You are on page 1of 6

LWT - Food Science and Technology 135 (2021) 109984

Contents lists available at ScienceDirect

LWT
journal homepage: www.elsevier.com/locate/lwt

Structure and properties of chitosan films: Effect of the type of solvent acid
Congde Qiao *, Xianguang Ma , Xujie Wang , Libin Liu
State Key Laboratory of Biobased Material and Green Papermaking, School of Materials Science and Engineering, Qilu University of Technology (Shandong Academy of
Sciences), Jinan, 250353, PR China

A R T I C L E I N F O A B S T R A C T

Keywords: The properties of chitosan films can be modified by changing the solvent type, attributing to the different
Hydrogen bonds interaction patterns between chitosan and acids. However, little is known about how these interactions affect the
Ionic interactions structure and properties of chitosan films. In this work, the influence of acid type on the structure and properties
Thermal transition
of chitosan films was studied by Fourier transform infrared spectroscopy (FTIR), X-ray diffraction (XRD) and
Mechanical properties
differential scanning calorimetry (DSC). FTIR spectra showed that the ionic interactions and hydrogen bonding
could occur between chitosan and acid ions. Structural analysis revealed that chitosan was partially crystalline in
hydrochloric and acetic acid films, whereas it was amorphous in lactic and citric acid films. DSC result indicated
that the glass transition temperature of chitosan was much lower in citric acid film with low water content than
in all the other films, suggesting that the citrate ions interacted more strongly with chitosan. A melting transition
appeared in hydrochloric and acetic acid films. In addition, the tensile strength of these films decreased with an
increase in the volume of acid. These observations indicate that the choice of a proper solvent for chitosan may
be desirable for certain special applications.

1. Introduction Domard, 1994; Zhong, Song, & Li, 2011). Up to now, acetic acid is the
most frequently used solvent. In chitosan solution, the chain confor­
Chitosan has been widely used in food, biomedicine and other in­ mation depends on the acidity of the -NH+ 3 (Park, Choi, & Park, 1983).
dustrial fields due to its distinctive advantages of easy processing, As a cationic polyelectrolyte, the chain flexibility of chitosan is also
biodegradability, low toxicity, biocompatibility, and availability (Hu sensitive to the type of counterions (i.e. the acid ions) (Chen, Wang, Hsu,
et al., 2020; Liu et al., 2020; Narasagoudr, Hegde, Chougale, Masti, & & Tsai, 2009). Thus the solvent type has an important influence on the
Dixit, 2020; Williams, 2011). Nonetheless, the strong inter- and intra­ rheological properties of chitosan solutions and chitosan hydrogels
molecular hydrogen bonds and semi-crystalline structure of chitosan (Lakehal et al., 2019; Soares et al., 2019). In addition, the structure of
render it difficult to be dissolved in water and most organic solvents chitosan in the solid state is closely related to the structural organization
(Tian et al., 2015). It is generally soluble in dilute acidic solutions below of the initial polymer solution (Popa-Nita, Alcouffe, Rochas, David, &
pH 6.0. The solubility of chitosan is attributed to the protonation of Domard, 2010). It has been reported that the chain flexibility is often
amino groups on the polymer chains in acid media, resulting in the manipulated to regulate the properties of chitosan films (Chen, Lin, &
destruction of the hydrogen-bonded networks within chitosan molecules Yang, 1994). Consequently, the solvent acid can alter the structure of
by the electrostatic repulsion between positive charges (Pillai, Paul, & chitosan films, which in turn affect the properties of polymer films. It is
Sharma, 2009). Parameters such as the degree of deacetylation (DD), generally observed that the chitosan films prepared from acetic acid
pH, polymer concentration, polymer molecular weight and temperature solutions have higher junction density and thus possess higher tensile
are known to influence the dissolution process of chitosan (Rinaudo, strength (TS) and Young’s modulus (E) compared to those cast from
2006). other monocarboxylic acids (Chenni, Djidjelli, Boukerrou, Grohens, &
Generally, chitosan is soluble in acid solutions such as hydrochloric Saulnier, 2018; Pavoni, Luchese, & Tessaro, 2019; Velásquez-Cock et al.,
acid and several organic acids including formic, acetic, butyric, malic, 2014; Zhong et al., 2011). As compared with films with monoacid sol­
citric, lactic, oxalic, propionic and succinic acids (Demarger-Andre & vent, the use of polyacid (malic acid and citric acid) solution as the

* Corresponding author. School of Materials Science and Engineering, Qilu University of Technology (Shandong Academy of Sciences), Postal address: Daxue Rd.
3501, Jinan, 250353, PR China.
E-mail address: cdqiao@qlu.edu.cn (C. Qiao).

https://doi.org/10.1016/j.lwt.2020.109984
Received 28 April 2020; Received in revised form 28 July 2020; Accepted 29 July 2020
Available online 12 August 2020
0023-6438/© 2020 Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
C. Qiao et al. LWT 135 (2021) 109984

solvent for chitosan could enhance the mechanical strength of the films, 40 ◦ C for 5 h to gain a homogenous solution, and then filtered through a
as a result of potential ionic crosslinking by polyacids (Chen et al., 2007; piece of PTFE membrane with a pore diameter of 0.45 μm to remove
Khouri, Penlidis, & Moresoli, 2020). Contrarily, it has been demon­ foam and any undissolved impurities. All the film-forming solutions
strated that chitosan films with lactic acid have lower TS and E, due to were casted on a PTFE mold, and dried at ambient temperature for about
the plasticization effect of citrate ions (Park, Marsh, & Rhim; Rhim, 48 h, yielding a series of chitosan films.
2002). In addition to mechanical properties, other properties such as To perform the conditioning experiment, the as-obtained films were
water vapor permeability, antimicrobial activity and thermal stability predried at 50 ◦ C for 24 h in a vacuum drying oven, and then equili­
are also affected by the acid type (Zhong et al.,2011; Pavoni et al.,2019; brated at 25 ◦ C in sealed chambers with relative humidity (RH) of 60%
Khouri et al., 2020). for at least one week prior to further analysis. In addition, a part of these
The solvent acid is known to interact with chitosan through elec­ conditioned films was heated at 110 ◦ C for 60 min in a vacuum oven.
trostatic interactions, as well as hydrogen bonds or hydrophobic in­ FTIR and DSC experiments were immediately performed for these heat-
teractions (Lakehal et al., 2019). All these interactions can influence the treated samples to evaluate the role of moisture in chitosan-acid sys­
structure and properties of chitosan films. Many studies highlight the tems. In this work, the films equilibrated at 60% RH and heated at 110
major impact that acids can have on chitosan through electrostatic in­ ◦
C are noted as hydrated and dry samples, respectively. In addition, the
teractions (Demarger-Andre & Domard, 1994; Khouri et al., 2020; chitosan films made with hydrochloric, acetic, lactic acid and citric acid
Velásquez-Cock et al., 2014; Zhong et al., 2011). In fact, this has been are noted as hydrochloric, acetic, lactic acid and citric acid films,
supported by the simple fact that most of the applications of chitosan are respectively.
based on its cationic nature. Apparently the size of acids, as counterions,
is expected to affect these ionic interactions between amine and car­ 2.3. Fourier-transform infrared (FTIR) spectroscopy
boxylic groups, and thus the properties of chitosan. It has been stated
that the film with larger counterions such as lactate is softer than that Fourier Transform Infrared (FTIR) data for the films were obtained
with acetate (Bégin & Van Calsteren, 1999). In the case of interactions by using a Nicolet iS10 spectrometer (Thermo Fisher Scientific Inc.,
between chitosan and polyacids, hydrogen bonds also contribute Waltham, MA, USA), equipped with an attenuated total reflection (ATR)
significantly. It has been demonstrated the presence of the hydrogen accessory. The FTIR spectra were performed from wavenumber
bonds between the second oxalic carboxyl group and the non-reactive 400–4000 cm− 1 with a resolution of 4 cm− 1 and 128 scans. The data
parts of chitosan in the chitosan-oxalic acid gel, which play a major were analyzed with an Omnic software.
role in its viscoelastic properties (Hamdine, Heuzey, & Begin, 2006).
However, another study indicated that only the ionic interaction pre­ 2.4. X-ray diffraction (XRD)
sented in chitosan–oxalic acid membranes (Fadzallah, Majid, Careem, &
Arof, 2014). Additionally, hydrophobic interactions between chitosan XRD patterns of chitosan films prepared from different acids were
and carboxylic acid could also take place, which could affect the sorp­ recorded by a Bruker.
tion of carboxylic acids on chitosan (Shamov, Bratskaya, & Avramenko, D8 Advance X-Ray Diffractomer equipped with a multichannel de­
2002). tector by use of a Cu Kα1 (λ = 0.1542 nm) monochromatic X-ray beam.
Since the influence of solvents on chitosan is attributed to the The relative intensity was recorded in the scattering range of (2θ) 5–50◦
different interaction patterns between chitosan and acids, the informa­ with a scan rate of 1◦ /min.
tion about these interactions is critical to understanding the relationship
between the structure and properties of chitosan films. It is reasonable to 2.5. Thermogravimetric analysis (TGA)
assume that the strength of these interactions is closely related to the
structure of acids. Although the effect of the acid type on the chitosan TGA measurements were carried out on a TGA-1 (Mettler Toledo
films has been studied extensively, little is known about how these in­ Instrument, Zurich, Switzerland) from 30 ◦ C to 800 ◦ C under nitrogen
teractions affect the structure and properties of chitosan films. In this atmosphere at a heating rate of 10 ◦ C/min.
study, the interactions between chitosan and various acids, including
hydrochloric acid, acetic acid, lactic acid and citric acid, were investi­ 2.6. Differential scanning calorimetry (DSC)
gated in details, along with an exploration of structure–property re­
lationships of chitosan films. The aim of this work was to determine the DSC experiments were performed on a Q 2000 DSC (TA Instrument,
effect of acid type on the interactions between chitosan and acids, and to New Castle, USA) equipped with a cooling accessory RCS90. As bio­
provide a deep understanding of how these interactions impacted this logical macromolecules are easy to absorb water to disturb the deter­
biopolymer at a molecular level. mination of thermal transitions, Tzero hermetic pans (withstand an
internal pressure of 3 atm) were used in this work. The hydrated samples
2. Experimental with mass about 5 mg were sealed in the Tzero hermetic pans with lids
and heated in the DSC cell from − 70 to 120 ◦ C at a heating rate of 5 ◦ C/
2.1. Materials min. In addition, the dry samples with mass about 5 mg were sealed in
the Tzero pans with lids and heated from − 70 to 180 ◦ C at a heating rate
Chitosan in the form of powder (90% DD, degree of deacetylation), of 5 ◦ C/min. All measurements were conducted under nitrogen flow of
hydrochloric acid, acetic acid, lactic acid and citric acid were all pur­ 50 mL/min. The experimental data were determined at least by tripli­
chased from Sinopharm Chemical Reagent Co., Ltd, China. The cate on each film sample.
viscosity-average molecular weight of chitosan is about 260,000 (Qiao,
Ma, Wang, & | Yao, 2019). All the reagents were used without further 2.7. Mechanical properties
purification. Ultrapure water (18 MΩ cm) was used to prepare various
acid solutions. The mechanical properties of chitosan films with different acids were
measured using an Universal testing machine (WDL-005, Jinan Xinshijin
2.2. Film preparation and conditioning Experimental Instrument Co., China). Test strip dimensions were 50 mm
× 10 mm (length × width). The tensile experiments were conducted at
Chitosan solutions (2%, w/v) were prepared by dissolving 0.6 g ambient temperature with a constant deformation rate of 20 mm/min.
chitosan powder in 30 mL of 0.15 mol/L hydrochloric, acetic, lactic or The measurements were run three times for each film and took the
citric acid aqueous solutions. Each film-forming solution was stirred at average as the results.

2
C. Qiao et al. LWT 135 (2021) 109984

2.8. Statistical analysis 1560 cm− 1. A plausible explanation for this phenomenon is that the
citric acid possessing two hydroxyl groups and three carboxyl groups
Statistical analyses were carried out using the SPSS software package can interact strongly with chitosan through hydrogen bonds and ionic
computer program (SPSS Statistical Software, Inc., Chicago, IL, USA). interactions. Additionally, a strong peak at 1705 cm− 1 appeared, indi­
Statistical significance was determined by analysis of variance cating the existence of free carboxyl groups in this film. Similar results
(ANOVA). The level of significance was set p < 0.05. All tests were were obtained by Tanigawa et al.(2008) working on chitosan films with
performed in triplicate. citric acid.
Contrary to carboxylic acids, the introduction of hydrochloric acid
3. Results and discussion will cause the amide II of chitosan shift to a lower wavenumber of 1519
cm− 1. It can be explained by the fact that hydrochloric acid is a strong
3.1. Interactions between chitosan and acids acid and chitosan is fully charged in film whereas it is only partially
charged in all the other films. Moreover, a sharp peak turned up at 1620
In order to study the interaction between chitosan and acids, FTIR cm− 1 in the spectrum of hydrochloric acid film. Obviously, this peak was
experiments were performed on chitosan films cast from different acid not correlated to amide II of chitosan, as the biopolymer had high degree
solutions. Fig. 1 displays FTIR spectra of the films with different acids. of deacetylation. It has been demonstrated that the moisture content of
For all the samples, the broad absorption band between 3600 and 3000 chitosan film with hydrochloric acid is higher than that of films with
cm− 1 is assigned to the stretching of hydrogen bonded hydroxyl groups carboxylic acids (Demarger-Andre & Domard, 1994; Gartner et al.,
and amino groups of the carbohydrate ring. The peak at 1536 cm− 1 for 2011). Furthermore, once the sample was heat treated at 110 ◦ C for 60
the polymer film with acetic acid corresponds to amide II, suggesting the min, this peak was significantly weakened (Fig. S1). Thus this peak is
amino group in chitosan is -NH+ 3 instead of -NH2 (Kumar-Krishnan et al., very likely to be associated with the residual water that is tightly bound
2014). This result confirmed the presence of electrostatic interaction to the chitosan. A similar observation was also reported by Zawadzki
between chitosan and the acid, and this film is essentially in the acetate and Kaczmarek (2010), who attributed this peak to the water physically
form (Demarger-Andre & Domard, 1994). This peak shifted slightly to a adsorbed.
higher wavenumber of 1552 cm− 1 for film with lactic acid. A similar
observation was reported by Niamsa and Baimark (2009) that the amide 3.2. Crystalline structure of chitosan films with different acids
II of chitosan was shifted to higher frequency when the lactic acid was
used instead of acetic acid. The shift of this band indicates the content of To investigate the influence of acid type on the structural changes of
the -NH+ 3 groups is low in lactic acid film. It has been reported that the chitosan films, the wide X-ray diffraction patterns of the chitosan films
interactions between chitosan and the acid increase as the pKa of the with different acids are compared, as shown in Fig. 2. It is seen that
acid decreases (Shamov et al., 2002). As the pKa of lactic acid (~3.86) is acetic acid film displays three main diffraction peaks at around 2θ = 12◦ ,
lower than that of acetic acid (~4.8), the ionic interactions between 20◦ and 23◦ , respectively, which is associated with the hydrated
amine and carboxylic groups are stronger in lactic acid film (Velás­ conformation of chitosan (Kobaisi, Murugaraj, & Mainwaring, 2012). It
quez-Cock et al., 2014). Furthermore, the possible hydrogen bonds is known that the regenerated chitosan is more amorphous than raw
formed between the hydroxyl groups of lactic acid and the hydroxyl chitosan due to a partial reconstitution of crystal region of raw chitosan
groups of chitosan could contribute to the interactions in this system. during the regeneration process (Zhong et al., 2011). The pattern of
The enhancement of interactions between chitosan and acid may result hydrochloric acid film was similar to that of acetate sample, except that
in a decrease of the concentration of protonated amine in polymer film. the diffraction peaks at around 2θ = 12◦ and 20◦ became stronger for the
As a consequence, the amide II of chitosan moved to a high frequency. former. It has been stated that the XRD pattern of chitosan acetate is very
This phenomenon was more obvious in citric acid film, in which the similar to that of chitosan salts with some inorganic acids, such as HF,
amide II became broadened and shifted to a higher wavenumber of HCl, and H2SO4 (Yamamoto, Kawada, Yui, & Ogawa, 1997). These re­
sults indicate chitosan molecules take up a similar crystalline

Fig. 1. FTIR spectra of the chitosan films prepared from different acid solu­ Fig. 2. Wide angle X-ray diffraction patterns of the chitosan films prepared
tions. All the samples are conditioned at 60% RH. Acetic acid ( ); Lactic acid from different acid solutions. All the samples are conditioned at 60% RH. Acetic
( ); Citric acid( ); Hydrochloric acid ( ). acid ( ); Hydrochloric acid ( ); Lactic acid ( ); Citric acid ( ).

3
C. Qiao et al. LWT 135 (2021) 109984

conformation in these systems. In addition, the reflections are more the interaction between chitosan and acid, the lower water content in
intense for hydrochloric acid film, suggesting a better organized the films. For the film with hydrochloric acid, it failed to accurately
long-range structure in this sample than in acetic acid film. However, an determine the water content from the weight loss at 120 ◦ C, as the loss of
opposite result was reported by Gartner et al. (2011), who investigated moisture was very limited around this temperature as a result of the tight
the role of the counterion (acetate vs Cl− ) on the structure of chitosan binding of water molecules to the chitosan and its Cl− counterion
films and found that the large acetate counterion with a delocalized (Gartner et al., 2011). The notable weight loss from 190 to 400 ◦ C was
charge plasticized the films enabling a good long range molecular or­ related to a complex process involving the decomposition of the acety­
ganization. It should be noted that the concentration of the two dilute lated as well as deacetylated units of chitosan (Khouri et al., 2020). In
acids used in their work was different ([HAc]~0.17M, [HCl]~0.12M). It addition, this weight loss may partly correspond to the evaporation of
was found that an increase in acetic acid concentration could enhance residual acids in films, which has been suggested from coupled
the crystallinity of chitosan film (Abdolrahimi, Seifi, & Zadeh, 2018). TGA-mass spectrometry measurements (Quijada-Garrido, Igle­
Thus an opposite result may be obtained when the two dilute acids are sias-González, Mazón-Arechederra, & Barrales-Rienda, 2007).
used at the same concentration. The DTG curves of the samples in Fig. 3 are shown in Fig. S2. It is
For lactic acid film, only one broad scattering peak occurred at 2θ = seen that there are two main peaks for hydrochloric acid, acetic acid and
21◦ , indicating that the molecules in this film were in amorphous state. lactic acid films. One peak is around 100 ◦ C, corresponding to the water
Similar results have been reported by other authors (Zhong et al., 2011). evaporation, and another peak is related to the decomposition of chi­
The strong interactions between chitosan and lactic acid may hinder the tosan, which is around 200, 280 and 280 ◦ C, respectively. For citric acid
movements of the chain segments, and hence restrains the crystalliza­ film, two overlapping peaks with centers at about 190 and 210 ◦ C may
tion process. Similarly, no crystalline diffraction peak was found in the be due to the evaporation of citric acid derivatives (Khouri et al., 2020),
diffractogram of citric acid film due to the fact that citric acid could and the third peak with center at about 350 ◦ C is associated to the
interact strongly with chitosan. This result was accordance with previ­ decomposition of chitosan. These observations suggest that the thermal
ous work by Tanigawa, Miyoshi, and Sakurai (2008). stability of chitosan films is in the order of citric acid > lactic acid ~
acetic acid > hydrochloric acid.
3.3. Thermogravimetric analysis (TGA)
3.4. Differential scanning calorimetry (DSC)
Fig. 3 displays the TGA curves of chitosan films cast from solutions of
different acids. All the samples show a similar thermal degradation The glass transition temperature (Tg) is an important parameter of
behavior, which suffer an initial small weight loss at around 120 ◦ C, and biopolymers because their physical properties will undergo significant
then a sharp weight loss from 190 to 400 ◦ C. The first weight loss at changes at the glass transition. In this study the influence of acid type on
approximately 120 ◦ C is associated with the removal of absorbed the Tg of chitosan films was investigated. The DSC thermograms of the
moisture in the films (Pavoni et al., 2019). For the three films with chitosan films with different acids are shown in Fig. 4, and the results are
carboxylic acids, the weight loss follows the order acetic acid > lactic listed in Table 1. It is seen that the glass transition temperature of acetic
acid > citric acid. This result indicates that the moisture content of acid film was about 54 ◦ C. This result was almost completely consistent
samples conforms to the same orders. It is generally accepted that the with Tg values of about 55 ◦ C previously reported for chitosan films
water content of the films is related to the degree of protonation of the prepared from solutions of acetic acid (Hosseini, Rezaei, Zandi, & Ghavi,
chitosan, due to the hydrophilicity of the NH+ 3 groups (Gartner et al., 2013; Kumar-Krishnan et al., 2014). It is worthy to note that the glass
2011). From the FTIR results (Fig. 1), it was found that the interactions transition of chitosan is closely related to the water content of polymer
between chitosan and acids followed citric acid > lactic acid > acetic films. The moisture content of our sample should be comparable to that
acid. Furthermore, these strong interactions may result in a reduced of their samples because all films were similarly conditioned at relative
charge density of the chitosan chains. This assumption was supported by humidities of 50–60%. In addition, both hydrochloric and lactic acid
Amorim et al. (2016) who demonstrated that citrate anions could be films exhibit a similar Tg to that of acetic film. In these hydrated
tightly retained in the electrical double layer of chitosan particles,
partially screening the biopolymer positive charges. Thus, the stronger

Fig. 3. TGA curve of the chitosan films prepared from different acid solutions. Fig. 4. DSC thermograms of the chitosan films prepared from different acid
All the samples are conditioned at 60% RH. Acetic acid ( ); Hydrochloric solutions. All the samples are conditioned at 60% RH. Acetic acid ( ); Lactic
acid ( ); Lactic acid ( ); Citric acid ( ). acid ( ); Citric acid ( ); Hydrochloric acid ( ).

4
C. Qiao et al. LWT 135 (2021) 109984

Table 1 highest tensile strength and lowest elongation at the break, a typical
Effect of solvent system on the glass transition temperature (Tg) and the melting pattern of brittle materials. Similar results were observed for the chi­
temperature (Tm) of chitosan films at wet and dry states (1: wet state; 2: dry tosan film made with acetic acid (Suyatma, Tighzert, Copinet, & Coma,
state; N: not found). 2005). Compared to acetic acid film, the lactic acid film has a low tensile
Solvents Tg1(∘C) Tm1(∘C) Tg2(∘C) Tm2(∘C) strength and high elongation at the break. This result was in accord with
HA 56 ±3 104 ± 3 125 ± 5 N previous work (Velásquez-Cock et al., 2014). In addition, the citric acid
AA 54 ±2 103 ± 2 123 ± 4 N film has the lowest tensile strength and highest elongation at break. As
LA 53 ±2 N 124 ± 4 N the Tg of citric acid film is about 18 ◦ C (Table 1), it exhibits a typical
CA 18 ±1 N 28 ± 2 N rubberlike behavior at room temperature. This rubberlike behavior was
HA: Hydrochloric acid, AA: Acetic acid, LA: Lactic acid, CA: Citric acid. frequently observed in chitosan-citric acid systems (Park, Marsh, &
Values are mean ± SD (n = 3). Rhim, 2002; Bégin, &; Tanigawa et al., 2008; Bégin and Van Calsteren,
1999).
chitosan salt films, water molecules are known to act as plasticizers, and In the present work, the tensile strength of these films was found to
the Tg of chitosan films decreases with an increase in water content. In be in the order of acetic acid > lactic acid > citric acid, which decreased
addition to moisture, the counterions such as lactate and Cl− also with the molecular weight of the acids. A similar correlation was found
contribute to the plasticization of chitosan due to their interactions with between Young’s modulus and volume of acid for chitosan films (Bégin
polymer (Gartner et al., 2011). The interactions of these counterions & Van Calsteren, 1999). It could be explained in term of the interactions
with chitosan were in the order of lactate > acetate > Cl− , whereas the between chitosan and acids. Compared to acetate and lactate ions, the
water content followed a reverse order in these films. Thus the citrate ions could interact more strongly with chitosan, and the inter­
compromise involving the plasticizing effects of water and counterions chain and intrachain hydrogen bonds in chitosan were destroyed,
led to a similar Tg for the three samples. For citric acid film, the glass resulting in a lower mechanical strength. In the case of lactic acid and
transition occurred at a low temperature of about 18 ◦ C. A much lower acetic films, although their Tg are similar, their aggregation structures
Tg with values of − 22.5 ◦ C was observed by Tanigawa et al. (2008). It are different. The former is amorphous, whereas the latter is partially
should be noted that the concentration of citric acid solution used in crystalline and thus has a higher mechanical strength. These observa­
their experiments was much higher than that of the acid solution used in tions indicate that the mechanical properties of the chitosan films are
this work. These results suggest that the citrate ions play a major role in quite different depending on the counter ions.
the plasticization of polymer film due to the strong interactions of chi­
tosan with citrate ions. 4. Conclusions
In the DSC thermogram of acetic acid film, an endothermic peak
appeared at about 103 ◦ C. This thermal event has been demonstrated to Interactions between chitosan and acid ions depend on the nature of
be associated with the melting transition of chitosan (Qiao et al., 2019). acids. In hydrochloric and acetic acid films, the ionic interactions occur
A very similar result was reported by Liu, Zhou, Zhang, Yu, and Cao and the chitosan is partially crystalline. While in lactic and citric acid
(2014) who stated that the Tm of chitosan film with acetic acid was 105 films, the ionic interactions as well as hydrogen bonding exist and the

C. It is worthy to note that the determination of the thermal transitions biopolymer is amorphous. The hydrochloric, acetic and lactic films
in hydrated chitosan films is affected by the integrity of the seal of the exhibit similar Tg values at around 54 ◦ C, whereas a lower glass tran­
DSC pans. The endothermic peak appeared at around 100 ◦ C in her­ sition occurs at 18 ◦ C in citric acid film. Furthermore, a melting transi­
metically sealed pans is attributed to the melting transition of chitosan, tion of chitosan is observed at about 103 ◦ C in hydrochloric and acetic
whereas it is assigned to moisture evaporation in open systems (Qiao acid film. However, this thermal transition is absent when the moisture
et al., 2019). Additionally, hydrochloric film displayed a similar Tm to is removed from the sample films. The stress-strain curve of acetic acid
that of acetic acid film. These observations indicated that chitosan was film shows a typical pattern of brittle materials, while the citric acid film
partially crystalline in both polymer films, which was confirmed by the shows characteristic features of rubber-like materials. In addition, the
presence of crystalline diffraction peaks in their diffractograms. On the
contrary, the melting behaviors were not observed for lactic and citric
acid films, suggesting that chitosan molecules were in the amorphous
state. This was in accordance with the X-ray spectra shown in Fig. 3,
which displayed completely amorphous patterns.
As the moisture is removed from the polymer films, the Tg will in­
crease accordingly (Fig. S3). Interestingly, hydrochloric, acetic and
lactic acid films also display a similar Tg with values of about 123 ◦ C,
much higher than that of the hydrated samples. However, there is a little
increase of Tg for citric acid film, which varied from 18 to 28 ◦ C. These
results indicate that only citrate ions act as plasticizers in the dry chi­
tosan salt films. In addition, the melting transition was not observed in
these films. In the absence of moisture, chitosan usually undergoes
decomposition prior to melting due to its rigid-rod polymer backbone
with strong inter/-intrachain hydrogen bonds (Murray & Dutcher,
2006).

3.5. Mechanical properties

The influence of acid type on the mechanical properties of chitosan


films was investigated. The stress-strain diagram of hydrochloric acid
film was not shown since it could not be tested due to its extreme brit­ Fig. 5. The stress-strain curves of the chitosan films prepared from different
tleness. Fig. 5 exhibits the stress-strain curves of the chitosan films with acid solutions. All the samples are conditioned at 60% RH. Acetic acid ( );
different acids at room temperature. The acetic acid film presents Lactic acid ( ); Citric acid ( ).

5
C. Qiao et al. LWT 135 (2021) 109984

lactic acid film exhibits an intermediate behavior. These results suggest Khouri, J., Penlidis, A., & Moresoli, C. (2020). Heterogeneous method of chitosan film
preparation: Effect of multifunctional acid on film properties. Journal of Appllied
that it may be possible to improve the performances of chitosan films by
Polymer Science, 137, 48648.
the choice of a proper solvent. Kobaisi, M. A., Murugaraj, P., & Mainwaring, D. E. (2012). Origin and influence of water
induced chain relaxation phenomena in chitosan biopolymers. Journal of Polymer
CRediT authorship contribution statement Science Part B: Polymer Physics, 50, 403–414.
Kumar-Krishnan, S., Prokhorov, E., Ramírez, M., Hernandez-Landaverde, M. A., Zarate-
Triviño, D. G., & Kovalenko, Y. (2014). Novel gigahertz frequency dielectric
Congde Qiao: Conceptualization, Supervision, Data curation, relaxations in chitosan films. Soft Matter, 10, 8673–8684.
Formal analysis, Validation, Visualization, Writing - original draft, Lakehal, I., Montembault, A., David, L., Perrier, A., Vibert, R., Duclaux, L., et al. (2019).
Prilling and characterization of hydrogels and derived porous spheres from chitosan
Writing - review & editing. Xianguang Ma: Methodology, Investigation, solutions with various organic acids. International Journal of Biological
Formal analysis, Writing - review & editing. Xujie Wang: Investigation, Macromolecules, 129, 68–77.
Writing - review & editing. Libin Liu: Formal analysis, Writing - review Liu, X. L., Xu, Y. X., Zhan, X. B., Xie, W. C., Yang, X. H., Cui, S. W., et al. (2020).
Development and properties of new kojic acid and chitosan composite biodegradable
& editing. films for active packaging materials. International Journal of Biological
Macromolecules, 144, 483–490.
Declaration of competing interests Liu, M., Zhou, Y., Zhang, Y., Yu, C., & Cao, S. (2014). Physicochemical, mechanical and
thermal properties of chitosan films with and without sorbitol. International Journal
of Biological Macromolecules, 70, 340–346.
The authors declare that they have no known competing financial Murray, C. A., & Dutcher, J. R. (2006). Effect of changes in relative humidity and
interests or personal relationships that could have appeared to influence temperature on ultrathin chitosan films. Biomacromolecules, 7, 3460–3465.
Narasagoudr, S. S., Hegde, V. G., Chougale, R. B., Masti, S. P., & Dixit, S. (2020).
the work reported in this paper.
Influence of boswellic acid on multifunctional properties of chitosan/poly (vinyl
alcohol) films for active food packaging. International Journal of Biological
Acknowledgement Macromolecules, 154, 48–61.
Niamsa, N., & Baimark, Y. (2009). Preparation and characterization of highly flexible
chitosan films for use as food packaging. American Journal of Food Technology, 4,
The work was supported by the Foundation (No. ZZ20190209) of 162–169.
State Key Laboratory of Biobased Material and Green Papermaking, Qilu Park, J. W., Choi, K. H., & Park, K. P. (1983). Acid-base equilibria and related properties
University of Technology, Shandong Academy of Sciences. of chitosan. Bulletin of the Korean Chemical Society, 4, 68–72.
Park, S. Y., Marsh, K. S., & Rhim, J. W. (2002). Characteristics of different molecular
weight chitosan films affected by the type of organic solvents. Journal of Food Scince
Appendix A. Supplementary data E: Food Engineering and Physical Properties, 67, 194–197.
Pavoni, J. M. F., Luchese, C. L., & Tessaro, I. C. (2019). Impact of acid type for chitosan
dissolution on the characteristics and biodegradability of cornstarch/chitosan based
Supplementary data to this article can be found online at https://doi. films. International Journal of Biological Macromolecules, 138, 693–703.
org/10.1016/j.lwt.2020.109984. Pillai, C. K. S., Paul, W., & Sharma, C. P. (2009). Chitin and chitosan polymers:
Chemistry, solubility and fiber formation. Progress in Polymer Science, 34, 641–678.
Popa-Nita, S., Alcouffe, P., Rochas, C., David, L., & Domard, A. (2010). Continuum of
References structural organization from chitosan solutions to derived physical forms.
Biomacromolecules, 11, 6–12.
Abdolrahimi, M., Seifi, M., & Zadeh, M. H. R. (2018). Study the effect of acetic acid on Qiao, C. D., Ma, X. G., Wang, X. J., & Yao, J. S. (2019). Effect of water on the thermal
structural, optical and mechanical properties of PVA/chitosan/MWCNT films. transition in chitosan films. Polymer Crystallization, 2, 10092.
Chinese Journal of Physics, 56, 221–230. Quijada-Garrido, I., Iglesias-González, V., Mazón-Arechederra, J. M., & Barrales-
Amorim, M. L., Dias Ferreira, G. M., Soares, L. D. S., et al. (2016). Physicochemical Rienda, J. M. (2007). The role played by the interactions of small molecules with
aspects of chitosan dispersibility in acidic aqueous media: Effects of the food acid chitosan and their transition temperatures. Glass-forming liquids: 1,2,3-Propantriol
counter-anion. Food Biophysics, 11, 388–399. (glycerol). Carbohydrate Polymers, 68, 173–186.
Bégin, A., & Van Calsteren, M. R. (1999). Antimicrobial films produced from chitosan. Rinaudo, M. (2006). Chitin and chitosan: Properties and applications. Progress in Polymer
International Journal of Biological Macromolecules, 26, 63–67. Science, 31, 603–632.
Chen, R. H., Chen, W. Y., Wang, S. T., Hsu, C. H., & Tsai, M. L. (2009). Changes in the Shamov, M. V., Bratskaya, S. Y., & Avramenko, V. A. (2002). Interaction of carboxylic
mark-Houwink hydrodynamic volume of chitosan molecules in solutions of different acids with chitosan: Effect of pK and hydrocarbon chain length. Journal of Colloid and
organic acids, at different temperatures and ionic strengths. Carbohydrate Polymers, Interface Science, 249, 316–321.
78, 902–907. Soares, L. S., Perim, R. B., Alvarenga, E. S., Guimarães, L. M., Teixeira, A. V. N. C.,
Chen, P. H., Hwang, Y. H., Kuo, T. Y., Liu, F. H., Lai, J. Y., & Hsieh, H. J. (2007). Coimbra, J. S. R., et al. (2019). Insights on physicochemical aspects of chitosan
Improvement in the properties of chitosan membranes using natural organic acid dispersion in aqueous solutions of acetic, glycolic, propionic or lactic acid.
solutions as solvents for chitosan dissolution. Journal of Medical and Biological International Journal of Biological Macromolecules, 128, 140–148.
Engineering, 27, 23–28. Suyatma, N. E., Tighzert, L., Copinet, A., & Coma, V. (2005). Effects of hydrophilic
Chen, R. H., Lin, J. H., & Yang, M. H. (1994). Relationships between the chain plasticizers on mechanical, thermal, and surface properties of chitosan films. Journal
flexibilities of chitosan molecules and the physical properties of their casted films. of Agricultural and Food Chemistry, 53, 3950–3957.
Carbohydrate Polymers, 24, 41-16. Tanigawa, J., Miyoshi, N., & Sakurai, K. (2008). Characterization of chitosan/citrate and
Chenni, A., Djidjelli, H., Boukerrou, A., Grohens, Y., & Saulnier, B. (2018). chitosan/acetate films and applications for wound healing. Journal of Applied
Thermomechanical and acidic treatments to improve plasticization and properties of Polymer Science, 110, 608–615.
chitosan films: A comparative study of acid types and glycerol effects. Materials Tian, Q., Liu, S., Sun, X., Sun, H., Xue, Z., & Mu, T. (2015). Theoretical studies on the
Testing, 60, 93–101. dissolution of chitosan in 1-butyl-3-methylimidazolium acetate ionic liquid.
Demarger-Andre, S., & Domard, A. (1994). Chitosan carboxylic acid salts in solution and Carbohydrate Research, 408, 107–113.
in the solid state. Carbohydrate Polymers, 23, 211–219. Velásquez-Cock, J., Ramírez, E., Betancourt, S., Putaux, J. L., Osorio, M., Castro, C., et al.
Fadzallah, I. A., Majid, S. R., Careem, M. A., & Arof, A. K. (2014). A study on ionic (2014). Influence of the acid type in the production of chitosan films reinforced with
interactions in chitosan–oxalic acid polymer electrolyte membranes. Journal of bacterial nanocellulose. International Journal of Biological Macromolecules, 69,
Membrane Science, 463, 65–72. 208–213.
Gartner, C., Lopez, B. L., Sierra, L., Graf, R., Spiess, H. W., & Gaborieau, M. (2011). Williams, P. A. (2011). Renewable resources for functional polymers and biomaterials:
Interplay between structure and dynamics in chitosan films investigated with solid Polysaccharides, proteins and polyesters. RSC Publishing.
state NMR, dynamic mechanical analysis, and X-ray diffraction. Biomacromolecules, Yamamoto, A., Kawada, J., Yui, T., & Ogawa, K. (1997). Conformational behavior of
12, 1380–1386. chitosan in the acetate salt: An X-ray study. Bioscience Biotechnology And
Hamdine, M., Heuzey, M. C., & Begin, A. (2006). Viscoelastic properties of phosphoric Biochemistry, 61, 1230–1232.
and oxalic acid-based chitosan hydrogels. Rheologica Acta, 45, 659–675. Zawadzki, J., & Kaczmarek, H. (2010). Thermal treatment of chitosan in various
Hosseini, S. F., Rezaei, M., Zandi, M., & Ghavi, F. F. (2013). Preparation and functional conditions. Carbohydrate Polymers, 80, 394–400.
properties of fish gelatin-chitosan blend edible films. Food Chemistry, 136, Zhong, Y., Song, X. Y., & Li, Y. F. (2011). Antimicrobial, physical and mechanical
1490–1495. properties of kudzu starch–chitosan composite films as a function of acid solvent
Hu, F., Sun, T., Xie, J., Xue, B., Li, X. H., Gan, J. H., et al. (2020). Functional properties types. Carbohydrate Polymers, 84, 335–342.
and preservative effect on Penaeus vannamei of chitosan films with conjugated or
incorporated chlorogenic acid. International Journal of Biological Macromolecules,
159, 333–340.

You might also like