14.07.2018 Views

Agar-lignin-silver

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

See discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/316772412<br />

Preparation and characterization of<br />

agar/<strong>lignin</strong>/<strong>silver</strong> nanoparticles composite<br />

films with ultraviolet lig....<br />

Article in Food Hydrocolloids · October 2017<br />

DOI: 10.1016/j.foodhyd.2017.05.002<br />

CITATIONS<br />

2<br />

READS<br />

184<br />

2 authors:<br />

Shiv Shankar<br />

46 PUBLICATIONS 558 CITATIONS<br />

SEE PROFILE<br />

Jong-Whan Rhim<br />

Mokpo National University<br />

168 PUBLICATIONS 5,951 CITATIONS<br />

SEE PROFILE<br />

Some of the authors of this publication are also working on these related projects:<br />

Presently, I am performing a project on "the devlopment of functional bionanocomposite packaging<br />

films and their applications" View project<br />

Biodegradable multifunctional nanocomposite films View project<br />

All content following this page was uploaded by Jong-Whan Rhim on 09 October 2017.<br />

The user has requested enhancement of the downloaded file.


Food Hydrocolloids 71 (2017) 76e84<br />

Contents lists available at ScienceDirect<br />

Food Hydrocolloids<br />

journal homepage: www.elsevier.com/locate/foodhyd<br />

Preparation and characterization of agar/<strong>lignin</strong>/<strong>silver</strong> nanoparticles<br />

composite films with ultraviolet light barrier and antibacterial<br />

properties<br />

Shiv Shankar a , Jong-Whan Rhim b, *<br />

a Department of Food Engineering and Bionanocomposite Research Institute, Mokpo National University, 61 Dorimri, Chungkyemyon, Muangun, 534-729<br />

Jeonnam, Republic of Korea<br />

b Center for Humanities and Sciences, and Department of Food and Nutrition, Kyung Hee University, 26 Kyungheedae-ro, Dongdaemun-gu, Seoul 120-701,<br />

Republic of Korea<br />

article<br />

info<br />

abstract<br />

Article history:<br />

Received 3 April 2017<br />

Received in revised form<br />

2 May 2017<br />

Accepted 2 May 2017<br />

Available online 3 May 2017<br />

Keywords:<br />

<strong>Agar</strong><br />

Lignin<br />

Silver nanoparticles<br />

Composite film<br />

Antibacterial activity<br />

Lignin was used to prepare <strong>silver</strong> nanoparticles (AgNPs), and they were incorporated into agar-based<br />

films. The composite films were characterized using UVevisible spectroscopy, FE-SEM, FTIR, XRD, and<br />

TGA. The color, mechanical, water vapor barrier, and antibacterial properties of the composite films were<br />

also evaluated. The composite films exhibited characteristic light absorption peaks at 250e300 nm and<br />

430 nm due to the <strong>lignin</strong> and AgNPs, respectively. The interaction of <strong>lignin</strong> and AgNPs with agar<br />

biopolymer was analyzed using FTIR. XRD results showed the characteristic peaks of crystalline AgNPs.<br />

Incorporation of <strong>lignin</strong> and AgNPs increased the mechanical, UV-light barrier, and water vapor barrier<br />

properties of the composite films. The loading of 1 wt% of AgNPs was the optimum concentration to<br />

improve the mechanical and water vapor barrier properties of the films. The films containing AgNPs<br />

exhibited antibacterial activity against food-borne pathogenic bacteria, Escherichia coli and Listeria<br />

monocytogenes.<br />

© 2017 Elsevier Ltd. All rights reserved.<br />

1. Introduction<br />

There have been serious concerns about the contamination of<br />

food by micro-organisms which can reduce the shelf-life of food as<br />

well as increase the risks of various food-borne infections causing<br />

serious illness to human beings (Devlieghere, Vermeiren, &<br />

Debevere, 2004). Traditionally, various physical and chemical<br />

preservation methods have been used in the food industry to<br />

reduce the spoilage of food, maintain food quality, and extend the<br />

shelf-life of food products. The most common method to control<br />

the growth of micro-organisms on the surface of the food is to<br />

apply antimicrobial agents or to dip the food materials into an<br />

antimicrobial solution (Sung et al., 2013). However, concerns on the<br />

diffusion of such antimicrobial agents to the inside of foodstuffs<br />

have limited the use of this technology. Also, recently increased<br />

consumer demands for the minimally processed and ready-to-eat<br />

fresh foods have inspired researchers to develop new<br />

* Corresponding author.<br />

E-mail address: jwrhim@khu.ac.kr (J.-W. Rhim).<br />

technologies for securing food safety and maintaining food quality<br />

(Bhat, 2013). Packaging materials with antimicrobial function have<br />

been recognized as one of the most promising active packaging<br />

systems for extending shelf-life of food, maintaining food safety<br />

and quality, and improving storage stability by destroying or<br />

inhibiting the spoilage and pathogenic microorganisms (Falguera,<br />

Quintero, Jimenez, Mu~noz, & Ibarz, 2011; Han, 2000). Antimicrobial<br />

packaging is often achieved by incorporation of antimicrobial<br />

agents into the packaging system. Some organic and inorganic<br />

materials have been used as antimicrobial agents for the antimicrobial<br />

packaging application (Kanmani & Rhim, 2014; Shankar,<br />

Teng, Li, & Rhim, 2015; Shankar, Teng, & Rhim, 2014). However, a<br />

low thermal stability of organic antimicrobial materials have<br />

limited their broad use in food packaging applications. In contrast,<br />

the high thermostability of inorganic antimicrobial agents such as<br />

metallic nanoparticles has opened the way for their use in the food<br />

packaging industries (Llorens, Lloret, Picouet, Trbojevich, &<br />

Fernandez, 2012).<br />

Among the metallic nanoparticles, <strong>silver</strong> nanoparticles (AgNPs)<br />

have been most widely used for the preparation of nanocomposite<br />

in the food packaging and biomedical applications due to their high<br />

http://dx.doi.org/10.1016/j.foodhyd.2017.05.002<br />

0268-005X/© 2017 Elsevier Ltd. All rights reserved.


S. Shankar, J.-W. Rhim / Food Hydrocolloids 71 (2017) 76e84 77<br />

surface area, unique optical, magnetic, electric, and catalytic<br />

properties with high thermal stability and broad-spectrum of<br />

antimicrobial activity (Duncan, 2011; Rhim, Park, & Ha, 2013). The<br />

antimicrobial effect of AgNPs is known to be greatly influenced by<br />

the size, shape, and method of preparation of the AgNPs (Lischer<br />

et al., 2011). The AgNPs synthesized by green methods are recognized<br />

as safe for the use in the food packaging applications. For the<br />

green synthesis of AgNPs, various biopolymers and biomolecules<br />

such as gelatin (Kanmani & Rhim, 2014), banana powder<br />

(Orsuwana, Shankar, Wang, Sothornvit, & Rhim, 2016), chitosan<br />

(Huang & Yang, 2004), vitamins (Shankar & Rhim, 2016), aminoacids<br />

(Shankar & Rhim, 2015), and plant extracts (Shankar,<br />

Chorachoo, Jaiswal, & Voravuthikunchai, 2014) have been used.<br />

As one of such biopolymeric materials with high potential for the<br />

preparation of AgNPs, <strong>lignin</strong> is interesting since it can be used not<br />

only as reducing and stabilizing agent for the AgNPs, but also as a<br />

filler for the preparation of nanocomposite films.<br />

Lignin is a complex phenolic compound which is abundantly<br />

present in the cell wall of plants in association with cellulose and<br />

hemicellulose (Tuomela, Vikman, Hatakka, & Itavaara, 2000).<br />

Although <strong>lignin</strong> is chemically heterogeneous and structurally<br />

complex material, it possesses multiple functional groups such as<br />

reductive aliphatic hydroxyls, phenolic hydroxyls, and thiols,<br />

which can serve as reducing and stabilizing agents for the synthesis<br />

of <strong>silver</strong> nanoparticles. Thiols have shown high affinity to<br />

<strong>silver</strong> nanoparticles, hence facilitating its adsorption and capping<br />

onto particle surface (Laibinis et al., 1991). The polar sulfonate<br />

groups in <strong>lignin</strong> can also help to disperse the formed nanoparticles<br />

in aqueous solutions. Also, <strong>lignin</strong> possesses a strong UV<br />

light absorption property (Monties, 1991; Shankar, Reddy, &<br />

Rhim, 2015).<br />

Recently, biopolymers from various natural resources have been<br />

used as eco-friendly packaging materials to substitute the nonbiodegradable,<br />

petroleum-based plastic based packaging materials<br />

(Gimenez, Lopez de Lacey, Perez-Santín, Lopez-Caballero, &<br />

Montero, 2013). Among biopolymers, agar has been widely used for<br />

the preparation of biodegradable packaging films due to its good<br />

film forming property with abundance, renewability, and biocompatibility<br />

(Shankar & Rhim, 2016; Shankar, Reddy et al., 2015;<br />

Shankar, Teng et al., 2014). <strong>Agar</strong> is a hydrophilic polysaccharide<br />

extracted from the Gelidiaceae and Gracilariaceae families of seaweeds<br />

and mainly composed of alternating repeating units of D-<br />

galactose and 3, 6-anhydro-ß-galactopyranose (Tako, Higa,<br />

Medoruma, & Nakasone, 1999).<br />

Therefore, the present study was aimed to develop agar-based<br />

biodegradable food packaging films incorporated with <strong>lignin</strong> and<br />

<strong>lignin</strong> capped <strong>silver</strong> nanoparticles. The films were characterized<br />

using various analytical techniques. The UV-light barrier, mechanical,<br />

thermal, water vapor barrier and antibacterial properties of the<br />

films were also evaluated.<br />

2. Materials and methods<br />

2.1. Materials<br />

The food grade agar was obtained from Fine <strong>Agar</strong> Co., Ltd.<br />

(Damyang, Jeonnam, Korea). Glycerol and alkali-soluble low sulfonate<br />

<strong>lignin</strong> were procured from Sigma-Aldrich Co. (St. Louis, MO,<br />

USA). The <strong>lignin</strong> was dried at 80 C for 6 h before use. Silver nitrate<br />

(AgNO 3 ), brain heart infusion broth (BHI), and tryptic soy broth<br />

(TSB) were obtained from Duksan Pure Chemicals Co., Ltd. (Ansan,<br />

Gyeonggi-do, Korea). Escherichia coli O157: H7 ATCC 43895 and<br />

Listeria monocytogenes ATCC 15313 were procured from the Korean<br />

Collection for Type Cultures (KCTC, Seoul, Korea).<br />

2.2. Synthesis of agar/<strong>lignin</strong>/AgNPs composite films<br />

Silver nanoparticles were synthesized using <strong>lignin</strong> as reducing<br />

and capping agents. 0.12 g of <strong>lignin</strong> (3 wt% based on agar) was<br />

dispersed in 150 mL of distilled water and stirred for 20 min at<br />

90 C using a magnetic stirrer. Different amounts of AgNO 3 (0.5, 1.0,<br />

1.5, and 2 wt% based on agar) were added dropwise into the <strong>lignin</strong><br />

solution and continued heating the mixture for 20 min. 1.2 g of<br />

glycerol (30 wt% based on agar) was added into the mixture as a<br />

plasticizer with vigorous stirring for 10 min followed by addition of<br />

4 g of agar and heated continuously with stirring at 90 C for 20 min<br />

using a hot plate. The fully solubilized film forming solution was<br />

cast evenly onto a leveled Teflon film-coated glass plate<br />

(24 cm 30 cm) and allowed to dry at room temperature for about<br />

48 h. The dried films were peeled off from the plate and conditioned<br />

in a humidity chamber set at 25 C and 50% RH for 48 h<br />

before further analysis. Additionally, neat agar and agar with 3 wt%<br />

of <strong>lignin</strong> films were prepared by the same procedure except adding<br />

AgNO 3 .<br />

2.3. Characterization of agar/<strong>lignin</strong>/AgNPs composite films<br />

2.3.1. Surface morphology and optical properties<br />

The surface morphology of the neat agar, agar/<strong>lignin</strong>, and agar/<br />

<strong>lignin</strong>/AgNPs composite films was observed using a field emission<br />

scanning electron microscope (FE-SEM, S-4800, Hitachi Co., Ltd.,<br />

Matsuda, Japan). The film samples were attached on the specimen<br />

holder, sputter coated with platinum, and the image was analyzed<br />

at an accelerating voltage of 3 kV.<br />

The optical properties of the composite films were determined<br />

by measuring the absorption of light between 200 and 700 nm<br />

using UVevisible spectrophotometer (Mecasys Optizen POP Series<br />

UV/Vis, Seoul, Korea).<br />

2.3.2. FTIR and XRD<br />

Fourier transform infrared (FTIR) spectra of the composite films<br />

were measured using an attenuated total Reflectance-Fourier<br />

Transform infrared (ATR-FTIR) spectrophotometer (TENSOR 37<br />

spectrophotometer with OPUS 6.0 software, Billerica, MA, USA).<br />

The FTIR spectra were recorded as 32 scans per samples at a resolution<br />

of 4 cm 1 .<br />

X-ray diffraction (XRD) pattern of the films was analyzed using<br />

an X-ray diffractometer (PANalytical X’ pert pro MRD diffractometer,<br />

Amsterdam, Netherlands). The film samples (2.5 cm 2.5 cm)<br />

were placed on a glass slide, and the XRD spectra were recorded<br />

using Cu Ka radiation (wavelength of 0.1541 nm) and a nickel<br />

monochromator filtering wave at 40 kV and 30 mA. The diffraction<br />

pattern was recorded at 2q ¼ 30e80 with a scanning speed of 0.4 /<br />

min at room temperature.<br />

2.3.3. Thermogravimetric analysis<br />

Thermal stability of the composite films was determined using a<br />

thermogravimetric analyzer (Hi-Res TGA 2950, TA Instrument, New<br />

Castle, DE, USA). The film samples were taken in a standard<br />

aluminum pan and heated from 30 to 600 C at the rate of 10 C/<br />

min under a nitrogen flow of 50 mL/min. A derivative form of TGA<br />

(DTG) was obtained and the maximum decomposition temperature<br />

(T max ) of the composite films was obtained from DTG curve.<br />

2.3.4. Surface color and light transmittance<br />

The Hunter color values (L, a, and b) of the composite films were<br />

measured using a Chroma meter (Minolta, CR-200, Tokyo, Japan). A<br />

white color standard plate (L ¼ 97.75, a ¼ - 0.49, and b ¼ 1.96) was<br />

used as a background for the color measurements. The total color<br />

difference (DE) was calculated as follows:


78<br />

S. Shankar, J.-W. Rhim / Food Hydrocolloids 71 (2017) 76e84<br />

qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi<br />

DE ¼ ðDLÞ 2 þðDaÞ 2 þðDbÞ 2<br />

where DL, Da, and Db are differences between each color value of<br />

the standard color plate and film specimen. Five measurements<br />

were taken for each film, and the average values are reported.<br />

The UV-barrier property and transparency of the films was<br />

determined by measuring percent transmittance of light at 280 and<br />

660 nm, respectively.<br />

2.3.5. Thickness and mechanical properties of the composite films<br />

The films were cut into 2.54 cm 15 cm strips using a precision<br />

double blade cutter (model LB.02/A, Metrotec, S.A., San Sebastian,<br />

Spain) and the thickness of films was measured using a digital<br />

micrometer (Digimatic Micrometer, QuantuMike IP 65, Mitutoyo,<br />

Japan) with an accuracy of 0.001 mm. Five random locations around<br />

each film sample were measured, and the average thickness was<br />

used for the calculation of tensile strength.<br />

The mechanical properties of the composite films in terms of<br />

tensile strength (TS), elongation at break (EAB), and elastic modulus<br />

(EM) were determined using Instron Universal Testing Machine<br />

(Model 5565, Instron Engineering Corporation, Canton, MA, USA) in<br />

a tensile mode with an initial grip separation and a cross-head<br />

speed set at 50 mm and 50 mm/min, respectively (Shankar &<br />

Rhim, 2016). The TS was calculated in MPa by dividing the<br />

maximum load (N) by the initial cross-sectional area (m 2 ) of the<br />

film sample. The EAB was calculated in % by dividing the extension<br />

at the rupture of the film by the initial length of the film (50 mm)<br />

and multiplied by 100. The EM was determined in GPa from the<br />

slope of a linear portion of the stress-strain curve.<br />

2.3.6. Water vapor permeability and water contact angle of<br />

composite films<br />

The water vapor permeability (WVP) of films was determined by<br />

the standard method of ASTM E96-95 with modification<br />

(Gennadios, Weller, & Goodings, 1994). The WVP measuring cup<br />

was filled with 18 mL of distilled water and placed film sample<br />

(7.5 cm 7.5 cm) on the top of the cup and sealed tightly to prevent<br />

the leakage of water vapor. The assembled WVP cup was weighed<br />

and subsequently placed in a controlled environmental chamber<br />

set at 25 C and 50% RH. Weight change of the cup was determined<br />

every 1 h for 8 h. The water vapor transmission rate (WVTR; g/m 2 s)<br />

of the film was calculated by using the slope of the steady-state<br />

(linear) portion of weight loss versus time plot. Then, the WVP<br />

(g.m/m 2 .s.Pa) of the film was calculated as follows:<br />

ðWVTR LÞ<br />

WVP ¼<br />

Dp<br />

where L was the mean thickness of the film (m) and Dp was water<br />

vapor partial pressure difference (Pa) across the film, which was<br />

calculated by the method of Gennadios et al. (1994).<br />

Surface hydrophobicity of the film was evaluated by measuring<br />

the water contact angle (WCA) of the film using a water contact<br />

angle analyzer (model Phoenix 150, Surface Electro-Optics Co., Ltd.,<br />

Kunpo, Korea). The film was cut into 3 cm 10 cm strips and placed<br />

on the horizontal movable stage (black Teflon coated steel,<br />

7cm 11 cm) fitted with the WCA analyzer. About 10 mL of water<br />

was dropped on the surface of the film using a microsyringe, and<br />

the contact angle on both sides of the water droplet was measured.<br />

Five measurements were taken for each sample, and average values<br />

were presented.<br />

2.3.7. Water solubility, and swelling ratio of composite films<br />

The water solubility (WS) of the film samples was determined as<br />

the percentage of dissolved dry matter after immersion in water.<br />

Three randomly selected specimens of each type of film<br />

(3 cm 3 cm) were first dried at 60 C for 24 h to determine the<br />

initial dry matter (W 1 ). Each film was immersed into 30 mL of<br />

distilled water in a 50 mL beaker with a gentle stirring for 24 h. The<br />

film samples were removed after 24 h and dried in a drying oven at<br />

105 C for 24 h to determine the undissolved final dry weight (W 2 ).<br />

The WS of the sample was calculated as follows:<br />

WS ¼ W 1 W 2<br />

W 1<br />

100<br />

The swelling ratio (SR) of the films was determined gravimetrically.<br />

Pre-weighed specimens (2.5 cm 5 cm) were immersed in<br />

distilled water for 1 h, and then the film samples were removed<br />

from the water and weighed after removing the surface water with<br />

blotting paper. The percent SR of the films was calculated as<br />

follows:<br />

SRð%Þ ¼ W 2 W 1<br />

W 1<br />

100<br />

where W 1 and W 2 are the weight of the film samples before and<br />

after soaking in the water, respectively.<br />

2.3.8. Antibacterial activity<br />

The antibacterial activity of composite films was examined<br />

against food-borne pathogenic Gram-positive bacteria,<br />

L. monocytogenes and Gram-negative bacteria, E. coli (Shankar &<br />

Rhim, 2016). The test bacteria were aseptically inoculated in the<br />

TSB and BHI broth, respectively, and subsequently incubated at<br />

37 C for 16 h. The inoculum was diluted, and 200 mL of diluted<br />

inoculum (10 8 e10 9 CFU/mL) was aseptically transferred to 50 mL of<br />

TSB and BHI broth containing 100 mg film samples and incubated at<br />

37 C for 15 h with mild shaking. The samples were taken out at 3, 6,<br />

9, 12, and 15 h of incubation, and the viable cell count was determined<br />

by diluting and plating the samples on agar plates. Broth<br />

without film was used as a positive control. Antimicrobial tests<br />

were performed in triplicate with individually prepared films.<br />

2.3.9. Statistical analysis<br />

Measurement of each property of the films was performed in<br />

triplicate with individually prepared film samples as the replicated<br />

experimental units, and the results were presented as mean<br />

values ± standard deviations (SD). One-way analysis of variance<br />

(ANOVA) was performed, and the significance of each mean property<br />

value was determined (p < 0.05) with Duncan’s multiple range<br />

tests using the SPSS statistical analysis computer program for<br />

Windows (SPSS Inc., Chicago, IL, USA).<br />

3. Results and discussion<br />

3.1. Preparation of AgNPs<br />

The <strong>lignin</strong> solution was yellow, but it changed to dark brown<br />

after the addition of <strong>silver</strong> nitrate solution. The change in color of<br />

the solution indicated the formation of AgNPs, which was formed<br />

through the reduction of Ag ions by the <strong>lignin</strong>. The solution<br />

exhibited the maximum absorption peak around 420 nm (data not<br />

shown), which is also evidence for the formation of AgNPs (Shankar<br />

& Rhim, 2015). The color intensity of the solution was dependent on<br />

the amount of <strong>silver</strong> nitrate used. Lignin possesses multiple functional<br />

groups such as aliphatic hydroxyl, phenolic hydroxyls, and


S. Shankar, J.-W. Rhim / Food Hydrocolloids 71 (2017) 76e84 79<br />

thiols, which has potential as reducing and capping agents during<br />

<strong>silver</strong> nanoparticle synthesis (Hu & Hsieh, 2016). A thiol group of<br />

<strong>lignin</strong> has a strong affinity to <strong>silver</strong> nanoparticles. Hence it facilitated<br />

to adsorb AgNPs and to stabilize AgNPs (Laibinis et al., 1991).<br />

The polar sulfonate groups of <strong>lignin</strong> can also help to disperse the<br />

formed nanoparticles in aqueous solutions.<br />

3.2. Characterization of agar/<strong>lignin</strong>/AgNPs composite films<br />

3.2.1. Morphology and optical properties<br />

All the films prepared were free standing, intact, and smoothsurfaced<br />

without any apparent defects. The microstructure of the<br />

composite films at the surface and cross-section was observed using<br />

FE-SEM as shown in Fig. 1. As expected, the surface of the neat<br />

agar and agar/<strong>lignin</strong> composite films was smooth and compact.<br />

However, the surface of composite films incorporated with AgNPs<br />

was slightly rough compared with the control films. As shown in<br />

both the surface and the cross-sectional views, the AgNPs were<br />

uniformly distributed through the agar/<strong>lignin</strong>/AgNPs composite<br />

films. The size of AgNPs increased with increase in the concentration<br />

of the precursor of AgNPs, AgNO 3 . This result indicated that the<br />

size of the AgNPs was greatly influenced by the concentration of<br />

<strong>silver</strong> precursor.<br />

Light absorption characteristics of the films were evaluated by<br />

measuring light absorption spectra in the wavelength of<br />

200e700 nm as shown in Fig. 2a. The neat agar film did not show<br />

any light absorption peak. On the other hand, <strong>lignin</strong> incorporated<br />

films exhibited broad maximum peaks in the range of 250e300 nm,<br />

which was attributed to the chromophore groups in the <strong>lignin</strong><br />

(Chaochanchaikul, Jayaraman, Rosarpitak, & Sombatsompop,<br />

2012). In addition to the peak of <strong>lignin</strong>, the agar/<strong>lignin</strong>/AgNPs<br />

composite films exhibited the light absorption peak at around<br />

430 nm, which was due to the surface Plasmon resonance of AgNPs<br />

in the composite films (Shankar, Chorachoo et al., 2014; Shankar,<br />

Jaiswal, Selvakannan, Ham, & Rhim, 2016).<br />

3.2.2. FTIR and XRD<br />

The FTIR spectra of the composite films were shown in Fig. 2b.<br />

The neat agar film showed broad bands at 3330 cm 1 and 2930,<br />

which were due to the stretching of hydroxyl (OeH) group and<br />

stretching of CeH that was associated with the ring of methane<br />

hydrogen atoms, respectively (Tako et al., 1999). The peak at<br />

1640 cm 1 was attributed to the stretching vibration of the peptide<br />

bond formed by an amine (NH) and acetone groups in agar (Kumar,<br />

Negi, Choudhary, & Bhardwaj, 2014). The peak at 1372 cm 1 was<br />

attributed to the ester sulfate group of agar (Volery, Besson, &<br />

Schaffer-lequart, 2004). The bands at 1040 and 930 cm 1 were<br />

associated with the CO stretching group of 3,6-anhydrogalactose<br />

(Shankar, Teng et al., 2014). The peak around 890 cm 1 was due<br />

to the CeH residual carbon of b-galactose. The <strong>lignin</strong> exhibit<br />

characteristic peak around 1515 cm 1 , however, there was no such<br />

peak observed when <strong>lignin</strong> was composite with agar (Oun & Rhim,<br />

2016). It was presumably because the peak of <strong>lignin</strong> was masked by<br />

the peak of agar between 1500 and 1650 cm 1 . All the FTIR spectra<br />

of composite films were similar to those of agar film, suggesting<br />

that there were no chemical bonds formed among agar, <strong>lignin</strong>, and<br />

AgNPs.<br />

The crystalline nature of <strong>lignin</strong> and AgNPs in the composite<br />

films was tested using the X-ray diffraction (XRD) analysis. The<br />

results of XRD of the agar, agar/<strong>lignin</strong>, and agar/<strong>lignin</strong>/AgNP 1.0 films<br />

were shown in Fig. 2c. The XRD patterns of agar/AgNPs composite<br />

films exhibited distinctive diffraction peaks at 2q values around 32,<br />

48, and 68 corresponding to the (111), (200), and (220) lattice<br />

planes of AgNPs in the composite films, which were attributed to<br />

the formation of crystalline AgNPs (Martínez-Casta~non, Ni~no-<br />

Martinez, Martínez-Gutierrez, Martínez-Mendoza, & Ruiz, 2008).<br />

However, the neat agar and agar/<strong>lignin</strong> composite film did not<br />

show any diffraction peak in the 2q range of 30e80 .<br />

3.2.3. Thermogravimetric analysis<br />

The TGA and DTG thermograms of the composite films were<br />

shown in Fig. 3. TGA curves showed the weight loss pattern of the<br />

films and DTG curves clearly showed the maximum decomposition<br />

temperature at each step of thermal decomposition. All the films<br />

showed multi-step thermal degradation. The first step of thermal<br />

degradation occurred at 80e120 C, which was attributed to the<br />

evaporation of moisture in the composite films (Shankar, Teng et al.,<br />

2014). The subsequent thermal degradation patterns were varied<br />

depending on the type of films. The main steps of thermal degradation<br />

were observed between 210 and 325 C which was due to<br />

the decomposition of glycerol (a plasticizer) and polymers (agar<br />

and <strong>lignin</strong>) (El-Hefian, Nasef, & Yahaya, 2012; Martucci &<br />

Ruseckaite, 2010). <strong>Agar</strong>/<strong>lignin</strong>/AgNPs composite films exhibited<br />

higher degradation temperature compared with the neat agar and<br />

agar/<strong>lignin</strong> composite films. The residual char content at 600 Cof<br />

the neat agar film was 10.5%, and it increased to 24.8, 25.6, and<br />

24.3% in the agar/<strong>lignin</strong>, agar/<strong>lignin</strong>/AgNP 1.0 , and agar/<strong>lignin</strong>/<br />

AgNP 2.0 composite films, respectively. The increased thermostability<br />

of the composite films was mainly due to the thermostable<br />

AgNPs.<br />

Fig. 1. FE-SEM micrographs of surface and cross-section of the composite films.<br />

3.2.4. Apparent color and light transmittance<br />

The apparent color and light transmittance of the films were<br />

presented in Table 1. The neat agar film was transparent without<br />

any color tint, but it became yellow after addition of <strong>lignin</strong>. The


80<br />

S. Shankar, J.-W. Rhim / Food Hydrocolloids 71 (2017) 76e84<br />

(a)<br />

(b)<br />

(c)<br />

Fig. 2. (a) UVevisible, (b) FTIR, and (c) XRD spectra of the composite films.<br />

surface color of agar/<strong>lignin</strong>/AgNPs composite films exhibited dark<br />

brown. The intensity of color from yellow to brown increased with<br />

the increase in the concentration of <strong>silver</strong> nitrate. The Hunter L-<br />

value (lightness) decreased significantly (p < 0.05) with increasing<br />

AgNPs content. However, Hunter-a and b and total color difference<br />

(DE) values of the composite films were higher than those of the<br />

neat agar film. This result indicates that the lightness of the film<br />

decreased with increased redness and yellowness after blending<br />

with <strong>lignin</strong> and AgNPs. Similar results of decreased L values and<br />

increased a and b values were found when <strong>lignin</strong> (Shankar, Reddy<br />

et al., 2015) and AgNPs (Shankar & Rhim, 2015) were incorporated<br />

into the agar films.<br />

The light transmission characteristics of the composite films<br />

were measured using a UV/Vis spectrophotometer at 280 and<br />

660 nm (Table 1). The T 660 (a measure of transparency) of the neat<br />

agar film were 88.8 ± 0.5% indicating the high transparency of the<br />

film. The T 660 decreased slightly to 85.5 ± 1.3 after incorporation of<br />

<strong>lignin</strong> and decreased significantly after composite formation with<br />

AgNPs. On the other hand, the light transmittance at 280 nm (T 280, a<br />

measure of UV-barrier) decreased significantly (p < 0.05) after<br />

incorporation of <strong>lignin</strong> and AgNPs. The nanocomposite film incorporated<br />

with <strong>lignin</strong> and AgNPs prevented UV-light passage almost<br />

completely. Since <strong>lignin</strong> is composed of a complex aromatic polymer<br />

of aromatic alcohols, it absorbs light in the ultraviolet region<br />

due to the presence of aromatic rings in its structure (Monties,<br />

1991). This result indicated that the composite film had a UVscreening<br />

property with a slight sacrifice of the transparency. The<br />

composite film has a high potential for the UV-screening food<br />

packaging film with the see-through property.<br />

3.2.5. Mechanical properties<br />

The mechanical properties of the films determined in term of<br />

the tensile strength (TS), elongation at break (EAB), and elastic<br />

modulus (EM) and the results were presented in Table 2. The<br />

thickness of the biopolymer films increased after incorporation of<br />

<strong>lignin</strong> and AgNPs. The increase in the thickness of the composite<br />

films was mainly due to the increased solid content by the <strong>lignin</strong><br />

and AgNPs (Shankar, Teng et al., 2014). The mechanical properties<br />

were greatly influenced by the addition of <strong>lignin</strong> and AgNPs. The TS<br />

of neat agar film was 40.3 ± 4.0 MPa, which was increased to<br />

44.1 ± 3.6 MPa after blending with <strong>lignin</strong> and it increased further<br />

after blending with AgNPs. The TS of agar/<strong>lignin</strong>/AgNPs composite<br />

films was also influenced by the concentration of AgNO 3 . The TS of<br />

the composite films increased up to 1 wt% of AgNO 3 , then<br />

decreased when a higher concentration of AgNO 3 was used. <strong>Agar</strong>/<br />

<strong>lignin</strong>/AgNP 1.0 exhibited the highest TS (49.7 ± 3.3 MPa) among the<br />

composite films. The increase in the TS of the composite films<br />

might be due to a high compatibility between the nanofiller and<br />

agar biopolymer. As shown in the SEM images (Fig. 1), the <strong>lignin</strong><br />

and <strong>silver</strong> nanoparticles were dispersed well in the agar<br />

biopolymer, indicating they were highly compatible. Since the<br />

<strong>lignin</strong> served as a reducing as well as capping agent for the <strong>silver</strong><br />

nanoparticles, the interaction between AgNPs stabilized by the<br />

<strong>lignin</strong> and agar polymer matrix was presumably increased to form<br />

strong composite films. The effectiveness of the reinforcement and<br />

dispersion of nanofillers in the polymer matrix functions controlling<br />

role for the effective stress transfer at the interface of the<br />

matrix and the filler and ultimately increase the tensile strength of<br />

the polymeric biocomposite materials (Ahmed, Arfat, Castro-


S. Shankar, J.-W. Rhim / Food Hydrocolloids 71 (2017) 76e84 81<br />

(a)<br />

Table 2<br />

Tensile properties of agar, agar/<strong>lignin</strong>, and agar/<strong>lignin</strong>/AgNPs composite films. 1)<br />

Film Thickness (mm) TS (MPa) EAB (%) EM (GPa)<br />

<strong>Agar</strong> 45.1 ± 2.3 a 40.3 ± 4.0 a 19.4 ± 4.2 a 1.39 ± 0.23 a<br />

<strong>Agar</strong>/<strong>lignin</strong> 55.9 ± 2.6 a 44.1 ± 3.6 b 16.1 ± 4.3 a 1.48 ± 0.13 ab<br />

<strong>Agar</strong>/<strong>lignin</strong>/AgNP 0.5 54.9 ± 7.4 b 45.5 ± 4.6 bc 16.2 ± 5.1 a 1.49 ± 0.15 ab<br />

<strong>Agar</strong>/<strong>lignin</strong>/AgNP 1.0 54.3 ± 2.9 a 49.7 ± 3.3 d 16.7 ± 4.7 a 1.60 ± 0.08 c<br />

<strong>Agar</strong>/<strong>lignin</strong>/AgNP 1.5 57.2 ± 1.9 a 47.0 ± 3.5 c 16.2 ± 3.4 a 1.52 ± 0.23 ab<br />

<strong>Agar</strong>/<strong>lignin</strong>/AgNP 2.0 53.0 ± 1.4 a 45.6 ± 2.2 bc 16.7 ± 2.9 a 1.50 ± 0.13 ab<br />

1) Each value is the mean of three replicates with the standard deviation. Any two<br />

means in the same column followed by the same letter are not significantly<br />

(p > 0.05) different by Duncan’s multiple range test.<br />

the EAB of composite films was slightly decreased after incorporation<br />

of <strong>lignin</strong> and AgNPs, though the decrease was not significantly<br />

different (p > 0.05). The results showed that the strength and<br />

stiffness of the composite films increased without sacrificing the<br />

flexibility of the films.<br />

(b)<br />

Fig. 3. (a) TGA and (b) DTG of the composite films.<br />

Aguirre, & Auras, 2016; Díez-Pascual & Díez-Vicente, 2014).<br />

Shankar, Reddy et al. (2015) also reported that the TS of agar film<br />

increased with increase in the concentration of <strong>lignin</strong> up to 3 wt%.<br />

Chen et al. (2009) also reported that the addition of <strong>lignin</strong> into<br />

chitosan increased the tensile strength of the film. The effect of<br />

<strong>lignin</strong> and AgNPs on the EM of the film was similar to the TS. The<br />

EM of agar film increased after <strong>lignin</strong> incorporation, and it<br />

increased further with the incorporation of AgNPs up to 1 wt%, then<br />

started to decrease with a higher concentration of AgNPs. The<br />

decreased mechanical strength of the composite films above 1 wt %<br />

of AgNPs might be due to aggregation of AgNPs. On the contrary,<br />

3.2.6. Water vapor permeability (WVP) and water contact angle<br />

(WCA)<br />

The WVP of composite films were presented in Table 3. The WVP<br />

of the neat agar film was 1.99 ± 0.26 10 9 g m/m 2 .Pa.s, which was<br />

decreased to 1.85 ± 0.13 10 9 g m/m 2 .Pa.s after incorporation of<br />

<strong>lignin</strong>. The WVP of the film decreased further when AgNPs were<br />

incorporated, however, the degree of decrease was dependent on<br />

the content of AgNPs. The WVP of the composite films decreased<br />

down to 1.56 ± 0.11 10 9 g m/m 2 .Pa.s when 1 wt% of AgNPs was<br />

incorporated, then increased with further increase in the AgNPs.<br />

The decreased WVP of the nanocomposite films was mainly due to<br />

the increased tortuous path for water vapor diffusion by the<br />

impermeable nanoparticles distributed in the polymer matrix (Yu,<br />

Yang, Liu, & Ma, 2009). Also, the compatibility between <strong>lignin</strong> and<br />

agar might help to form a strong intermolecular interaction and to<br />

reduce water vapor penetration through the composite films. The<br />

increase in the WVP of the composite films incorporated with more<br />

than 1 wt% was probably due to the aggregation of the nanoparticles<br />

at a higher concentration as observed in the SEM result.<br />

This result also indicates 1 wt% of AgNPs is optimum to increase<br />

water vapor barrier property of the composite film. It has<br />

frequently been observed that the incorporation of nanofillers<br />

decreased the WVP of the composite films within a certain level of<br />

concentration, but the effect was reversed when the concentration<br />

of nanofiller was higher than the optimum level (Shankar & Rhim,<br />

2016; Shankar, Reddy et al., 2015).<br />

The hydrophobicity and wettability of the films are usually<br />

evaluated by the water contact angle (WCA). The WCA of the test<br />

films were also shown in Table 3. The WCA of the neat agar film was<br />

54.3 ± 1.3 , it increased to 55.6 ± 0.7 and 59.4 ± 0.6 for the agar/<br />

<strong>lignin</strong> and agar/<strong>lignin</strong>/AgNP 1.0 composite films, respectively. The<br />

hydrophobicity of the composite film increased due to the hydrophobicity<br />

of the AgNPs. It was interesting to note that the WCA of<br />

the agar/<strong>lignin</strong>/AgNPs composite films also increased with the<br />

Table 1<br />

Apparent color and light transmittance of agar, agar/<strong>lignin</strong> and agar/<strong>lignin</strong>/AgNPs biocomposite films. 1)<br />

Film L a b DE T 660 (%) T 280 (%)<br />

<strong>Agar</strong> 94.81 ± 0.29 d 0.99 ± 0.02 a 3.67 ± 0.06 a 1.05 ± 0.09 a 88.8 ± 0.5 b 52.9 ± 1.7 b<br />

<strong>Agar</strong>/<strong>lignin</strong> 84.30 ± 1.74 c 0.98 ± 0.41 a 23.57 ± 2.08 b 23.08 ± 2.62 b 85.5 ± 1.3 b 0.1 ± 0.0 a<br />

<strong>Agar</strong>/<strong>lignin</strong>/AgNP 0.5 56.19 ± 5.16 b 5.51 ± 1.58 b 29.07 ± 3.18 d 47.05 ± 2.43 c 60.6 ± 7.7 a 0.1 ± 0.0 a<br />

<strong>Agar</strong>/<strong>lignin</strong>/AgNP 1.0 49.11 ± 2.13 a 17.81 ± 0.31 c 28.73 ± 1.75 cd 55.38 ± 0.97 d 65.8 ± 7.0 a 0.1 ± 0.0 a<br />

<strong>Agar</strong>/<strong>lignin</strong>/AgNP 1.5 46.05 ± 5.55 a 17.86 ± 0.31 c 26.11 ± 4.35 bcd 57.02 ± 2.81 de 66.6 ± 5.3 a 0.1 ± 0.0 a<br />

<strong>Agar</strong>/<strong>lignin</strong>/AgNP 2.0 43.77 ± 1.39 a 24.43 ± 3.35 d 24.08 ± 1.49 bc 60.42 ± 2.05 e 68.2 ± 0.9 a 0.1 ± 0.0 a<br />

1) Each value is the mean of three replicates with the standard deviation. Any two means in the same column followed by the same letter are not significantly (p > 0.05)<br />

different by Duncan’s multiple range test.


82<br />

S. Shankar, J.-W. Rhim / Food Hydrocolloids 71 (2017) 76e84<br />

Table 3<br />

Water vapor permeability (WVP), water contact angle (WCA), water solubility (WS), and swelling ratio (SR) of agar, agar/<strong>lignin</strong>, and agar/<strong>lignin</strong>/AgNPs composite films. 1)<br />

Film WVP (10 9 g m/m 2 Pa s) WCA (degree) WS (%) SR (%)<br />

<strong>Agar</strong> 1.99 ± 0.26 d 54.3 ± 1.3 a 27.3 ± 0.8 bc 2785.6 ± 343.2 c<br />

<strong>Agar</strong>/<strong>lignin</strong> 1.85 ± 0.13 c 55.6 ± 0.7 ab 31.2 ± 0.6 d 2587.4 ± 483.8 bc<br />

<strong>Agar</strong>/<strong>lignin</strong>/AgNP 0.5 1.71 ± 0.09 bc 58.3 ± 0.8 c 28.4 ± 0.9 c 2268.7 ± 453.9 bc<br />

<strong>Agar</strong>/<strong>lignin</strong>/AgNP 1.0 1.56 ± 0.11 a 59.4 ± 0.6 d 25.5 ± 0.8 ab 1873.8 ± 559.4 ab<br />

<strong>Agar</strong>/<strong>lignin</strong>/AgNP 1.5 1.62 ± 0.09 a 56.7 ± 1.2 b 24.0 ± 0.5 a 1428.1 ± 289.6 a<br />

<strong>Agar</strong>/<strong>lignin</strong>/AgNP 2.0 1.67 ± 0.07 ab 54.2 ± 0.8 a 23.6 ± 1.1 a 1398.7 ± 195.4 a<br />

1) Each value is the mean of three replicates with the standard deviation. Any two means in the same column followed by the same letter are not significantly (p > 0.05)<br />

different by Duncan’s multiple range test.<br />

concentration of AgNPs up to 1 wt%, then decreased. The decrease<br />

in WCA at higher concentration of AgNPs might be due to the<br />

increased surface roughness of the films (Orsuwana et al., 2016).<br />

3.2.7. Water solubility and swelling ratio<br />

The results of water solubility (WS) and swelling ratio (SR) of the<br />

composite films were also shown in Table 3. The agar/<strong>lignin</strong> composite<br />

film showed the highest WS among the tested films. The WS<br />

of the composite films increased after incorporation of <strong>lignin</strong>, but it<br />

decreased when AgNPs were incorporated. The WS of the composite<br />

films decreased linearly with the increasing concentration of<br />

AgNPs. The higher WS of agar/<strong>lignin</strong> composite films compared to<br />

the neat agar films might be due to the higher water solubility of<br />

<strong>lignin</strong> than agar. On the other hand, the decrease in WS of AgNPs<br />

incorporated films might be due to the presence of water-insoluble<br />

AgNPs in the composite films.<br />

The SR of the composite films was also greatly affected by the<br />

incorporation of the fillers. The neat agar film exhibited the highest<br />

SR (2785.6 ± 343.2%), which was reduced to 2587.4 ± 483.8% in<br />

agar/<strong>lignin</strong> composite films. The SR of AgNPs incorporated films<br />

was decreased linearly with the increase in the concentration of<br />

AgNPs.<br />

3.2.8. Antibacterial activity<br />

The antibacterial activity of agar, agar/<strong>lignin</strong>, and agar/<strong>lignin</strong>/<br />

AgNPs composite films was tested against food-bore pathogenic<br />

Gram-positive (L. monocytogenes) and Gram-negative (E. coli) bacteria<br />

(Fig. 4). The neat agar and agar/<strong>lignin</strong> films did not show any<br />

antibacterial activity against both Gram-positive and Gramnegative<br />

bacteria. Similar to the control, they showed exponential<br />

growth of both bacteria from an initial concentration of<br />

10 6 e10 7 CFU/mL to 10 10 CFU/mL within 12e15 h of incubation. On<br />

the contrary, the AgNPs incorporated composite films exhibited<br />

distinctive antimicrobial activity against both Gram-positive and<br />

Gram-negative bacteria, and their activity was dependent on the<br />

concentration of AgNPs. Here, we used two composite films, agar/<br />

<strong>lignin</strong>/AgNP 1.0 and agar/<strong>lignin</strong>/AgNP 2.0 to test the antibacterial activity<br />

of the composite films. The one was chosen as the optimum<br />

concentration of AgNPs based on the result of the mechanical and<br />

water vapor barrier properties. The other was chosen to test the<br />

effect of the concentration of AgNPs. Both composite films exhibited<br />

stronger antimicrobial activity against E. coli than<br />

L. monocytogenes. The composite films with 1 wt% AgNPs reduced<br />

the bacterial population of L. monocytogenes from 10 6 CFU/mL to<br />

10 2 CFU/mL after 15 h. However, it destroyed E. coli completely after<br />

9 h of incubation. The antimicrobial activity increased when the<br />

concentration of AgNPs increased, i.e., the composite films with<br />

2 wt% AgNPs destroyed L. monocytogenes and E. coli completely in<br />

12 h and 6 h, respectively. It is well known that the AgNPs have<br />

stronger antimicrobial activity against Gram-negative than Grampositive<br />

bacteria (Shankar & Rhim, 2015), which agrees well with<br />

the present results. The difference in the antimicrobial activity of<br />

Fig. 4. Antibacterial activity of composite films.<br />

AgNPs between Gram-positive and Gram-negative bacteria might<br />

be due to the difference in the cell wall structure. Gram-positive<br />

bacteria have thick peptidoglycan layer and complex cell wall<br />

structure that makes more difficult for AgNPs to penetrate into the<br />

cell (Priyadarshini, Gopinath, Priyadharsshini, Ali, & Velusamy,<br />

2013). On the contrary, Gram-negative bacteria have thin peptidoglycan<br />

layer coated with the negatively charged outer membrane,<br />

which makes the entry of AgNPs easier. Several probable<br />

mechanisms have been reported on the antimicrobial activity of


S. Shankar, J.-W. Rhim / Food Hydrocolloids 71 (2017) 76e84 83<br />

AgNPs, though a precise mechanism has not been established yet.<br />

The most widely known mechanism is the interaction between<br />

positively charged <strong>silver</strong> ions and negatively charged phosphorus<br />

or sulfur-containing biomolecules such as proteins and nucleic<br />

acids, which causes structural changes in the bacterial cell walls<br />

leading to disruption of metabolic processes and cell death (Butkus,<br />

Edling, & Labare, 2003). The other antibacterial mechanism of <strong>silver</strong><br />

nanoparticles is related to membrane damage due to free radicals<br />

derived from the surface of the AgNPs, increasing the membrane<br />

permeability and cell death (Kim et al., 2007). Above all, the antibacterial<br />

action of the composite films would be initiated by the<br />

contact of AgNPs, exposed to the surface of the film or free AgNPs or<br />

Ag ions released from the film, with the bacterial cells. As seen from<br />

the SR test results, it is evident that the water molecules diffused<br />

into the film will swell the film and facilitate the release of AgNPs<br />

into the liquid medium to be oxidized into <strong>silver</strong> ions.<br />

The composite films with strong antimicrobial activity have a<br />

high potential for the application in an antimicrobial active food<br />

packaging. However, there is a safety issue on AgNPs with possible<br />

migration from packaging film to food materials or the environment.<br />

Presently, very little information is available to assess the<br />

migration of nanoparticles from food packaging materials to the<br />

packaged food and possible hazard on human beings and environment<br />

(Rhim et al., 2013). However, it has been reported that the<br />

impact of AgNPs synthesized by the green methods is much lower<br />

than that synthesized by chemical or physical methods (Shankar,<br />

Prasad, Selvakannan, Jaiswal, & Laxman, 2015). Therefore, the<br />

nanocomposite films developed in the present study have the potential<br />

for their applications in food packaging to prolong the shelflife<br />

of packed food.<br />

4. Conclusion<br />

Lignin and AgNPs were used to prepare antimicrobial agarbased<br />

films with improved film properties. The <strong>lignin</strong> was used as<br />

reducing agent for the synthesis of AgNPs and a reinforcing filler for<br />

the preparation of the composite films. The incorporation of <strong>lignin</strong><br />

increased the tensile strength, water vapor barrier, and UV-light<br />

barrier properties. The inclusion of AgNPs further increased the<br />

tensile strength and water vapor barrier properties of the composite<br />

films and provided strong antibacterial property. The combined<br />

use of <strong>lignin</strong> and AgNPs enhanced the functional properties<br />

of the composite films. The incorporation of AgNPs up to 1 wt%<br />

increased the mechanical and water vapor barrier properties of the<br />

composite films. The composite films exhibited potent antibacterial<br />

activity against E. coli and L. monocytogenes. These results suggest<br />

that the agar/<strong>lignin</strong>/AgNPs nanocomposite films with UV-light<br />

barrier property, antimicrobial activity, and improved film properties<br />

have a high potential for their application in active food<br />

packaging to secure the food safety and to prolong the shelf-life of<br />

the packaged foods.<br />

Acknowledgements<br />

This research was supported by the Agricultural Research<br />

Council (ARC 710003) program of the Ministry of Agriculture, Food<br />

and Rural Affairs, Korea, and Korea Research Fellowship Program<br />

through the National Research Foundation of Korea (NRF) funded<br />

by the Ministry of Science, ICT and Future Planning<br />

(2016H1D3A1903910).<br />

References<br />

Ahmed, J., Arfat, Y. A., Castro-Aguirre, E., & Auras, R. (2016). Mechanical, structural<br />

and thermal properties of Ag-Cu and ZnO reinforced polylactide nanocomposite<br />

films. International Journal of Biological Macromolecules, 86, 885e892.<br />

Bhat, S. (2013). Minimal processing and preservation of fruits and vegetables by<br />

active packaging. International Journal of Herbal Medicine, 1, 131e138.<br />

Butkus, M. A., Edling, L., & Labare, M. P. (2003). The efficacy of <strong>silver</strong> as a bactericidal<br />

agent: Advantages, limitations, and considerations for future use. Journal of<br />

Water Supply: Research and Technology-AQUA, 52, 407e416.<br />

Chaochanchaikul, K., Jayaraman, K., Rosarpitak, V., & Sombatsompop, N. (2012).<br />

Influence of <strong>lignin</strong> content on photodegradation in wood/HDPE composites<br />

under UV weathering. Bioresources, 7, 38e55.<br />

Chen, L., Tang, C., Ning, N. Y., Wang, C. Y., Fu, Q., & Zhang, Q. (2009). Preparation and<br />

properties of chitosan/<strong>lignin</strong> composite films. Chinese Journal of Polymer Science,<br />

27, 739e746.<br />

Devlieghere, F., Vermeiren, L., & Debevere, J. (2004). New preservation technologies:<br />

Possibilities and limitations. International Dairy Journal, 14, 273e285.<br />

Díez-Pascual, A. M., & Díez-Vicente, A. L. (2014). ZnO-reinforced poly(3-<br />

hydroxybutyrate-co-3-hydroxyvalerate) bionanocomposites with antimicrobial<br />

function for food packaging. Applied Materials & Interfaces, 6, 9822e9834.<br />

Duncan, T. V. (2011). Application of nanotechnology in food packaging and food<br />

safety: Barrier materials, antimicrobials, and sensors. Journal of Colloid and<br />

Interface Science, 363, 1e24.<br />

El-Hefian, E. A., Nasef, M. M., & Yahaya, A. H. (2012). Preparation and characterization<br />

of chitosan/agar blended films: Part 2. Thermal, mechanical, and surface<br />

properties. E-Journal of Chemistry, 9, 510e516.<br />

Falguera, V., Quintero, J. P., Jimenez, A., Mu~noz, J. A., & Ibarz, A. (2011). Edible films<br />

and coatings: Structures, active functions, and trends in their use. Trends in Food<br />

Science and Technology, 22, 292e303.<br />

Gennadios, A., Weller, C. L., & Goodings, C. H. (1994). Measurement errors in water<br />

vapor permeability of high permeable hydrophilic edible films. Journal of Food<br />

Engineering, 21, 395e409.<br />

Gimenez, B., Lopez de Lacey, A., Perez-Santín, E., Lopez-Caballero, M. E., &<br />

Montero, P. (2013). Release of active compounds from agar and agar-gelatin<br />

films with green tea extract. Food Hydrocolloids, 30, 264e271.<br />

Han, J. H. (2000). Antimicrobial food packaging. Food Technology, 54, 56e65.<br />

Huang, H., & Yang, X. (2004). Synthesis of polysaccharide-stabilized gold and <strong>silver</strong><br />

nanoparticles: A green method. Carbohydrate Research, 339, 2627e2631.<br />

Hu, S., & Hsieh, Y. L. (2016). Silver nanoparticle synthesis using <strong>lignin</strong> as reducing<br />

and capping agents: A kinetic and mechanistic study. International Journal of<br />

Biological Macromolecules, 82, 856e862.<br />

Kanmani, P., & Rhim, J. W. (2014). Physical, mechanical and antimicrobial properties<br />

of gelatin based active nanocomposite films containing AgNPs and nanoclay.<br />

Food Hydrocolloids, 35, 644e652.<br />

Kim, J. S., Kuk, K. E., Yu, K. N., Kim, J. H., Park, S. J., Lee, H. J., et al. (2007). Antimicrobial<br />

effects of <strong>silver</strong> nanoparticles. Nanomedicine: Nanotechnology, Biology<br />

and Medicine, 3, 95e101.<br />

Kumar, A., Negi, Y. S., Choudhary, V., & Bhardwaj, N. K. (2014). Effect of modified<br />

cellulose nanocrystals on microstructural and mechanical properties of polyvinyl<br />

alcohol/ovalbumin biocomposite scaffolds. Materials Letters, 129, 61e64.<br />

Laibinis, P. E., Whitesides, G. M., Allara, D. L., Tao, Y. T., Parikh, A. N., & Nuzzo, R. G.<br />

(1991). Comparison of the structures and wetting properties of self-assembled<br />

monolayers of normal-alkane thiols on the coinage metal-surfaces, Cu, Ag, Au.<br />

Journal of American Chemical Society, 113, 7152e7167.<br />

Lischer, S., Korner, E., Balazs, D. J., Shen, D., Wick, P., Grieder, K., et al. (2011).<br />

Antibacterial burst-release from minimal Ag-containing plasma polymer coatings.<br />

Journal of the Royal Society Interface, 8, 1019e1030.<br />

Llorens, A., Lloret, E., Picouet, P. A., Trbojevich, R., & Fernandez, A. (2012). Metallic<br />

based micro and nanocomposites in food contact materials and active food<br />

packaging. Trends in Food Science and Technology, 24, 19e29.<br />

Martínez-Casta~non, G. A., Ni~no-Martinez, N., Martínez-Gutierrez, F., Martínez-<br />

Mendoza, J. R., & Ruiz, F. (2008). Synthesis and antibacterial activity of <strong>silver</strong><br />

nanoparticles with different sizes. Journal of Nanoparticle Research, 10,<br />

1343e1348.<br />

Martucci, J. F., & Ruseckaite, R. A. (2010). Three-layer sheets based on gelatin and<br />

poly(lactic acid), Part 1: Preparation and properties. Journal of Applied Polymer<br />

Science, 118, 3102e3110.<br />

Monties, B. (1991). Plant cell walls as fibrous lignocellulosic composites: Relations<br />

with <strong>lignin</strong> structure and function. Animal Feed Science and Technology, 32,<br />

159e175.<br />

Orsuwana, A., Shankar, S., Wang, L. F., Sothornvit, R., & Rhim, J. W. (2016). Preparation<br />

of antimicrobial agar/banana powder blend films reinforced with <strong>silver</strong><br />

nanoparticles. Food Hydrocolloids, 60, 476e485.<br />

Oun, A., & Rhim, J. W. (2016). Isolation of cellulose nanocrystals from grain straws<br />

and their use for the preparation of carboxymethyl cellulose-based nanocomposite<br />

films. Carbohydrate Polymers, 150, 187e200.<br />

Priyadarshini, S., Gopinath, V., Priyadharsshini, N. M., Ali, M. D., & Velusamy, P.<br />

(2013). Synthesis of anisotropic <strong>silver</strong> nanoparticles using novel strain. Colloids<br />

and Surfaces B: Biointerfaces, 102, 232e237.<br />

Rhim, J. W., Park, H. M., & Ha, C. S. (2013). Bio-nanocomposites for food packaging<br />

applications. Progress in Polymer Science, 38, 1629e1652.<br />

Shankar, S., Chorachoo, J., Jaiswal, L., & Voravuthikunchai, S. P. (2014). The effect of<br />

reducing agent concentrations and temperature on characteristics and antimicrobial<br />

activity of <strong>silver</strong> nanoparticles. Materials Letters, 137, 160e163.<br />

Shankar, S., Jaiswal, L., Selvakannan, P. R., Ham, K. S., & Rhim, J. W. (2016). Gelatinbased<br />

dissolvable antibacterial films reinforced with metallic nanoparticles. RSC<br />

Advances, 6, 67340e67352.<br />

Shankar, S., Prasad, R. G. S. V., Selvakannan, P. R., Jaiswal, L., & Laxman, R. S. (2015).


84<br />

S. Shankar, J.-W. Rhim / Food Hydrocolloids 71 (2017) 76e84<br />

Green synthesis of <strong>silver</strong> nanoribbons from waste X-ray films using alkaline<br />

protease. Materials Express, 5, 165e170.<br />

Shankar, S., Reddy, J. P., & Rhim, J. W. (2015). Effect of <strong>lignin</strong> on water vapor barrier,<br />

mechanical, and structural properties of agar/<strong>lignin</strong> composite films. International<br />

Journal of Biological Macromolecules, 81, 267e273.<br />

Shankar, S., & Rhim, J. W. (2015). Amino acid mediated synthesis of <strong>silver</strong> nanoparticles<br />

and preparation of antimicrobial agar/<strong>silver</strong> nanoparticles composite<br />

films. Carbohydrate Polymers, 130, 353e363.<br />

Shankar, S., & Rhim, J. W. (2016). Tocopherol-mediated synthesis of <strong>silver</strong> nanoparticles<br />

and preparation of antimicrobial PBAT/<strong>silver</strong> nanoparticles composite<br />

films. LWT-Food Science Technology, 72, 149e156.<br />

Shankar, S., Teng, X., Li, G., & Rhim, J. W. (2015). Preparation, characterization, and<br />

antimicrobial activity of gelatin/ZnO nanocomposite films. Food Hydrocolloids,<br />

45, 264e271.<br />

Shankar, S., Teng, X., & Rhim, J. W. (2014). Properties and characterization of agar/<br />

CuNPs bionanocomposite films prepared with different copper salts and<br />

reducing agents. Carbohydrate Polymers, 114, 484e492.<br />

Sung, S. Y., Sin, L. T., Tee, T. T., Bee, S. T., Rahmat, A. R., Rahman, W. A. W. A., et al.<br />

(2013). Antimicrobial agents for food packaging applications. Trends in Food<br />

Science and Technology, 33, 110e123.<br />

Tako, M., Higa, M., Medoruma, K., & Nakasone, Y. (1999). A highly methylated agar<br />

from red seaweed, Gracilaria arcuata. Botanica Marina, 42, 513e518.<br />

Tuomela, M., Vikman, M., Hatakka, A., & Itavaara, M. (2000). Biodegradation of<br />

<strong>lignin</strong> in a compost environment: A review. Bioresource Technology, 72,<br />

169e183.<br />

Volery, P., Besson, R., & Schaffer-lequart, C. (2004). Characterization of commercial<br />

carrageenans by Fourier transform infrared spectroscopy using single-reflection<br />

attenuated total reflection. Journal of Agricultural and Food Chemistry, 52,<br />

7457e7463.<br />

Yu, J., Yang, J., Liu, B., & Ma, X. (2009). Preparation and characterization of glycerol<br />

plasticized-pea starch/ZnO-carboxymethyl cellulose sodium nanocomposites.<br />

Bioresource Technology, 100, 284e2832.<br />

View publication stats

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!