Next Article in Journal
Nanopillar Diffraction Gratings by Two-Photon Lithography
Previous Article in Journal
Laser-Patternable Graphene Field Emitters for Plasma Displays
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

(Bio)polymer/ZnO Nanocomposites for Packaging Applications: A Review of Gas Barrier and Mechanical Properties

1
Packaging Technology Center, IMO-IMOMEC, Hasselt University, Wetenschapspark 27, 3590 Diepenbeek, Belgium
2
Functional Materials Engineering Group, IMO-IMOMEC, Hasselt University, Wetenschapspark 1, 3590 Diepenbeek, Belgium
3
IMEC vzw-Division IMOMEC, Wetenschapspark 1, 3590 Diepenbeek, Belgium
*
Author to whom correspondence should be addressed.
Nanomaterials 2019, 9(10), 1494; https://doi.org/10.3390/nano9101494
Submission received: 5 September 2019 / Revised: 9 October 2019 / Accepted: 15 October 2019 / Published: 19 October 2019

Abstract

:
Nanotechnology is playing a pivotal role in improving quality of life due to its versatile applications in many areas of research. In this regard, nanoparticles have gained significant importance. Zinc oxide nanoparticles (ZnO NPs) amongst other nanoparticles are being used in producing nanocomposites. Methods like solvent casting, solution casting, solvent volatilization, twin-screw extrusion, melt compounding and extrusion blow molding have been applied to produce ZnO NPs based (bio)polymer composites. These composites are of great interest in the research area of food packaging materials due to their improved multifunctional characteristics like their mechanical, barrier and antimicrobial properties. This paper gives an overview of the main methods to synthesize ZnO NPs, methods to incorporate ZnO NPs in (bio)polymers, and finally, the gas barrier and mechanical properties of the nanocomposites. As a conclusion, a maximum decline in oxygen, carbon dioxide and water vapor permeability was reported as 66%, 17% and 38% respectively, while tensile strength and young’s modulus were observed to increase by 32% and 57% respectively, for different (bio)polymer/ZnO nanocomposites.

1. Introduction

Nanotechnology combined with other areas of research such as the life sciences, physics, chemistry, medicine, engineering, the cognitive sciences, and information technology has gained significant importance due to its broad applications [1]. It has also proven its significance in the field of packaging technology. Its applications have also become a good basis to bring large benefits to the food and nutrition sector [2,3,4]. In recent years, nanotechnology has emerged as a multibillion-dollar global industry expected to surpass US $125 billion mark by 2024 [5]. The world’s leading food companies like Nestle, Unilever, Kraft, and Heinz are exploring nanotechnology in food processing and packaging areas [6]. Food packaging materials implementing nanotechnology are the largest category of application of nanotechnology for the food sector [7]. Smart and intelligent packaging systems, the use of biosensors in bacteria identification, and monitoring of food quality are some of the emerging applications of nanotechnology [8]. Food is a necessary part of our daily life, which has to be protected from undesirable factors that can potentially decrease its nutritional value and shelf-life. Depending on the food to be preserved, packaging requires specific moisture, gas and/or antimicrobial barriers. Food packaging is the sole consumer of more than 20% of the plastics produced [9]. However, the mainstream of plastic packaging materials has originated from petroleum resources. Non-sustainability and non-biodegradability are two important issues related to these plastic materials. Poor barrier properties against moisture and gases are also severe issues in food packaging as food products need optimal conditions to preserve them. Biopolymers are environmentally friendly materials and have the potential to improve food protection while retaining its quality and safety. These materials have also mitigated the environmental issues that are related to conventional polymers [10]. However, the main disadvantages associated with these materials are their weak barriers and mechanical properties which make them unfavorable for many food packaging applications [11]. Recent developments in (bio)polymer nanocomposites have assisted in expanding their functional properties [12]. Extensive research and development activities both in academia and industry can be helpful to cope with the above-mentioned challenges.
A variety of nanoparticles (NPs) reinforced composites have been developed that typically consist of up to 5% w/w NPs [13]. The addition of NPs into films has made end-products fire resistant, lightweight, stronger in thermal and mechanical performance and less permeable to moisture and gases. When incorporated in plastic films, nanoparticles-such as zinc oxide (ZnO) and silica and titanium dioxide (TiO2) are capable of reducing the flow of oxygen inside packaging containers. They also act as a barrier against moisture to keep food fresh for an extended period [14]. Silicon dioxide (SiO2) and TiO2 are the most commonly used nanoparticles in food packaging. SiO2 is used as a drying and anticaking agent [15]. TiO2 can be used as a barrier against UV protection for food packaging. These nanoparticles have also been reported for their antimicrobial activity, however, their applications are limited due to their photocatalytic behavior [16]. Silver nanoparticles can protect food from microbial invasion [17]. Copper and copper oxide, cadmium, ZnO, telluride, magnesium oxide, single-walled carbon nanotubes are also reported for their antimicrobial activity [18].
Nanomaterials in the field of food packaging are creating a tremendous impact by improving the numerous properties of packaging films. Zinc oxide nanoparticles (ZnO NPs) amongst other nanoparticles have gained a vital position in the enhancement of packaging properties like barrier, mechanical and antimicrobial properties [19]. These ZnO NPs consisting of various particle sizes and shapes have been synthesized via different routes. Their availability at commercial scale has made their access easy to use them directly in research experiments. Packaging films consisting of ZnO NPs along with (bio)polymers have been made through diverse processing techniques, such as solvent casting [20,21], melt compounding [22], solution casting [23,24,25,26], twin-screw extrusion [27], extrusion blow molding [28], and solvent volatilizing [29].
This article reviews the methods used to synthesize ZnO NPs and the methods to incorporate them in (bio)polymers, and then gives an overview of the use of ZnO NPs in packaging materials in order to improve their gas barrier and mechanical properties.

2. Zinc Oxide Nanoparticles

2.1. Synthesis Methods of ZnO NPs

A variety of synthesis methods to produce ZnO NPs have been published worldwide, e.g., sol-gel process, hydrothermal process, wet chemical method, etc. Different orientations, morphologies, size, and shapes of ZnO NPs can be obtained by varying a set of parameters such as type of solvent, reaction temperature, etc. Some important categories of synthesis methods of ZnO NPs are briefly described here.
The sol-gel method is a widely used method by scientists to synthesize ZnO NPs [30,31,32,33,34,35,36,37,38,39]. The sol-gel method is based on colloidal chemistry. Sols are referred to as the colloidal solution that consists of solid particles suspended in a liquid phase. Gels are typically formed by the polycondensation or polyesterification methods followed by aging to achieve phase transformations. This transformation is a very important step in the development of particle formation [40].
Among other methods, this technique has high reliability, low production cost, good repeatability, low process temperature, simplicity of the process, ease of control on morphology of nanoparticles and good homogeneity [41,42].
The hydrothermal method requires no organic solvents. The autoclave is used in the synthesis process, where a mixture of substrates is gradually heated to a temperature of 100–300 °C and then left for several days. As a result, crystal nuclei with nano dimensions are formed. Depending on the process conditions and materials used, crystals with a variety of shapes and dimensions can be obtained [43,44].
Sonochemical technology has gained significant attention in the synthesis of different compounds in material science. It is based on acoustic cavitation which results from the continuous formation, growth and implosive collapse of bubbles in a liquid [45]. Some researchers have also used this method to prepare different shapes of ZnO nanocrystallites [46].
The microemulsion method can be defined as an isotropic, thermodynamically stable system constituting smaller droplets dispersed in an immiscible solvent and an amphiphilic surfactant species on the surface of the micelle. For self-assembly and size-selective preparation of nanoparticles, the microemulsion mediated synthesis has been used due to excellent control on the particle size [47,48,49,50].
The solvothermal method is an appropriate method for the preparation of both crystalline oxide and non-oxide materials. Crystalline solids, including silicate materials with high porosity like zeolites [51] and oxide or non-oxide nanoparticles [52] can be produced by this method. The main advantage of this method is that the problems like solvents’ toxicity and the inability of dissolving the salts can be overtaken [53].
The chemical vapor deposition (CVD) method involves the deposition of a solid film on a heated surface by a chemical reaction from the gas or vapor phase. Different types of CVDs include plasma CVD, photo-laser CVD, thermal CVD, metal-organic chemical vapor deposition (MOCVD), and laser CVD, etc. For example, in thermal CVD, a temperature above 900 °C is used to activate the reaction [54].
The co-precipitation method normally involves the reaction of inorganic alkalis with zinc (Zn) salt. Precipitates produced from this reaction are washed and then calcined. In this method, calcination is done at different temperatures in order to achieve the nanoparticles with required morphology [55].
The microwave-assisted synthesis method is clean, simple, and without the problems of thermal gradient effects [56]. This process is widely used to produce hydroxide, oxide, and sulfide nanoparticles, etc. [57].
The green synthesis method is an eco-friendly, energy-efficient and cost-effective approach to synthesize nanoparticles. Plants, bacteria, fungi, algae, and viruses are used to synthesize ZnO NPs from this method [58].
In general, the preparation of nanoparticles is a complex process that contains different variables in the chemical reactions. It is difficult to control all the variables in the synthesis process. Each of these variables might have an effect on the properties of the end product. The most important variables are the type of solvent and precursor as well as physical and chemical conditions like pH and temperature [59]. Table 1 represents an overview of some synthesis methods and the respective size and morphology of the obtained ZnO NPs.

2.2. Commercial Grade ZnO NPs

There are several companies manufacturing ZnO NPs with different particle sizes, shapes, purity and phase states. Commercial, highly purified ZnO NPs are available in the range from 10 to 800 nm and they exist in powder, aqueous, suspension, emulsion and dispersion forms. Table 2 covers a list of vendors of ZnO NPs along with the information about the characteristics of these nanoparticles.

3. Production of (Bio)Polymer/ZnO Nanocomposites

Many (bio)polymers such as linear low density polyethylene (LLDPE), poly(lactic acid) (PLA), chitosan, poly(3-hydroxybutyrate-co-3-hydroxyvalerate) (PHBV), poly(3-hydroxybutyrate) (PHB), poly(butylene adipate-co-terephthalate) (PBAT), low density polyethylene (LDPE), semolina flour and bovine skin gelatin type-B (BSG) have been incorporated with ZnO NPs with different particle shapes and sizes through different processing methods to produce (bio)polymer nanocomposites such as solvent casting, solution casting, melt processing, and extrusion process, etc.
The solvent casting process contains ultrasonication or vigorous stirring of the nanoparticles in a polymer solution before casting in a mold and then evaporation of the solvent. Both organic solvents and water can be used to produce nanocomposites with either thermosets or thermoplastics [88]. In solution casting, a polymer is dissolved in a suitable solvent and then the active compound of interest is incorporated, then this mixed solution is poured on an inert surface. Upon solvent evaporation, films with the desired functionalities are formed [89]. In an extrusion process, the active compound and polymers are melted in a single or twin-screw extruder by heat transfer that leads to a blend formation from which films can be obtained [88]. Melt compounding or melt blending is a simple and cost-effective technique that is free from using toxic solvents. Blending can be done by using extrusion or injection molding methods [90].
Table 3 provides the information about ZnO NPs, (bio)polymers, characteristics and their manufacturers; and the processing methods to develop ZnO NPs based (bio)polymer composites.

4. Barrier Properties of (Bio)Polymer/ZnO Nanocomposites

Nanocomposites can enhance barrier properties by providing multipurpose chemical functionality. Significant improvement in the barrier properties on the incorporation of nanoparticles into the polymers has been reported [95,96,97,98]. The oxygen transmission rate (OTR) is defined as the amount of oxygen gas passing through an area in a certain time under specified conditions of temperature, humidity and pressure and its units are cm3/(m2.day) [99]. Permeability coefficients of oxygen (PO2-cm3.mm/(m2.day.atm)) and carbon dioxide (PCO2-cm3.mm/(m2.day.atm)) are obtained by multiplying the OTR and CO2-TR with the thickness (mm) of the sample, respectively [100]. While the water vapor transmission rate (WVTR) is the quantity of water vapor transmitted through an area in a certain time under specified conditions of humidity and temperature and it is presented in g/(m2.day) [101]. Water vapor permeability (WVP-g.mm/(m2.day)) is the product of permeance and thickness (mm) of the sample [102].
Jasim et al. [28] reported that the OTR has decreased by 23.2% upon 10 wt% ZnO NPs loading for LLDPE films reinforced with ZnO NPs. A uniform distribution of ZnO NPs in the polymer matrix could contribute to this decrease in OTR. For a PLA/ZnO bionanocomposite, an 18% and 17% decrease in PO2 and PCO2 were shown, respectively, due to the uniform distribution of ZnO nanoparticles in the PLA matrix, while WVP was increased by 16% [27].
For ZnO NPs based biocomposites, on the addition of 4 wt% ZnO in PHBV, the PO2 value was observed to decrease up to 35% due to strong interfacial adhesion with the polyester matrix, which caused chain immobilization [23]. In the case of ZnO/PHB bionanocomposites, upon 5 wt% ZnO NPs loading-PO2 value was decreased by about 53% and a decreasing trend was also observed for WVP value by up to 38% for the same ZnO NPs loading [24]. For ZnO/PBAT composite, there was a decreasing trend in OTR values upon increasing ZnO NPs loading range from 0 to 10 wt%. The lowest value of OTR was observed for 10 wt% ZnO NPs loading [25].
In a research conducted on untreated and 3-methacryloxypropyltrimethoxysilane treated ZnO nanoparticle reinforced-PLA nanocomposites, PO2 values of plasticized PLA film decreased by 36.07% and 55.1% by adding 10% ZnO (untreated) and ZnO (3-methacryloxypropyltrimethoxysilane treated) NPs, respectively [20]. This enormous decrease in PO2 values could be ascribed to the reinforcement of ZnO NPs with its high aspect ratio and its nice distribution throughout the polymer matrix [103].
In a study on Semolina reinforced with nanofillers (ZnO-nanorod/nano-kaolin), PO2 was decreased by up to 66% [21]. Wenhui Li et al. [29] reported in their research on ZnO based PLA nanocomposite that OTR was decreased on increasing ZnO content. The tortuous pathway prolongs the oxygen pathway and is the main reason for the improvement of oxygen resistance in the nano-blend films [104]. However, there was an increasing trend for WVP that could be due to the hydrophilicity of nano-ZnO and the improved hydrophilic interaction of the films.
For bovine skin gelatin type-B (BSG) composite films incorporated with ZnO nanorods and clove essential oil (CEO), PO2 was decreased by 32.27% by the addition of 2 wt% into the polymer matrix. This reduction in PO2 value might be due to the uniform dispersion and high aspect ratio of ZnO nanorods into the gelatin matrix [26].
It can be concluded from permeability tests of ZnO NPs based (bio)polymer composites that the maximum decrease in PO2 value by up to 66% is observed for semolina reinforced with nanofillers [21] where ZnO particles have been used in nanorods shape. In addition, a 17% decrease in PCO2 upon the incorporation of 1 wt% ZnO in ZnO/PLA biocomposites [27] and 38% decrease in WVP value at 5 wt% ZnO NPs loading in ZnO/PHB nanobiocomposites [24] are reported.
In summary, the barrier properties of the nanocomposites can be improved by a high aspect ratio, uniform dispersion and low incorporation (up to 5 wt%) of nanoparticles within the polymer matrix.
The barrier properties regarding various ZnO NPs based (bio)polymer composites are represented in Table 4.

5. Mechanical Properties of (Bio)Polymer/ZnO Nanocomposites

Incorporation of ZnO NPs into the polymer matrix has shown a significant effect on the mechanical properties of the composites, which include tensile strength, impact strength, young’s modulus, stress at yield, strain at break, elastic strength and elongation at break.
Roberto Pantani et al. [22] reported that PLA-ZnO nanocomposite films showed a slight increase in Young’s modulus and low elongation at break in tensile tests. The addition of a plasticizer into PLA-ZnO nanocomposite films could be considered for the improvement of these parameters. Tensile tests on ZnO/PHBV nanobiocomposites showed that Young’s modulus was increased by 57% on 4 wt% ZnO NPs loading. An increase in the crystallinity of PHBV, homogeneous distribution of nanoparticles and strong interfacial adhesion between the phases could be the possibilities for the significant increase in Young’s modulus. However, the strain at break was decreased by 30% with increasing ZnO NPs content. Due to filler reinforcement, the ductile flow of the polymer chains was restricted, which was also depicted from low strain at break values [23].
In the case of ZnO/PHB nanobiocomposites, the tensile strength, young’s modulus, and impact strength were increased up to 32%, 43% and 26% respectively due to a strong interfacial adhesion between the matrix and the nanofiller via interactions by hydrogen bonding [24]. Tensile strength and elongation at break values were increased in the case of ZnO/PBAT nanocomposites. This could be due to the fine distribution of ZnO NPs in the polymer matrix [25]. Songee Beak et al. [92] reported the analysis of the tensile tests on an olive flounder bone gelatin-ZnO nanocomposite. They observed that the tensile strength of the nanocomposite films was increased while elongation at break was decreased by 37%. Intermolecular interactions of ZnO NPs with gelatin molecules were responsible for improving the rigidity of the gelatin films.
In a study on LDPE/ZnO nanocomposites, the values of elongation at break and tensile strength were decreased due to agglomeration and poor interfacial adhesion in the polymer nanocomposites [93]. A similar trend of decreasing elongation at break and tensile strength was observed for ZnO based PP nanocomposites [94].
It can be concluded from the above-mentioned analyses on mechanical properties of ZnO nanoparticles based (bio)polymer composites that a maximum increase in tensile strength up to 32% has been observed for ZnO/PHB bionanocomposites [24], while a 57% increase in young’s modulus upon 4 wt% ZnO loading in ZnO/PHBV bionanocomposites [23] is reported.
In summary, the uniform distribution of nanoparticles within the polymer matrix and strong interfacial adhesion can lead to enhanced mechanical properties of nanocomposites. The mechanical properties of ZnO nanoparticles based (bio)polymer composites are summarized in Table 5.

6. Conclusions

Nanotechnology is providing new possibilities for controlling properties, adding value and introducing new features in almost all scientific and technological fields to improve the quality of life. In this regard, nanocomposites are contributing significantly as they offer excellent opportunities to explore new functionalities as compared to conventional materials. Owing to the multifunctional properties in (bio)polymer composites, ZnO NPs have drawn interest of a lot of researchers in the diversified area of food packaging. The aim of this paper was to summarize the main methods to synthesize ZnO NPs, the methods to incorporate ZnO NPs in (bio)polymer matrices and the impact of the ZnO NPs incorporation on the gas barrier and mechanical properties of the (bio)polymer nanocomposites. These innovative (bio)polymer/ZnO nanocomposite films can be potentially applied as food packaging materials with improved properties. Based on comprehensive investigations on ZnO NPs based (bio)polymer nanocomposites, a well-defined feature of these nanocomposites is the small size of nanoparticles that leads to an increased interfacial area as compared to the traditional composites. The interfacial interaction between ZnO NPs and the polymer matrix plays an important role in improving the final properties. Upon surface modification of these nanoparticles, the desired properties of the matrix can be improved further even at lower wt% of nanoparticles. However, there are certain challenges associated with the production of nanocomposites such as: (1) a uniform dispersion of the nanoparticles in the matrix and (2) a proper selection of manufacturing techniques in order to regulate the level of dispersion of these nanoparticles which also create a significant impact on reinforcement as well as end-product properties. Moreover, tremendous research efforts are essential in moving towards the industrialization of ZnO/(bio)polymer nanocomposites to fulfill the requirements of numerous technological fields.

Author Contributions

M.A. did the writing and literature research for the paper. R.P., M.B., and W.D. contributed to the review and editing. R.P. supervised the work.

Funding

This research was funded by the Higher Education Commission (HEC), Pakistan, under award letter No. PD (HRDI-UESTPs)/Batch-IV/2016/1450.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Rossi, M.; Cubadda, F.; Dini, L.; Terranova, M.L.; Aureli, F.; Sorbo, A.; Passeri, D. Scientific basis of nanotechnology, implications for the food sector and future trends. Trends Food Sci. Technol. 2014, 40, 127–148. [Google Scholar] [CrossRef] [Green Version]
  2. Cushen, M.; Kerry, J.; Morris, M.; Cruz-Romero, M.; Cummins, E. Nanotechnologies in the food industry–Recent developments, risks and regulation. Trends Food Sci. Technol. 2012, 24, 30–46. [Google Scholar] [CrossRef]
  3. Silvestre, C.; Duraccio, D.; Cimmino, S. Food packaging based on polymer nanomaterials. Prog. Polym. Sci. 2011, 36, 1766–1782. [Google Scholar] [CrossRef]
  4. Weiss, J.; Takhistov, P.; McClements, D.J. Functional Materials in Food Nanotechnology. J. Food Sci. 2006, 71, R107–R116. [Google Scholar] [CrossRef] [Green Version]
  5. Global Nanotechnology Market 2018–2024: Market is Expected to Exceed US$ 125 Billion. Available online: https://www.prnewswire.com/news-releases/global-nanotechnology-market-2018-2024-market-is-expected-to-exceed-us-125-billion-300641054.html (accessed on 4 September 2019).
  6. Bernardes, P.C.; de Andrade, N.J.; Soares, N.D.F.F. Nanotechnology in the food industry. Biosci. J. 2014, 30, 1919–1932. [Google Scholar]
  7. Bradley, E.L.; Castle, L.; Chaudhry, Q. Applications of nanomaterials in food packaging with a consideration of opportunities for developing countries. Trends Food Sci. Technol. 2011, 22, 604–610. [Google Scholar] [CrossRef]
  8. Yousefi, H.; Su, H.-M.; Imani, S.M.; Alkhaldi, K.M.; Filipe, C.D.; Didar, T.F. Intelligent food packaging: A review of smart sensing technologies for monitoring food quality. ACS Sens. 2019, 4, 808–821. [Google Scholar] [CrossRef]
  9. Rhim, J.-W.; Park, H.-M.; Ha, C.-S. Bio-nanocomposites for food packaging applications. Prog. Polym. Sci. 2013, 38, 1629–1652. [Google Scholar] [CrossRef]
  10. Garavand, F.; Rouhi, M.; Razavi, S.H.; Cacciotti, I.; Mohammadi, R. Improving the integrity of natural biopolymer films used in food packaging by crosslinking approach: A review. Int. J. Biol. Macromol. 2017, 104, 687–707. [Google Scholar] [CrossRef]
  11. Cacciotti, I.; Mori, S.; Cherubini, V.; Nanni, F. Eco-sustainable systems based on poly (lactic acid), diatomite and coffee grounds extract for food packaging. Int. J. Biol. Macromol. 2018, 112, 567–575. [Google Scholar] [CrossRef]
  12. Vahedikia, N.; Garavand, F.; Tajeddin, B.; Cacciotti, I.; Jafari, S.M.; Omidi, T.; Zahedi, Z. Biodegradable zein film composites reinforced with chitosan nanoparticles and cinnamon essential oil: Physical, mechanical, structural and antimicrobial attributes. Colloids Surf. B Biointerfaces 2019, 177, 25–32. [Google Scholar] [CrossRef] [PubMed]
  13. Llorens, A.; Lloret, E.; Picouet, P.A.; Trbojevich, R.; Fernandez, A. Metallic-based micro and nanocomposites in food contact materials and active food packaging. Trends Food Sci. Technol. 2012, 24, 19–29. [Google Scholar] [CrossRef]
  14. Horner, S.R.; Mace, C.R.; Rothberg, L.J.; Miller, B.L. A proteomic biosensor for enteropathogenic E. coli. Biosens. Bioelectron. 2006, 21, 1659–1663. [Google Scholar] [CrossRef] [PubMed]
  15. Zhao, R.; Torley, P.; Halley, P.J. Emerging biodegradable materials: Starch-and protein-based bio-nanocomposites. J. Mater. Sci. 2008, 43, 3058–3071. [Google Scholar] [CrossRef]
  16. Arora, A.; Padua, G.W. Nanocomposites in food packaging. J. Food Sci. 2010, 75, R43–R49. [Google Scholar] [CrossRef]
  17. Cacciotti, I.; Fortunati, E.; Puglia, D.; Kenny, J.M.; Nanni, F. Effect of silver nanoparticles and cellulose nanocrystals on electrospun poly (lactic) acid mats: Morphology, thermal properties and mechanical behavior. Carbohydr. Polym. 2014, 103, 22–31. [Google Scholar] [CrossRef]
  18. Arshak, K.; Adley, C.; Moore, E.; Cunniffe, C.; Campion, M.; Harris, J. Characterisation of polymer nanocomposite sensors for quantification of bacterial cultures. Sens. Actuators B Chem. 2007, 126, 226–231. [Google Scholar] [CrossRef]
  19. Huang, Y.; Mei, L.; Chen, X.; Wang, Q. Recent developments in food packaging based on nanomaterials. Nanomaterials 2018, 8, 830. [Google Scholar] [CrossRef]
  20. Arfat, Y.A.; Ahmed, J.; Al Hazza, A.; Jacob, H.; Joseph, A. Comparative effects of untreated and 3-methacryloxypropyltrimethoxysilane treated ZnO nanoparticle reinforcement on properties of polylactide-based nanocomposite films. Int. J. Biol. Macromol. 2017, 101, 1041–1050. [Google Scholar] [CrossRef]
  21. Jafarzadeh, S.; Ariffin, F.; Mahmud, S.; Alias, A.K.; Hosseini, S.F.; Ahmad, M. Improving the physical and protective functions of semolina films by embedding a blend nanofillers (ZnO-nr and nano-kaolin). Food Packag. Shelf Life 2017, 12, 66–75. [Google Scholar] [CrossRef]
  22. Pantani, R.; Gorrasi, G.; Vigliotta, G.; Murariu, M.; Dubois, P. PLA-ZnO nanocomposite films: Water vapor barrier properties and specific end-use characteristics. Eur. Polym. J. 2013, 49, 3471–3482. [Google Scholar] [CrossRef]
  23. Díez-Pascual, A.; Díez-Vicente, A. Poly (3-hydroxybutyrate)/ZnO bionanocomposites with improved mechanical, barrier and antibacterial properties. Int. J. Mol. Sci. 2014, 15, 10950–10973. [Google Scholar] [CrossRef] [PubMed]
  24. Díez-Pascual, A.M.; Diez-Vicente, A.L. ZnO-reinforced poly (3-hydroxybutyrate-co-3-hydroxyvalerate) bionanocomposites with antimicrobial function for food packaging. ACS Appl. Mater. Interfaces 2014, 6, 9822–9834. [Google Scholar] [CrossRef] [PubMed]
  25. Venkatesan, R.; Rajeswari, N. ZnO/PBAT nanocomposite films: Investigation on the mechanical and biological activity for food packaging. Polym. Adv. Technol. 2017, 28, 20–27. [Google Scholar] [CrossRef]
  26. Ejaz, M.; Arfat, Y.A.; Mulla, M.; Ahmed, J. Zinc oxide nanorods/clove essential oil incorporated Type B gelatin composite films and its applicability for shrimp packaging. Food Packag. Shelf Life 2018, 15, 113–121. [Google Scholar] [CrossRef]
  27. Marra, A.; Silvestre, C.; Duraccio, D.; Cimmino, S. Polylactic acid/zinc oxide biocomposite films for food packaging application. Int. J. Biol. Macromol. 2016, 88, 254–262. [Google Scholar] [CrossRef]
  28. Ahmed, J.; Arfat, Y.A.; Al-Attar, H.; Auras, R.; Ejaz, M. Rheological, structural, ultraviolet protection and oxygen barrier properties of linear low-density polyethylene films reinforced with zinc oxide (ZnO) nanoparticles. Food Packag. Shelf Life 2017, 13, 20–26. [Google Scholar] [CrossRef]
  29. Li, W.; Li, L.; Cao, Y.; Lan, T.; Chen, H.; Qin, Y. Effects of PLA film incorporated with ZnO nanoparticle on the quality attributes of fresh-cut apple. Nanomaterials 2017, 7, 207. [Google Scholar] [CrossRef]
  30. Sharma, A.; Singh, B.; Dhar, S.; Gondorf, A.; Spasova, M. Effect of surface groups on the luminescence property of ZnO nanoparticles synthesized by sol–gel route. Surf. Sci. 2012, 606, L13–L17. [Google Scholar] [CrossRef]
  31. Erol, A.; Okur, S.; Comba, B.; Mermer, Ö.; Arıkan, M. Humidity sensing properties of ZnO nanoparticles synthesized by sol–gel process. Sens. Actuators B Chem. 2010, 145, 174–180. [Google Scholar] [CrossRef]
  32. Tokumoto, M.S.; Pulcinelli, S.H.; Santilli, C.V.; Briois, V. Catalysis and temperature dependence on the formation of ZnO nanoparticles and of zinc acetate derivatives prepared by the sol−gel route. J. Phys. Chem. B 2003, 107, 568–574. [Google Scholar] [CrossRef]
  33. Chu, S.-Y.; Yan, T.-M.; Chen, S.-L. Characteristics of sol-gel synthesis of ZnO-based powders. J. Mater. Sci. Lett. 2000, 19, 349–352. [Google Scholar] [CrossRef]
  34. Chen, Z.; Li, X.; Chen, N.; Wang, H.; Du, G.; Suen, A.Y. Effect of annealing on photoluminescence of blue-emitting ZnO nanoparticles by sol–gel method. J. Sol Gel Sci. Technol. 2012, 62, 252–258. [Google Scholar] [CrossRef]
  35. Cheetham, A.K.; Mellot, C.F. In Situ Studies of the Sol−Gel Synthesis of Materials. Chem. Mater. 1997, 9, 2269–2279. [Google Scholar] [CrossRef]
  36. Niederberger, M. Nonaqueous sol–gel routes to metal oxide nanoparticles. Acc. Chem. Res. 2007, 40, 793–800. [Google Scholar] [CrossRef]
  37. Liu, Y.; Morishima, T.; Yatsui, T.; Kawazoe, T.; Ohtsu, M. Size control of sol–gel-synthesized ZnO quantum dots using photo-induced desorption. Nanotechnology 2011, 22, 215605. [Google Scholar] [CrossRef]
  38. Nipane, D.; Thakare, S.; Khati, N. ZnO nanoparticle by sol-gel and its UV application in cosmetics formulation. Int. J. Knowl. Eng. 2012, 3, 168–169. [Google Scholar]
  39. Ba-Abbad, M.M.; Kadhum, A.A.H.; Mohamad, A.B.; Takriff, M.S.; Sopian, K. Optimization of process parameters using D-optimal design for synthesis of ZnO nanoparticles via sol–gel technique. J. Ind. Eng. Chem. 2013, 19, 99–105. [Google Scholar] [CrossRef]
  40. Khan, M.F.; Ansari, A.H.; Hameedullah, M.; Ahmad, E.; Husain, F.M.; Zia, Q.; Baig, U.; Zaheer, M.R.; Alam, M.M.; Khan, A.M. Sol-gel synthesis of thorn-like ZnO nanoparticles endorsing mechanical stirring effect and their antimicrobial activities: Potential role as nano-antibiotics. Sci. Rep. 2016, 6, 27689. [Google Scholar] [CrossRef]
  41. Vafaee, M.; Ghamsari, M.S. Preparation and characterization of ZnO nanoparticles by a novel sol–gel route. Mater. Lett. 2007, 61, 3265–3268. [Google Scholar] [CrossRef]
  42. Köse, H.; Karaal, Ş.; Aydın, A.O.; Akbulut, H. A facile synthesis of zinc oxide/multiwalled carbon nanotube nanocomposite lithium ion battery anodes by sol–gel method. J. Power Sources 2015, 295, 235–245. [Google Scholar]
  43. Djurisic, A.B.; Chen, X.Y.; Leung, Y.H. Recent progress in hydrothermal synthesis of zinc oxide nanomaterials. Recent Pat. Nanotechnol. 2012, 6, 124–134. [Google Scholar] [CrossRef] [PubMed]
  44. Innes, B.; Tsuzuki, T.; Dawkins, H.; Dunlop, J.; Trotter, G.; Nearn, M.; McCormick, P. Nanotechnology and the cosmetic chemist. Cosmet. Aeorosol Toilet. Aust. 2002, 15, 10–24. [Google Scholar]
  45. Suslick, K.S. Sonochemistry. Science 1990, 247, 1439–1445. [Google Scholar] [CrossRef]
  46. Askarinejad, A.; Alavi, M.A.; Morsali, A. Sonochemically assisted synthesis of ZnO nanoparticles: A novel direct method. Iran. J. Chem. Chem. Eng. 2011, 30, 75–81. [Google Scholar]
  47. Kim, B.-H.; Kim, J.-H.; Kwon, I.-H.; Song, M.-Y. Electrochemical properties of LiNiO2 cathode material synthesized by the emulsion method. Ceram. Int. 2007, 33, 837–841. [Google Scholar] [CrossRef]
  48. Balint, I.; You, Z.; Aika, K.-I. Morphology and oxide phase control in the microemulsion mediated synthesis of barium stabilized alumina nanoparticles. Phys. Chem. Chem. Phys. 2002, 4, 2501–2503. [Google Scholar] [CrossRef]
  49. Han, D.; Yang, H.; Shen, C.; Zhou, X.; Wang, F. Synthesis and size control of NiO nanoparticles by water-in-oil microemulsion. Powder Technol. 2004, 147, 113–116. [Google Scholar] [CrossRef]
  50. Lu, C.-H.; Chang, H.-H.; Lin, Y.-K. Preparation and characterization of nanosized lithium cobalt oxide powders for lithium-ion batteries. Ceram. Int. 2004, 30, 1641–1645. [Google Scholar] [CrossRef]
  51. Byrappa, K.; Yoshimura, M. Handbook of Hydrothermal Technology; William Andrew Publishing: New York, NY, USA, 2001. [Google Scholar]
  52. Rajamathi, M.; Seshadri, R. Oxide and chalcogenide nanoparticles from hydrothermal/solvothermal reactions. Curr. Opin. Solid State Mater. Sci. 2002, 6, 337–345. [Google Scholar] [CrossRef] [Green Version]
  53. Charitidis, C.A.; Georgiou, P.; Koklioti, M.A.; Trompeta, A.-F.; Markakis, V. Manufacturing nanomaterials: From research to industry. Manuf. Rev. 2014, 1, 11. [Google Scholar] [CrossRef]
  54. Rajput, N. Methods of preparation of nanoparticles—A review. Int. J. Adv. Eng. Technol. 2015, 7, 1806. [Google Scholar]
  55. Mantzaris, N.V. Liquid-phase synthesis of nanoparticles: Particle size distribution dynamics and control. Chem. Eng. Sci. 2005, 60, 4749–4770. [Google Scholar] [CrossRef]
  56. Pimentel, A.; Nunes, D.; Duarte, P.; Rodrigues, J.; Costa, F.; Monteiro, T.; Martins, R.; Fortunato, E. Synthesis of long ZnO nanorods under microwave irradiation or conventional heating. J. Phys. Chem. C 2014, 118, 14629–14639. [Google Scholar] [CrossRef]
  57. Zhao, J.; Yan, W. Chapter 8-Microwave-assisted Inorganic Syntheses. In Modern Inorganic Synthetic Chemistry; Xu, R., Pang, W., Huo, Q., Eds.; Elsevier: Amsterdam, The Netherlands, 2011. [Google Scholar]
  58. Naveed Ul Haq, A.; Nadhman, A.; Ullah, I.; Mustafa, G.; Yasinzai, M.; Khan, I. Synthesis Approaches of Zinc Oxide Nanoparticles: The Dilemma of Ecotoxicity. J. Nanomater. 2017, 2017, 8510342. [Google Scholar] [CrossRef]
  59. Espitia, P.J.P.; Soares, N.D.F.F.; Coimbra, J.S.D.R.; De Andrade, N.J.; Cruz, R.S.; Medeiros, E.A.A.; Andrade, N.J. Zinc oxide nanoparticles: Synthesis, antimicrobial activity and food packaging applications. Food Bioprocess Technol. 2012, 5, 1447–1464. [Google Scholar] [CrossRef]
  60. Lepot, N.; Van Bael, M.; Van den Rul, H.; D’Haen, J.; Peeters, R.; Franco, D.; Mullens, J. Synthesis of ZnO nanorods from aqueous solution. Mater. Lett. 2007, 61, 2624–2627. [Google Scholar] [CrossRef]
  61. Elen, K.; Van den Rul, H.; Hardy, A.; Van Bael, M.K.; D’Haen, J.; Peeters, R.; Franco, D.; Mullens, J. Hydrothermal synthesis of ZnO nanorods: A statistical determination of the significant parameters in view of reducing the diameter. Nanotechnology 2009, 20, 055608. [Google Scholar] [CrossRef]
  62. Jalal, R.; Goharshadi, E.K.; Abareshi, M.; Moosavi, M.; Yousefi, A.; Nancarrow, P. ZnO nanofluids: Green synthesis, characterization, and antibacterial activity. Mater. Chem. Phys. 2010, 121, 198–201. [Google Scholar] [CrossRef]
  63. Bhadra, P.; Mitra, M.; Das, G.; Dey, R.; Mukherjee, S. Interaction of chitosan capped ZnO nanorods with Escherichia coli. Mater. Sci. Eng. C 2011, 31, 929–937. [Google Scholar] [CrossRef]
  64. Kumar, V.; Wariar, P.; Prasad, V.; Koshy, J. A novel approach for the synthesis of nanocrystalline zinc oxide powders by room temperature co-precipitation method. Mater. Lett. 2011, 65, 2059–2061. [Google Scholar] [CrossRef]
  65. Lim, S.K.; Hwang, S.-H.; Kim, S.; Park, H. Preparation of ZnO nanorods by microemulsion synthesis and their application as a CO gas sensor. Sens. Actuators B Chem. 2011, 160, 94–98. [Google Scholar] [CrossRef]
  66. Elen, K.; Kelchtermans, A.; Van den Rul, H.; Peeters, R.; Mullens, J.; Hardy, A.; Van Bael, M. Comparison of two novel solution-based routes for the synthesis of equiaxed ZnO nanoparticles. J. Nanomater. 2011, 2011, 390621. [Google Scholar] [CrossRef]
  67. Šarić, A.; Štefanić, G.; Dražić, G.; Gotić, M. Solvothermal synthesis of zinc oxide microspheres. J. Alloys Compd. 2015, 652, 91–99. [Google Scholar] [CrossRef]
  68. Ghosh, S.; Majumder, D.; Sen, A.; Roy, S. Facile sonochemical synthesis of zinc oxide nanoflakes at room temperature. Mater. Lett. 2014, 130, 215–217. [Google Scholar] [CrossRef]
  69. Laurenti, M.; Garino, N.; Porro, S.; Fontana, M.; Gerbaldi, C. Zinc oxide nanostructures by chemical vapour deposition as anodes for Li-ion batteries. J. Alloys Compd. 2015, 640, 321–326. [Google Scholar] [CrossRef]
  70. Anand, V.; Srivastava, V.C. Zinc oxide nanoparticles synthesis by electrochemical method: Optimization of parameters for maximization of productivity and characterization. J. Alloys Compd. 2015, 636, 288–292. [Google Scholar] [CrossRef]
  71. Meliorum Technologies, Inc. Available online: https://www.meliorum.com/zinc-oxide-nanoparticles (accessed on 4 September 2019).
  72. Sukgyung AT Co., Ltd. Available online: http://www.sukgyung.com/1-3cosmetic-2zinc.php?dept3=8&2 (accessed on 4 September 2019).
  73. US Research Nanomaterials, Inc. Available online: https://www.us-nano.com/?gclid=EAIaIQobChMIlIW85cSO2gIVBbcbCh1CQwF9EAMYAiAAEgLkJvD_BwE (accessed on 4 September 2019).
  74. SkySpring Nanomaterials, Inc. Available online: https://www.ssnano.com/nanopowders---micron-powders (accessed on 4 September 2019).
  75. Stanford Advanced Materials. Available online: http://www.samaterials.com/zinc/1200-zinc-oxide-powder-zno-powder.html (accessed on 4 September 2019).
  76. Ultrananotech Materials Beyond Imagination. Available online: http://ultrananotec.com/products/metal-oxide-nanopowder/metal-oxide-nanopowders (accessed on 4 September 2019).
  77. Advanced Nano Products. Available online: http://www.anapro.com/eng/product/uv_blocker_zinc_oxide_powder_slurry.html (accessed on 4 September 2019).
  78. mkNANO. Available online: https://mknano.com/Nanoparticles/Single-Element-Oxides (accessed on 4 September 2019).
  79. Nanophase™ Nanoengineered Products. Available online: http://nanophase.com/products/zinc-oxide/ (accessed on 4 September 2019).
  80. Inframat® Advanced Materials™. Available online: http://www.advancedmaterials.us/30N-0801.htm (accessed on 4 September 2019).
  81. TECNAN. Available online: https://tecnan-nanomat.es/nanoparticulas-en-polvo/ (accessed on 4 September 2019).
  82. Micronisers. Available online: http://www.micronisers.com/products/additives-intermediates/powders/nanosun-zinc-oxide-p99-30/ (accessed on 4 September 2019).
  83. EPRUI Nanoparticles & Microspheres Co. Ltd. Available online: https://www.nanoparticles-microspheres.com/Products/ZnO-nanoparticles.html (accessed on 4 September 2019).
  84. American Elements. Available online: https://www.americanelements.com/zinc-oxide-nanoparticle-dispersion-1314-13-2 (accessed on 4 September 2019).
  85. Linari NanoTech. Available online: http://linaribiomedical.com/index.php/nanoparticles/zno-colloid/zno-deg-detail (accessed on 4 September 2019).
  86. Sigma Aldrich®. Available online: https://www.sigmaaldrich.com/materials-science/material-science-products.html?TablePage=18010478 (accessed on 4 September 2019).
  87. NYACOL® Nano Technologies, Inc. Available online: http://www.nyacol.com/products/zinc-oxide/ (accessed on 4 September 2019).
  88. Vasile, C. Polymeric Nanocomposites and Nanocoatings for Food Packaging: A Review. Materials 2018, 11, 1834. [Google Scholar] [CrossRef]
  89. Riccardi, C.; Zanini, S.; Tassetti, D. A Polymeric Film Coating Method on a Substrate by Depositing and Subsequently Polymerizing a Monomeric Composition by Plasma Treatment. Patent WO2014191901 A1, 4 December 2014. [Google Scholar]
  90. Ravichandran, K.; Praseetha, P.K.; Arun, T.; Gobalakrishnan, S. Chapter 6—Synthesis of Nanocomposites. In Synthesis of Inorganic Nanomaterials; Mohan Bhagyaraj, S., Oluwafemi, O.S., Kalarikkal, N., Thomas, S., Eds.; Woodhead Publishing: Cambridge, UK, 2018; pp. 141–168. [Google Scholar] [CrossRef]
  91. Noshirvani, N.; Ghanbarzadeh, B.; Mokarram, R.R.; Hashemi, M. Novel active packaging based on carboxymethyl cellulose-chitosan-ZnO NPs nanocomposite for increasing the shelf life of bread. Food Packag. Shelf Life 2017, 11, 106–114. [Google Scholar] [CrossRef]
  92. Beak, S.; Kim, H.; Song, K.B. Characterization of an Olive Flounder Bone Gelatin-Zinc Oxide Nanocomposite Film and Evaluation of Its Potential Application in Spinach Packaging. J. Food Sci. 2017, 82, 2643–2649. [Google Scholar] [CrossRef]
  93. Polat, S.; Fenercioğlu, H.; Güçlü, M. Effects of metal nanoparticles on the physical and migration properties of low density polyethylene films. J. Food Eng. 2018, 229, 32–42. [Google Scholar] [CrossRef]
  94. Polat, S.; Fenercioglu, H.; Unal Turhan, E.; Guclu, M. Effects of nanoparticle ratio on structural, migration properties of polypropylene films and preservation quality of lemon juice. J. Food Process. Preserv. 2018, 42, e13541. [Google Scholar] [CrossRef]
  95. Lange, J.; Wyser, Y. Recent innovations in barrier technologies for plastic packaging—A review. Packag. Technol. Sci. 2003, 16, 149–158. [Google Scholar] [CrossRef]
  96. Duncan, T.V. Applications of nanotechnology in food packaging and food safety: Barrier materials, antimicrobials and sensors. J. Colloid Interface Sci. 2011, 363, 1–24. [Google Scholar] [CrossRef] [PubMed]
  97. Smolander, M.; Chaudhry, Q. Nanotechnologies in Foods; RSC Publishing: Cambridge, UK, 2010; pp. 86–101. [Google Scholar]
  98. Avella, M.; De Vlieger, J.J.; Errico, M.E.; Fischer, S.; Vacca, P.; Volpe, M.G. Biodegradable starch/clay nanocomposite films for food packaging applications. Food Chem. 2005, 93, 467–474. [Google Scholar] [CrossRef]
  99. Available online: http://www.polyprint.com/flexographic-otr.htm (accessed on 29 September 2019).
  100. Available online: http://en.labthink.com/en-us/literatures/gas-transmission-rate-permeance-and-permeability-coefficient-application-guide.html (accessed on 29 September 2019).
  101. Food Packaging Permeability Behaviour: A Report. Int. J. Polym. Sci. 2012, 2012, 302029. [CrossRef]
  102. Patricia Miranda, S.; Garnica, O.; Lara-Sagahon, V.; Cárdenas, G. Water vapor permeability and mechanical properties of chitosan composite films. J. Chil. Chem. Soc. 2004, 49, 173–178. [Google Scholar] [CrossRef]
  103. Alebooyeh, R.; MohammadiNafchi, A.; Jokr, M. The Effects of ZnOnanorodson the Characteristics of Sago Starch Biodegradable Films. J. Chem. Health Risks 2018, 2. [Google Scholar] [CrossRef]
  104. Marra, A.; Rollo, G.; Cimmino, S.; Silvestre, C. Assessment on the effects of ZnO and Coated ZnO particles on iPP and PLA properties for application in food packaging. Coatings 2017, 7, 29. [Google Scholar] [CrossRef]
Table 1. Overview of synthesis methods and morphology of zinc oxide (ZnO) nanoparticles (NPs).
Table 1. Overview of synthesis methods and morphology of zinc oxide (ZnO) nanoparticles (NPs).
MethodMaterialsSize (nm)ShapeReference
HydrothermalZinc acetate dihydrate, polyvinylpyrrolidone (PVP)L: 5000,
D: 50–200
Nanorods[60]
Zinc acetate dihydrate, zinc chloride, sodium hydroxide 60Nanorods[61]
Microwave decomposition1-Butyl-3-methylimidazolium bis (trifluoromethylsulfonyl) imide [bmim][NTf2], zinc acetate dehydrate37–47Sphere[62]
Co-precipitationZinc acetate, double distilled waterD: 30–60,
L: 80
Nanorods[63]
Tetrahydrated zinc nitrate, ammonium hydroxide20–40Crystals[64]
Micro-emulsionZinc acetate dihydrate, ethylbenzene acid sodium salt (EBS), xylene, dodecylbenzene sulfonic acid sodium salt (DBS), ethanol and hydrazine DDBS: 300
DEBS: 80
Nanorods[65]
Zn(AOT)2, heptane, diethyl oxalate, chloroform, methanol10–20Quasispherical[66]
SolvothermalZinc acetate dihydrate, polyethylene glycol, absolute ethanol10–20Quasispherical[66]
Triethanolamine, zinc acetylacetonate monohydrate, 1-octanol, and absolute ethanol Lrod ~100
Dsphere ~20
Rods (ethanol without triethanolamine—TEA) spherical (ethanol with TEA)[67]
Sol-gelOxalic acid dihydrate, zinc acetate dihydrate, hydrochloric acid, ammonia, and absolute ethanol20Spherical[39]
SonochemicalPotassium hydroxide, zinc nitrate hexahydrate, and cetyltrimethylammonium bromide200–400 wide, a few nm thickFlakes[68]
Chemical vapor depositionZinc acetate dihydrate, ethanolAverage D: 90 and L: 564Nanorods[69]
ElectrochemicalOxalic acid dihydrate purified, Zn electrode, potassium chloride, nitric acid, and sodium hydroxide Dspherical: 50–100
Lcylindrical: 150–200
Spherical and cylindrical particles[70]
Table 2. Some of the major ZnO NPs vendors.
Table 2. Some of the major ZnO NPs vendors.
Sr. No.CompanyPhysical CharacteristicsPhaseCountryReference
1Meliorum Technologies Inc.10 nmPowder, aqueous, dispersionUnited States[71]
2Sukgyung AT Co., Ltd.10–20 nm, 20–40 nmPowderKorea, Republic[72]
3US Research Nanomaterials, Inc.10–30, 18, 20, 35–45, 30–40, 50–80, 80–200 nmPowder, dispersionUnited States[73]
4SkySpring Nanomaterials, Inc.10–30 nm, <30 nm, 200 nm, 200–800 nmPowderUnited States[74]
5Stanford Advanced Materials17–27 nm, 30–50 nm, 70–90 nmPowderUnited States[75]
6Ultrananotech≥20 nmPowderIndia[76]
7Advanced Nano Products 20–30 nmPowder, emulsionKorea, Republic[77]
8MKnano20, 30, 40, 50–150 nmPowderCanada[78]
9Nanophase™ Technologies20, 40, 60 nm (elongated)Powder, dispersionUnited States[79]
10Inframat® Advanced Materials™~30 nmPowderUnited States[80]
11TECNAN30–40 nmPowderSpain[81]
12Micronisers30–50 nmPowderAustralia[82]
13EPRUI Nanoparticles & Microspheres Co. Ltd.30, 50, 200 nm (nearly spherical)PowderChina[83]
14American Elements≤40 nmPowder, dispersionUnited States[84]
15Linari NanoTech45 ± 5 nmSuspensionItaly[85]
16Sigma Aldrich®<50 nm, <100 nm, <110 nm, <130 nmPowder, dispersionUnited States[86]
17Nyacol® Nanotechnologies, Inc.50, 125 nmSuspensionUnited States[87]
Table 3. ZnO NPs, (bio)polymers and incorporation/deposition methods to produce ZnO NPs based (bio)polymer composites.
Table 3. ZnO NPs, (bio)polymers and incorporation/deposition methods to produce ZnO NPs based (bio)polymer composites.
Characteristics-ZnO NPsSupplierPolymerBiopolymerSupplierIncorporation MethodReference
Particle size <100 nmSigma Aldrich (St. Louis, MO, USA)Linear low density polyethylene (LLDPE)-Equate Petrochemical Co. (Kuwait)Extrusion blowing method[28]
Particles size 100–500 nmPylote SAS in Dremil-Lafage, France-Poly(lactic acid) (PLA)NatureWorks LLC (USA)Twin-screw extrusion[27]
Rod-like particles, ZnO content: 96.2 ± 0.5% Umicore, Belgium-PLANatureWorksMelt compounding[22]
Average particle size <25 nmSigma Aldrich-ChitosanSigma AldrichNA[91]
Particle size <100 nm Sigma Aldrich-Poly(3-hydroxybutyrate-co-3-hydroxyvalerate) (PHBV)Goodfellow Corp.Solution casting[23]
Particle size <100 nmSigma Aldrich-Poly(3-hydroxybutyrate) (PHB)Biomer Ltd. (Krailling, Germany)Solution casting[24]
Particle size 60 nm, spherical in shapeCo-precipitation method using zinc acetate and sodium hydroxide precursors-Poly(butylene adipate-co-terephthalate) (PBAT)BASF Ltd., JapanSolution casting[25]
Nanopowder, <50 nm particle sizeSigma Aldrich Co. (St. Louis, MO, USA)Olive flounder (Paralichthys olivaceus) bones-A seafood restaurant in Daejeon, KoreaNA[92]
Particle size 30 nmUmicore, Belgium-PLA Ingeo™ 4043DNatureWorks LLC (Minnetonka, MN, USA)Solvent casting[20]
An average particle diameter of 70 nmPars Nanonasb (Persia)Low density polyethylene (LDPE)-Petkim (Turkey)Melt blending method[93]
An average particle diameter of 70 nmPars Nanonasb (Persia)Polypropylene (PP)-Borealis (Vienna, Austria)Melt blending method[94]
NanorodsCatalyst-free combust oxidized mesh process-Semolina flourLocal market in Tehran, IranSolvent casting[21]
Purity = 99.9%MaiKun Industrial Co., Ltd. (Shanghai, China)-PLANatureWorks LLC (Lincoln, NE, USA)Solvent volatilizing method[29]
Particle size <100 nmUmicore, BelgiumBovine skin gelatin type-B (BSG)-Sigma-Aldrich (St. Louis, MO, USA)Solution casting[26]
Table 4. Barrier properties of ZnO NPs based (bio)polymer composites.
Table 4. Barrier properties of ZnO NPs based (bio)polymer composites.
ZnO NPs Based CompositesOxygen Transmission Rate (OTR)/Oxygen Permeability (PO2)Water Vapor Permeability (WVP)Carbon Dioxide Permeability (PCO2)Reference
LLDPE films reinforced with ZnO NPsOTR decreased by 23.2% for 10 wt% ZnO incorporationNANA[28]
PLA/ZnO biocompositeFor 1 wt% ZnO incorporation, oxygen permeability (PO2) decreased by 18%. Then there is no further decrease for addition up to 5 wt%For 1 wt% ZnO incorporation, water vapor permeability (WVP) increased by 16%. Then there is no change for higher ZnO content.For 1 wt% ZnO incorporation, carbon dioxide (CO2) permeability decreased by about 17%. Then there is no further decrease for higher ZnO content.[27]
PLA-ZnO nanocomposite filmsNAWVP decreased on increasing ZnO NP concentration from 1 to 3 wt%NA[22]
Carboxymethyl cellulose-chitosan-ZnO NPs nanocompositesNAWVP decreased on increasing ZnO NP concentration up to 2 wt%NA[91]
ZnO-reinforced PHBV bionanocompositesPO2 decreased up to 35% with 4 wt% ZnO loading NANA[23]
ZnO/PHB bionanocompositesPO2 decreased by about 53% at 5 wt% ZnO NPs loadingWVP decreased by up to 38% at 5 wt% ZnO NPs loadingNA[24]
ZnO/PBAT nanocomposite filmsLowest value of OTR observed for 10 wt% ZnO NPs loading (for 0-10 wt% ZnO NPs loading range)NANA[25]
An olive flounder bone gelatin-ZnO nanocompositeNAWVP decreasedNA[92]
Untreated and 3-methacryloxypropyltrimethoxysilane treated ZnO nanoparticle reinforced-PLA nanocompositesPO2 values of plasticized PLA film reduced by 36.07 and 55.1% with the incorporation of 10% ZnO (untreated) and ZnO (3-methacryloxypropyltrimethoxysilane treated) NPsNANA[20]
ZnO based LDPE nanocompositesOTR decreased by 17% on adding 5 wt% ZnO NPsWVTR decreased by 22% on adding 5 wt% ZnO NPsNA[93]
ZnO based PP nanocompositesOTR decreased by 22% on adding 5 wt% ZnO NPsWVTR decreased by 12% on adding 5 wt% ZnO NPsNA[94]
Semolina reinforced with nanofillers (ZnO-nanorod/nano-kaolin)PO2 decreased by up to 66%NANA[21]
ZnO based PLA nanocompositeDecreasedIncreasedNA[29]
Bovine skin gelatin type-B (BSG) composite films incorporated with ZnO nanorods and clove essential oil (CEO)PO2 decreased by 32.27% with the addition of 2 wt% ZnO NPsNANA[26]
Table 5. Mechanical properties of ZnO NPs based (bio)polymer composites.
Table 5. Mechanical properties of ZnO NPs based (bio)polymer composites.
ZnO NPs Based CompositesEffect on Mechanical PropertiesReference
LLDPE films reinforced with ZnO NPsElongation at break: decreased; Tensile strength: increased[28]
PLA/ZnO biocompositeStress at yield: higher in the machine direction; Elongation at break: higher in the machine direction[27]
PLA-ZnO nanocomposite filmsYoung’s modulus: slightly increased; Elongation at break: decreased[22]
ZnO-reinforced PHBV bionanocompositesYoung’s modulus: increased by ~57% on 4 wt% ZnO NPs loading; Strain at break: decreased by ~30% on increasing ZnO NPs content[23]
ZnO/PHB bionanocompositesTensile strength: increased up to 32%; Young’s modulus: increased up to 43%; Impact strength: increased up to 26%[24]
ZnO/PBAT nanocomposite filmsElongation at break: increased; Tensile strength: increased[25]
An olive flounder bone gelatin-ZnO nanocompositeElongation at break: decreased by ~37%; Tensile strength: increased[92]
Untreated and 3-methacryloxypropyltrimethoxysilane treated ZnO nanoparticle reinforced-PLA nanocompositesTensile strength: decreased in ZnO NPs (untreated) composites; increased in ZnO NPs (3-methacryloxypropyltrimethoxysilane treated) composites; Elongation at break: decreased in ZnO NPs (untreated) composites; a marginal drop in ZnO NPs (3-methacryloxypropyltrimethoxysilane treated) composites[20]
ZnO based LDPE nanocompositesElongation at break: decreased; Tensile strength: decreased[93]
ZnO based PP nanocompositesElongation at break: decreased; Tensile strength: decreased[94]
ZnO based PLA nanocompositeElongation at break: increased; Tensile strength: decreased[29]
Bovine skin gelatin type-B (BSG) composite films incorporated with ZnO nanorods and clove essential oil (CEO)Elongation at break: increased; Tensile strength: increased[26]

Share and Cite

MDPI and ACS Style

Abbas, M.; Buntinx, M.; Deferme, W.; Peeters, R. (Bio)polymer/ZnO Nanocomposites for Packaging Applications: A Review of Gas Barrier and Mechanical Properties. Nanomaterials 2019, 9, 1494. https://doi.org/10.3390/nano9101494

AMA Style

Abbas M, Buntinx M, Deferme W, Peeters R. (Bio)polymer/ZnO Nanocomposites for Packaging Applications: A Review of Gas Barrier and Mechanical Properties. Nanomaterials. 2019; 9(10):1494. https://doi.org/10.3390/nano9101494

Chicago/Turabian Style

Abbas, Mohsin, Mieke Buntinx, Wim Deferme, and Roos Peeters. 2019. "(Bio)polymer/ZnO Nanocomposites for Packaging Applications: A Review of Gas Barrier and Mechanical Properties" Nanomaterials 9, no. 10: 1494. https://doi.org/10.3390/nano9101494

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop