You are on page 1of 12

Journal of Inorganic and Organometallic Polymers and Materials (2018) 28:1528–1539

https://doi.org/10.1007/s10904-018-0848-1

Effect of Biopolymer Blend Matrix on Structural, Optical and Biological


Properties of Chitosan–Agar Blend ZnO Nanocomposites
G. Magesh1 · G. Bhoopathi1 · N. Nithya1 · A. P. Arun2 · E. Ranjith Kumar3

Received: 19 December 2017 / Accepted: 7 April 2018 / Published online: 9 April 2018
© Springer Science+Business Media, LLC, part of Springer Nature 2018

Abstract
The present research is focused on the development of ecofriendly biopolymer blend based nanocomposites to enhance
the effect of cytotoxic activity. Novel eco-friendly synthesis of pure Chitosan–Agar blend and Chitosan–Agar/ZnO nano-
composites was successfully synthesized by in-situ chemical synthesis method. The influence of Chitosan–Agar (1:1 wt/
wt%) concentrations (0.1, 0.5, 1 and 3 g) was studied. The presence of ZnO nanoparticles in Chitosan–Agar polymer matrix
was confirmed by UV, FTIR, XRD, FESEM, EDAX and TEM. The crystallite size of the nanocomposites in the range of
12–17 nm is observed from XRD analysis. PL and UV reveal that Nanocomposites shows an blue shift by increase in the
blend concentrations. TEM analysis shows that 0.1 and 3 g of Chitosan–Agar/ZnO Nanocomposites are in spindle and spheri-
cal shape with polycrystalline nature. The prepared Nanocomposites shows the respectable Antibacterial activity against
Gram-positive (Staphylococcus aureus and Bacillus subtilis) and Gram-negative (Pseudomonas aureginosa and Klebsilla
pneumonia) bacteria. The potential toxicity of Chitosan–Agar/ZnO nanocomposites was studied for normal (L929) and
breast cancer cell line (MB231). The result of this investigation shows that the Chitosan–Agar/ZnO nanocomposites deliver
a dose dependent toxicity in normal and cancer cell line.

Keywords  Nanocomposites · Structural analysis · Morphological analysis · Antibacterial activity · Anticancer activity

1 Introduction biopolymers was increased considerably due to their excep-


tional property in biomedical application [3].
Recently, bio polymer blending has been widely used for the Chitosan is a cationic linear polysaccharide comprised of
development of drug delivery systems and eco-friendly anti- β (1–4) 2-acetamido-2-deoxy-d-glucosamine [4]. Chitosan
bacterial packaging materials as an alternative for synthetic has tremendous application in the biomedical field, owing
polymers [1]. Biopolymers are a versatile carrier due to their to its gel forming capability, solubility in dilute acidic acids,
enhanced biodegradability, nontoxic, biocompatibility, eco- biocompatibility, nontoxicity and chemical resistant proper-
friendly property etc. [2]. Polysaccharide based bio polymers ties [5]. The binding ability of chitosan with other polymer
are readily available, biocompatible, environmental friendly which is attributed to the amino (–NH2) and hydroxyl (–OH)
and are also cheap. The performance of the polysaccharide functional groups [6]. To improve the biological properties
of Chitosan, several bio polymers are blended with chitosan
such as methyl cellulose [7], sodium alginate [8], gelatin [9],
* G. Bhoopathi
bbijjju03@gmail.com collogen [10] and corn starch [11]. Agar is an unbranched
polysaccharide extracted from the marine algae that consists
* E. Ranjith Kumar
ranjueaswar@gmail.com of a mixture of agarose and agaropectin. It is one of the most
attractive material because of its low cost, good gel strength,
1
Department of Physics, PSG College of Arts and Science, abundant in nature, non-toxicity, eco-friendly and recycla-
Coimbatore 641 014, Tamilnadu, India bility [12]. Saxena et al. studied that, Agar–Gelatin blend
2
Department of Mechanical Engineering, Kumaraguru had potential effectiveness of drug release.The characteris-
College of Technology, Coimbatore 641035, Tamilnadu, tics with a large array of functions and molecular diversity
India
structure of agar made it as a good material for biopoly-
3
Department of Physics, Dr. N.G. P. Institute of Technology, mer in composites and blend [13]. Chitosan–Agar blend
Coimbatore 643 048, India

13
Vol:.(1234567890)
Journal of Inorganic and Organometallic Polymers and Materials (2018) 28:1528–1539 1529

has been already reported that, they have high miscibility temperature. The agar was prepared by preheated ultrapure
strength and good mechanical and chemical properties [14]. water followed by stirring at a room temperature of 90 °C
Hu et al., reported that, Chitosan–Agar blend has potential for 2 h until a clear solution obtained. The aqueous Agar
application in food packaging material and targeted drug solution was added drop by drop to Chitosan solution,
delivery system [15]. The properties of biopolymer blend under constant stirring until a homogeneous solution were
(Chitosan–Agar) may lead it as biomaterial for the develop- obtained. The homogeneous of Chitosan/Agar (1:1 wt/
ment of new bio polyblend Nanocomposites. wt%) solution was cast into a petri dish. The cast solution
Recently, polyblend nanocomposites have a great research was allowed to dry in oven at 60 °C for 24 h. The dried
interest in the field of nanotechnology [16]. In-situ synthesis, film was carried and stored in a desiccator for futher char-
metal oxide nanoparticles were prepared from their corre- acterization. The blended films were prepared by ionically
sponding precursors in the presence of the polymer blend cross-linked method is described by El-hefian et al. [14].
matrix [17]. This successful method is used to prevent
agglomeration and improve the size distribution in a poly-
mer matrix which is attributed to the interfacial interaction 2.3 Preparation of Chitosan–Agar /ZnO
between organic and inorganic material [18]. It enhances Nanocomposites
the biological properties because of their synergistic effect.
Among the metal oxide nanoparticles, ZnO nanoparticles 0.2M of zinc acetate was added into the equal quantity of
show a potential antibacterial activity against gram negative Chitosan–Agar blend under constant stirring for 2 h. The
and gram positive bacteria [19, 20] and cytotoxicity against required amount of NaOH solution was added dropwise to
cancer cell line with lower toxicity towards the normal cell the mixture with constant stirring until pH 12 is attained.
line [21]. Alireza et al. observed that, PVA/PANI/ZnO nano- The transparent solution turned to milky white and the
composites exhibits a strong antibacterial activity against reaction continued for 2 h. The white precipitate formed
gram negative and gram positive bacteria [20]. Shekhar et al. was allowed to settle by keeping the mixture for 24  h
suggested Chitosan–PVA hydrogel as a perfect applicant for aging. The precipitate was separated by centrifugation
the effective in-situ synthesis of silver nanoparticles [22]. and then rinsed with distilled water for several times to
In this work, different concentration (0.1, 0.5, 1, 3 g) of remove the excess of by-products. Finally, Chitosan–Agar/
Chitosan–Agar blend ZnO nanocomposites were prepared ZnO Nanocomposites were dried at 60 °C for 24 h. A set
by in-situ chemical synthesis method. Here, Chitosan–Agar of four homogeneous samples named: (a) 0.1 g, (b) 0.5 g,
blend was used to achieve controlled synthesis of ZnO nano- (c) 1 g, (d) 3 g was prepared by varying the concentration
particles. The Chitosan–Agar blend ZnO nanocomposites of Chitosan and Agar blend at a fixed concentration of
were characterized by XRD, FESEM and EDX, HRTEM, ZnO and NaOH.
UV, FTIR, PL analysis and their biological properties (anti-
bacterial and anticancer activity) were investigated.
2.4 Material Characterization

2 Materials and Methods The Chitosan–Agar /ZnO nanocomposites were character-


ized by XRD, FESEM with EDX, HRTEM, UV, FTIR and
2.1 Chemicals PL. The phase purity of the samples was confirmed by X-ray
Diffraction (X-PERT-3, Philips PW3064/60) with CuKα1
Chitosan with 85% degree of deacetylation and Agar was (λ = 1.54060 Å) and CuKα2 (λ = 1.54443 Å) radiation oper-
purchased from Sigma Aldrich Co. Ltd. Acetic acid (Gla- ating at 30 mA and 45 kV. Data were collected over a 2Θ
cial 100%, Pro analysis), sodium hydroxide and zinc acetate range of 10°–90°. The structure morphology and composi-
dehydrate (Zn ­(CH3COO)2·2H2O) were purchased from tion of the samples were analyzed by (SIGMA HV-CARL
Merck. All chemicals were used without further purification ZEISS) FESEM with Bruker Quantax-200-Z10 EDX detec-
and freshly prepared solutions were used in all experiments. tor and HRTEM images were taken in (JEOL–JEM-2010,
The detailed sample preparation techniques are described Japan) with an accelerating voltage of 200 kV using SAED
below. pattern were also performed. The optical properties of the
samples were characterized by UV spectrometer (JASCO-
2.2 Preparation of Chitosan and Agar Blend V-770 Spectroscopy) at a wavelength range between 200 and
900 nm. PL spectrum was performed with FP-3800 spectro-
The preparation of Chitosan–Agar Blend was carried at flurometer. FTIR spectroscopic analysis were recorded using
(1:1 wt/wt%) ratio. Briefly, Chitosan was prepared by dis- (Bruker Alpha FTIR spectrometer) at a wavenumber range
solving 0.5 g in 100 ml of acetic acid solution at room of 400–4000 cm−1.

13

1530 Journal of Inorganic and Organometallic Polymers and Materials (2018) 28:1528–1539

2.5 Antibacterial Activity 1 and 3 g) of Chitosan–Agar /ZnO nanocomposites were


studied by XRD analysis. The Chitosan–Agar blend shows
The antibacterial activity of pure Chitosan–Agar blend and amorphous feature characterized by two halos around Bragg
various concentration of Chitosan–Agar/ZnO was performed angles 10.67° and 19.96° is shown in Fig. 1a. This property
against Gram-positive (Bacillus subtilis and Staphylococcus is attributed by the formation of intermolecular hydrogen
aureus) and Gram-negative (Pseudomonas aureginosa and bonding interaction occurred between the two polymers
Klebsilla pneumonia) bacteria by disc diffusion method. The [23]. The peaks become weak in the XRD pattern of Chi-
microorganisms were inoculated in the fresh nutrient broth and tosan–Agar /ZnO Nanocomposites as shown in Fig. 1b–e.
incubated to get required turbidity. The nutrient agar plate was All the XRD peaks of Chitosan–Agar /ZnO nanocomposites
swabbed with the test bacterial culture and various concentra- corresponds to the hexagonal wurtzite structure. The peaks
tions of sample discs (100 µg) (discs are soaked overnight in for 2θ values at 36.3°, 34.4°, 31.8°, 47.6°, 56.67°, 62.93°,
nanoparticle suspension) was added into the well. After 24 h 68.0° and 69.15° were pertaining to ZnO which indicate the
incubation at 37 °C, the antibacterial activity was measured reduction of ZnO nanoparticles in the polymer matrix. The
based on the diameter of the zone of clearance. discernible peaks can be indexed to the planes (101), (002),
(100), (102), (110), (103). (112) and (220) and these are
2.6 Cytoxicity of the Samples perfectly matched with JCPDS Card No. (36-1451). From
Fig. 1b–e shows that, with an increase in the concentration of
2.6.1 Cell Line and Cell Culturing Chitosan–Agar blend the peak becomes broader and weaker
which implies the possibility of changes in grain size. This
Breast cancer cell line (MB-231) and Normal Fibro Blast might be due to the lower crystallinity and the formation
(L929) cell line were obtained from National Centre for of smaller nanoparticles. However, it confirms the role of
Cell Science (NCCS), Pune and grown in Eagles Minimum Chitosan–Agar blend as reducing and stabilizing agents. The
Essential Medium (EMEM) containing 10% Fetal Bovine crystallite size and the strain analysis of the Nanocomposites
Serum (FBS). The cells were cultured in a humidified atmos- are carried out using De-bye Scherrer method and W–H plot
phere with 5% ­CO2 at 37% in 96 well tissue plates. method which is discussed below [24].

2.6.2 MTT Assay 3.1.1 Debye Scherer Method

The Chitosan–Agar/ZnO Nanocomposites were evaluated The crystallite size and strain of the Chitosan–Agar blend
for cytotoxic activity of MB-231 (cancer cell) and L929 ZnO nanocomposites was calculated by Debye Scherer
(Normal cell). The cells were plated in 96 well cell plates Eq. (1).
(1 × 105) cells/well and left overnight at 37 °C. The medium
was then discarded and the cells were treated with differ-
ent concentrations of ZnO (25, 50, 75, 100 and 125 µg/ml).
Incubate the samples at 37° ± 1 °C for 18 h, then MTT was
added to all the wells and incubated for 4 h. After incuba-
tion, DMSO was added into the wells and the absorbance
read at 570 nm using a photometer. Cytotoxicity and Cell
viability of the samples were calculated using the formula
given below. The concentrations required for 50% Inhibition
of Viability ­(IC50) were determined graphically.
[ ]
Cytotoxicity = (Control − Treated)∕Control × 100

Cell Viability =(Treated∕Control) × 100

3 Result and Disccussion

3.1 Structural Evolution of Nanoparticles Through


X‑ray Diffraction Analysis
Fig. 1  X-ray diffraction patterns a pure Chitosan–Agar blend and
Figure 1 shows the structural properties of the pure Chi- b–e 0.1, 0.5, 1 and 3 g of Chitosan–Agar/ZnO nanocomposites (inset
tosan–Agar blend and different concentrations (0.1, 0.5, shows peak broadening)

13
Journal of Inorganic and Organometallic Polymers and Materials (2018) 28:1528–1539 1531

k𝜆 is called W–H plot equations. Figure 2 shows the W–H


D= , (1) plot of Chitosan–Agar blend ZnO nanocomposites. From
𝛽 cos 𝜃
the linear fit data, the Y-intercept and slope represents the
where D is the crystallite size (nm), λ is the wavelength of crystallite size, strain respectively.
the radiation (λ = 1.541 Å), k is a constant (0.94), β is the peak
width at half maximum (FWHM) in radian and θ is the peak 3.2 FESEM and EDX Analysis
position. Micro strain is calculated using the Eq. (2).
𝛽 cos 𝜃 The surface morphology (0.1, 0.5, 1, and 3  g) of Chi-
𝜀= . (2) tosan–Agar/ZnO samples have been characterized by
4
FESEM. Figure 3a–d shows the well distributed ZnO in
The calculated crystallite size and strain were shown in biopolymer matrix. From the SEM images, it was clear that
Table 1. The unit cell volume, bond length and lattice param- they are in nanosize with uniform spindle shape morphol-
eters of the Nanocomposites are calculated from the equation ogy. While increasing the concentration of Chitosan–Agar
stated by Bindu and Sabu [25]. blend the size distribution of the particles is different. There
From Table 1 it can be found that the crystallite size was was no agglomeration observed between the particles. It is
found to be 17.07–12.98  nm. While increasing the Chi- suggested that the Chitosan–Agar blend played a crucial
tosan–Agar concentration, the intensity of the peaks is found role in stabilizing and terminates the growth of the particle
to decrease gradually and becomes relatively broad suggesting which correlates with XRD. The elemental composition of
the decrease in the degree of crystallinity [26]. The decrease in the (0.1, 0.5, 1 and 3 g) Chitosan–Agar/ZnO nanocomposites
crystallite size (D) is mainly because of the Biopolymers (Chi- were analyzed by EDX in the range between 0 and 20 Kev
tosan–Agar) which adsorbs to the surface and it restricts the binding energy. Figure 4a–d shows the quantitative atomic
growth of ZnO nanoparticles [27]. The microstarin increases and weight percentage (%) of Zn and O. The EDX analysis
with decrease in particle size. The calculated lattice constants confirms the presence of Zinc and Oxygen which indicates
‘a’ and ‘c’ is very close to the standard values [36-1451]. There the purity of the samples.
is no significant change in lattice constants. The bond length
values of Chitosan–Agar /ZnO Nanocomposites are 1.973, 3.3 HRTEM and SAED Analysis
1.975, 1.976 and 1.979 Å respectively. The reduction of ZnO
in the biopolymer matrix may be the reason for changes on The structural morphology of the samples was character-
Zn–O bond length. ized by HRTEM and SAED pattern. Figure 5 shows the
HRTEM morphology with different magnification of 0.1 g
3.1.2 Williamson–Hall (W–H) Method of Chitosan–Agar /ZnO nanocomposites. It is observed that
at 0.1 g of Chitosan–Agar/ZnO nanocomposites (Fig. 5a,
In order to understand the strain associated with the sample b) were spindle in shape with well defined edges. Figure 5c
due to lattice deformation and line broadening analysis were indicates the lattice fringes of interplanar distance d = 2.80 Å
carried out by Williamson–Hall (W–H) method [28]. William- that corresponds to miller indices of (200) crystal planes.
son and Hall proposed a good method for separation of strain Figure 5d exhibits the SAED pattern and the particle size
and size effects on broadening by the peak width as a function distribution histogram. An average particle size of 15–20 nm
of diffraction angle. The addition of the Scherrer equation and is found from the particle size histogram which is consist-
the volume-weighted average strain εstr = βhkl/4 tan θ results in ent with the crystallite size obtained from XRD. The bright
the following Eq. (3), spots in SAED pattern show the polycrystalline nature of the
k𝜆 sample with symmetrical orientation. The HRTEM image of
𝛽hkl = + 4𝜀 tan 𝜃 (3) 3 g Chitosan–Agar/ZnO nanocomposites (Fig. 6a, b) were
D cos 𝜃

Table 1  Structural parameters of Chitosan–Agar/ZnO nanocomposites


Chitosan–Agar Debye Scherer method W–H plot method Lattice parameters Atomic packing Bond length
concentration (g) factor c/a ratio (Zn–O) L (Å)
Crystallite size ‘D’ (nm) D (nm) Strain ɛ × 10− 3 a = b (Å) c (Å)

0.1 17.07 17.34 0.5 3.241 5.197 1.6036 1.9730


0.5 16.87 16.31 0.9 3.245 5.200 1.6032 1.9750
1 15.08 15.23 1.1 3.246 5.201 1.6030 1.9755
3 12.98 13.34 3.2 3.252 5.214 1.6023 1.9795

Standard values: a = b = 3.498 and c = 5.2066 Å (JCPDS card no. 36-1451)

13

1532 Journal of Inorganic and Organometallic Polymers and Materials (2018) 28:1528–1539

Fig. 2  a–d Williamson and Hall (W–H) plot of (0.1, 0.5, 1, and 3 g) Chitosan–Agar/ZnO nanocomposites

Fig. 3  a–d SEM images of (0.1, 0.5, 1 and 3 g) Chitosan–Agar/ZnO nanocomposites

13
Journal of Inorganic and Organometallic Polymers and Materials (2018) 28:1528–1539 1533

Fig. 4  a–d EDX spectrum of (0.1, 0.5, 1 and 3 g) Chitosan–Agar/ZnO nanocomposites

in spherical and spindle shape morphology. Chitosan–Agar n − π* transistion and semicrystalline nature of the sample
blend formed a layer around the ZnO nanoparticles is clearly [30]. This peak comes from the presence of C=O strech-
visible as a layer of amorphous phase. Figure 6c indicates ing bond which were observed in FTIR at about 1646 cm−1
the lattice fringes of interplanar distance d = 2.24 Å cor- [31]. From Fig. 7a–d, the absorption band is obtained at
responding to the Miller indices of (101) crystal plane. 367 nm for low concentration (0.1 g) of Chitosan–Agar/ZnO
Figure 6d exhibits the SAED pattern and particle size dis- Nanocomposites. By increasing the concentration (0.5, 1 and
tribution histogram. An average particle size of 14–15 nm 3 g) of Chitosan–Agar, the absorption peak shifts to lower
is found from the histogram which is consistent with the wavelength 359, 354 and 349 nm which can be assigned
crystalline size obtained form XRD. The rings with discrete to the intrinsic bandgap absorption of ZnO due to electron
spots shows the polycrystalline nature of the sample. The d transistion from the valence band to the conduction band.
spacing values calculated from the circular ring pattern cor- This indicates the formation of ZnO nanoparticles in the
responds to the lattice planes of (100), (002), (101), (102), biopolymer blend matrix.
(110) and (103) which well agrees with the XRD results. In The optical bandgap of the nanocomposites were deter-
this case, Chitosan–Agar biopolymer blend may enhance mined by using the tauc relation (4) [32].
the Ostwald ripening kinetics by forming the inter and intra
molecular interaction between metal oxides ­(Zn2+) ions and
)n
(4)
(
(𝛼h𝜐) = 𝛽 h𝜐 − Eg .
the amine, hydroxyl group of the blends which terminate the
rate of particle growth [29]. The value of n = 1/2, 3/2, 2 or 3 depends on the nature of
the electronic transistion responsible for absorption. Here
3.4 UV–Vis Analysis (UV–Vis) n = 1/2 corresponds to the direct bandgap of semiconduc-
tors [33].
Figure 7 shows the room temperature UV–Vis spectra of Where, α = absorption coefficient, hυ = photon energy,
pure Chitosan–Agar blend and Chitosan–Agar/ZnO nano- ­Eg = direct bandgap energy.
composite for various concentrations of (0.1, 0.5, 1 and 3 g) The bandgap energy of the Chitosan–Agar/ZnO nano-
Chitosan–Agar blend. Figure 2(insert) shows sharp peak at composites were calculated by the extrapolation of a straight
220 nm for Chitosan–Agar blend which was attributed to the line graph between the (αhυ)2 verses photon energy (hυ)

13

1534 Journal of Inorganic and Organometallic Polymers and Materials (2018) 28:1528–1539

Fig. 5  HRTEM images a, b 0.1 g of Chitosan–Agar/ZnO nanocomposites for different magnification, c interplanar lattice fringes, d SAED pat-
tern and the particle size distribution histogram

Fig. 6  HRTEM images a, b
3 g of Chitosan–Agar/ZnO
nanocomposites for different
magnification, c interplanar
lattice fringes, d SAED pattern
and the particle size distribution
histogram

13
Journal of Inorganic and Organometallic Polymers and Materials (2018) 28:1528–1539 1535

Fig. 7  a–d UV–Vis absorption spectra of (0.1, 0.5, 1 and 3  g) Chi- Fig. 8  a–d Tauc plot of (0.1, 0.5, 1 and 3  g) Chitosan–Agar/ZnO
tosan–Agar/ZnO nanocomposites (insert shows the UV–Vis spectra nanocomposites
of pure Chitosan–Agar blend)

as shown in Fig. 8a–d. The bandgap energy of the Nano- shifts to lower energy side from 3400 to 3366 and 1662 to
composites was calculated as 3.14, 3.21, 3.23 and 3.26 eV 1646 cm− 1 with increase in Chitosan–Agar concentration.
respectively. While increasing the concentration of Chi- This observation results confirm the reduction of ZnO nano-
tosan–Agar blend the band gap also increases. The observed particles in the Chitosan–Agar blend matrix.
red shift in bandgap is due to smaller crystallites and strain.
3.6 Photoluminescence Spectroscopic
3.5 FTIR Spectrum Characterization

Figure 9 shows the FTIR spectrum of pure Chitosan–Agar The effect of Chitosan–Agar/ZnO nanocomposites at vari-
blend and Chitosan–Agar/ZnO nanocomposites between ous concentrations (0.1, 0.5, 1 and 3 g) of Chitosan–Agar
4000 and 400 cm−1 using KBr pellet method at room tem- blend ratio was studied using PL under photon excitation
perature. Figure 9a exhibits a characteristics broadband at at 325 nm. Figure 10 shows the changes in the PL spec-
3400 cm−1 which is attributed to (–OH) stretching vibration trum of Chitosan–Agar/ZnO nanocomposites. The emission
modes of the adsorded water on the surface [34]. While the
absorption peak at 1647, 1559, 1417 and 1376 cm−1 are
ascribed to bending vibration of C=O, –NH2, C–C, and CH
(amide II) group. Based on the observation Chitosan–Agar
is perfectly compatible and miscible polymers by intramo-
lecular hydrogen bends between the amino and the hydroxyl
group of Chitosan and Agar [35]. The IR spectrum of nano-
composites contains all the characteristics peak of Chitosan,
Agar and ZnO (Fig. 9b–e). On comparision with blend a
new intense absorption peaks at 584–512 and 456–452 cm−1
are assigned to stretching mode of Zn–O bonding is due
to the inter molecular interaction between Amide group
of Chitosan–Agar blend and ZnO. The weak band around
898–877 cm−1 is ascribed to the Zn–O vibration frequency.
This absorption bands is well matched with the literature
[36, 37]. While increasing the Chitosan–Agar concentra-
tion, the intensity of the Zn–O vibration frequency become
weak and shifts to lower frequency region which is due to
the changes in the nanostructural features. The intensity of Fig. 9  FTIR spectrum of a pure Chitosan–Agar blend and (0.1, 0.5, 1
the characteristics band become broader and considerably and 3 g) Chitosan–Agar/ZnO nanocomposites (b–e)

13

1536 Journal of Inorganic and Organometallic Polymers and Materials (2018) 28:1528–1539

spectrum exhibits a weak blue broadband which is observed the Chitosan–Agar blend loaded with ZnO exhibited potent
at 446 nm due to surface defects in ZnO, mainly due to Zn antibacterial activity against Gram-positive and Gram-neg-
vacancy [38]. A sharp emission peak is observed at 545 nm ative bacteria. As the concentration of Chitosan–Agar blend
in the green region of the spectrum which is due to the sin- increases from 0.1 to 3 g the zone of inhibition also increases
gly ionized oxygen vacancy in ZnO and recombination of (Fig. 12). Enhanced bioactivity is due to the generation of
a photo generated hole in the single ionized charge state ­H2O2 molecules on the surface of ZnO which can penetrate
of the particular defect [39]. On varying the concentra- through the cell membrane and inhibit the growth of bac-
tions, we observe a small blue shift in the blue emission teria. In addition, it is clear that the bactericidal activity
along with a systematic blue shift in the green emission of increases with decrease in particle size which might be due
Chitosan–Agar/ZnO nanocomposites. On interaction with to the large surface to volume ratio [19]. While comparing
3 g of Chitosan–Agar blend, a large blue shift of the green Gram-positive and Gram-negative bacteria, it has been fre-
emission from 547 to 535.5 nm is observed as shown in quently observed that Chitosan–Agar/ZnO Nanocomposites
Fig. 10. But in this case intensities of both blue emission and shows strong antibacterial activity towards Gram-negative
green emission of Chitosan–Agar/ZnO nanocomposites are than Gram-positive which may be attiributed to the different
quenched significantly while increasing the concentrations structure and thickness of peptide glycans in the cell wall
of Chitosan–Agar [40]. [42]. Ahmed et al. reported that CMC/PVA films loaded
Figure 10 inset shows concentration vs intensity graph with ZnO nanocomposites shows the respectable antibacte-
of Chitosan–Agar/ZnO nanocomposites. It is evident that at rial activity against Gram-positive and Gram-negative bac-
545 nm, the PL peak intensity decreases with increase in the teria [43]. Based on the experimental data (Table 2) it was
concentration of Chitosan–Agar blend which might be due observed that, the bactericidal activity of Chitosan–Agar/
to the decrease of particle size. These changes of emission ZnO nanocomposites was greater than Chitosan–Agar blend
confirms the reduction of ZnO nanoparticles in the Biopoly- which may attribute to the synergistic biological properties
mer blend matrix [41]. of Chitosan, Agar and ZnO.

3.7 Antibacterial Activity 3.8 Anticancer Activity

Figure  11 shows the antibacterial activity of Pure Chi- 3.8.1 MTT Assay


tosan–Agar blend and (0.1, 0.5, 1 and 3 g) Chitosan–Agar/
ZnO nanocomposites against Gram-positive (B. subtilis In vitro cytotoxic activity of Chitosan–Agar/ZnO nanocom-
and S. aureus) and Gram-negative (P. aeruginosa and K. posites (Control, 25, 50, 75, 100, 125 µg/ml) against MB-231
pneumonia) bacteria respectively. Chitosan–Agar blend (cancer cell line) was evaluated and compared with L929
(control) did not reveal any antibacterial activity, whereas (normal fibroblast cell line) after 24 h incubation. Figure 13
shows the anticancer activity of the Nanocomposites treated
with MB-231 cell line. The highest cell inhibition 87% was
observed in the concentration of 125 µg/ml. Cytotoxicity
rate against MB-231 cell line increases with an increase in
the concentration of the sample. The L929 normal cell line
similarly treated with ZnO Nanocomposites shows low tox-
icity 17% at 125 µg/ml concentration as shown in Fig. 14a.
The cytotoxicity of the nanocomposites might be due to the
electrostatic interaction between negatively charged anionic
phospholipids on the cancer cells with positively charged
surface of the ZnO. The size of the nanocomposites plays
a great role in the cytotoxicity [44].The cytotoxicity of the
nanocomposites against the cancer cell line is due to the
generation of reactive oxygen species (ROS). The anticancer
activity of ZnO might be due to the functionalization of bio-
active compounds present in the Chitosan–Agar biopolymer
blend. Similar reports of cytotoxicity of ZnO nanoparticles
against A549 and normal cell line were discussed by Balraj
et al. [45] and Mahendran et al. [46].
Fig. 10  a–d Photoluminescence spectra of (0.1, 0.5, 1 and 3 g) Chi-
tosan–Agar/ZnO nanocomposites [insert shows the PL peak intensity
(545 nm) variation with the concentration]

13
Journal of Inorganic and Organometallic Polymers and Materials (2018) 28:1528–1539 1537

Fig. 11  Antibacterial activity
of pure Chitosan–Agar blend
­(B0), 0.5 g (­ B1), 1 g (­ B2), g (­ B3)
and 5 g ­(B4) Chitosan–Agar/
ZnO nanocomposites against
a, b Gram-negative bacteria (P.
aeruginosa and K. neumoniae)
and c, d Gram-positive bacteria
(B. subtilis and S. aureus)

25 Table 2  Antibacterial activity of Chitosan–Agar/ZnO nanocompos-


ites
20
Samples (g) Zone of inhibition (mm)
Zone of Inhibtion (mm)

15 P. aeruginosa K. pneumonia S. aureus B. subtilis


0.1 g

10
0.5 g Control – – – –
1g O 0.1 14 12 15 14
3g O
5 0.5 16 15 16 14
1 18 15 17 15
0 3 20 18 17 16
P.aeruginosa K.neumoniae S.aureaus B.subtillis
Gram Negative Gram Positive

Fig. 12  Graphical representation of zone of inhibition (mm) of (0.1, 4 Conclusion


0.5, 1 and 3 g) Chitosan–Agar/ZnO nanocomposites
Chitosan–Agar blend ZnO Nanocomposites with varied
Chitosan–Agar blend (0.1, 0.5, 1 and 3 g) concentration
The dose response curves for the MB-231 and L929 cell were prepared by simple, facile and green in-situ approach.
lines are shown in Fig. 14b. The inhibitory concentration XRD result exhibits the Chitosan–Agar blend ZnO Nano-
IC(50) values for ZnO Nanocomposites against MB-231 composites were in the hexagonal wurtzite structure. On
cell line was 48.82 µg/ml, while it was found to be signifi- increasing the concentration of Chitosan–Agar blend, the
cantly higher (> 100 µg/ml) for the normal cell line. The crystallite size reduced from 17.07 to 12.98 nm. FESEM
IC(50) value of the MB-231 cancer cell line is lower than and HRTEM images reveal the spindle shape morphology
L929 cell line indicating the ZnO Nanocomposites as a and the size distribution of Chitosan/Agar blend nanocom-
potential dose dependent anticancer agent. posites. The presence of Zn, O and the composition of

13

1538 Journal of Inorganic and Organometallic Polymers and Materials (2018) 28:1528–1539

Fig. 13  a–f Cytotoxic activity


of Chitosan–Agar/ZnO nano-
composites on MB-231 cell line
at different concentrations

Fig. 14  a In vitro cytotoxic


effect of Chitosan–Agar/ZnO
nanocomposites (bar diagram).
b Percentage of viability versus
concentration Chitosan–Agar/
ZnO nanocomposites against
MB-231 (Cancer) and L929
(Normal) cell line

the nanocomposites were confirmed by EDX spectrum. References


FTIR spectrum reveals the intraction between the func-
tional group of ZnO and the polymer blend matrix. From 1. L.A. Utracki (ed.) Polymer Blends Handbook (Kluwer Academic
UV, the optical bandgap energy of nanocomposites was Publication Print, Dordrecht, 2003), pp. 1–122
2. R. Tong, J. Cheng, Anticancer Polymeric Nanomedicines. Polym.
found to vary from 3.14, 3.21, 3.23 and 3.26  eV. Chi- Rev. 47, 345–381 (2007)
tosan–Agar blend ZnO nanocomposites exhibit excellent 3. L. Yu, Biodegradable Polymer Blends and Composites from
antibacterial activity against Gram-positive and Gram- Renewable Resources. (Wiley, Hoboken, 2009)
negative bacteria which may due to the synergitic biologi- 4. E.A. El-hefian, M.M. Nasef, A.H. Yahaya, Chitosan-based poly-
mer blends: current status and applications. J. Chem. Soc. Pak.
cal property of ZnO,Chitosan and Agar. The antibacterial 36(1), 11–27 (2014)
activity increases with an increase in the concentration 5. H. Honarkar, M. Barikani, Applications of biopolymers I: Chi-
of Chitosan–Agar blend. Furthermore, the Chitosan–Agar tosan. Monatsh. Chem. 140, 1403–1420 (2009)
blend ZnO nanocomposites exhibit in vitro cytotoxic effect 6. J. Li, Y. Wu, L. Zhao, Antibacterial activity and mechanism of
Chitosan with ultra high molecular weight. Carbohydr. Polym.
against breast cell line (MB231) and with lower toxic- 148, 200–205 (2016)
ity against fibroblast normal cell line (L929). Therefore, 7. O. Gh, A. Rawad, R.H. Yahya, Structural, optical, and electri-
from the above studies, the synthesized Chitosan–Agar cal characterization of Chitosan: methylcellulose polymer blends
blend ZnO nanocomposites could be applicable for the based film. J. Mater. Sci. Mater. Electron. 28, 10283–10294
(2017)
biomedical field.

13
Journal of Inorganic and Organometallic Polymers and Materials (2018) 28:1528–1539 1539

8. V.K. Malesu, D.Sahoo and A.P.L. Nayak, Chitosan–sodium algi- the non enzymatic glucose biosensor. Mater. Sci. Eng. C 75,
nate nanocomposites blended with cloisite 30b as a novel drug 1472–1479 (2017)
delivery system for anticancer drug curcumin. Int. J. Appl. Biol. 29. G. Shan, H. Hao, X. Wang, Z. Shang, Y. Chen, Y. Liu, The effect
Pharm. Technol. 2(3), 402–411 (2011) of PVP on the formation and optical properties ZnO/Ag nanocom-
9. E. Pulieri et al., Chitosan/gelatin blends for biomedical applica- posites. Colloids Surf. A 405, 1–5 (2012)
tions. J. Biomed. Mater. Res. A 86, 311–322 (2007) 30. S. Elashmawi, A.M. Abdelghany, N.A. Hakeem, Quantum con-
10. J. F. Mendes, J.E. Oliveira, Biodegradable polymer blends based finement effect of CdS nanoparticles dispersed within PVP/PVA
on cornstarch and thermoplastic Chitosan processed by extrusion. nanocomposites. J. Mater. Sci. 24, 2956–2961 (2013)
Carbohydr. Polym. 137, 1–19, (2015) 31. N.S. Alghunaim, Results in physics spectroscopic analysis of
11. C. Tangsadthakun, S. Kanokpanont, R. Pichyangkura, T. Banap- PMMA/PVC blends containing C ­ oCl2. Results Phys. 5, 331–336
rasert, Y. Tabata, The influence of molecular weight of Chitosan (2015)
on the physical and biological properties of collagen/Chitosan 32. I. Pankove, Optical Process in Semiconductors. (Prentice-Hall,
scaffolds. J. Biomater. Sci. Polym. 18(2), 147–163 (2012) New Jersey, 1971)
12. H.P.S.A. Khalil et al., Biodegradable polymer films from seaweed 33. K.M. Reddy, S.V. Manorama, A.R. Reddy, Bandgap studies on
polysaccharides: a review on cellulose as a reinforcement mate- anatase titanium dioxide nanoparticles. Mat. Chem. Phys. 78,
rial. Express Polym. Lett. 11(4), 244–265 (2017) 239–245 (2002)
13. A. Saxena, A. Tahir, M. Kaloti, J. Ali, H.B. Bohidar, Effect of 34. M.M. AbdElhady, Preparation and characterization of Chitosan/
agar–gelatin compositions on the release of salbutamol tablets. Zinc oxide nanoparticles for imparting antimicrobial and UV pro-
Int. J. Pharm. Investig. 1(2), 93–98 (2011) tection to cotton fabric. Int. J. Carbohydr. Chem. 2012, 1–6 (2012)
14. E.A. El-hefian, M.M. Nasef, A. Hamid, Preparation and character- 35. E.A. El-Hefian, M.M. Nasef, Preparation and characterization of
ization of Chitosan/Agar blended films: part 1. chemical structure Chitosan/Agar blends: rheological and thermal studies. J. Chilean
and morphology. E-Journal Chem. 9(3), 1431–1439 (2012) Chem. Soc. 55, 130–136 (2010)
15. Z. Hu, P. Hong, M. Liao, S. Kong, N. Huang, C. Ou, Prepara- 36. G.V.R. Murugan, T.M.G. Ravi, Comparative study of structural
tion and characterization of Chitosan–Agarose composite films. and magnetic properties of transition metal (Co, Ni) doped ZnO
Materials. 9, 1–9 (2016) nanoparticles. J. Mater. Sci. Mater. Electron. 26, 7205–7213
16. S.I. Thakore, Role of biopolymers in green nanotechnology. Prod. (2015)
Appl. Biopolym. 125, 121–140 (2015) 37. M. Varghese, P. Periyat, Silver-doped zinc oxide as a nanofiller for
17. Q. Guo et al., Comparison of in situ and ex situ methods for syn- development of poly(vinyl alcohol)/poly(vinyl pyrrolidone) blend
thesis of two-photon polymerization polymer nanocomposites. nanocomposites. Adv. Polym. Technol. 37, 137–143 (2016)
Polymers 6, 2037–2050 (2014) 38. Z.J. Yu, M.R. Kumar, D.L. Sun, L.T. Wang, R.Y. Hong, Large
18. D. Feldman, Review polyblend nanocomposites. J. Macromol. scale production of hexagonal ZnO nanoparticles using PVP as a
Sci. A 52, 648–658 (2015) surfactant. Mater. Lett. 166, 284–287 (2016)
19. A. Sirelkhatim, S. Mahmud, A. Seeni, Review on zinc oxide nano- 39. S.B. Rana, A. Singh, N. Kaur, Structural and optoelectronic
particles: antibacterial activity and toxicity mechanism. Nano- characterization of prepared and Sb doped ZnO nanoparticles. J.
Micro, Lett. 7(3), 219–242 (2015) Mater. Sci. 24, 44–52 (2013)
20. G.K. Prashanth et al., In vitro antibacterial and cytotoxicity studies 40. R. Singh, S.J. Dhoble, Combustion synthesis of ­Eu2+ and ­Dy3+
of ZnO nanopowders prepared by combustion assisted facile green activated ­Sr3(VO4)2 phosphor for LEDs. Bull. Mater. Sci. 34(3),
synthesis. Karbala Int. J. Mod. Sci. 1, 67–77 (2015) 557–562 (2011)
21. Z. Emami-karvani, P. Chehrazi, Antibacterial activity of ZnO 41. S. Dey, B. Das, R. Voggu, A. Nag, D.D. Sarma, C.N.R. Rao,
nanoparticle on gram-positive and gram-negative bacteria anti- Interaction of CdSe and ZnO nanocrystals with electron-donor
bacterial activity of ZnO nanoparticle on gram-positive and gram- and -acceptor molecules. Chem. Phys. Lett. 556, 200–206 (2013)
negative bacteria. Afr. J. Microbiol. Res. 5(12), 1368–1373 (2011) 42. R. Bomila, S. Srinivasan, A. Venkatesan, B. Bharath, K. Perinbam,
22. S. Agnihotri, S. Mukherji, Antimicrobial Chitosan–PVA hydrogel Structural, optical and antibacterial activity studies of Ce-doped
as a nanoreactor and immobilizing matrix for silver nanoparticles. ZnO nanoparticles prepared by wet-chemical method. Mater. Res.
Appl. Nanosci. 2, 179–188 (2012) Innov. (2017). https​://doi.org/10.1080/14328​917.2017.13243​79
23. E.A. El-Hefian, M.M. Nasef, Preparation and characterization of 43. A.M. Youssef, I.E. El-nagar, E. Amm, A.E. Aa, A.M. Youssef,
Chitosan/Agar blends: rheological and thermal studies. J. Chilean D.P. Chemica, Development and characterization of CMC/PVA
Chem. Soc. (2016) https​://doi.org/10.4067/S0717​-97072​01000​ films loaded with zno-nanoparticles for antimicrobial packaging
01000​31 application. Der Pharm. Chem. 9(9), 157–163 (2017)
24. M.H. Najar, K. Majid, Synthesis and characterization of nanocom- 44. K. Vidhya, M. Saravanan, V.P. Devarajan, S. Satheeskumar, M.A.
posite of polythiophene with ­Na2[Fe(CN)3(OH)(NO)C6H12N4] Ali, Green synthesis and anti-cancer activity of multifunctional
­H2O: a potent material for EMI shielding applications. J. Mater. ZnO: Mn-natural biomolecule quantum dots system. Malaya J.
Sci. Mater. Electron. 26, 6458–6470 (2013) Biosci. 1(4), 267–272 (2014)
25. P. Bindu, S. Thomas, Estimation of lattice strain in ZnO nanoparti- 45. B. Balraj et al., Synthesis and characterization of zinc oxide nano-
cles: X-ray peak profile analysis. J. Theor. Appl. Phys. 8, 123–134 particles using marine Streptomyces sp. with its investigation on
(2014) anticancer and antibacterial activity. Res. Chem. Intermed. 43,
26. R.M. Hodge, G.H. Edward, G.P. Simon, Water absorption and 2367–2376 (2016)
states of water in semicrystalline poly(vinyl alcohol) films. Poly- 46. D. Mahendiran et al., Biosynthesis of zinc oxide nanoparticles
mer 37, 1371 (1996) using plant extracts of aloe vera and Hibiscus sabdariffa: phyto-
27. R. Javed, M. Usman, S. Tabassum, M. Zia, Effect of capping chemical, antibacterial, antioxidant and anti-proliferative studies.
agents: structural, optical and biological properties of ZnO nano- BioNanoScience. 7, 530–545 (2017)
particles. Appl. Surf. Sci. 386, 319–326 (2016)
28. T. Dayakar, K.V. Rao, K. Bikshalu, V. Rajendar, S.-H. Park, Novel
synthesis and structural analysis of zinc oxide nanoparticles for

13

You might also like