TS increased and EAB reduced. 5.3. Thermal properties
Interestingly, decrease in EAB is typically attended by increase
in TS, which apparently conflicts with some results reported in Thermodynamic characteristics, established by thermal anal-
Table 4. Controversial results regard the effect of introducing ysis, are important to comprehend the quality of substances
crosslinkers on EAB values can be interpreted and justified with dif- under cooling/heating rates, inert/reduction/oxidation atmosphere
ferent reasons. It seems that the crosslinker amount and the nature or under gas pressures. Recently, several techniques were used to
of film-forming constitutes are important items which strongly determine the thermo-physical properties of materials, including
influence EAB. For example, Gonzalez et al. [78] observed two dif- differential scanning calorimetry (DSC), thermogravimetric analy-
ferent behaviors of soy protein crosslinked by genipin, in terms of sis (TGA), differential thermal analysis (DTA), dynamic mechanical
EAB. When the genipin amount was fewer than or equal to 1%, more analysis (DMA), evolved gas analysis (EGA), dilatometry (DIL),
expansible networks (high EAB) were established because of some dielectric analysis (DEA). The principal application of the TGA and
intrinsic interruptions in protein interactions as a result of genipin DSC is to study phase transitions under different atmospheric con-
ditions, cooling/heating rates and temperatures [86]. Table 5 shows
700 F. Garavand et al. / International Journal of Biological Macromolecules 104 (2017) 687–707

Table 5
Thermal properties of various crosslinked biopolymer films.

Biopolymer(s) Crosslinking agent DSC TGA Reference

Gelatin Immersing films into Increased thermal stability- – [32]


glutaraldehyde solutions endothermic peak at 72 ◦ C, and
(0.125–2.5% (w/w)) the denaturation enthalpy
amounts to about 6 J/g
Gelatin Immersing films into genipin Endothermic peak at 86–98 ◦ C – [33]
solutions (0.07–2.0% (w/w))
Galactomannan/collagen Immersing films into First transition at 48 ◦ C – [16]
glutaraldehyde solutions second transition at
(0.001–1.5%) 58.7–66.7 ◦ C
Gelatin Transglutaminase (10 U/g of Melting temperature – [30]
protein) decreased to 62.95 ◦ C
Gelatin Formaldehyde Melting temperature increased – [30]
(8.8 mmol/100 ml of filmogenic to 77.43 ◦ C
solution)
Gelatin Glyoxal (26.5 mmol/100 ml of Melting temperature increased – [30]
filmogenic solution) to 87.09 ◦ C

Fish gelatin Microbial transglutaminase The melting temperature – [35]


(5–20 mg/g of biopolymer) increased from 124 ◦ C to 158 ◦ C
Chitosan/poly(vinyl Immersing films into Broad exothermic peak at Weight losses at 40–120 ◦ C [48]
alcohol) glutaraldehyde solution (0.01% 310 ◦ C and 170–300 ◦ C
(w/w)) containing sulfuric acid
(0.5% (w/w)) as the catalyst
Chitosan Immersing films into sulfuric Sharp exothermic peaks at Weight losses at 40–150◦C [48]
acid solution (2% (w/w)) 290 ◦ C, and 200–300 ◦ C
endothermic peak at 100 ◦ C

Starch Citric acid 5% (w/w) of starch Increase the temperature of Lower weight loss between [42]
water evaporation 120 and 220 ◦ C, higher
endothermic peak weight loss between 220
and 320 ◦ C, 20% lower
weight loss at 600 ◦ C
Soy protein Maillard reaction A single Tg in the range A single and higher [38]
isolate/Carboxymethyl 75–100 ◦ C and melting decomposition
cellulose between 140 and 160 ◦ C temperature in the range
250–350 ◦ C
Chitosan/corn cob Glutaraldehyde – Weight losses at 50 ◦ C, [51]
300 ◦ C and 310–350 ◦ C

selected publications focused on the thermal properties, including ing comparable values. In details, both genipin and glutaraldehyde
TGA and DSC, of crosslinked biopolymeric films. generally induce a decrease in the denaturation enthalpy [32,33].
Furthermore, the thermal stability was increased by 2% (w/w)
genipin in the crosslinked treatment, due to the slight but sig-
5.3.1. Differential scanning calorimetry
nificant increase of the denaturation temperature from 91 ◦ C to
DSC is a thermal analysis instrument which is widely applied to
98 ◦ C. However, the denaturation temperature of gelatin films
determine the difference in the amount of heat required to increase
showed non-significant changes with 2.5% (w/w) glutaraldehyde,
the temperature of a sample and reference as a function of time and
exhibiting a lower denaturation temperature and a less thermal
temperature [87]. Both the sample and reference are maintained at
stability than films crosslinked by genipin. In consistent with
equal temperature throughout the analysis. The program of a DSC
these results, de Carvallho and Grosso [30] observed that the
analysis is usually designed such that the temperature of sample
Tm increased from 65.06 ◦ C in native gelatin films to 87.09 ◦ C
and reference holders linearly increases as a function of time.
and 77.43 ◦ C in glyoxal- and formaldehyde-treated gelatin films,
For example, in a paper focused on chitosan and poly(vinyl alco-
respectively. They reported a decrease in ␧-amino groups in the
hol) (PVA) blend films crosslinked by glutaraldehyde [48], DSC was
transglutaminase-modified film, because of an increase in molecu-
used to examine their thermal properties. An endothermic peak at
lar weight. However, no change was observed in the melting point
about 60 ◦ C was related to water loss, whereas a broad exothermic
of the resulting film compared to the native film, and the values of
peak at 310 ◦ C was ascribed to the chitosan thermal degradation
melting enthalpy (Hm ) showed a decrease as compared to those
that occurred at 290 ◦ C in neat chitosan films, suggesting that the
of the native film. Increase in the number of crosslinking covalent
thermal stability of crosslinked chitosan–PVA film was significantly
bonds, which exothermically break, resulted in a decrease of the
improved.
hydrogen bonds number, which endothermically break, demon-
Yi et al. [35] investigated the thermal properties of micro-
strating the decrease in the Hm , despite the increase in Tm [32,88].
bial transglutaminase crosslinked fish gelatin films. Their results
Considering collagenous materials, an endothermic peak in the
showed an increase in melting temperature (Tm ) caused by the
DSC thermogram is usually recorded and ascribed to the helix-coil
incorporation of crosslinker. The Tm measured by DSC increased
transition of collagen. The amount of the denaturation enthalpy
from 124.78 ◦ C to 158.49 ◦ C as the enzyme reaction time increased
related to this peak is depended on the relative amount of triple
from 0 to 50 min. There was no significant difference of Tm and
helical structure in the samples, and is significantly higher for col-
crosslinking degree between 10 and 30 min reactions. Bigi et al.
lagen compared to gelatin [88,89]. Rault et al. [90] studied collagen
investigated the influence of both genipin and glutaraldehyde
films crosslinked by different chemical crosslinking agents, observ-
crosslinkers on the thermal properties of gelatin films, evidenc-
F. Garavand et al. / International Journal of Biological Macromolecules 104 (2017) 687–707 701

ing an increase in the Tm of the film modified with glutaraldehyde. desolation and solid-gas reactions (such as oxidation or reduction)
Usha and Ramasami [91] also reported an increase in the Tm of col- [99]. TGA is commonly used to measure characteristics of sub-
lagen films processed with formaldehyde due to the formation of stances that show either mass loss or gain because of oxidation,
crosslinks. Kopp et al. [92] observed a decrease in the Hm caused decomposition or loss of volatiles (such as moisture), and it is also
by crosslinking of collagen fibers. As well, Figueiró et al. [16] stud- applicable to study the polymers [100].
ied the effects of glutaraldehyde crosslinking agent on the thermal For example, Tripathi et al. [48] investigated the thermal decom-
properties of galactomannan-collagen films, using DSC. The pres- position of chitosan-PVA films crosslinked with glutaraldehyde.
ence of two transitions in modified galactomannan–collagen films Thermographs of chitosan films showed that the weight loss
indicated two structures with different thermal stability because of between 40 and 150 ◦ C was as a result of the moisture vaporiza-
the heterogeneous reaction of glutaraldehyde in a mixture of galac- tion. The weight loss between 200 and 300 ◦ C was associated to
tomannan and collagen. The first transition at low temperature the chitosan thermal degradation. Furthermore, two considerable
was due probably to the not-occurring the crosslinking in collagen weight losses were exhibited in the TGA plot of chitosan–PVA film
by glutaraldehyde. Therefore, they concluded that the presence of crosslinked by glutaraldehyde: one at 40–120 ◦ C was ascribed to
galactomannan could avoid the glutaraldehyde access to collagen the moisture vaporization and the other at 170–300 ◦ C was due to
molecules in the internal of film matrix [93]. The second transition the degradation of crosslinked chitosan–PVA film. They concluded
at higher temperature, such as 60.7 and 66.7 ◦ C (samples treated that introducing crosslinker in the network of the composite causes
with 0.05 and 1% glutaraldehyde, respectively), showed that the a remarkable improvement in the thermal stability. Furthermore,
fixation of collagen molecules on the film surface was a function of Yeng et al. [51] prepared the unmodified and modified chitosan-
glutaraldehyde concentration. The galactomannan blocked by the corn cob composite films with glutaraldehyde as a crosslinking
network of collagen fibers stabilized by glutaraldehyde. The tran- agent and performed TGA measurements, identifying three main
sition at 58.7 ◦ C for the galactomannan–collagen film treated with weight losses: the first one started at about 50 ◦ C and was related
1.5% glutaraldehyde could present the occurrence of polymeriza- to the evaporation of absorbed water; the second one at 300 ◦ C
tion reactions by higher concentrations of glutaraldehyde, resulting was ascribed to the thermal decomposition of hemicellulose, cel-
glutaraldehyde-induced polymers which could also block the pen- lulose and lignin; the third one took place in the temperature of
etration of glutaraldehyde into the internal regions of the collagen 310–350 ◦ C due to polymer decomposition. The thermal stability
matrix, correspondingly decreasing the denaturation temperature. of chitosan-corn cob films reduced with increasing corn cob con-
Glutaraldehyde stabilized the collagen molecules and resulted in tent. However, they showed that glutaraldehyde as a crosslinker
an increase of the denaturation temperature [94]. increased the thermal stability of films due to the presence of imine
It has been reported that citric acid could act as a crosslinker linkages.
by forming strong hydrogen bond interactions with starch, which Su et al. [38] studied the effects of the Maillard reaction on
improve its thermal stability, and prevent retrodegradation [42,95]. the thermal stability of edible films based on CMC-SPI blends.
According to DSC data reported by Reddy et al. [42], the DSC plot of CMC-SPI blend films exhibited about 15% weight loss in the tem-
the non-crosslinked film during the first heating cycle showed an perature range between 0 and 150 ◦ C as a result of the water
endothermic peak at about 100 ◦ C, whereas that of the crosslinked evaporation. However, these samples showed different decompo-
film with 5% citric acid had an endothermic peak at about 150 ◦ C. sition ratios in the range of 30–80% residual mass. The slope of
This peak in the both non-crosslinked and crosslinked dried films the weight–temperature line was reduced with increasing CMC
disappeared after the second heating cycle, suggesting that it can content of the CMC-SPI blends, indicating greater thermal sta-
be ascribed to the moisture in the films. Moreover, the shift of the bility compared to neat SPI that decomposed at a relatively low
identified endothermic peak at higher temperatures in the case of temperature (about 110 ◦ C), as demonstrated by the related TGA
crosslinked films confirmed that the crosslinking is able to act as curve. In details, the CMC-SPI blends decomposed at temperatures
obstruction to water evaporation in the film. 250–350 ◦ C, temperatures higher than those of both neat SPI and
The Maillard reaction can take place when protein is mixed CMC films. The increase in thermal stability of the blend films
with carbohydrates at high temperatures [96]. Lii et al. [97] formed could be assigned to crosslinking stimulated by Maillard reaction
covalent bonds between carboxymethyl cellulose (CMC) and casein between CMC and SPI. Finally, TGA curves of all blends showed a
by electro-synthesis. These CMC-casein complexes exhibited good single decomposition temperature, indicating good compatibility
emulsifying properties and thermal stability. Moreover, edible and miscibility between CMC and SPI.
films based on CMC and soy protein isolate (SPI), plasticized by
glycerol and crosslinked by Maillard reactions were studied [38].
Investigation of films by DSC showed that pure SPI films had glass
transition temperature (Tg ) at about 60 ◦ C, in agreement with their
previously observations [98]. CMC-SPI blends exhibited a Tm value 6. Complementary techniques for identification of
in the range 140–160 ◦ C and a single Tg between 75 and 100 ◦ C, crosslinking occurrence in biopolymers
indicating that SPI and CMC are highly compatible and miscible.
All crosslinked films showed an endothermic Tg curve peaked at Identification of biopolymeric microstructures is essential since
a higher temperature than that of SPI and CMC. Consequently, it determines their mechanical and barrier properties. Moreover,
increase in Tg with increasing proportion of CMC was due to an structural studies of polymers would improve our concep-
increasing extent of crosslinking. tion about crosslinking reactions. The investigation of physical,
mechanical and/or thermal properties of polymers exposed to
5.3.2. Thermogravimetric analysis crosslinking agents is considered an acceptable way to deter-
TGA is a technique in which changes in weight of materials mine whether reticulated network in polymers has formed
are determined as a function of temperature with constant heat- or not. However, as shown in Table 6, some complementary
ing rate, or isothermally as a function of time [99]. TGA allows methods, such as X-ray diffraction (XRD), SDS-PAGE (sodium
to study physical phenomena, such as second-order phase tran- dodecyl sulfate–polyacrylamide gel electrophoresis), FTIR (Fourier
sitions, including desorption, absorption, adsorption, vaporization transform infrared) and Nuclear magnetic resonance (NMR) spec-
and sublimation. Moreover, TGA can supply information about troscopy, can approve the crosslinking reaction occurrence in
chemical phenomena, including decomposition, chemisorption, polymers.
702
Table 6
Complementary techniques for identification of crosslinking occurrence in some biopolymers.

Biopolymer(s) Crosslinking agent Crosslinling bond XRD SDS-PAGE FTIR NMR Reference

Hydroxyl propyl methyl Citric acid Ester bond – – Increase in absorption peak – [46]
cellulose intensity of the ester band
vibration (1735 cm−1 )
Casein/gelatin Transglutaminase Isopeptide bond – Formation of new high – – [31]
molecular mass
polymers (≥212 kDa)
Methyl Cellulose Glutaraldehyde Hydrogen bond – – Decrease in the peak of – [45]
C O stretching from
aldehyde group of free
Glutaraldehyde

F. Garavand et al. / International Journal of Biological Macromolecules 104 (2017) 687–707


(1710 cm−1 )
Starch/poly(vinyl alcohol) Citric acid Ester bond Decrease in the crystal – Increase in the peak of the – [82]
Hydrogen bond peak at 19.3◦ C O stretching vibration
(1729 cm−1 )
Increase in the peak of the
stretching vibration of the
hydroxyl groups
(3340 cm−1 )
Fish gelatin/nanoclay Transglutaminase Isopeptide bond – Decrease in band – – [36]
intensity <75 kDa
Increase in band
intensity >250 kDa
Carboxymethyl Maillard reaction C N – – Decrease in the intensity of – [38]
cellulose/soy protein C N the bands at 3100–2900,
isolate 1600–1400 and 890 cm−1
(amide I, П and Ш)
The absorption band at
1241–1472 cm−1 (C N)
and the new absorption
band appeared at
1647 cm−1 (C N)
Chitosan Tannic acid Hydrogen bond Shift the peaks toward – Decrease in the intensity of – [49]
lower 2␪ values absorption peaks
Starch Citric acid Hydrogen bond Increase in the diffraction – The additional peak in the – [42]
intensity of peaks at 17◦ crosslinked film of the
and 22◦ , decrease in the% carboxyl and ester carbonyl
crystallinity bands (1724 cm−1 )
Soy protein Maillard reaction C N – – Decrease in the intensity of Decrease in the [17]
isolate/carboxymethyl C N the bands at 3100–2900, intensity of 13 C peak at
cellulose 1600–1400 and 822 cm−1 173 ppm (COO
(amide I, П and Ш) backbone carbonyl
The n band at group)
1278–1296 cm−1 (C N) No signal at 50 ppm
and the new n band at (NH2 group)
1645 cm−1 (C N)
Bacterial cellulose Boric acid Hydrogen bond – – Increase in the wave – [11]
nanocrystals/Poly(vinyl B O C ether bond number of OH stretching
alcohol) peak (3000–3550 cm−1 )
A new and weak
absorption peak at about
1035 cm−1 attributed to
the stretching vibration of
B OC
F. Garavand et al. / International Journal of Biological Macromolecules 104 (2017) 687–707 703

6.1. X-ray diffraction 6.2. Sodium dodecyl sulfate–polyacrylamide gel electrophoresis

X-ray crystallography allows identifying the atomic and molec- Protein based polymerization can be evaluated by means of SDS-
ular structure of crystals and, thus, can be used to investigate the PAGE method, a very common technique for separating proteins by
influence of the crosslinking agent on the polymer crystallization. electrophoresis, as described by Laemmli [108].
For example, in the case of chitosan, according to Wang et al., In a study by Chambi and Grosso [31], the molecular weight
the two peaks at 2␪ 11.6◦ and 8.4◦ correspond to the hydrated crys- profile of the films treated with transglutaminase was evaluated.
talline structure, while the broad peak at around 2␪ 23◦ indicates In the absence of transglutaminase, the casein films showed intense
the presence of an amorphous structure [101]. XRD of tannic acid bands around 30 kDa, related to ␣s1-, ␤- and ␬-casein [109], and the
crosslinked chitosan films was investigated by Rivero et al. [49]. prominent band of the gelatin film had a molecular mass of 212 kDa,
They reported that neat chitosan films exhibited a slight trend to ascribed to the ␤-gelatin subunit [110]. In the casein/gelatin blend
higher 2␪ values during the storage, which indicates that acetic films, the intensities of the bands of the individual proteins differed
acid, used in order to completely solubilize chitosan, along with as a function of the gelatin and casein contents. The occurrence of
water molecules, was spontaneously removed from the chitosan crosslinking in the proteins resulted in the formation of new high
matrix. Ogawa et al. ascribed this shift to a gradual conformational molecular weight polymers (≥212 kDa) and a partial disappearance
evolution toward the anhydrous polymorph of chitosan, related of the bands related to the non-crosslinked casein. Oh et al. achieved
to the equilibrium moisture [102]. Brown et al. reported a simi- complete crosslinking of the casein fractions using casting method
lar shift [103]. Films containing plasticizer, tannic acid or both of soon after the addition of transglutaminase, which could allow the
them, showed similar XRD patterns, such that a shift of their char- polymerization during drying at 23 ◦ C for 12 h, despite of the high
acteristic peaks was observed during the storage. The magnitude moisture content [111].
of these shifts was different depending on the used additives. The Gomez-Guillen et al. [112] stated that crosslinking with micro-
incorporation of tannic acid shifted the peaks toward lower 2␪ val- bial transglutaminase (MTGase) could lead to an increase in the
ues, while the effect was more considerable with the addition of molecular weight of fish gelatin. Kolodzlejska et al. [113] reported
plasticizer. However, a lower percentage of displacement of the that reaction time and concentration of MTGase-treated Baltic cod
aforementioned peaks was observed with the addition of both tan- skin gelatin considerably increased, leading to an increment in the
nic acid and glycerol as a plasticizer, than in the case of plasticized speed of gel formation and of the gelatin viscosity during MTGase
chitosan films, attributing to a synergic effect of tannic acid and treatment. Bae et al. [36] investigated the SDS–PAGE of neat fish
glycerol. gelatin and MTGase-treated fish gelatin. In MTGase-treated fish
XRD was also performed to investigate the physical structure of gelatin, the band intensity with molecular masses below 75 kDa
citric acid crosslinked starch films by Reddy and Yang [42]. They decreased with increasing the treatment time, whereas the band
reported that XRD patterns of the crosslinked and non-crosslinked intensity above 250 kDa increased. They indicated that crosslinking
starch films showed similarly two distinguished peaks at about 2␪ reaction in the fish gelatin matrix was induced by MTGase, lead-
17◦ and 22◦ . However, both films did not present the typical starch ing to increased molecular mass. Indeed, the molecular weight of
peaks at about 2␪ 5◦ and 14◦ , suggesting changes in crystal struc- MTGase is 38 kDa with 331 amino acid residues [114,115] and, the
ture and/or the crystallinity. The non-crosslinked and crosslinked probability of band overlap is very low; thus the revealed bands
starch films had comparable crystallinity of about 17% and 14%, between 37 kDa and 75 kDa have to be associated to fish gelatin
respectively. Also the crosslinked starch films with epichlorohy- molecules. Similarly, other authors observed the decrease of band
drin compared to native starch, did not show any change in peak intensity at about 100 kDa and the increase of band intensity in the
positions or crystallinity [104,105]. Some of the amorphous regions upper position when treated with MTGase [112,116].
were oriented well after crosslinking that was indicated by the
increase in the diffraction intensity followed by a decrease in the 6.3. Fourier transform infrared spectroscopy
crystallinity of the crosslinked starch films.
The effect of citric acid on crystalline structure of the crosslinked FTIR can be considered as a practical and supplementary method
starch-polyvinyl films was studied by Shi et al. [82]. XRD patterns to characterize the microstructures of crosslinked biopolymer
showed a high percentage of crystallinity (␹c = 54%) in PVA which films. In fact, it allows identifying the possible functional chemical
presents the typical diffraction peak at 2␪ 19.3◦ [106], while corn groups and molecular interactions [11].
starch presents peaks at 2␪ 15◦ , 17◦ , 18◦ , 23◦ and 26.5◦ [107]. Coma et al. [46] assessed the degree of reticulation of hydrox-
Most of the diffraction peaks in the PVA-starch-glycerol blend films ypropyl methyl cellulose (HPMC) with citric acid, comparing the
were more similar to PVA ones than those of the neat corn starch. results of FTIR analysis at ester band (1735 cm−1 ) with acid–base
This indicated that the crystalline structure of starch granules was titration method, to determine the rate of crosslinking and ester
decomposed due to gelatinization during the blending with PVA bond in crosslinked cellulosic derivatives. The comparison between
at temperatures >100 ◦ C. XRD patterns exhibited no retrogradation the reticulation rates calculated from both methods showed a pos-
of starch, indicating that the PVA and glycerol molecules strongly itive linear relationship. They concluded that FTIR spectroscopy,
prevented the crystalline structure of starch and PVA diffraction compared to titration method, is an easy, rapid and acceptable
dominated in the blends. The intensity of the peak related to the method for the approximate determination of the ester bond
PVA at 2␪ 19.3◦ decreased with the addition of glycerol and cit- formation rate directly on cellulosic films crosslinked by polycar-
ric acid. This could be due to plasticizing and crosslinking effects boxylic acid. In another work, Reddy and Yang investigated the
of citric acid that reduced the crystallinity of the PVA. Moreover, crosslinking of starch films by citric acid through FTIR [42]. The
XRD pattern of PVA-starch-citric acid blend showed a new peak comparison between the FTIR spectra of the non-crosslinked films
at around 2␪ 31.5◦ , ascribable to the residual unreacted citric acid. and films crosslinked with 5% citric acid revealed an extra peak
On the other hand, this peak was not detected in the XRD curve in the crosslinked film at around 1724 cm−1 that is related to the
of PVA-starch-citric acid with glycerol, probably because glycerol C O stretching vibrational mode [117]. It was improbably that all
kept the citric acid from assembles, and the citric acid attended in the carboxyl groups are esterified and this peak could be a coales-
the blends in the molecular level. cent one due to both the ester bond and carboxyl carbonyl groups
in citric acid. However, all the samples were washed before the
test, to remove thoroughly the effects of unbound citric acid in the
704 F. Garavand et al. / International Journal of Biological Macromolecules 104 (2017) 687–707

blend. Therefore, the presence of this peak confirmed the chemical weak absorption peak related to the stretching vibration of B OC
linkages between starch and citric acid. Similarly, Shi et al. [82] at wave number of about 1035 cm−1 . Moreover, the peak of OH
investigated the effects of citric acid as a crosslinker on chemi- stretching (3000–3550 cm−1 ) was shifted to a higher wave number.
cal structure of starch-PVA films. The IR spectra of PVA-citric acid These results indicated the occurrence of the crosslinking reaction
and starch-citric acid exhibited the peak at 1729 cm−1 , demonstrat- between boric acid with PVA and/or BCNC.
ing the occurrence of esterification between citric acid and either
starch or PVA. However, the peak at 1729 cm−1 of PVA-citric acid 6.4. Nuclear magnetic resonance (NMR) spectroscopy
was very weaker with respect to that revealed in the case of starch-
citric acid, suggesting more esterification between the citric acid NMR spectroscopy uses the nuclear magnetic resonance of spe-
and starch than PVA. In addition, as expected, the carbonyl peak cific atoms to provide information about the chemical and physical
intensity increased with increasing the citric acid concentration. properties of atoms or their related molecules, as well as reaction
On the other hand, the carbonyl peak of PVA-starch-citric acid with state, dynamics, structure and chemical environment. This tech-
glycerol was lower than the blend without glycerol. The reason was nique can be used to verify the occurred crosslinking of polymers
probably that some citric acid molecules reacted with glycerol to and to investigate the Maillard reaction. Su et al. used solid-state
obtain glycerol citrates which were solubilized into the water dur- 13 C NMR analysis to prove crosslinking by Maillard reaction in SPI-
ing immersion. As a result, the addition of glycerol prevented the CMC blend films [17]. 13 C NMR spectra of pure SPI showed that a
crosslinking reaction between citric acid and starch or PVA. Fur- characteristic sharp peak of C1 (the carboxylic carbon of SPI, back-
thermore, it is worthy to underline that the intensity of the peak bone carbonyl group) could be clearly identified at approximately
around 3340 cm−1 (stretching vibration of the hydroxyl groups) of 173 ppm [123]. The C2 carbon spectrum contained mainly bands in
PVA-starch-glycerol was lower than that of PVA-starch-citric acid, the alkyl region, with a prominent peak at about 50 ppm from non-
probably due to the fact that citric acid can form stronger hydro- transformed NCH2 moieties [123], implying that Maillard reaction
gen bonds with PVA or starch compared to glycerol. As expected, occurred by consuming −NH2 . The intensity of the characteristic
higher citric acid concentrations caused an increase in the number 13 C peak of pure SPI at 173 ppm gradually decreased as the propor-
of the hydroxyl and carboxyl groups and, thus, of the hydroxyl peak tion of CMC in the blends increased. There was no signal at 50 ppm,
height. approving occurrence of Maillard reactions between SPI and CMC
The rate of Maillard reactions in CMC-SPI films plasticized by blends.
glycerol was studied by Su et al. using FTIR spectroscopy [38].
The FTIR spectra of various weight ratios of CMC-SPI blend films
showed that the intensity of the bands at 3100–2900, 1600–1400, 7. Cytotoxicity
and 890 cm−1 gradually decreased as CMC was added to SPI. These
changes revealed modifications of the amide I, II and III vibra- Several studies have shown that the synthetic crosslinking
tions. These results showed that the hydroxyl group in CMC and agents, including carbodiimide and hexamethylenediisocyanate,
amino groups in SPI were consumed during the blending process are all cytotoxic and dangerous for human body and nature. There-
at high temperature. Furthermore, a new band at 1647 cm−1 indi- fore, it is delightful to provide a safe crosslinking agent to be used
cated a C N stretching vibration assigned to the Maillard reaction in packaging materials that has low cytotoxicity and to construct a
[118–120]. Therefore, complex Maillard reactions are expected to stable and biocompatible film or coating [124]. Natural crosslinking
occur between CMC and SPI. agents, such as those originated from microbial fermentation and
Rimdusit et al. [45] characterized the structure of methyl cellu- those from plant materials, show appropriate potential both in case
lose (MC)-glutaraldehyde crosslinked films by FTIR spectroscopy. of constructing powerful bonds between biopolymer chains and
For crosslinked MC, the peaks at 1710 cm−1 (carbonyl stretch- safe or less-cytotoxic effects. Genipin, transglutaminase, carboxylic
ing band from aldehyde group of glutaraldehyde) indicated the acids, and crosslinking via Maillard reaction are examples of safe
crosslinkage between MC and glutaraldehyde. With an increase and naturally-occurred crosslinking agents applied in biopolymers
in the glutaraldehyde above 6.0 wt%, the peak at 1710 cm−1 sub- till now.
stantially increased, due to an increment in unbound aldehyde. For example poly(carboxylic acids), including boric acid and
Therefore, the optimal glutaraldehyde content was below 6.0 wt% citric acid, are cost-effective and non-toxic components which
and this peak was not found at the concentration of 4.5 wt%. More- improve the functionality of a wide range of biopolymer materials,
over, the appearance of this absorption spectrum suggested the from proteins to polysaccharides [42]. Some of the crosslinkers, like
formation of inter-molecular hydrogen bonding, similar to the borate, not only are accessible and cost-effective but also introduce
results reported for chitosan crosslinked with glutaraldehyde [121]. some advantages to the soil and environment [125]. As well, some
The potential crosslinking activity of tannic acid was studied in of the complications, such as toxicity and calcification as a result of
the chitosan films by FTIR [49]. Rivero et al. [49] applied the sec- crosslinking by aldehyde groups, especially glutaraldehyde, have
ond derivative method in FTIR spectra to enhance the resolution been effectively lightened by applying glycine or glutamic acid
and verify the peak positions because the exact wavenumber posi- as post-treatment. Glutamic acid treatment is also supposed to
tion of biological samples is sometimes difficult to be identified. quench reactive aldehyde groups by connecting amino groups from
The IR spectrum of crosslinked chitosan films was similar to that the acid end [126].
of non-crosslinked ones. Nevertheless, the intensity of absorption
peaks decreased as a result of crosslinking reaction; similarly, the 8. Concluding remarks
second derivative method proved this result. According to Aelenei
et al. [122] typical bands of the tannic acid were either weakened Biopolymer films, based on various carbohydrates and proteins
or covered, as shoulders under other bands, which confirmed that and, in a fewer extent, on some lipids, possess favorable barrier
chitosan matrix consumed the free tannic acid. Furthermore, a syn- characteristics, but, in some cases, exhibit weak mechanical, struc-
ergistic effect of components glycerol and tannic acid was shown in tural and thermal attributes, along with their moisture-sensitivity,
the spectra of crosslinked chitosan films plasticized with glycerol. compared to well-established commercial polymers. Several
Effect of boric acid (as a crosslinker) on FTIR spectrum of endeavors were recently devoted to investigate the possibility
poly(vinyl alcohol) (PVA)/bacterial cellulose nanocrystals (BCNC) to improve permeability and/or thermo-mechanical attributes of
films was studied by Rouhi et al. [11]. They observed a new and film-forming biopolymers by introducing crosslinks into the film
F. Garavand et al. / International Journal of Biological Macromolecules 104 (2017) 687–707 705

network. Crosslinking reactions enable the preparation of materials [24] I. Migneault, C. Dartiguenave, M.J. Bertrand, K.C. Waldron, Glutaraldehyde:
with enhanced mechanical, thermal and physicochemical prop- behavior in aqueous solution, reaction with proteins, and application to
enzyme crosslinking, Biotechniques 37 (5) (2004) 790–806.
erties. Particularly, naturally-occurred crosslinking agents have [25] G. Tillet, B. Boutevin, B. Ameduri, Chemical reactions of polymer
attracted more attention in order to take into account environmen- crosslinking and post-crosslinking at room and medium temperature, Prog.
tal and health concerns, along with the economic issues. It can be Polym. Sci. 36 (2) (2011) 191–217.
[26] G. Galietta, L. Di Gioia, S. Guilbert, B. Cuq, Mechanical and
concluded that crosslinking approach is a cost-effective and effi- thermomechanical properties of films based on whey proteins as affected by
cient method compared to other reinforcement strategies, e.g. the plasticizer and crosslinking agents, J. Dairy Sci. 81 (12) (1998) 3123–3130.
addition of nanomaterials in order to obtain biopolymeric matrix [27] T. Tanabe, N. Okitsu, K. Yamauchi, Fabrication and characterization of
chemically crosslinked keratin films, Mater. Sci. Eng.: C 24 (3) (2004)
composites which present inherent drawbacks.
441–446.
[28] N. Cao, Y. Fu, J. He, Mechanical properties of gelatin films cross-linked,
respectively, by ferulic acid and tannin acid, Food Hydrocoll. 21 (4) (2007)
References 575–584.
[29] C. Mu, J. Guo, X. Li, W. Lin, D. Li, Preparation and properties of dialdehyde
[1] E.L. Bradley, L. Castle, Q. Chaudhry, Applications of nanomaterials in food carboxymethyl cellulose crosslinked gelatin edible films, Food Hydrocoll. 27
packaging with a consideration of opportunities for developing countries, (1) (2012) 22–29.
Trends Food Sci. Technol. 22 (11) (2011) 604–610. [30] R. De Carvalho, C. Grosso, Characterization of gelatin based films modified
[2] D. Dehnad, H. Mirzaei, Z. Emam-Djomeh, S.-M. Jafari, S. Dadashi, Thermal with transglutaminase, glyoxal and formaldehyde, Food Hydrocoll. 18 (5)
and antimicrobial properties of chitosan-nanocellulose films for extending (2004) 717–726.
shelf life of ground meat, Carbohydr. Polym. 109 (2014) 148–154. [31] H. Chambi, C. Grosso, Edible films produced with gelatin and casein
[3] J.-W. Rhim, H.-M. Park, C.-S. Ha, Bio-nanocomposites for food packaging cross-linked with transglutaminase, Food Res. Int. 39 (4) (2006) 458–466.
applications, Prog. Polym. Sci. 38 (10) (2013) 16–1629. [32] A. Bigi, G. Cojazzi, S. Panzavolta, K. Rubini, N. Roveri, Mechanical and
[4] M. Avella, R. Avolio, E. Di Pace, M.E. Errico, G. Gentile, M.G. Volpe, thermal properties of gelatin films at different degrees of glutaraldehyde
Polymer-based nanocomposites for food packaging applications, crosslinking, Biomaterials 22 (8) (2001) 763–768.
bio-nanotechnology: a revolution in food, Biomed. Health Sci. (2013) [33] A. Bigi, G. Cojazzi, S. Panzavolta, N. Roveri, K. Rubini, Stabilization of gelatin
212–226. films by crosslinking with genipin, Biomaterials 23 (24) (2002) 4827–4832.
[5] S.F. Hosseini, M. Rezaei, M. Zandi, F. Farahmandghavi, Fabrication of [34] R. Bhat, A. Karim, Towards producing novel fish gelatin films by combination
bio-nanocomposite films based on fish gelatin reinforced with chitosan treatments of ultraviolet radiation and sugars (ribose and lactose) as
nanoparticles, Food Hydrocoll. 44 (2015) 172–182. cross-linking agents, J. Food Sci. Technol. 51 (7) (2014) 1326–1333.
[6] D. Dehnad, Z. Emam-Djomeh, H. Mirzaei, S.M. Jafari, S. Dadashi, Optimization [35] J. Yi, Y. Kim, H. Bae, W. Whiteside, H. Park, Influence of
of physical and mechanical properties for chitosan-nanocellulose transglutaminase-induced cross-linking on properties of fish gelatin films, J.
biocomposites, Carbohydr. Polym. 105 (2014) 222–228. Food Sci. 71 (9) (2006) E376–E383.
[7] S. Tajik, Y. Maghsoudlou, F. Khodaiyan, S.M. Jafari, M. Ghasemlou, M. Aalami, [36] H.J. Bae, D.O. Darby, R.M. Kimmel, H.J. Park, W.S. Whiteside, Effects of
Soluble soybean polysaccharide: a new carbohydrate to make a transglutaminase-induced cross-linking on properties of fish
biodegradable film for sustainable green packaging, Carbohydr. Polym. 97 gelatin-nanoclay composite film, Food Chem. 114 (1) (2009) 180–189.
(2) (2013) 817–824. [37] N. Reddy, Q. Jiang, Y. Yang, Preparation and properties of peanut protein
[8] L. Cabedo, J. Luis Feijoo, M. Pilar Villanueva, J.M. Lagarón, E. Giménez, films crosslinked with citric acid, Ind. Crops Prod. 39 (2012) 26–30.
Optimization of biodegradable nanocomposites based on aPLA/PCL blends [38] J.-F. Su, Z. Huang, X.-Y. Yuan, X.-Y. Wang, M. Li, Structure and properties of
for food packaging applications, Macromol. Symp. (2006) 191–197, Wiley carboxymethyl cellulose/soy protein isolate blend edible films crosslinked
Online Library. by Maillard reactions, Carbohydr. Polym. 79 (1) (2010) 145–153.
[9] B. Ghanbarzadeh, A. Oromiehi, Biodegradable biocomposite films based on [39] E. Zactiti, T. Kieckbusch, Potassium sorbate permeability in biodegradable
whey protein and zein: barrier, mechanical properties and AFM analysis, Int. alginate films: effect of the antimicrobial agent concentration and
J. Biol. Macromol. 43 (2) (2008) 209–215. crosslinking degree, J. Food Eng. 77 (3) (2006) 462–467.
[10] X. Tang, P. Kumar, S. Alavi, K. Sandeep, Recent advances in biopolymers and [40] S. Benavides, R. Villalobos-Carvajal, J. Reyes, Physical, mechanical and
biopolymer-based nanocomposites for food packaging materials, Crit. Rev. antibacterial properties of alginate film: effect of the crosslinking degree
Food Sci. Nutr. 52 (5) (2012) 426–442. and oregano essential oil concentration, J. Food Eng. 110 (2) (2012) 232–239.
[11] M. Rouhi, S.H. Razavi, S.M. Mousavi, Optimization of crosslinked poly(vinyl [41] A.C. Bierhalz, M.A. da Silva, M.E. Braga, H.J. Sousa, T.G. Kieckbusch, Effect of
alcohol) nanocomposite films for mechanical properties, Mater. Sci. Eng.: C calcium and/or barium crosslinking on the physical and antimicrobial
71 (2017) 1052–1063. properties of natamycin-loaded alginate films, LWT-Food Sci. Technol. 57
[12] A.P. Kumar, R.P. Singh, Biocomposites of cellulose reinforced starch: (2) (2014) 494–501.
improvement of properties by photo-induced crosslinking, Bioresour. [42] N. Reddy, Y. Yang, Citric acid cross-linking of starch films, Food Chem. 118
Technol. 99 (18) (2008) 8803–8809. (3) (2010) 702–711.
[13] H.M. Azeredo, K.W. Waldron, Crosslinking in polysaccharide and protein [43] B. Ghanbarzadeh, H. Almasi, A.A. Entezami, Improving the barrier and
films and coatings for food contact—a review, Trends Food Sci. Technol. 52 mechanical properties of corn starch-based edible films: effect of citric acid
(2016) 109–122. and carboxymethyl cellulose, Ind. Crops Prod. 33 (1) (2011) 229–235.
[14] M.A. da Silva, A.C.K. Bierhalz, T.G. Kieckbusch, Alginate and pectin composite [44] L. Goetz, A. Mathew, K. Oksman, P. Gatenholm, A.J. Ragauskas, A novel
films crosslinked with Ca2+ ions: effect of the plasticizer concentration, nanocomposite film prepared from crosslinked cellulosic whiskers,
Carbohydr. Polym. 77 (4) (2009) 736–742. Carbohydr. Polym. 75 (1) (2009) 85–89.
[15] J. Delville, C. Joly, P. Dole, C. Bliard, Solid state photocrosslinked starch based [45] S. Rimdusit, S. Jingjid, S. Damrongsakkul, S. Tiptipakorn, T. Takeichi,
films: a new family of homogeneous modified starches, Carbohydr. Polym. Biodegradability and property characterizations of methyl cellulose: effect
49 (1) (2002) 71–81. of nanocompositing and chemical crosslinking, Carbohydr. Polym. 72 (3)
[16] S. Figueiró, J.C. Góes, R. Moreira, A. Sombra, On the physico-chemical and (2008) 444–455.
dielectric properties of glutaraldehyde crosslinked galactomannan-collagen [46] V. Coma, I. Sebti, P. Pardon, F. Pichavant, A. Deschamps, Film properties from
films, Carbohydr. Polym. 56 (3) (2004) 313–320. crosslinking of cellulosic derivatives with a polyfunctional carboxylic acid,
[17] J.-F. Su, X.-Y. Yuan, Z. Huang, X.-Y. Wang, X.-Z. Lu, L.-D. Zhang, S.-B. Wang, Carbohydr. Polym. 51 (3) (2003) 265–271.
Physicochemical properties of soy protein isolate/carboxymethyl cellulose [47] H.M. Azeredo, C. Kontou-Vrettou, G.K. Moates, N. Wellner, K. Cross, P.H.
blend films crosslinked by Maillard reactions: color, transparency and Pereira, K.W. Waldron, Wheat straw hemicellulose films as affected by citric
heat-sealing ability, Mater. Sci. Eng.: C 32 (1) (2012) 40–46. acid, Food Hydrocoll. 50 (2015) 1–6.
[18] M. Khanzadi, S.M. Jafari, H. Mirzaei, F.K. Chegini, Y. Maghsoudlou, D. [48] S. Tripathi, G. Mehrotra, P. Dutta, Physicochemical and bioactivity of
Dehnad, Physical and mechanical properties in biodegradable films of whey cross-linked chitosan–PVA film for food packaging applications, Int. J. Biol.
protein concentrate-pullulan by application of beeswax, Carbohydr. Polym. Macromol. 45 (4) (2009) 372–376.
118 (2015) 24–29. [49] S. Rivero, M. García, A. Pinotti, Crosslinking capacity of tannic acid in
[19] V. Siracusa, P. Rocculi, S. Romani, M. Dalla Rosa, Biodegradable polymers for plasticized chitosan films, Carbohydr. Polym. 82 (2) (2010) 270–276.
food packaging: a review, Trends Food Sci. Technol. 19 (12) (2008) 634–643. [50] A.A. Alhwaige, T. Agag, H. Ishida, S. Qutubuddin, Biobased
[20] P. Bordes, E. Pollet, L. Avérous, Nano-biocomposites: biodegradable chitosan/polybenzoxazine cross-linked films: preparation in aqueous media
polyester/nanoclay systems, Prog. Polym. Sci. 34 (2) (2009) 125–155. and synergistic improvements in thermal and mechanical properties,
[21] F. Chivrac, E. Pollet, L. Avérous, Progress in nano-biocomposites based on Biomacromolecules 14 (6) (2013) 1806–1815.
polysaccharides and nanoclays, Mater. Sci. Eng. R: Rep. 67 (1) (2009) 1–17. [51] C.M. Yeng, S. Husseinsyah, S.S. Ting, Chitosan/corn cob biocomposite films
[22] S.M. Jafari, M. Khanzadi, H. Mirzaei, D. Dehnad, F.K. Chegini, Y. by cross-linking with glutaraldehyde, BioResources 8 (2) (2013) 2910–2923.
Maghsoudlou, Hydrophobicity, thermal and micro-structural properties of [52] K.S. Mikkonen, M.I. Heikkilä, S.M. Willför, M. Tenkanen, Films from
whey protein concentrate-pullulan-beeswax films, Int. J. Biol. Macromol. 80 glyoxal-crosslinked spruce galactoglucomannans plasticized with sorbitol,
(2015) 506–511. Int. J. Polym. Sci. 2012 (2012).
[23] R. Tharanathan, Biodegradable films and composite coatings: past, present
and future, Trends Food Sci. Technol. 14 (3) (2003) 71–78.
706 F. Garavand et al. / International Journal of Biological Macromolecules 104 (2017) 687–707

[53] K.S. Mikkonen, J. Schmidt, A.-H. Vesterinen, M. Tenkanen, Crosslinking with [81] W. Ning, Y. Jiugao, M. Xiaofei, W. Ying, The influence of citric acid on the
ammonium zirconium carbonate improves the formation and properties of properties of thermoplastic starch/linear low-density polyethylene blends,
spruce galactoglucomannan films, J. Mater. Sci. 48 (12) (2013) 4205–4213. Carbohydr. Polym. 67 (3) (2007) 446–453.
[54] M. Abdollahi, M. Rezaei, G. Farzi, Improvement of active chitosan film [82] R. Shi, J. Bi, Z. Zhang, A. Zhu, D. Chen, X. Zhou, L. Zhang, W. Tian, The effect of
properties with rosemary essential oil for food packaging, Int. J. Food Sci. citric acid on the structural properties and cytotoxicity of the polyvinyl
Technol. 47 (4) (2012) 847–853. alcohol/starch films when molding at high temperature, Carbohydr. Polym.
[55] S.F. Hosseini, M. Rezaei, M. Zandi, F. Farahmandghavi, Preparation and 74 (4) (2008) 763–770.
characterization of chitosan nanoparticles-loaded fish gelatin-based edible [83] X. Ma, P.R. Chang, J. Yu, M. Stumborg, Properties of biodegradable citric
films, J. Food Process Eng. 39 (5) (2015) 521–530. acid-modified granular starch/thermoplastic pea starch composites,
[56] S. Ou, Y. Wang, S. Tang, C. Huang, M.G. Jackson, Role of ferulic acid in Carbohydr. Polym. 75 (1) (2009) 1–8.
preparing edible films from soy protein isolate, J. Food Eng. 70 (2) (2005) [84] C.-H. Tang, Y. Jiang, Modulation of mechanical and surface hydrophobic
205–210. properties of food protein films by transglutaminase treatment, Food Res.
[57] M. Lacroix, T. Le, B. Ouattara, H. Yu, M. Letendre, S. Sabato, M. Mateescu, G. Int. 40 (4) (2007) 504–509.
Patterson, Use of ␥-irradiation to produce films from whey, casein and soya [85] J. Fang, P. Fowler, C. Escrig, R. Gonzalez, J. Costa, L. Chamudis, Development
proteins: structure and functionals characteristics, Radiat. Phys. Chem. 63 of biodegradable laminate films derived from naturally occurring
(3) (2002) 827–832. carbohydrate polymers, Carbohydr. Polym. 60 (1) (2005) 39–42.
[58] B. Ouattara, L. Canh, C. Vachon, M. Mateescu, M. Lacroix, Use of ␥-irradiation [86] G. Klančnik, J. Medved, P. Mrvar, Differential thermal analysis (DTA) and
cross-linking to improve the water vapor permeability and the chemical differential scanning calorimetry (DSC) as a method of material
stability of milk protein films, Radiat. Phys. Chem. 63 (3) (2002) 821–825. investigation Diferenčna termična analiza (DTA) in diferenčna vrstična
[59] M. Lee, S. Lee, K.B. Song, Effect of ␥-irradiation on the physicochemical kalorimetrija (DSC) kot metoda za raziskavo materialov, RMZ-Mater.
properties of soy protein isolate films, Radiat. Phys. Chem. 72 (1) (2005) Geoenviron. 57 (1) (2010) 127–142.
35–40. [87] P.J. Haines, M. Reading, F.W. Wilburn, Differential thermal analysis and
[60] M.R. de Moura, M.V. Lorevice, L.H. Mattoso, V. Zucolotto, Highly stable, differential scanning calorimetry, in: E.B. Michael (Ed.), Handbook of
edible cellulose films incorporating chitosan nanoparticles, J. Food Sci. 76 (2) Thermal Analysis and Calorimetry, Elsevier Science B.V., 1998 (Chapter 5,
(2011) N25–N29. pp. 279–361).
[61] C. Menzel, E. Olsson, T.S. Plivelic, R. Andersson, C. Johansson, R. Kuktaite, L. [88] D. Achet, X. He, Determination of the renaturation level in gelatin films,
Järnström, K. Koch, Molecular structure of citric acid cross-linked starch Polymer 36 (4) (1995) 787–791.
films, Carbohydr. Polym. 96 (1) (2013) 270–276. [89] A. Bigi, B. Bracci, G. Cojazzi, S. Panzavolta, N. Roveri, Drawn gelatin films with
[62] S. Wang, J. Ren, W. Li, R. Sun, S. Liu, Properties of polyvinyl alcohol/xylan improved mechanical properties, Biomaterials 19 (24) (1998) 2335–2340.
composite films with citric acid, Carbohydr. Polym. 103 (2014) 94–99. [90] I. Rault, V. Frei, D. Herbage, N. Abdul-Malak, A. Huc, Evaluation of different
[63] X. Ma, P.R. Chang, J. Yu, Properties of biodegradable thermoplastic pea chemical methods for cros-linking collagen gel, films and sponges, J. Mater.
starch/carboxymethyl cellulose and pea starch/microcrystalline cellulose Sci. 7 (4) (1996) 215–221.
composites, Carbohydr. Polym. 72 (3) (2008) 369–375. [91] R. Usha, T. Ramasami, Effect of crosslinking agents (basic chromium sulfate
[64] T.J. Gutiérrez, M.S. Tapia, E. Pérez, L. Famá, Structural and mechanical and formaldehyde) on the thermal and thermomechanical stability of rat
properties of edible films made from native and modified cush–cush yam tail tendon collagen fibre, Thermochim. Acta 356 (1) (2000) 59–66.
and cassava starch, Food Hydrocoll. 45 (2015) 211–217. [92] J. Kopp, M. Bonnet, J. Renou, Effect of collagen crosslinking on collagen-water
[65] A.C.K. Bierhalz, M.A. da Silva, T.G. Kieckbusch, Natamycin release from interactions (a DSC investigation), Matrix 9 (6) (1990) 443–450.
alginate/pectin films for food packaging applications, J. Food Eng. 110 (1) [93] G. Goissis, S.D. Figueiró, D.M. Braile, R.B. de Araujo, V.D. Ramirez, Reticulação
(2012) 18–25. Progressiva de Pericárdio Bovino com Glutaraldeído para Confecção de
[66] B.-S. Chiou, R.J. Avena-Bustillos, P.J. Bechtel, H. Jafri, R. Narayan, S.H. Imam, Válvulas Cardíacas Biológicas, Polím.: Ciênc. Tecnol. 8 (1998) 46–54.
G.M. Glenn, W.J. Orts, Cold water fish gelatin films: effects of cross-linking [94] E. Casagranda, J.A. Werkmeister, J. Bamshaw, G. Ellender, Evaluation of
on thermal, mechanical, barrier, and biodegradation properties, Eur. Polym. alternative glutaraldehyde stabilization strategies for collagenous
J. 44 (11) (2008) 3748–3753. biomaterials, J. Mater. Sci.: Mater. Med. 5 (6–7) (1994) 332–337.
[67] H. Li, X. Gao, Y. Wang, X. Zhang, Z. Tong, Comparison of chitosan/starch [95] Y. Jiugao, W. Ning, M. Xiaofei, The effects of citric acid on the properties of
composite film properties before and after cross-linking, Int. J. Biol. thermoplastic starch plasticized by glycerol, Starch-Starke 57 (2005)
Macromol. 52 (2013) 275–279. 494–504.
[68] C. Bilbao-Sainz, J. Bras, T. Williams, T. Sénechal, W. Orts, HPMC reinforced [96] J.-S. Kim, Y.-S. Lee, Effect of reaction pH on enolization and racemization
with different cellulose nano-particles, Carbohydr. Polym. 86 (4) (2011) reactions of glucose and fructose on heating with amino acid enantiomers
1549–1557. and formation of melanoidins as result of the Maillard reaction, Food Chem.
[69] A. Leerahawong, R. Arii, M. Tanaka, K. Osako, Edible film from squid 108 (2) (2008) 582–592.
(Todarodes pacificus) mantle muscle, Food Chem. 124 (1) (2011) 177–182. [97] C.-y. Lii, H.-H. Chen, S. Lu, P. Tomasik, Electrosynthesis of ␬-carrageenan
[70] J. Guo, L. Ge, X. Li, C. Mu, D. Li, Periodate oxidation of xanthan gum and its complexes with gelatin, J. Polym. Environ. 11 (3) (2003) 115–121.
crosslinking effects on gelatin-based edible films, Food Hydrocoll. 39 (2014) [98] J.-F. Su, Z. Huang, K. Liu, L.-L. Fu, H.-R. Liu, Mechanical properties,
243–250. biodegradation and water vapor permeability of blend films of soy protein
[71] Y. Dou, X. Huang, B. Zhang, M. He, G. Yin, Y. Cui, Preparation and isolate and poly(vinyl alcohol) compatibilized by glycerol, Polym. Bull. 58
characterization of a dialdehyde starch crosslinked feather keratin film for (5–6) (2007) 913–921.
food packaging application, RSC Adv. 5 (34) (2015) 27168–27174. [99] A. Coats, J. Redfern, Thermogravimetric analysis. A review, Analyst 88 (1053)
[72] C. Schramm, S.B. Vukušić, D. Katović, Non-formaldehyde durable press (1963) 906–924.
finishing of dyed fabrics: evaluation of cotton-bound polycarboxylic acids, [100] T. Uragami, T. Matsuda, H. Okuno, T. Miyata, Structure of chemically
Color. Technol. 118 (5) (2002) 244–249. modified chitosan membranes and their characteristics of permeation and
[73] B. Kottes Andrews, C. Welch, B. Trask-Morell, Efficient ester crosslink separation of aqueous ethanol solutions, J. Membr. Sci. 88 (2) (1994)
finishing for formaldehyde-free durable press cotton fabrics, Am. Dyest. 243–251.
Rep. (June) (1989) 15–23. [101] S.-F. Wang, L. Shen, W.-D. Zhang, Y.-J. Tong, Preparation and mechanical
[74] P. Hernández-Muñoz, R. Villalobos, A. Chiralt, Effect of cross-linking using properties of chitosan/carbon nanotubes composites, Biomacromolecules 6
aldehydes on properties of glutenin-rich films, Food Hydrocoll. 18 (3) (2004) (6) (2005) 3067–3072.
403–411. [102] K. Ogawa, T. Yui, K. Okuyama, Three D structures of chitosan, Int. J. Biol.
[75] S. Park, D. Bae, K. Rhee, Soy protein biopolymers cross-linked with Macromol. 34 (1) (2004) 1–8.
glutaraldehyde, J. Am. Oil Chem. Soc. 77 (8) (2000) 879–884. [103] C.D. Brown, L. Kreilgaard, M. Nakakura, N. Caram-Lelham, D.K. Pettit, W.R.
[76] G. Makishi, R. Lacerda, A. Bittante, H. Chambi, P. Costa, C. Gomide, R. Gombotz, A.S. Hoffman, Release of PEGylated granulocyte-macrophage
Carvalho, P. Sobral, Films based on castor bean (Ricinus communis L.) colony-stimulating factor from chitosan/glycerol films, J. Control. Release 72
proteins crosslinked with glutaraldehyde and glyoxal, Ind. Crops Prod. 50 (1) (2001) 35–46.
(2013) 375–382. [104] S. Garg, A.K. Jana, Studies on the properties and characteristics of
[77] C. Jo, H. Kang, N.Y. Lee, J.H. Kwon, M.W. Byun, Pectin-and gelatin-based film: starch-LDPE blend films using cross-linked, glycerol modified, cross-linked
effect of gamma irradiation on the mechanical properties and and glycerol modified starch, Eur. Polym. J. 43 (9) (2007) 3976–3987.
biodegradation, Radiat. Phys. Chem. 72 (6) (2005) 745–750. [105] M. Kim, S.-J. Lee, Characteristics of crosslinked potato starch and
[78] A. González, M.C. Strumia, C.I.A. Igarzabal, Cross-linked soy protein as starch-filled linear low-density polyethylene films, Carbohydr. Polym. 50 (4)
material for biodegradable films: synthesis, characterization and (2002) 331–337.
biodegradation, J. Food Eng. 106 (4) (2011) 331–338. [106] L. García-Cerda, M. Escareno-Castro, M. Salazar-Zertuche, Preparation and
[79] Y. Wang, A. Liu, R. Ye, W. Wang, X. Li, Transglutaminase-induced characterization of polyvinyl alcohol-cobalt ferrite nanocomposites, J.
crosslinking of gelatin-calcium carbonate composite films, Food Chem. 166 Non-Cryst. Solids 353 (8) (2007) 808–810.
(2015) 414–422. [107] R. Shi, Q. Liu, T. Ding, Y. Han, L. Zhang, D. Chen, W. Tian, Ageing of soft
[80] W. Weng, H. Zheng, Effect of transglutaminase on properties of tilapia scale thermoplastic starch with high glycerol content, J. Appl. Polym. Sci. 103 (1)
gelatin films incorporated with soy protein isolate, Food Chem. 169 (2015) (2007) 574–586.
255–260. [108] U.K. Laemmli, Cleavage of structural proteins during the assembly of the
head of bacteriophage T4, Nature 227 (5259) (1970) 680–685.
F. Garavand et al. / International Journal of Biological Macromolecules 104 (2017) 687–707 707

[109] L. Creamer, Casein nomenclature, structure and association properties, in: H. [119] E.H. Ajandouz, V. Desseaux, S. Tazi, A. Puigserver, Effects of temperature and
Roginski, J. Fulquay, P. Fox (Eds.), Encyclopedia of Dairy Sciences: Milk pH on the kinetics of caramelisation, protein cross-linking and Maillard
Proteins, Academic Press, New York, 2002, pp. 1895–1902. reactions in aqueous model systems, Food Chem. 107 (3) (2008) 1244–1252.
[110] J. Poppe, Gelatin, Thickening and Gelling Agents for Food, Springer, 1997 [120] M. Alaiz, F.J. Hidalgo, R. Zamora, Effect of pH and temperature on
(pp. 144–168). comparative antioxidant activity of nonenzymatically browned proteins
[111] J.H. Oh, B. Wang, P.D. Field, H.A. Aglan, Characteristics of edible films made produced by reaction with oxidized lipids and carbohydrates, J. Agric. Food
from dairy proteins and zein hydrolysate cross-linked with Chem. 47 (2) (1999) 748–752.
transglutaminase, Int. J. Food Sci. Technol. 39 (3) (2004) 287–294. [121] K. Kurita, Y. Koyama, A. Taniguchi, Studies on chitin. IX. Crosslinking of
[112] M.C. Gómez-Guillén, A.I. Sarabia, M.T. Solas, P. Montero, Effect of microbial water-soluble chitin and evaluation of the products as adsorbents for cupric
transglutaminase on the functional properties of megrim (Lepidorhombus ion, J. Appl. Polym. Sci. 31 (5) (1986) 1169–1176.
boscii) skin gelatin, J. Sci. Food Agric. 81 (7) (2001) 665–673. [122] N. Aelenei, M.I. Popa, O. Novac, G. Lisa, L. Balaita, Tannic acid incorporation
[113] I. Kołodziejska, K. Kaczorowski, B. Piotrowska, M. Sadowska, Modification of in chitosan-based microparticles and in vitro controlled release, J. Mater.
the properties of gelatin from skins of Baltic cod (Gadus morhua) with Sci.: Mater. Med. 20 (5) (2009) 1095–1102.
transglutaminase, Food Chem. 86 (2) (2004) 203–209. [123] X. Fang, K. Schmidt-Rohr, Fate of the amino acid in glucose- glycine
[114] E. Dickinson, Enzymic crosslinking as a tool for food colloid rheology control melanoidins investigated by solid-state nuclear magnetic resonance (NMR),
and interfacial stabilization, Trends Food Sci. Technol. 8 (10) (1997) J. Agric. Food Chem. 57 (22) (2009) 10701–10711.
334–339. [124] R.A. Muzzarelli, Genipin-crosslinked chitosan hydrogels as biomedical and
[115] M. Motoki, K. Seguro, Transglutaminase and its use for food processing, pharmaceutical aids, Carbohydr. Polym. 77 (1) (2009) 1–9.
Trends Food Sci. Technol. 9 (5) (1998) 204–210. [125] T. Matsunaga, T. Ishii, Borate cross-linked/total rhamnogalacturonan II ratio
[116] L.T. Lim, Y. Mine, M. Tung, Barrier and tensile properties of transglutaminase in cell walls for the biochemical diagnosis of boron deficiency in
cross-linked gelatin films as affected by relative humidity, temperature, and hydroponically grown pumpkin, Anal. Sci. 22 (8) (2006) 1125–1127.
glycerol content, J. Food Sci. 64 (4) (1999) 616–622. [126] J.E. Gough, C.A. Scotchford, S. Downes, Cytotoxicity of glutaraldehyde
[117] C.Q. Yang, B. Andrews, Infrared spectroscopic studies of the crosslinked collagen/poly (vinyl alcohol) films is by the mechanism of
nonformaldehyde durable press finishing of cotton fabrics by use of apoptosis, J. Biomed. Mater. Res. 61 (1) (2002) 121–130.
polycarboxylic acids, J. Appl. Polym. Sci. 43 (9) (1991) 1609–1616.
[118] M.D. Aaslyng, L.M. Larsen, P.M. Nielsen, The influence of maturation on
flavor and chemical composition of hydrolyzed soy protein produced by
acidic and enzymatic hydrolysis, Z. Lebensm. Forsch. A 208 (5–6) (1999)
355–361.

You might also like