You are on page 1of 16

Accepted Manuscript

Title: Liquefied Chitin/Polyvinyl Alcohol Based Blend


Membranes: Preparation and Characterization and
Antibacterial Activity

Authors: Jie Zhang, Wen-Rong Xu, Yucang Zhang, Wei Li,


Jiadan Hu, Fengyi Zheng, Yangtian Wu

PII: S0144-8617(17)31160-8
DOI: https://doi.org/10.1016/j.carbpol.2017.10.014
Reference: CARP 12863

To appear in:

Received date: 8-7-2017


Revised date: 11-9-2017
Accepted date: 3-10-2017

Please cite this article as: Zhang, Jie., Xu, Wen-Rong., Zhang, Yucang., Li,
Wei., Hu, Jiadan., Zheng, Fengyi., & Wu, Yangtian., Liquefied Chitin/Polyvinyl
Alcohol Based Blend Membranes: Preparation and Characterization and Antibacterial
Activity.Carbohydrate Polymers https://doi.org/10.1016/j.carbpol.2017.10.014

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

Jie Zhang, Wen-Rong Xu*a, Yucang Zhang*b, Wei Li, Jiadan Hu, Fengyi Zheng, Yangtian Wu
Key Laboratory of Advanced Materials of Tropical Island Resources of Ministry of Education,
College of Materials and Chemical Engineering, Hainan University, Haikou 570228, China.

*a corresponding author, E-mail: xuwr2016@hainu.edu.cn (W.-R. Xu)

*b corresponding author, E-mail: yczhang@hainu.edu.cn (Y. Zhang)

Highlights

 Liquefied chitin products were first used to modify the polymer materials

 The influences of LBMC on characteristics and performances of the LBMC/PVA

blend membranes were evaluated.

 Mechanical property and thermal stability were remarkably strengthened under

optimal blend membrane.

 The incorporation of LBMC obviously enhanced the antibacterial activity of

blend membranes.

Abstract: To expand the applications of fishing industrial wastes, the liquefaction technique was
employed to convert chitin into liquids, which were first further used for the modification of polymer
materials. Ball-mill treated chitin was effectively liquefied into polyols in polyethylene glycol
400/glycerin mixed solvent. FTIR, 1H NMR and size exclusion chromatography analyses of
liquefied chitin turned out that depolymerization and deacetylation reactions occurred during
liquefaction process. The liquefied chitin/polyvinyl alcohol blend membranes with various mixing
ratios were prepared and characterized by FTIR and SEM. In addition, transparency, tensile strength,
elongation at break, water absorption, water retention and antibacterial properties were thoroughly
discussed. The mechanical property and thermal stability were greatly enhanced under the optimized

1
conditions of 25 wt% liquefied chitin incorporation to polyvinyl alcohol. Moreover, the blend
membranes showed improvements in water absorption and water retention capacity. In particular, the
antibacterial activity was significantly improved after incorporation of liquefied chitin.
Keywords: chitin; liquefaction; polyvinyl alcohol; blend membrane; antibacterial activity

Chemical compounds studied in this article


Polyethylene glycol (PubChem CID: 174); Glycerin (PubChem CID: 753); Polyvinyl Alcohol
(PubChem CID: 11199); Ethanol (PubChem CID: 702); Sulfuric acid (PubChem CID: 1118); KBr
(PubChem CID: 253877); LiCl (PubChem CID: 433294) ; DMAc (PubChem CID: 31374)

1. Introduction
During the past decade, increasing scientific efforts have been dedicated to biomass conversion
due to the slow but continued fossil fuel depletion around the world (Xu, Xie, Wang & Jiang, 2016;
Yamada & Ono, 2001). As one of the simplest and most effective biomass conversion strategies,
liquefaction is widely practiced to convert solid biomass into fuels, chemicals and raw materials for
polymer synthesis and modification (Karger, 2011). It is usually conducted by using polyhydric
alcohols as the solvent and acids as the catalyst under atmospheric pressure to break the chemical
backbones and functionalize the resulting fragments. Lignocellulose, consisting of cellulose,
hemicellulose and lignin, plays an important role in biomass because it is the most abundant
renewable resource in nature. The researches on the utilization of lignocellulose through liquefaction
process have been well established. Especially in recent years, lignocellulose has been converted
into liquid polyols (Jasiukaitytė, Kunaver & Strlič, 2009) and further applied in the preparation of
various polymer materials, such as polyurethane foams (Chen & Lu, 2009; Hu, Wan & Li, 2012; Luo,
Hu, Zhang & Li, 2013), epoxy resin (Kuo, Sain & Yan, 2014; Zhang, Zhang & Zhao, 2012)
andadhesive (dos Santos et al., 2016; Lee, Lee, Hsu & Hsieh, 2014). Indeed, our group has
previously reported the preparation of blend membranes by incorporating liquefied banana
pseudo-stem into polyvinyl acetate (PVAc). The addition of liquefied products significantly
enhanced the tensile strength and elongation at break of membranes (Li, Zhang & Lao, 2015).
Further addition of nano-silica sol to the polymer membranes resulted in a visible increase in tensile
strength and break elongation by 425% and 250%, respectively (Fan-rong, Yu-cang, Ji-hui,
Wen-xing & Jie, 2016).
In recent years, Yan et al. began to employ liquefaction method in the conversion of animal
cellulose, chitininto small molecular weight chemicals. Chitin is the second most abundant natural
biopolymer after cellulose. The chemical structure of chitin is comparable to that of cellulose, and
the 2-hydroxy group on each sugar unit of cellulose is replaced with acetamido to give chitin, i.e.,
β-(1→4)-linked N-acetylglucosamine polymer (Chen, Liu, Kerton & Yan, 2015; Cho, Jang, Park &
Ko, 2000; Yan & Chen, 2015). Chitin is found in many places throughout the natural world,
especially in the shells of crustaceans such as crabs, lobsters and shrimps. So far, researches on the
isolation of chitin from fishery wastes have been very mature (Bahasan, Satheesh & Baakdah, 2016;
Benhabiles et al., 2013; Hongkulsup, Khutoryanskiy & Niranjan, 2016). However, most of the shells Commented [A1]: Corresponding to question 4 of
are now regarded as wastes in the fishing industry, and their further disposal consumes more reviewer#3
financial and human resources. In fact, chitin and its deacetylated derivative chitosan areversatile
materials that can be used in many fields such as pharmaceuticals, cosmetics, textiles , water
treatment and coatings (Davila, 2009; Morganti et al., 2008; Revoredo et al., 2006; Wawrzkiewicz,

2
Bartczak & Jesionowski, 2017; Benhabiles, Drouiche, Lounici, Pauss & Mameri, 2013). Making full Commented [A2]: Corresponding to question 4 of
use of discarded shells can turn waste into treasure. As reported by Yan et al., liquefaction of chitin reviewer#3
in acid-catalyzed conditions in small molecular solvents such as ethylene glycol and formic acid can
convert chitin into low molecular weight chemicals, including HADP, HAADP, acetic acid, pyrrole,
etc (Pierson, Chen, Bobbink, Zhang & Yan, 2014; Zhang & Yan, 2016). The researches not only
provide a new way for the application of liquefaction technology, but also open up new ways for the
high value-added use of fishery industrial wastes. However, the employments of chitin liquefaction
in the preparation and modification of polymer materials still remain empty.
Polyvinyl alcohol (PVA) is a biocompatible, non-toxic and chemical resistant polymer that is
extensively used in blend membranes (Ifuku et al., 2013), hydrogels (Wan, Campbell, Zhang, Hui &
Boughner, 2010) and food packaging (Liu, Han, Zhao, Wei, Zhao & Yuan, 2017). However, the
limited mechanical properties and thermal stability of PVA membrane obviously restrict its
heavy-duty applications, especially in high-temperature environments. These properties are usually
modified by many physical and chemical methods such as blending (Svang-Ariyaskul et al., 2006),
crosslinking (Ahmad, Ooi & BS, 2005) and grafting (Wongthep, Ruksakulpiwat & Amnuaypanich,
2010). Among these methods, the blending process is turned out to be an effective approach by
forming strong intermolecular interactions between the polymers (Deng, Li, Yang & Li, 2014). For
instance, Sharma et al. reported a series of blend membranes of chitosan/PVA with varying weight
ratios, which showed improved physicochemical properties (Sharma et al., 2016). Interestingly, the
liquefied chitin based PVA blend membranes have not been reported yet. This is mainly due to the
lack of research on chitin liquefaction.
In the present work, we propose to study the liquefaction behavior of chitin and blend the
liquefied products with PVA to modify the mechanical and thermal properties, as well as some other
potential properties. According to report, a mixed solvent including polyhydric alcohols and low
equivalent of small molecular alcohol is conducive to improving the liquefaction yield (D’Souza,
Song, Camargo & Ning, 2016). In addition, the use of ball-mill treated chitin (BM-chitin) as a
starting material is also advantageous for liquefaction production (Chen, Gao, Wang, Chen & Yan,
2015). Therefore, polyethylene glycol 400 (PEG 400) and glycerin (4:1, w/w) were selected as the
solvents, and BM-chitin was used as raw material to liquefy chitin at the presence of sulfuric acid at
160 ºC under atmosphere pressure. The resulting liquefied products of ball-mill treated chitin
(LBMC) were fully characterized by FTIR and 1H NMR spectroscopy and size exclusion
chromatography (SEC). The hydroxyl value of LBMC was measured by Hassan method (Hassan &
Shukry, 2008). Moreover, LBMC based PVA blend membranes with different weight ratios were
prepared and their physical and chemical properties were further characterized. Furthermore, the
antibacterial activities of blend membranes against MRSA were also evaluated by detecting the
inhibition zones of cell growth and proliferation.
2. Material and methods
2.1 Materials
Chitin (white powder) was purchased from Zhejiang Gold-shell Biochemical Co., Ltd.
Polyethylene glycol 400 (PEG 400) and glycerin were purchased from Xilong Chemical Co., Ltd.
PVA (1750 ± 50, 99% hydrolyzed) was purchased from Sinopharm Chemical Reagent Co., Ltd. All
commercially available chemicals and reagents were used as received without further purification.
2.2 Ball-mill pretreatment of chitin
Ball-mill treated chitin (BM-chitin) was obtained as follows: chitin was ground with a planetary

3
ball mill (Fritsch, Germany) at 350 rpm using zirconia balls (diameter 3 mm, 0.5 kilogram) in a
zirconia pot (500 ml volume). The grinding process was made up of 4 circles, each of which
included 20 min milling time and 40 min resting time, and the rotation direction was reversed after
each circle (Zhang & Yan, 2016).
2.3 Liquefaction of BM-chitin
In a three-necked glass flask equipped with a stirrer, condenser, and thermometer, chemicals
were added in the order of mixed solvent of PEG 400 and glycerin (28 g, 4:1, w/w), sulfuric acid (2
g) and BM-chitin (4 g). The mixture was stirred at 160 ºC for 90 min and then cooled down to room
temperature and diluted with ethanol. Subsequently, the suspension was filtered through a Buchner
funnel to separate the residues and filtrate. Ethanol in filtrate was removed over a rotary evaporator
to afford the liquid fractions, which denoted as LBMC and used for -subsequent study. Liquefaction
yield was calculated by the following eq. (1).

Starting chitin (mg)−Residual solid (mg)


Liquefaction yield (%) = × 100 % (1)
Starting chitin (mg)

2.4 Preparation of PVA/LBMC blend membranes


Blend membranes of LBMC and PVA were prepared by solution casting method. PVA was
stirred in deionized water at 90 ºC for 2 h to obtain 4 wt% of homogeneous solution. The blend
membranes were prepared with different weight percentage of LBMC to PVA (0, 12.5, 25, 50 and
100 wt%). The solution was mixed uniformly with vigorous stirring at 65 ºC for 3 h and then poured
onto a clean glass plate, subsequently natural dried at room temperature for 48 h. The dried blend
membranes were peeled off carefully from the glass plate and further measured as approximately 0.1
mm thick. The blend membranes with 0, 12.5, 25, 50 and 100 wt% of LBMC to PVA were Commented [A3]: Corresponding to question 1 of
designated as PVA, 8PVA-LBMC, 4PVA-LBMC, 2PVA-LBMC and 1PVA-LBMC, respectively. reviewer#1
2.5 Structure and morphology characterization
The structure of LBMC was analyzed by 1H NMR and FTIR spectroscopy combined with
hydroxyl number and average molecular weight (Mw) measurement. 1H NMR (400 MHz) spectra
was conducted on a 400 MHz spectrometer (Bruker). Chemical shift values were given in ppm.
Residual solvent signals in the 1H spectra were used as the internal reference (D2O: δH = 4.79). The
hydroxyl number of LBMC was determined according to ISO 14900-2001. SEC was performed on
Styragel HT-2THF and Styragel HT-4THF columns using a Waters 1515 chromatographic apparatus
equipped with a refractive index detector (Waters 2414) at 45 ºC to determine the Mw of chitin and
LBMC. A N, N-dimethylacetamide (DMAc) solution containing 0.5 wt% of LiCl was used as eluent
and the flow rate was 1.0 ml/min. FTIR spectra were recorded on a Fourier transform infrared
spectrometer (TENSOR27, Bruker Optics; Germany). The test specimens were prepared using the
regular KBr-disk method. All spectra were recorded over 4000–650 cm–1 at a resolution of 4 cm–1
with 32 scans. The deacetylation degree (DD) of BM-chitin was calculated by FTIR spectra (Baxter,
Dillon, Taylor & Roberts, 1992; Bharty, 2015). following eq. (2), where Aі is the area of the peak at
wavenumber і.
DD (%) = (100 – A1655/A3450 × 115) % (2)
The functional groups in blend membranes were characterized by FTIR and surface
morphology was investigated under scanning electron microscopy (SEM). SEM images were taken
with a Hitachi S3500N microscope operating at 10 kV. The membranes were coated with gold prior
to the SEM observation. The regular light transmittances of the blend membranes were measured by
a UV-vis spectrophotometer (LAMBDA 750s) in the wavelength range of 400–800 nm.
4
2.6 Mechanical property
The tensile strength and elongation at break of the membranes were measured using a universal
testing machine (WDW-1, JiNan; China) at a crosshead speed of 50 mm/min. The membranes were
cut into dumbbell, which shapes with length 75 mm and width 4 mm. Tensile strength and
elongation at break were obtained from the stress–strain curves. A batch of five specimens was tested
for each sample and the average values and standard deviations were calculated.
2.7 Thermogravimetric analysis
The thermal decomposition behavior of the membranes was performed using a
thermogravimetric analysis (TGA) devise (NETZSCH5 209F3) in a temperature range from 40 ºC to
800 ºC at a heating rate of 20 ºC/min under nitrogen flow.
2.8 Immersion test
The pre-weighted membranes (W) with 20 × 20 mm2 excised pieces were immersed in
deionized water at 25 ºC for 24 hours. The membranes were taken out from the water and weighted
(Ws) after wiping off the surface water with filter paper. The membranes were then dried at 37 ºC for
24 h and weighed again (Wd). The average value of three measurements was accounted for each
sample, and the water absorption ratio and weight loss ratio were calculated by the formulas (eq. (3)
and (4)).
Water absorption ratio (%) = [(Ws–W)/W]×100 % (3)
Weight loss ratio (%) = [(W–Wd)/W]×100 % (4)
2.9 Antibacterial activity
Methicillin-resistant Staphylococcus aureus (MRSA) was used to examine the antibacterial
activity of the membranes by the disk inhibition zone assay according to Pelissari method (Pelissari,
Grossmann, Yamashita & Pineda, 2009). The blend membranes were placed on plates containing
solid agar culture, which had been previously spread with 0.1 mL of MRSA inoculums. After
incubation at 37 ºC for 24 h, the diameters of inhibition zones were measured using a vernier caliper.
All the tests were carried out in triplicate for each sample.
3. Results and discussion
3.1 Characterization of LBMC
As shown in Table 1, up to 81% of chitin (Mw=498316) was converted into liquid products
with sulfuric acid as catalyst at 160 ºC for 90 min under a PEG 400 and glycerin (4:1, w/w) mix
solvent system. The LBMC showed as a dark brown homogeneous solution with good mobility and
water solubility. The hydroxyl number and average molecular weight (Mw) of LBMC were
measured as 285 and 5145, respectively. The results indicated that chitin was converted into low
molecular weight polyols through liquefaction process.

To compare the chemical compositions of LBMC with BM-chitin, FTIR investigation was
employed. As shown in Fig. A1, the OH and NH stretching bands appeared around 3450 and 3270
cm–1, and the peak at 1655 cm–1 was assigned to Amide I band. The DD value of BM-chitin was
calculated as 15% for the baseline reference. The band at 1655 cm–1 was apparently weakened in the
spectra of LBMC, which suggested that some of the C–N bonds were broken. A small absorption
appeared at 1960 cm–1 could be ascribe to the –NH3+ group in LBMC due to the protonation of
primary amine in the existence of acid (Rao, Subha, Sairam, Mallikarjuna & Aminabhavi, 2010).
1
H NMR spectrum of LBMC showed two groups of well separated proton signals at δ ~ 4.27

5
and 2.19 ppm (Fig. A2), which could be attributed to H-1 resonances of GlcN residue and GlcNAc
residue, respectively (Kubota & Eguchi, 1997; Kubota, Tatsumoto, Sano & Toya, 2000). Therefore,
the deacetylation degree (DD’) of LBMC could be estimated from the ratio of the integral intensity
of H-1 from GlcN (IH1-N) to the total that of H-1 from GlcN and GlcNAc (IH1-N + IH1-NAc), following
the formula (eq. (5)), where IH1-N was 0.13 and IH1-NAc was 0.21. As a result, the DD’ value of LBMC
was calculated as 38%, which was higher than that of BM-chitin. The 1H NMR result gave a further
verification that deactylation reaction happened during liquefaction process. According to the
previously proposed mechanism of chitin liquefaction (Pierson, Chen, Bobbink, Zhang & Yan, 2014),
a reasonable pathway was proposed for BM-chitin liquefaction in PEG400/glycerol under acid
condition (Fig. A3). Commented [A4]: Corresponding to question 2 of
DD’ (%) = (IH1-N / (IH1-N + IH1-NAc)) ×100 % (5) reviewer#2
3.2 Characterization of blend membranes
3.2.1 FTIR analysis
FTIR spectra of pure PVA, PVA/LBMC membranes and pure LBMC were compared for their
signature peaks and interactions (Fig. 1a–f). FTIR spectra of pure PVA membrane (Fig. 1a) exhibited
the characteristic absorption bands at 3443 cm–1(OH), 2925 cm–1(CH2) and 1095 cm–1 (CO) (Deng,
Li, Yang & Li, 2014). The blend membranes displayed absorption bands of both PVA and LBMC.
The shifting and width variation of ~3440 cm–1 indicated a possible overlapped stretching between
the –NH2 (in LBMC) and the –OH (in PVA) groups and therefore, suggesting some physical and
chemical interactions between PVA and LBMC (Sharma et al., 2016).

3.2.2 Surface morphology of membranes


The selected membranes were used to examine the fracture surface morphology by SEM
investigation. As shown in the SEM images (Fig. 2), distinct changes appeared in the cross section
morphology of PVA membranes after blending with LBMC. The SEM images of pure PVA
membrane showed an extremely rough cross section with dense texture (Fig. 2a, b). For
4PVA-LBMC membrane, the network structures formed by crosslinking of micro-rods could be
observed (Fig. 2c, d). The formation of micro-rods and network structures might be connected with
the increasing bonding between LBMC and PVA (Wu & Hakkarainen, 2014). However, when the
LBMC content increased to 100 wt%, it could be observed that closely packed grains and apparently
holes displayed in the SEM images of 1PVA-LBMC membrane (Fig.2e, f). This morphology might
be caused by the self-assembly of superfluous LBMC (Song & Zheng, 2013).

3.2.3 Transparency of the composite membranes


The photographs of pure PVA and PVA/LBMC blend membranes are arranged in Fig.A4a–A4e.
The LBMC showed as a dark brown homogeneous solution (Fig. A4f). After the addition of LBMC, Commented [A5]: Corresponding to question 2 of reviewer
a light brown color was visible in the blend membranes. And the background could be observed 1
clearly through all membranes, indicating their highly transparency. In order to better understand the
transparency of membranes, regular light transmittance (Tr) was measured (Fig.A5). At a visible
wavelength of 600 nm, the Tr value of the 4PVA-LBMC membrane was 90%, only 9% lower than
that of the pure PVA film. Increasing amount of LBMC in PVA membrane leaded to the decreasing
of Tr value. The results suggested that the membranes could be used directly without further

6
decoloration.
3.3 Mechanical performance

The mechanical properties of the blend membranes under different proportions were examined
by tensile strength and elongation at break tests. The addition of LBMC into PVA was basically
beneficial to improve the mechanical properties of PVA membrane (Fig. 3). When the concentration
of LBMC was 25 wt% (4PVA-LBMC), the tensile strength and elongation at break reached the
maximum values, i.e. 31 MPa and 988%. This result should be contributed to the network structure
formed in this blend membrane. However, the mechanical strength of the blend membranes declined
when the content was higher than 25 wt%, especially for the decreasing of tensile strength. This
probably caused by the formation of closely packed grains and holes observed in SEM investigations
with the over addition of LBMC. As a result, the optimum mechanical performance was observed in
4PVA-LBMC membrane, in which tensile strength and elongation at break were increased
significantly by 39% and 108%, respectively.
3.4 Thermal properties of composite membranes
TGA and DTG analyses were conducted to examine the thermal stability of the membranes. Fig.
4a displays the TGA plots of PVA/LBMC blend membranes with various LBMC compositions.
Three weight loss stages were observed, which corresponded to: (i) the decomposition of residual
solvents, water and short oligomers at a temperature range of 55–170 ºC; (ii) the elimination of side
groups of PVA and LBMC at a temperature range of 170–356 ºC; (iii) the breakdown of polymer
backbones of PVA and LBMC at a temperature range of 400–500 ºC (García-Cruz, Casado-Coterillo,
Irabien, Montiel & Iniesta, 2016). By contrast, the major weight loss distributed in the latter two
stages. For LBMC and PVA membrane, the second stage showed 54% and 67% mass loss, and the
third stage showed 19% and 13% mass loss, respectively. However, the mass loss of 4PVA-LBMC at
the second stage decreased to 36% and at the third stage increased to 33%. The TGA curves revealed
that after blending with LBMC, the thermal stability of membranes was apparently improved,
particularly for that of 4PVA-LBMC membrane. The results provided more evidence for the
successful blending of PVA and LBMC, and the further addition of LBMC would restrict the
performance of blend membranes.

Fig. 4b displayed the DTG plots of various blend membranes. The temperature for the
maximum weight-loss rate (TDTGMAX) of 4PVA-LBMC membrane increased to 289 ºC compared
with that of PVA membrane at 264 ºC. However, further increasing the content of LBMC leaded to
the declination of the maximum weight-loss rate. Besides, blend membranes exhibited the second
degradation peaks at 410–450 ºC in the curves. In addition, the statistic heat-resistant index
temperature (Ts) is given in Table 2. The results were again consistent with the previous
demonstrations. This data was determined from the temperatures at 5 % weight loss (Td5) and 30 %
weight loss (Td30) of the sample obtained by TGA, following Eq. (6):
Ts = 0.49 (Td5 + 0.6 (Td30–Td5)) (6)

Commented [A6]: Corresponding to question 4 of


3.5 Wettability and swelling performance reviewer#2

7
The wettability and swelling performance is associated with the structure and composition of
the polymer membrane. In Fig. 5, the water absorption ratio and weight loss ratio are plotted against
the blend membranes with different LBMC content. The water absorption ratio declined with the
increasing of LBMC content, which might be attributed to the increasing density caused by the
interaction between the PVA and LBMC, as well as the self-interaction of LBMC. However, the
weight loss ratio increased with increasing the LBMC content. It probably caused by the weight loss
from the dissolving of dissociated LBMC with hydrophilic character in water (Li, Zhang & Lao,
2015).

3.6 Antibacterial activity of the blend membranes


To assess the antibacterial activity of the PVA membranes after blending with LBMC, MRSA
was selected to perform the test by the disk method. Besides, the antibacterial activity of chitin
membrane was also conducted as a contrast. As shown in Fig. 6, a negligible inhibition zone and an
obvious inhibition zone were found in pure PVA membrane and chitin membrane, respectively. For
blend membranes, the sizes of corresponding inhibition zones increased dramatically as the LBMC
content increased from 8PVA-LBMC to PVA-LBMC. It was surprisingly that the blend membranes
containing high levels of LBMC showed stronger antibacterial activity than chitin membrane. These
results suggested that the addition of LBMC was beneficial for antibacterial activity, which was
probably attributed to its abundant protonated amine groups in the structure (Yang, Chou & Li,
2005), as well as its improved solubility.

Conclusions
BM-chitin was effectively converted into LBMC by liquefaction method in PEG 400/glycerin
mixed solvent system. Combined with FTIR, 1H NMR and SEC analyses of LBMC, it was turned
out that depolymerization and deacetylation reactions occurred during the liquefaction process, as
well as the Mw and DD’ values of LBMC were determined as 5145 and 38%, respectively. The
results of FTIR suggested that LBMC was further successfully blended with PVA through some
chemical and physical interactions between –OH, –NH2 and –NHAc groups in LBMC and –OH
groups in PVA to form high-performance membranes. The optimum blend membrane with high
tensile and heat resistance was determined as 4PVA-LBMC membrane by the mechanical
performance and thermal stability tests. Moreover, the blend membranes expressed higher weight
loss ratio and better water resistance than pure PVA membrane. Especially, blending of LBMC into
PVA significantly enhanced the antibacterial activity of the membranes. In summary, chitin was
successfully converted into polyols through simple liquefaction process, and the polyols were further
used for the first time in the preparation of polymer materials. Due to the good mechanical and
thermal performance, fine water absorption and retention capacity, as well as the superior
antibacterial properties of blend membranes, they are supposed to have potential application value in
functional packaging materials such as food packaging and medical packaging. Commented [A7]: Corresponding to question 3 of
reviewer#1

8
Acknowledgments
We acknowledge the Key Scientific Research Project Funding of Hainan Province
(ZDYF2017005), the Scientific Research Project Funding of Hainan Higher Education Institution
(Hnky2017-13) and the Graduate Student Innovation Project of Hainan Province (Hys2016-01). We
would like to thank the Analytical and Testing Center of Hainan University for the technical support
and guidance.

References
Ahmad, Ooi, & BS. (2005). Properties–performance of thin film composites membrane: study on
trimesoyl chloride content and polymerization time. Journal of Membrane Science, 255(1), 67-77.
Bahasan, S. H. O., Satheesh, S., & Baakdah, M. A. (2016). Extraction of Chitin from the Shell Wastes of
Two Shrimp Species Fenneropenaeus semisulcatus and Fenneropenaeus indicus using Microorganisms.
Journal of Aquatic Food Product Technology, 26(4), 390-405.
Baxter, A., Dillon, M., Taylor, K. D., & Roberts, G. A. (1992). Improved method for i.r. determination of the
degree of N-acetylation of chitosan. International Journal of Biological Macromolecules, 14(3), 166-169.
Benhabiles, M. S., Abdi, N., Drouiche, N., Lounici, H., Pauss, A., Goosen, M. F. A., & Mameri, N. (2013).
Protein recovery by ultrafiltration during isolation of chitin from shrimp shells Parapenaeus longirostris.
Food Hydrocolloids, 32(1), 28-34.
Benhabiles, M. S., Drouiche, N., Lounici, H., Pauss, A., & Mameri, N. (2013). Effect of shrimp chitosan
coatings as affected by chitosan extraction processes on postharvest quality of strawberry. Journal of
Food Measurement & Characterization, 7(4), 215-221.
Bharty, M. K. (2015). Extraction and characterization of chitin and chitosan from fishery waste by
chemical method. Environmental Technology & Innovation, 3, 77-85.
Chen, F., & Lu, Z. (2009). Liquefaction of wheat straw and preparation of rigid polyurethane foam from
the liquefaction products. Journal of Applied Polymer Science, 111(1), 508-516.
Chen, X., Gao, Y., Wang, L., Chen, H., & Yan, N. (2015). Effect of Treatment Methods on Chitin Structure
and Its Transformation into Nitrogen‐Containing Chemicals. Chempluschem, 80(10), 1565-1572.
Chen, X., Liu, Y., Kerton, F. M., & Yan, N. (2015). Conversion of chitin and N-acetyl-d-glucosamine into a
N-containing furan derivative in ionic liquids. RSC Adv., 5(26), 20073-20080.
Cho, Y. W., Jang, J., Park, C. R., & Ko, S. W. (2000). Preparation and solubility in acid and water of partially
deacetylated chitins. Biomacromolecules, 1(4), 609-614.
D’Souza, J., Song, Z. W., Camargo, R., & Ning, Y. (2016). Solvolytic Liquefaction of Bark: Understanding the
Role of Polyhydric Alcohols and Organic Solvents on Polyol Characteristics. ACS Sustainable Chemistry &
Engineering, 4(3), 851–861.
Davila, J. L. (2009). Synthesis of a chitin-based biocomposite for water treatment: Optimization for
fluoride removal. Journal of Fluorine Chemistry, 130(8), 718-726.
Deng, Q., Li, J., Yang, J., & Li, D. (2014). Optical and flexible α-chitin nanofibers reinforced poly(vinyl
alcohol) (PVA) composite film: Fabrication and property. Composites Part A: Applied Science and
Manufacturing, 67, 55-60.
dos Santos, R. G., Carvalho, R., Silva, E. R., Bordado, J. C., Cardoso, A. C., do Rosário Costa, M., & Mateus,
M. M. (2016). Natural polymeric water-based adhesive from cork liquefaction. Industrial Crops and
Products, 84, 314-319.
Fan-rong, M., Yu-cang, Z., Ji-hui, L., Wen-xing, F., & Jie, Z. (2016). Preparation of a liquefied banana
pseudo-stem based PVAc-nanosilica hybrid membrane and its modification by octadecyltrichlorosilane.

9
RSC Adv., 6(96), 94170-94176.
García-Cruz, L., Casado-Coterillo, C., Irabien, Á., Montiel, V., & Iniesta, J. (2016). High Performance of
Alkaline Anion-Exchange Membranes Based on Chitosan/Poly (vinyl) Alcohol Doped with Graphene
Oxide for the Electrooxidation of Primary Alcohols. Journal of Carbon Research, 2(2), 10.
Hassan, E. B. M., & Shukry, N. (2008). Polyhydric alcohol liquefaction of some lignocellulosic agricultural
residues. Industrial Crops & Products, 27(1), 33-38.
Hongkulsup, C., Khutoryanskiy, V. V., & Niranjan, K. (2016). Enzyme assisted extraction of chitin from
shrimp shells (Litopenaeus vannamei). Journal of Chemical Technology & Biotechnology, 91(5),
1250-1256.
Hu, S., Wan, C., & Li, Y. (2012). Production and characterization of biopolyols and polyurethane foams
from crude glycerol based liquefaction of soybean straw. Bioresource Technology, 103(1), 227-233.
Ifuku, S., Ikuta, A., Egusa, M., Kaminaka, H., Izawa, H., Morimoto, M., & Saimoto, H. (2013). Preparation
of high-strength transparent chitosan film reinforced with surface-deacetylated chitin nanofibers.
Carbohydrate Polymers, 98(1), 1198-1202.
Jasiukaitytė, E., Kunaver, M., & Strlič, M. (2009). Cellulose liquefaction in acidified ethylene glycol.
Cellulose, 16(3), 393-405.
Karger, J. (2011). Liquefaction of plant biomass for use in polymers - is it the right strategy? Express
Polymer Letters, 5(2), 92-92.
Kubota, N., & Eguchi, Y. (1997). Facile Preparation of Water-Soluble N-Acetylated Chitosan and Molecular
Weight Dependence of Its Water-Solubility. Polymer Journal, 29(2), 123-127.
Kubota, N., Tatsumoto, N., Sano, T., & Toya, K. (2000). A simple preparation of half N-acetylated chitosan
highly soluble in water and aqueous organic solvents. Carbohydrate Research, 324(4), 268-274.
Kuo, P. Y., Sain, M., & Yan, N. (2014). Synthesis and characterization of an extractive-based bio-epoxy
resin from beetle infested Pinus contorta bark. Green Chemistry, 16(7), 3483-3493.
Lee, Y.-Y., Lee, W.-J., Hsu, L.-Y., & Hsieh, H.-M. (2014). Properties of molding plates made with various
matrices impregnated with PF and liquefied wood-based PF resins. Holzforschung, 68(1), 37-43.
Li, W., Zhang, Y., & Lao, B. (2015). Preparation and Properties of Liquefied Banana Pseudo-stem based
PVAc Membrane. bioresources, 10(3), 4519-4529.
Liu, B., Han, X., Zhao, H., Wei, L., Zhao, L., & Yuan, L. (2017). Preparation and characterization of
intelligent starch/PVA films for simultaneous colorimetric indication and antimicrobial activity for food
packaging applications. Carbohydrate Polymers, 157, 842-849.
Luo, X., Hu, S., Zhang, X., & Li, Y. (2013). Thermochemical conversion of crude glycerol to biopolyols for
the production of polyurethane foams. Bioresource Technology, 139, 323-329.
Morganti, P., Fabrizi, G., Palombo, P., Palombo, M., Ruocco, E., Cardillo, A., & Morganti, G. (2008).
Chitin-nanofibrils: a new active cosmetic carrier. Journal of Applied Cosmetology, 26(3), 113-128.
Pelissari, F. M., Grossmann, M. V., Yamashita, F., & Pineda, E. A. (2009). Antimicrobial, mechanical, and
barrier properties of cassava starch-chitosan films incorporated with oregano essential oil. Journal of
Agricultural & Food Chemistry, 57(16), 7499-7504.
Pierson, Y., Chen, X., Bobbink, F. D., Zhang, J., & Yan, N. (2014). Acid-Catalyzed Chitin Liquefaction in
Ethylene Glycol. ACS Sustainable Chemistry & Engineering, 2(8), 2081-2089.
Rao, K. S. V. K., Subha, M. C. S., Sairam, M., Mallikarjuna, N. N., & Aminabhavi, T. M. (2010). Blend
membranes of chitosan and poly(vinyl alcohol) in pervaporation dehydration of isopropanol and
tetrahydrofuran. Journal of Applied Polymer Science, 103(3), 1918-1926.
Revoredo, O. B., Nieto, O. M., Suarez, Y., Garcia, V., Fernandez, M., Iraizoz, A., & Henriques, R. D. (2006).

10
Applications of chitin and chitosan in pharmacy and cosmetology. European Journal of Pharmaceutical
Sciences, 28, S7-S8.
Sharma, P., Mathur, G., Dhakate, S. R., Chand, S., Goswami, N., Sharma, S. K., & Mathur, A. (2016).
Evaluation of physicochemical and biological properties of chitosan/poly (vinyl alcohol) polymer blend
membranes and their correlation for Vero cell growth. Carbohydrate Polymers, 137, 576-583.
Song, H., & Zheng, L. (2013). Nanocomposite films based on cellulose reinforced with nano-SiO2:
microstructure, hydrophilicity, thermal stability, and mechanical properties. Cellulose, 20(4), 1737-1746.
Svang-Ariyaskul, A., Huang, R. Y. M., Douglas, P. L., Pal, R., Feng, X., Chen, P., & Liu, L. (2006). Blended
chitosan and polyvinyl alcohol membranes for the pervaporation dehydration of isopropanol. Journal of
Membrane Science, 280(1-2), 815-823.
Wan, W. K., Campbell, G., Zhang, Z. F., Hui, A. J., & Boughner, D. R. (2010). Optimizing the tensile
properties of polyvinyl alcohol hydrogel for the construction of a bioprosthetic heart valve stent. J
Biomed Mater Res, 63(6), 854-861.
Wawrzkiewicz, M., Bartczak, P., & Jesionowski, T. (2017). Enhanced removal of hazardous dye form
aqueous solutions and real textile wastewater using bifunctional chitin/lignin biosorbent. International
Journal of Biological Macromolecules, 99, 754-764.
Wongthep, W., Ruksakulpiwat, C., & Amnuaypanich, S. (2010). Membranes Prepared from a Blend of
Poly(acrylic Acid) and Natural Rubber-Graft-Poly(vinyl Alcohol) (PAA/NR-g-PVA). Advanced Materials
Research, 93-94, 268-271.
Wu, D., & Hakkarainen, M. (2014). A Closed-Loop Process from Microwave-Assisted Hydrothermal
Degradation of Starch to Utilization of the Obtained Degradation Products as Starch Plasticizers. ACS
Sustainable Chemistry & Engineering, 2(2), 2172–2181.
Xu, J., Xie, X., Wang, J., & Jiang, J. (2016). Directional liquefaction coupling fractionation of lignocellulosic
biomass for platform chemicals. green chemical, 18(10), 3124-3138.
Yamada, T., & Ono, H. (2001). Characterization of the products resulting from ethylene glycol liquefaction
of cellulose. The Japan Wood Research Society, 47, 458-464.
Yan, N., & Chen, X. (2015). Don’t waste seafood waste. Nature, 524, 155-157.
Yang, T. C., Chou, C. C., & Li, C. F. (2005). Antibacterial activity of N-alkylated disaccharide chitosan
derivatives. International Journal of Food Microbiology, 97(3), 237-245.
Zhang, J., & Yan, N. (2016). Formic acid-mediated liquefaction of chitin. Green Chemistry, 18(18),
5050-5058.
Zhang, W., Zhang, Y., & Zhao, D. (2012). Synthesis of liquefied corn barn-based epoxy resins and their
thermodynamic properties. Journal of Applied Polymer Science, 125(3), 2304-2311.

11
Fig. 1. FTIR spectra of PVA, PVA/LBMC blend membranes and LBMC. (a) PVA, (b) 8PVA-LBMC,
(c) 4PVA-LBMC, (d) 2PVA-LBMC, (e) 1PVA-LBMC and (f) LBMC.

Fig. 2. SEM images of the cross section of (a, b) PVA, (c, d) 4PVA-LBMC and (e, f) 1PVA-LBMC.

12
Fig. 3. Effect of the LBMC content on the mechanical performance of the blend membranes.

Fig. 4. (a) TGA curves and (b) DTG curves of PVA/LBMC blend membranes. Commented [A8]: Corresponding to question 4 of
reviewer#2

13
Fig. 5. The water absorption and weight loss ratio of the hybrid membranes.

Fig. 6. Antibacterial activity of (a) blank, (b) PVA, (c) chitin, (d) 8PVA-LBMC, (e) 4PVA-LBMC, (f)
2PVA-LBMC and (g) 1PVA-LBMC membranes. Commented [A9]: Corresponding to question 3 of
reviewer#2

14
Table 1
Characteristics data of chitin and LBMC
Acid Average Average
number of Hydroxyl Average molecular molecular
LBMC number of molecular weight of weight of
Liquefaction (mg LBMC (mg weight of BM-chitin LBMC
yield (%) KOH/g) KOH/g) chitin (Mw) α (Mw) α (Mw) α
81 19 285 498316 99857 5145
α
The average molecular weight was calculated by using the calibration curve. Polystyrenes with
various molecular weights were used as the standards. Commented [A10]: Corresponding to question 1 of
reviewer#2

Table 2
TGA data of the membranes and LBMC
Codes Td5 (℃) Td30 (℃) Ts (℃) TDTGMAX (℃) Char800 (%)
PVA 108.9 257.0 96.9 264.2 3.12
8PVA-LBMC 90.0 277.5 99.2 275.5 1.41
4PVA-LBMC 86.3 294.5 103.5 289.0/439.6 3.76
2PVA-LBMC 72.8 251.7 88.3 247.5/443.1 5.88
PVA-LBMC 73.7 243.7 86.1 237.0/417.2 7.07
LBMC 124.9 237.0 94.2 243.9/413.0 6.64

15

You might also like