Explore

Communities in English

Advertise on Engormix

Copper nanoparticles as an alternative feed additive in poultry diet: a review

Published: June 2, 2023
By: Abdullah Scott 1, Krishna Prasad Vadalasetty 1, André Chwalibog 1 and Ewa Sawosz 2 / 1 Department of Veterinary and Animal Sciences, University of Copenhagen, 1870 Frederiksberg C, Denmark; 2 Department of Animal Nutrition and Biotechnology, Warsaw University of Life Sciences, 02-786 Warsaw, Poland.
1 Introduction
Copper (Cu) is a crucial trace element in animals; however, it cannot be stored in the body, thus, a regular dietary supplement is required. In addition, feed ingredients are commonly deficient in Cu; hence, the commercial diet should provide the essential amount of Cu in a biologically dynamic form, which depends on the physical and chemical properties of the form of the supplement in which the Cu is given in the diet [1].
Cu sulphate (CuSO4) is the main Cu source in the diet of chickens and other animals; however, the inorganic salt shows poor bioavailability caused by the presence of ingredients that can inhibit absorption. Thus, inorganic mineral administration in animal feed poses a risk to the environment, as excretion of high mineral levels contaminates the soil and water.
Organic sources of Cu show a reduction in excretion compared to inorganic forms, due to greater bioavailability and stability in the upper gastrointestinal tract of chickens; however, due to high cost and lower supplementation doses than inorganic Cu, results are still not consistent [2]. Therefore, efforts have been made to find alternative sources of Cu to increase bioavailability and absorption, and to avoid causing harmful effects on the health and performance of chickens.
Different sources such as Cu chloride, Cu oxide, Cu citrate, Cu sulphate and tribasic Cu chloride at different concentrations have been applied in poultry feed, depending on Cu bioavailability [3]. However, the feed industry still prefers CuSO4 for economic reasons.
Nanotechnology has modernised the commercial application of nanosized minerals, which have been used recently as a tool in the fields of biology, biotechnology, mineral nutrition, physiology, reproduction and pharmacology in animals [4].
Nanoparticles refer to a particle size of 1–100 nm, in which the physical, chemical and biological properties of materials differ fundamentally from their bulk form [5]. Furthermore, their small size increases the potency of active ingredients and potentially reduces the applied quantity [6].
The use of Cu nanoparticles (Cu-NP) has recently received much attention due to their high electrical and melting points, low electrochemical migration behaviour, high bioavailability and relatively low cost of production [7].
The reduced size and increased surface area of Cu-NP compared to the bulk material improve thermoconductivity and fluid viscosity, leading to a variety of applications like chemosensors, surfactants and antimicrobials [8]. However, little work has been done to explore the efficacy of Cu-NP as feed additives in animal feeding and particularly in poultry. It was hypothesised that Cu-NP, because of their high physical reactivity, could be used as an alternative, effective health and growth promoters in much smaller doses than bulk minerals in animal feed [9–11], in consequence, significantly reducing the excretion of these minerals into the environment. Moreover, Cu-NP are effective antibacterial, antifungal and antiviral agents [12]. However, their size might affect their toxicity by helping the cellular uptake and translocation of the particles in the animal’s body [13]. Therefore, the objectives of the review are to present the current knowledge about Cu-NP physicochemical properties, nutritional characteristics, antimicrobial activity, immunological and toxic effects, and the possibility of using Cu-NP as an alternative growth-promoting supplement in animal diets.
2 The importance of Cu in the diet
Cu is a key element required for animal growth and development of bones, connective tissue, the heart and several other organs [14]. Furthermore, Cu is involved in the stimulation of the immune system to combat infections and repair injured tissues [15]. Additionally, it supports neutralisation of free radicals that cause severe injury to cells [16]. Cu is shown to be absorbed partially by the stomach; however, the majority is absorbed in the small intestine. Cu is mostly excreted via bile that is released into the gastrointestinal tract (GIT), with marginal Cu reabsorbed by intestinal cells [17].
The biological functions of Cu are associated with Cu’s role in the active site of metalloenzymes. Cu is found in a large number of metalloenzymes such as cytochrome oxidase, superoxide dismutase (SOD), lysyl oxidase, dopamine hydroxylase and tyrosinase [18]. Cu is also needed for the development of antibodies and white blood cells, in addition to antioxidant enzyme production [19].
Ceruloplasmin is a Cu transport protein, which provides Cu to the cells and performs as an enzyme, which possesses oxidative activity [20]. However, organisms have developed a variety of antioxidant defences that include metal-sequestering proteins to protect against oxidative damage, using compounds such as vitamins C and E, and specified antioxidant enzymes [21].
A study conducted on mink [22] showed that Cu affects the regulation of ceruloplasmin concentration, SOD activity and the digestion of dietary fat. The SOD enzyme helps in removing the damage caused by reactive oxygen species (ROS) by catalysing the dismutation of two superoxide radicals to hydrogen peroxide and oxygen [20].
Cu is also important in the body as a component of enzyme systems involved in iron metabolism, red blood cell formation and immune function. Moreover, Cu supports the formation of connective tissues such as collagen and elastin, and enhances nervous system development by synthesis of dopamine [23].
Cu is involved in cellular respiration, energy production, synthesis of collagen, uptake and utilisation of other trace minerals, antioxidant activity, cardiac function, bone formation, keratinisation and pigmentation of tissue and myelination of the spinal cord [24]. In addition, it contributes to the regulation of glucose and cholesterol metabolism [23].
A deficiency of Cu supplementation in the diet can cause disturbances in reproduction and development of sperm, high mortality of embryos during hatching, poor pigmentation of feathers, slow growth and a reduction in body weight [23]. Furthermore, it may result in muscle weakness, anaemia, bone alterations that resemble scurvy, defective connective tissue synthesis, impaired myelinisation of nerve tissues and neurological defects, altered lipid metabolism and cardiac malfunction [23]. However, it has been demonstrated that ceruloplasmin mRNA in the liver is not affected by a dietary deficiency of Cu [25].
An excess of Cu may also have adverse effects on chicken performance [26]. Therefore, Cu must be provided to livestock in optimal concentrations and according to requirements that change during the growth and development of the animal.
The essentiality of Cu depends on the ability of Cu atoms to gain and lose electrons to form cuprous Cu+1 and cupric Cu+2 states. Therefore, this alteration is crucial for enzymes to support the metabolism of all major substrates such as proteins, lipids and carbohydrates [27].
The oxidation of cuprous to cupric Cu can generate free radicals from the oxidation of lipids, proteins and nucleic acids, causing extensive functional damage in tissues and organs. Consequently, organisms have developed systems involving specific protein carriers, which regulate the absorption, distribution, use and excretion of Cu, thereby, controlling the production of free Cu ions and minimising the risk of oxidative damage [27].
The assessment of Cu status in chickens can be examined using several methods. However, the most subtle indicator of Cu status is the concentration of Cu in the liver [28]. Liver Cu concentration in animals is correlated to the bioavailability of Cu in the diet [29]. In addition, when Cu intake is below physiological need, the stored Cu in the liver is released; thus, a reduction in liver Cu can be a sign of low Cu intake. The concentrations of Cu in feed ingredients and their bioavailability vary because of the differences in agronomic conditions and processing [30]. Thus, various sources of supplemental Cu (oxide, citrate, sulphate and tribasic Cu chloride) have been included in animal diets [31, 32].
The bioavailability of trace minerals for animals is defined as the degree to which the ingested trace minerals from a particular source are absorbed in a form that can be metabolised by the animal [33]. The emphasis on the bioavailability of Cu has become a concern of researchers because the use of CuSO4 as the main source in feed brings some adverse effects like the interaction with other ingredients, some physicochemical problems owing to high hygroscopicity and chemical reactivity [34]. Furthermore, it is an acidic compound and a strong electron acceptor. Hence, these properties can lead to solidification of the mixture in addition to oxidation, destroying its value and integrity. In chicken feed, the moisture content is about 20%, which can make CuSO4 crystal surfaces more concentrated for reactions that break down labile organic compounds such as vitamins, enzymes, fats and oils [35]. Many studies have compared the bioavailability of different Cu forms to CuSO4. It has been suggested that Cu citrate is more efficient than CuSO4 [36]. Moreover, Ledoux et  al. [37] showed the relative bioavailability of cupric oxide, cupric carbonate and cupric sulphate, using supplementation levels of 150, 300 and 450 mg Cu/ kg feed, based on Cu levels in the liver; the relative biological availability was estimated to be 88.5%, 54.3% and 0% for the sulphate, carbonate and oxide, respectively. Feeding supplementation up to 390 mg/kg feed [38] suggested that tribasic Cu chloride (TBCC) has a relative value of 134% compared to CuSO4, while Miles et al. [39] proposed a value of 112% for TBCC. In terms of liver Cu accumulation, TBCC was 109% as valuable as sulphate [34], which suggests that TBCC is less harmful to vitamins both in the feed and within the bird. The liver and plasma vitamin E levels were higher in birds fed TBCC compared to sulphate using a diet of 36 IU/kg. Additionally, TBCC and cupric sulphate (220 and 180 mg Cu/kg feed, respectively) equally improved the carcass weight of broiler chickens after 45 days of feeding, compared to the control group (30 mg Cu/kg) [40].
Alternatively, the organic form of Cu (chelated with amino acids, peptides, or proteins) appears to provide a pathway for minimising Cu level in the excreta [41]. It has been suggested that chelated Cu is better absorbed and metabolised, and may prevent antagonism of Cu with other minerals [42]. Furthermore, it has been investigated that chelated Cu can replace CuSO4 as a growth promoter for pigs [43].
Cu absorption in the stomach is lower than in the intestine and is further delayed by binding with phytic acid and forming insoluble compounds. This was reported when CuSO4 added at 250 mg/kg of feed caused a reduction in phosphor retention in broilers, by forming an insoluble Cu-phytate chelate complex [44]. In addition, it was demonstrated that organic Cu is more effective than CuSO4 in decreasing the plasma cholesterol of broilers [45]; however, a few studies have found that organic Cu is less efficient than CuSO4 [46].
Another way of predicting Cu availability in chickens was investigated by the solubility of different Cu combinations in several solutions for their ability in chicken liver [47]; the authors found that solubility at pH 2 was the best predictor. The same authors found that Cu chelated with lysine or amino acid is more bioavailable than CuSO4.
It has been demonstrated that supplementation with 4  mg/kg feed of Cu as a cupric chelate of amino acid hydrate may be sufficient for normal broiler growth to 29 days of age [48], while Jegede et al. [49] demonstrated a significantly higher daily weight gain in broilers fed Cu proteinate compared to cupric sulphate supplementation. Moreover, supplementation (4–8 mg/kg) with Cu glycine chelate did not affect the Cu concentration in the liver, but reduced the Cu concentration in the broiler faeces compared to CuSO4 [50].
Assessing the bioavailability of Cu proteinate relative to CuSO4 in broilers was studied by Liu et al. [51]. The estimated relative bioavailability of Cu proteinate was 78.8% and 79.3%, respectively, compared to that of CuSO4; however, the differences were not significant. Furthermore, it was demonstrated that the substitution of CuSO4 with Cu proteinate in chicken diet positively affects body gain and feed conversion ratio [52].
The outcome [53] indicated that Cu proteinate, compared to CuSO4, may enhance the detrimental effect of aflatoxicosis on broiler chickens, improving the growth performance by feeding a diet containing a high level of aflatoxin
Assessing 100 mg/kg feed (cumethionine or cuproteinate) as an alternative to antibiotics in broiler diet showed that both Cu sources improved the growth performance of broilers, and the enhancement was comparable with supplementation of an antibiotic (avilamycin), which increased populations of lactobacilli while decreasing Escherichia coli in the intestine [54]. On the other hand, Kwiecień et al. [55] did not reveal any differences in performance indices of chickens supplemented with CuSO4 or Cu glycinate chelate, but organic Cu positively affected the biomechanical properties of femur bones.
Diet supplementation with 50, 100 and 150 mg/kg levels of Cu proteinate or CuSO4 had no significant effect on growth; however, Cu proteinate decreased plasma cholesterol, low-density lipoprotein and triglyceride in comparison to CuSO4 in pullets [56].
Moreover, the comparison between CuSO4 and Culysine added to duck diet (4, 8, 12 and 150 ppm Cu) showed that the Cu source had no significant effect on weight gain, while feed intake was increased and feed efficiency was degraded by dietary organic Cu lysine. Compared to CuSO4, Cu lysine increased Cu excretion, Cu concentration in the liver and plasma cholesterol concentration, but decreased plasma triglycerides. Although organic Cu showed some improvement in chicken performance compared to inorganic forms, the results were not always consistent [57].
The effect of different dietary Cu sources on serum trace mineral and cholesterol status in broiler chickens revealed that higher Cu content with lower iron and manganese content in serum was noticed by increasing Cu levels in the diet [58]. Furthermore, the Cu propionate source had a significantly higher serum Cu compared to CuSO4. The provision of CuSO4 brought significantly higher serum zinc compared to that of Cu propionate; however, the total cholesterol content in serum was not affected by Cu diets. Excess Cu supplementation in broiler chickens showed that Cu increased haemoglobin, decreased plasma cholesterol and triglyceride significantly, decreased plasma proteins and its fraction did not change due to Cu supplementation [59]. From the studies mentioned above, it can be determined that Cu supplementation has an effect on the blood levels and erythropoietic system of the chicken, which could be used as an indicator of the impact of its toxicity in chickens. Furthermore, changes in the peripheral blood enable more accurate evaluation and explanation of the effect of Cu on the chicken’s body.
The growth-promoting function of Cu is related to the growth hormone axis [60] and hypothalamic appetite-regulating genes [61]. Moreover, it has been shown that adding Cu stimulates mitogen activity and also enhances the growth hormone (GH) mRNA level in the pituitary glands of pair-fed pigs. However, the production of GH is known to be affected by many factors. The influence of Cu on GH gene expression can be through either direct action on pituitary cells or indirect action on the factors influencing GH gene expression [62]. Therefore, stimulation of growth in chickens by GH association with Cu deficiency is still not clear.
It can be concluded that Cu is present within chicken tissues in very small and regular amounts, but it plays an essential role in chicken growth, acting as a catalyst in enzyme systems within cells [63]. However, a relatively constant concentration of Cu in the body of chickens suggests that the content of Cu increases with increasing body weight [2]. A deficiency of Cu will certainly affect chicken growth, while an excess of Cu is not recommended because either it will be excreted or will have an adverse effect on performance.
3 Physicochemical characteristics of Cu-NP The Cu-NP
possesses ultraviolet-visible sensitivity and electrical, catalytic, thermal and antibacterial properties due to their large surface-to-volume ratio [64]. Many atoms are present on the surface due to the smaller particle size. The surface-to-volume ratio of particles differs and depends on the shape and size of the nanoparticles. The characterisation of Cu-NP including the electronic energy levels, electron affinity, electronic transitions, magnetic properties, phase transition temperature, melting point and affinity for polymers, biological and organic molecules depends on the change in the surface area [65]. Consequently, functional activities such as the chemical, catalytic, or biological effects of nanoparticles are profoundly influenced by the particle size of the nanometals [66].
Quantum effects are due to a combination of quantum-size and Coulomb-charging effects that impart the charge to nanoparticles. When the Coulomb-charge effect is coupled with the quantum size, a range of properties are obtained that are not observed for the same bulk material. Quantum effects are prominent in spherical particles and in particles with sharp edges [67].
The Cu-NP can be characterised by different methods such as ultraviolet-visible absorption spectroscopy, X-ray diffraction [68], directly by scanning electron microscopy [69] for relatively coarse powders or transmission electron microscopy for fine powders [70, 71], atomic force microscopy [72, 73] and infrared spectroscopy. In addition, other methods have shown convenient ways of characterising Cu-NP such as dynamic light scattering, X-ray scattering at small angles and ultraviolet-visible spectroscopy (UVS) [12]. UVS is used because the absorption peak positions are dependent upon particle size and shape [74]. Cu-NP is normally absorbed from 280 to 360 nm. Ascorbic acid, a common reducing agent, provides a shorter peak from 240 to 280 nm. Nanoparticles are characterised by absorption spectroscopy with a peak at 580 nm. Infrared spectroscopy is used to characterise biomolecules interacting with Cu-NP [75]. Particles prepared by microwave irradiation are typically spherical and show surface plasmon resonance at 535 nm [76]. Moreover, Cu-NP prepared by microwave irradiation are influenced by the irradiation time; a linear relationship is observed between particle size and time [76]. In addition, ethers, alcohols and carbon-hydrogen bonds are responsible for the interaction of nanoparticles with biomolecules [77].
Many studies have shown different correlations between the various physicochemical properties of mineral nanoparticles linked with the health effects; the diversity of nanomaterials in terms of size, shape, aggregation and surface chemistry poses a challenge to those who are trying to characterise the animal health and environmental risks associated with incidental and unintentional exposure [78]. Evidence from several studies has revealed that several factors associated with sonication, such as temperature, sonication time, sonication methods, sonication power output, sample volume and concentration can influence the physicochemical properties of nanoparticle suspensions [79, 80]. In addition, other parameters like coating and surface roughness, solvent conditions, composition and crystalline structure will change depending on the nanoparticle preparation technique and the stage of formation [81, 82].
However, there are inconsistent reports of physicochemical properties of Cu-NP such as particulate core and outer shell chemical composition, surface oxidation state, surface charge, singlet and agglomerate sizes in relevant carriers, shape, solubility and surface area [83]. These properties could impact on the nanomaterial dose that reaches target organs. Moreover, it is believed that the surface properties are expected to change when Cu-NP enter the body and during transportation.
The morphology of Cu-NP is highly affected by the type of Cu salt used for synthesis. The Cu-NP is crystalline in nature and has strong antimicrobial activity against both, Gram-positive and Gram-negative bacteria. The different shape, size and strong antimicrobial activity of Cu-NP have higher prospective in the field of biomedical and food packaging [75].
The reducing agent affects the size of Cu-NP, as high concentrations may decrease the size while maintaining the concentration of the precursor [84]. It has been reported that an increasing concentration of a reducing agent causes a reduction in monodispersity and increases the number of nanoparticles [85]. Furthermore, it has been shown that the stabilising agent also affects the size of the nanoparticles. Micelle formation stabilises nanoparticles and makes the system monodispersive and stable in air.
The pH is the most important factor that affects the size of the nanoparticles, which increases with increasing pH of the reaction mixture. A marked increase in size appears above pH 5. It has been reported that the Cu-NP have a high solubility in an acidic environment (pH 5.5) with a high positive ζ potential [86, 87]. Cu-NP are not present at a higher pH; rather, the oxide form of the Cu-NP is present due to excess hydroxide. Further increases in pH produce Cu hydroxide without nanoparticle formation [85]. However, some works have reported the formation of Cu-NP at pH from 9 to 11. As the pH is increased, the concentration of the hydroxide increases, leading to the formation of Cu hydroxide. This explains the formation of Cu-NP in basic solution.
It has been revealed that various biological mechanisms including cellular uptake and efficiency of particle processing in the endocytic pathway are dependent on the size of the material [88–90]. The size and surface area of the particles will dictate how the system responds to, distributes and eliminates the materials [81]. However, the size of the Cu-NP could be also their main disadvantage, and it represents a challenge for the scientific community to achieve adequate physical and chemical characterisation [91].
The surface of the Cu-NP can influence cellular uptake, clearance and biocompatibility [77] and also the agglomeration of nanoparticles. Furthermore, the surface chemistry also determines in vivo dissolution rates and, therefore, contributes to distribution, retention and toxicity. Moreover, the surface composition of the Cu-NP can affect the generation of ROS via redox or catalytic activity [88, 92]. It has been found that the Cu-NP with longer chain ligands have surfaces that are better protected from oxidation and a corresponding lower ROS-generating capacity than the particles with shorter chain ligands. On the other hand, the Cu-NP with a greater surface oxidation also have a higher ROS-generating capacity [93]. The Cu-NP with high oxidant and cytotoxic potential may have more significant effects at the initial site of deposition and may also be cleared more rapidly due to an influx of inflammatory cells.
The surface charge of nanoparticles will influence nonspecific interactions with proteins that are present in the environment. These observations illustrate the difficulty in predicting nanoparticle uptake as a function of a single property.
The surface energy affects nanoparticle interaction with biomolecules and tissues [94]. Low-energy surfaces (hydrophobic) are particularly prone to nonspecific adsorption as proteins unfold to expose their hydrophobic core, which may disorganise the lipid components of cell membranes and enhance epithelial penetration [95]. High-energy surfaces (hydrophilic), particularly those that carry a weak negative or neutral charge, are ideal for resisting protein adsorption and cell uptake [96].
Different aspects of nanomaterials like selective adsorption of nanoparticles [97], colloidal behaviour, plasma protein binding [98], blood-brain barrier integrity and transmembrane permeability are primarily regulated by the surface charge of nanoparticles [99]. It is noteworthy that positively charged nanoparticles show significant cellular uptake compared to negatively charged and neutral nanoparticles, owing to their enhanced opsonisation by plasma proteins. As the surface charge is the main determinant of colloidal behaviour, it will affect the organism’s response upon exposure to nanoparticles by changing their shape and size through agglomerate formation [97].
The biological activity of Cu-NP depends directly on the physical and chemical properties that include size, shape, concentration, surface and charge [100]. However, these parameters change depending on the nanoparticle preparation technique and the stage of formation [101].
Cu-NP can be synthesised either by the mechanical grinding of bulk metals (physical method) or via chemical reduction of metal salts (nucleation and growth of metallic atoms). Recently, the green synthesis of Cu-NP has also emerged as a novel method and is gaining more importance among researchers [102, 103]. However, the synthesis of Cu-NP needs a variety of stabilisers like donor ligands, polymers and surfactants to prevent agglomeration [71].
In this regard, it is necessary to know the physical and chemical properties and antibacterial influence of the obtained Cu-NP to standardise this nanomaterial during subsequent use as a feed supplement for animals. In addition, the physicochemical characteristics of Cu-NP may affect the epithelial barrier penetration and have emerged as important determinants, including size, surface charge and surface energy [104]. Furthermore, the Cu-NP solubility and protein-binding capacity also play a key factor in the passage of the Cu-NP across epithelial barriers, cellular uptake, cytotoxicity and biodistribution [105, 106].
4 Biological properties of Cu-NP
4.1 Nutritional and physiological characteristics
The small size of the Cu-NP can increase the uptake from the GIT and, hence, make them more effective than the bulk Cu at lower doses [107]. In the animal body, the Cu-NP interacts more effectively with organic and inorganic materials due to their larger surface area [108]. The Cu-NP has the capability to cross the small intestine and distribute into the blood, brain, heart, kidney, spleen, liver and intestine [109]. In order to reach the target organs, the Cu-NP will interact first with the protective barriers of the GIT. Hence, those nanoparticles that reach the physiological barriers are highly determined by the properties of the particles and the barriers themselves [83]. Other properties of the Cu-NP like the composition, solubility and proteinbinding capacity also play a significant role in the passage of the nanoparticles across the epithelial barriers, cellular uptake, cytotoxicity and biodistribution. However, the current understanding regarding the passage of the Cu-NP across the epithelial tissue and the mechanisms of distribution and elimination is lacking.
The advantage of the Cu-NP is to prevent mineral dissociation with other nutrients and subsequently avoid antagonism. It is assumed that supplementing chicken feed with an adequate level of Cu-NP will help to reduce the Cu levels in the excreta, thereby, reducing environmental contamination. Many studies have compared the availability of the Cu-NP to CuSO4 when supplemented in the animal diet [110, 111], suggesting that the Cu-NP is better than CuSO4 in enhancing the growth and performance of animals.
Generally, nanoparticles can enter the GIT in many ways, such as ingestion directly from food and water, administration of therapeutic nanodrugs and oral delivery into the GIT; inhaled nanoparticles can also be swallowed and enter the GIT following clearance from the respiratory tract [112]. The uptake of particles in the GIT depends on diffusion and accessibility through mucus and contact with the cells of the GIT. Smaller particles will diffuse faster through the GIT mucus to reach the cells of the intestinal lining, followed by uptake through the GIT barrier to reach the blood. Uptake occurs variously by passive diffusion across the mucosal cells, via active transport mechanisms and intercellular [113]. It has been suggested that smaller particles that are capable of being taken up by the villus epithelium may directly enter the bloodstream, thereafter, being predominantly scavenged by the liver and the spleen [114].
Depending on the mass of insoluble nanoparticles, they can be more readily taken up across the intestinal barrier; therefore, they will be more bioavailable than their micro or macro equivalents [115]. However, the mechanism by which nanoparticles enter the cell has important implications for their impact on biological systems, and the challenge is to identify how an organism responds to these nanoparticles [116].
Nanoparticles can enter cells by transcytosis; particles are engulfed by Kupffer cells, of sizes 50–100  nm, resulting in smaller (2  nm) nanoparticles being filtrated out of the liver via the kidneys, whereas bigger particles (40 nm) are retained in the Kupffer cells [117].
According to the physicochemical characteristics of Cu-NP, uptake may occur via one of the different forms of endocytosis [106]. However, there is no evidence that those nanoparticles are taken up individually, and one may doubt the technical ability to capture this event. In the case of small nanoparticles, it is expected that uptake occurs only when a critical density is reached [118]. It has been reported that nanoparticles with a diameter up to several hundreds of nanometres preferentially enter cells via pino- or macropinocytosis; in addition, clathrin-dependent uptake is a route for nanoparticles on whose surface serum proteins are adsorbed. Finally, negatively charged nano-objects enter mostly by caveolin- or clathrin-mediated endocytosis [106, 119].
Small drug molecules enter the cell through passive diffusion, whereas most nanomaterials are taken up by active processes such as phagocytosis or pinocytosis depending on a dynamic series of physicochemical properties [120].
The uptake of nanoparticles is determined by particle solubility, concentration and size. However, the passive uptake of particles has been shown to occur transcellularly via the intestinal lining and, to a lesser extent, between epithelial cells [121].
Nevertheless, it is not known whether Cu-NP remain in the intestinal tract unabsorbed and are excreted in the faeces or go into the body system readily [122]. Some evidence indicates that particles smaller than 100  nm are absorbed in various tissues and organs [123]. Uptake of Cu-NP across these defensive barriers occurs through various transport mechanisms including receptor-mediated endocytosis and adsorptive endocytosis [124].
Particles taken by the villus epithelium can enter the bloodstream and, then, mostly will be foraged by the liver and spleen [114, 125].
Increasing the uptake of Cu-NP in the GIT might have an impact on the growth and health of animals. Cu-NP have been reported to enhance growth performance and improve feed utilisation compared to CuSO4 when provided as a feed supplement for piglets, poultry and fish [9, 126, 127]. The improvement was attributed to the better bioavailability of Cu-NP compared with CuSO4 salts. However, the mechanism behind this improvement is still not clear. Some studies indicated that the effect of Cu-NP can be ascribed to their antibacterial properties [40], while others suggested the better digestibility of energy and fat [9, 11]. Furthermore, some studies demonstrated that the activity of SOD was enhanced with Cu-NP supplementation in animal diet [128, 129].
It was reported that adding Cu-loaded chitosan nanoparticles to broiler feed enhanced growth performance and immunological capacity, influenced intestinal microbiota and improved protein synthesis [130]. Further, it has been reported that there is a connection between arginine levels in the liver and the growth of chickens, which was improved after one intramuscular injection with Cu-NP [131]. Moreover, in ovo injection of Cu-NP to chicken eggs has been found to improve the performance of chickens in the finishing period [110]. A similar effect was shown in fish treated with 2 and 4 mg/kg of Cu-NP [127].
Generally, there is some alteration in blood biomarkers after Cu provision in the animal diet. Accordingly, an increase in albumin, alanine transferase (ALT) and uric acid, but a reduction in alkaline phosphatase, was shown in chickens fed 100–400 mg/kg of CuSO4 [132]. Similarly, it was observed that in ovo injection of 50 mg/kg of Cu-NP decreased the levels of ALT, glucose and cholesterol [23]. The same author demonstrated a non-significant increase in calcium, magnesium and phosphorus levels in both Cu-NP and CuSO4 groups compared to the control. However, Scott et al. [11] did not indicate any significant alteration in the blood of hatchlings after administering 50 mg/kg of both forms of Cu.
Compared to CuSO4, dietary supplementation with Cu-NP causes a significant reduction in glucose levels in fish [127]. Additionally, Canli and Canli [133] revealed that 5 mg/kg of oxide Cu-NP significantly decreased glucose, total cholesterol and triglyceride levels in rats following oral administration compared to the control. This could be attributed to hyperglycaemia resulting from cortisolmediated gluconeogenesis [134].
The reduction of cholesterol levels after adding Cu to a poultry diet [135, 136] can be explained by Cu supplementation regulating cholesterol biosynthesis indirectly by decreasing the reduced form of glutathione and increasing the oxidised form of glutathione [137, 138].
Blood urea also shows some reduction with Cu supplementation in the diet. This might be an indicator that amino acids are utilised more efficiently for growth [139].
Similarly, it has been shown that supplementation of Culoaded chitosan nanoparticles in broiler feed alters the concentration of urea in the blood [130], which could be an indicator of the rate of protein synthesis [140].
A study by Mroczek-Sosnowska et al. [23] showed that in ovo injection of CuSO4 increased counts of white blood cells (WBCs), lymphocytes and eosinophils compared to Cu-NP. However, Cu-NP decreased the number of leucocytes compared to the CuSO4 group. Additionally, adding Cu significantly altered the number of erythrocytes in the peripheral blood of the birds. Similar results were observed in turkeys treated with Cu-lysine chelate in drinking water; however, it caused a reduction in the levels of haematocrit and haemoglobin [141].
Additionally, the concentration of haemoglobin was increased, which allowed the birds to keep oxygen transport at an appropriate level [23]. The increased level of haemoglobin could be due to its continued synthesis by erythrocytes already circulating in the peripheral blood. It is also likely that the increased level of haemoglobin could be linked with the homeopathic function of Cu. Cu has been shown to directly stimulate erythrocyte synthesis, as it determines iron absorption into the body and its incorporation in haemoglobin [142].
On the other hand, it was demonstrated that CuSO4 reduced the percentage of monocytes, which suppressed the metabolic activity of these cells. A reduction in phagocytic activity was also observed in the chickens, which indicates that CuSO4 could contribute to accelerated damage of phagocytes and inhibit the metabolic activity of the cells [23], while the administration of Cu-NP caused an increase in the blood level of uric acid.
The addition of the oxide form of Cu-NP to broiler chicken feed indicated that haemoglobin and albumin concentrations were reduced in comparison to the nontreated group [143]. However, the same authors showed some protein denaturation, which suggests that the oxide form of Cu-NP might have a toxic effect on blood proteins.
4.2 Antibacterial activity of Cu-NP
Chickens frequently receive additional Cu as a growth-promoting supplement. However, it has been shown that bacteria can develop resistance to Cu inside a chicken’s body, particularly enterococci, which are often associated with resistance to antimicrobial drugs like macrolides and glycopeptides [144]. Such resistant bacteria can be transferred from food-producing animals to humans. The resistance is more frequent at high Cu levels. Therefore, Cu supplementation for growth promotion is, today, one of the crucial health and environmental concerns that should be optimised. Although the misuse of antibiotics in animal feed is partially liable for the increased level of antibiotic resistance in bacteria, exposure to trace metals may also contribute to antibiotic resistance, even in the absence of the antibiotics, themselves [145]. However, information regarding mechanisms for inducing Cu resistance in bacteria and the dose-response relationship is limited.
Therefore, with the most concerning issue of microorganism resistance to antibiotics, developing new nanoagents with antibacterial effects is of great theoretical and practical concern.
Previously, the bactericidal effects of Cu-NP on strains of E. coli and Staphylococcus aureus were demonstrated [146]. Furthermore, the antibacterial activity of Cu-NP depends on the physical characteristics of the nanomaterial [101].
Cu-NP have been used for a long period as disinfectants due to their antibacterial properties [147]. Although the mechanism of action as an antibacterial agent has still not been clarified, reports have suggested that the mechanism of antibacterial activity of Cu-NP could be the same as that of silver nanoparticles [70].
Nanoparticles are capable of attaching to the bacterial membrane by electrostatic interaction and disturbing its integrity [148]. Thus, the antibacterial activity of nanoparticles is generally triggered by inducing ROS following the administration of Cu-NP [149]. Generally, three mechanisms of action of nanoparticles have been hypothesised. First, the accumulation of nanoparticles in the bacterial membrane, changing its permeability, with subsequent release of lipopolysaccharides, membrane proteins and intracellular biomolecules and dissipation of the proton motive force across the plasma membrane [150, 151]. Second, generating ROS and their corresponding ions from nanoparticles, with subsequent oxidative damage to cellular structures [152, 153]. Finally, the uptake of metallic ions derived from nanoparticles into the cells, followed by depletion of intracellular ATP production and disruption of DNA replication [154, 155].
The filamentation is caused by Cu-NP-mediated depolarisation of the cell membrane, while cell damage is caused by Cu-NP-mediated ROS generation in cells, resulting in lipid peroxidation, protein oxidation and DNA degradation [156].
The linoleic acid-capped Cu-NP are highly bactericidal for S. aureus, E. coli and Bacillus subtilis, suggesting that Cu-NP can be used as effective growth inhibitors in various microorganisms, making them applicable in diverse medical devices and antimicrobial control systems [157].
Based on oxidative stress, Chang et al. [149] discussed the coordination and non-homeostasis effects of Cu oxide nanoparticle toxicity on eukaryotic cells. The authors described that nanoparticles can diffuse into the cell directly through the pores present in the cell membrane or through ion channels and transporter proteins present on the plasma membrane, while other nanoparticles may enter the cells via endocytosis. Eventually, the nanoparticles that enter the cell may directly interact with other organelles.
The Cu-NP showed the most efficient form among the different Cu forms for antimicrobial activity due to their high surface area [158, 159]. Besides that, the crystal morphology of the Cu-NP results in several reactive sites, which simplifies the interaction between the Cu-NP and the microbial cell membrane [160]. Therefore, many studies have pointed out the antibacterial action of the metal nanoparticles [161–165].
Many investigations have focused on the bactericidal activity mechanism of the Cu-NP [156, 166]. They indicated that the antibacterial activity of the Cu-NP results from the ions released by the nanostructures. However, Ruparelia et al. [70] showed that Cu2+ ions are released because of a reaction with the nutrient components, suggesting that Cu ion release may be related to reaction with chloride from the nutrient or to the presence of a thin oxide layer on the nanoparticle surface.
The bactericidal activity of Cu-NP results not only from their particle size, high specific surface value and close interaction with microbial membranes but also from the formation of leached cuprum-peptide complexes that lead to an increase in ROS generation and decreased cell viability [167].
Lately, a study demonstrated that the antimicrobial properties of chitosan implanted with Cu-NP reduce gut bacteria such as E. coli, Enterococcus faecalis, S. aureus and, particularly, Lactobacillus fermentum, which is one of the primary targets of antibiotic growth promoters, suggesting that the Cu-NP could be used to minimise undesirable levels of microbial populations without causing cytotoxicity [168].
The antimicrobial activity of Cu seems to be similar to that of antibiotics, altering the gut microbiota and, hence, reducing fermentation loss of nutrients [169]. The reduced level of bacteria in the proximal part of the GIT may benefit the chicken by allocating more feed components for its growth performance. CuSO4 reduces the quantity of coliform in the large intestine, which might be a part of other mechanisms such as the inhibition of specific pathogens and inducing resistance of the chicken to pathogen adhesion and invasion as well as toxins [170].
Supplementing chicken diet with additional Cu reduces the population of clostridia in the GIT of chickens. Generally, the concentration of Cu salts that affect the populations of coliform bacteria is higher (up to 200 mg/kg) than Cu-NP (below 50  mg/kg). While the population of lactobacilli in the GIT contents was stimulated by 150  mg/kg of Cu, the higher concentration did not reveal any effect [27].
Cu may be toxic to bacteria; however, this depends on the solubility of the metal compounds under physiological conditions. To avoid cellular toxicity, bacteria have developed mechanisms of metal tolerance [171]. Furthermore, antimicrobial agents have more proficiency than antibiotics to support a co-selection process, ultimately selecting for antibiotic resistance [172].
Cu-NP is effective against Gram-positive and Gramnegative bacteria [173, 174].
Further, Cu-NP inhibits the growth of S. aureus, B. subtilis and E. coli bacteria [175], Micrococcus luteus, Klebsiella pneumoniae and Pseudomonas aeruginosa [176]. The subsistence rate of E. coli and B. subtilis bacteria is decreased by increasing Cu-NP concentrations [177]. However, higher susceptibility is shown to Gram-negative strains [178]. Furthermore, the activity of Cu-NP against E. coli has been demonstrated [179, 180].
The activity of Cu-NP depends on temperature, aeration, pH and concentration of nanostructures [181]. The main biocidal activity of Cu-NP is caused by oxidative damage to cellular structures and inhibition of DNA replication and amino acid synthesis in microorganisms [182]. The mechanism relates to the dissipation of the cell membrane under the influence of nanoparticle accumulation or the generation of ROS by highly concentrated Cu2+ ions [156].
A study by Ramyadevi et al. [176] carried out with M. luteus, S. aureus, E. coli, K. pneumoniae and P. aeruginosa, as well as the fungi Aspergillus flavus, Aspergillus niger and Candida albicans, indicated that the Cu-NP is more toxic to bacteria than fungi. Another study revealed the antibacterial activity of polyethylene-Cu nanoparticles against E. coli. The result revealed a complete suppression of the number of live bacteria, and the nanocomposites damaged the plasma membrane of the bacteria, revealing a bacteriolytic effect [183].
The bacteria in the chicken’s body may develop resistance to Cu and the resistance gene to this trace element is recognised in some bacterial species from animals. Resistance genes to Cu are often located on the plasmids, which may be transferable to other bacteria species [144].
From these findings, it can be concluded that the smallest nanoparticles, the highest antibacterial activity and Cu-NP have a better impact on bacteria than the other forms of Cu. In addition, the antibacterial activity of Cu-NP can allow more nutrient to be digested in the GIT, consequently improving the growth of chickens.
Table 1: Overview of antibacterial activity of different forms of Cu.
Copper nanoparticles as an alternative feed additive in poultry diet: a review - Image 1
Table 1 (continued)
Copper nanoparticles as an alternative feed additive in poultry diet: a review - Image 2
An overview of the antibacterial activity of the differ - ent forms of Cu is shown in Table 1 .
4.3 Immunological effects
Here, we primarily focus on the immunological impact of the most studied Cu-NP. The interactions of Cu-NP with the immune system are a key issue to guide the future use of Cu-NP in animal feed and medicine.
Exposure to nanoparticles may induce immune stimulation or suppression; inadvertent suppression of immunological function leads to infectious diseases. In contrast, inappropriate enhancement of immune reaction can result in autoimmune diseases [195]. To date, many studies have demonstrated that Cu-NP are involved in promoting inflammatory responses, due to their physico-chemical properties. However, it has been demonstrated that smaller nanoparticles (less than 70 nm) are not able to be recognised as foreign particles and are possibly later translocated through the capillary blood flow [196] and the lymphatic system [197] to the lymph nodes for antigen presentation [198]. While the recognised nanoparticles will be cleared by macrophage-mediated clearance, phagocytosed nanoparticles may be destroyed within the lysosomes of phagocytic cells [199]. If the nanoparticles exceed the size of the engulfing phagocyte, they take too long to be phagocytosed. Later, they may induce an inflammatory response, including cytokines, chemokines and ROS, which can result in inflammation [200] and DNA damage [201]. From the lymphatic and circulatory systems, nanoparticles may distribute to organs including the kidneys, from where partial or total clearance may occur. Inflammation in mice is increased when exposed to Cu-NP through inhalation [202]. Exposure to Cu-NP via inhalation and instillation in mice induces pro-inflam - matory cytokines and recruitment of neutrophils, and it is expected that Cu ions at high Cu-NP doses lead to an increase in inflammatory response [54, 202]. However, in ovo studies with a chicken embryo model, Cu-NP injected into chicken embryos did not show immune stimulatory properties, which may be due to the Cu-NP not being recognised by antigen-presenting cells (APCs) [11, 203], as immune responses depend on nanoparticle uptake by APCs. Exposure to nanoparticles may induce toxicity, resulting in detrimental effects on the immune function.
A study by Wang et al. [204] on fish reported that IL-1B and TNF-α expression increased in intestine samples with an increased concentration of Cu-NP, and also, IgM concentration was decreased with Cu-NP concentration.
Table 2: Toxic effect of Cu nanoparticles in different species of animals according to delivery type and dosage used.
Copper nanoparticles as an alternative feed additive in poultry diet: a review - Image 3
4.4 Toxicological effects
The predominant processes underlying nanoparticle-mediated toxicity includes oxidative stress, inflammation, DNA damage and inhibition of cell division and death. Despite the toxicity of bulk Cu being affected mainly by the salt composition, in the case of Cu-NP, additional physicochemical properties such as size, surface area, surface chemistry, surface roughness, dispersion medium and ability to agglomerate play a vital role in determining their toxicity [13].
There are many studies reporting on the toxicity of Cu-NP [205]. However, we could not find results regarding Cu-NP toxicity in chickens. The intranasal instillation of Cu-NP (23.5 nm) in mice resulted in the accumulation of Cu-NP in the liver and lung tissues, decreased body weight and dose-dependent lesions in the lung and liver [206]. Further, it was reported that Cu-NP induced apoptosis in the intestine of juvenile Epinephelus coioides via the mitochondrial pathway [207]. The toxicity may be due to the Cu-NP and released Cu ions [208]. Furthermore, it was demonstrated that the smaller Cu-NP are more toxic than the larger Cu-NP in zebrafish embryos [209]. The Cu-NP caused membrane damage and ROS formation in lung cells (A549 type II) [210] and mammalian cell lines (H4IIE and HepG2) [211]. The Cu-NP (10 μg/ml) also induced DNA strand breaks [212]. The cell viability effect of the Cu-NP embedded in the polymer matrices (hydrogel and polypropylene) strongly depends on the polymer characteristics, as hydrogels showed cytotoxicity at concentrations higher than 0.5% of that of the Cu-NP, whereas polypropylene did not show any effect even at higher concentrations (20% of Cu-NP) [213]. It was concluded that Cu-NP toxicity  is elucidated by Cu ions released from the composites.
Many studies have pointed out the cytotoxicity effect of Cu-NP, depending on the animal species, delivery, concentration and size. An overview of the toxic effects of Cu-NP is shown in Table 2.
5 Potential use of Cu-NP in animal diet
The main reason for considering the Cu-NP as an alternative growth and health promoter is the overuse of Cu in animal diets, causing severe environmental pollution. In addition, the objective of using nanoparticles in animal feed is to decrease the numbers of harmful bacteria and stimulate the growth of beneficial bacteria, which may improve growth performance of animals [221]. Herein, we provide an overview of the research on the potential use of mineral nanoparticles as poultry feed additives and propose novel strategies for nanoparticle use in animal nutrition. The hypothesis behind growth improvement with Cu-NP supplementation may be related to better absorption of nutrients, i.e. with Cu-NP more nutrients will be available in the GIT for absorption. Hence, using Cu-NP as a supplement in chicken diet might have beneficial effects on growth, feed efficiency and chicken health by causing damage to pathogens, effecting a reduction in the production of bacterial toxins, increasing the synthesis of vitamins and other growth factors, improving the absorption of nutrients by reducing the thickness of the intestinal epithelium and reducing intestinal mucosal epithelial cell turnover and motility [222].
Reduction in the size of materials to the nanorange increases the absorption of nutrients, as shown for iron and selenium [223]. Selenium nanoparticles are more effective and showed better performance of broilers supplied with 0.3  ppm [224] and 0.5  ppm [225] compared to both organic and inorganic selenium sources [226]. Feeding zinc nanoparticles to chickens have produced encouraging responses in growth, immunity and reproduction. These nanoparticles enhance growth and improve feed efficiency in piglets and poultry [227–229]. Chromium nanoparticles reduce heat stress in chickens, improve feed conversion and provoke hepatic-related alterations [230]. Silver nanoparticles have antibacterial activity in chickens and pigs [231, 232]. Furthermore, Sawosz et  al. [233] showed that silver nanoparticles potentially improve muscle morphology without affecting broiler performance at embryo growth. Moreover, a concentration of 50 mg/kg silver nanoparticles chelated with amino acids, threonine and cysteine improved immune competence in embryos and chickens [234]. Other metals have been synthesised as nanoparticles and supplemented to poultry feed. For example, metal oxide nanoparticles have shown antibacterial activity against Salmonella sp.: Al2 O3 , Fe3 O4 , CeO2 , ZrO2 and MgO [235]. The above studies of metal nanoparticles focused on reducing antibacterial activity and, to some extent, on their influence on animal performance. However, the application of Cu-NP as a feed additive appears to be in the development phase.
Cu-NP administered through in ovo injection at levels of 4, 8, 12 and 16 μg/egg did not have an adverse effect on embryo growth [10]. However, a concentration of 50 mg/kg of Cu-NP showed an increase in metabolic rate with regard to oxygen consumption and heat production, which are important regulators in the developmental stages of chicken embryos [11]. The same concentration was found to exhibit pro-angiogenic properties at a systemic level, with the promotion of blood vessel development during embryogenesis and, thereafter, increase the body weight, improve feed conversion ratio and increase breast and leg muscles of broiler chickens upon administering Cu-NP in ovo [110, 126]. Just recently, it was shown that in ovo injection of Cu-NP stimulates proliferating cell nuclear antigen (PCNA)-positive cells in the long bones of broiler chickens, indicating a stimulatory effect during embryogenesis [236].
Table 3: Advantages of using Cu nanoparticles in different animal species, according to the delivery and dosage used.
Copper nanoparticles as an alternative feed additive in poultry diet: a review - Image 4
Table 3 (continued)
Copper nanoparticles as an alternative feed additive in poultry diet: a review - Image 5
Some studies compared the inorganic forms of Cu with Cu-NP, and the latter showed an improvement in the growth performance of piglets [9]. Moreover, Cu availability was significantly improved, and the faecal Cu level was reduced compared to CuSO4. Additionally, Cu-NP improved the digestibility of crude fat and energy in pigs. Furthermore, IgGγ globulin and total globulin protein levels were improved, and SOD activity increased. The same authors observed that the Cu level and cholesterol concentrations in serum were not affected by Cu-NP supplementation. The reason for these improvements may be attributed to better bioavailability and antibacterial activity of Cu-NP than CuSO4.
Intramuscular injection of Cu-NP in chickens was investigated by Miroshnikov et al. [131], and they observed that Cu-NP stimulated chicken growth, increased red blood cell and haemoglobin levels, enriched Cu and protein levels in serum and also increased the arginine content in the chicken liver.
Another study reported that Cu-NP-loaded chitosan [130] improved growth performance and immune status, enhanced protein synthesis and was beneficial to the caecal microbiota of broiler chickens. Further, Zheng et al. [237] used Cu silicate nanoparticles in chicken diet and reported that adding 2 g/kg of Cu silicate nanoparticles could regulate the intestinal microflora, promote the growth of beneficial bacteria and inhibit harmful ones, enhance nitrogen metabolism and reduce ammonia emission from excreta. Recently, Nguyen et al. [238] added metal nanoparticles of iron, Cu, zinc oxide and selenium to supplement a chicken diet premix, which resulted in certain improvements of poultry farming depending on the quantity of nanocrystalline metal, which replaced the inorganic mineral component in the feed premix. The results also confirmed that nanocrystalline metals are able to decrease inorganic minerals in diet premixes by at least four times, allowing chickens to absorb feed minerals more efficiently and, consequently, decreasing the risk of environmental pollution. In addition, Cu silicate nanoparticles modified the intestinal microbiota of chicken, increasing counts of Lactobacillus species and decreasing E. coli [239].
Several studies have shown the advantages of adding Cu-NP to fish diet. A study by El Basuini et al. [127] investigated different concentrations of Cu-NP in fish diet, and the results demonstrated that 2 and 4 mg/kg of Cu-NP caused the highest final body weight, better feed efficiency, protein retention, immune response and antioxidant defence system compared to CuSO4 and the control groups. Another study was performed to assess the effect of Cu-NP on freshwater prawns by Muralisankar et al. [240]; the results demonstrated that adding 20  mg/kg of Cu-NP showed an improvement in growth, digestive enzyme activity, the concentration of biochemical constituents and haemocyte count. However, the same authors reported that 40–80 mg/kg of Cu-NP had a negative effect and might be toxic to prawns. Lately, El Basuini et al. [241] determined the effects of Cu-NP and vitamin C on red sea bream and found that feed supplemented with 2  mg/kg of Cu-NP significantly influenced body weight gain, final body weight, feed intake, feed and protein efficiency ratios, body protein and lipid content, protease and bactericidal activity and tolerance against stress compared to the control diet.
There was also a study conducted on rabbits fed diets supplemented with different concentrations of Cu-NP, which showed a significantly higher body weight and better performance in comparison to the control [129].
An overview of the effects of Cu-NP on animal performance is shown in Table 3.
6 Conclusions
The main purpose of using Cu-NP as feed additives in chickens is to improve growth and performance in addition to reducing the pathogen load to improve health and reduce the excretion of Cu into the environment. Their small size and large surface-to-volume ratio make Cu-NP biologically more active than bulk Cu, allowing more efficient interaction with biological systems, which may decrease the amount of Cu used in the diet without causing an adverse effect on chicken performance and the environment.
The majority of studies have pointed out the beneficial effects on performance, antibacterial activity and immune status of Cu-NP as a feed supplement for poultry and other animals. Furthermore, much work has been done regarding the synthesis and stability of Cu-NP.
However, it is important to control the size and concentration in order to achieve maximal benefits from Cu-NP and avoid their potential toxic effects. More studies should be conducted to describe the effects of Cu-NP on gut microbes and metabolite production. Along with these, studies should include more toxicology, histology and residue analysis of Cu-NP in the body and excreta before any wide adoption could take place.
Cu-NP incorporation in animal nutrition, which can enhance the performance and health of animals, should be conducted at a lower risk to consumers. However, a great amount of research is still required to support the safety of Cu-NP application in animal nutrition, avoiding any harm to livestock, the environment and human beings.
The inclusion of Cu-NP supplements in poultry feed seems possible in the near future; however, the use of Cu-NP is still in its infancy, but encouraging results from recent studies are driving further investigations.
      
This article was originally published in Nanotechnology Reviews, vol. 7, no. 1, 2018, pp. 69-93. https://doi.org/10.1515/ntrev-2017-0159. This is an Open Access article is distributed under the terms of the Creative Commons Attribution Non-Commercial License.

[1] European Commission Directorate C Opinion of the Scientific Committee for Animal Nutrition on the Use of Copper in Feeding Stuffs. European Commission Health and Consumer Protection Directorate-General 2003.

[2] Bao YM, Choct M. Trace mineral nutrition for broiler chickens and prospects of application of organically complexed trace minerals: a review. Anim. Prod. Sci. 2009, 49, 269–282.

[3] Pang Y, Patterson JA, Applegate TJ. The influence of copper concentration and source on ileal microbiota. Poul. Sci. 2009, 88, 586–592feed and food. Trends Food Sci. Technol. 2016, 54, 155–164.

[5] Feng M, Wang ZS, Zhou AG, Ai DW. The effects of different sizes of nanometer zinc oxide on the proliferation and cell integrity of mice duodenum- epithelial cells in primary culture. PJN 2009, 8, 1164–1166

[6] Sri Sindhura K, Prasad TNVKV, Panner Selvam P, Hussain OM. Synthesis, characterization and evaluation of effect of phytogenic zinc nanoparticles on soil exo-enzymes. Appl. Nanosci. 2014, 4, 819–827.

[7] Tamilvanan A, Balamurugan K, Ponappa K, Madhan Kumar B. Copper nanoparticles: synthetic strategies, properties and multifunctional application. Int. J. Nanosci. 2014, 13, 1430001

[8] Boyles, MSP, Ranninger, C, Reischl, R, Rurik, M, Tessadri, R,
Kohlbacher, O, Duschl, A, Huber, CG. Copper oxide nanoparticle toxicity profiling using untargeted metabolomics. Part.
Fibre Toxicol. 2016, 13, 49.
[9] Gonzales-Eguia A, Fu C-M, Lu F-Y, Lien T-F. Effects of nanocopper on copper availability and nutrients digestibility, growth performance and serum traits of piglets. Livest. Sci.
2009, 126, 122–129.
[10] Joshua PP, Valli C, Balakrishnan V. Effect of in ovo supplementation of nano forms of zinc, copper, and selenium on post-hatch performance of broiler chicken. Vet. World 2016, 9,
287–294.
[11] Scott A, Vadalasetty KP, Sawosz E, Łukasiewicz M,
Vadalasetty RKP, Jaworski S, Chwalibog A. Effect of copper nanoparticles and copper sulphate on metabolic rate and development of broiler embryos. Anim. Feed Sci. Technol.
2016, 220, 151–158.
[12] Camacho-Flores B.A, Martínez-Álvarez O, Arenas-Arrocena
MC, Garcia-Contreras R, Argueta-Figueroa L, de la FuenteHernández J, Acosta-Torres LS. Copper: synthesis techniques in nanoscale and powerful application as an antimicrobial agent.
J. Nanomater. 2015, 2015, 1–10.
[13] Gatoo MA, Naseem S, Arfat MY, Mahmood Dar A, Qasim K,
Zubair S. Physicochemical properties of nanomaterials: implication in associated toxic manifestations. Biomed Res. Int.
2014, 2014, 1–8.
[14] Hefnawy AE, El-khaiat H. Copper and animal health (importance, maternal fetal, immunity and DNA relationship, deficiency and toxicity). Int. J. Agro Vet. Med. Sci. 2015, 9,
195–211.
[15] Failla ML. Trace elements and host defense: recent advances and continuing challenges. J. Nutr. 2003, 133(5 Suppl 1),
1443S–1447S.
[16] Tapiero H, Townsend DM, Tew KD. Trace elements in the human physiology and pathology. Copper. Biomed. Pharmacother.
2003, 57, 386–398.
[17] Magaye J. Genotoxicity and carcinogenicity of cobalt-, nickeland copper-based nanoparticles (Review). Exp. Ther. Med.
2012, 4, 551–561.
[18] Makarski B, Zadura A. Wplyw chelatu miedzi z lizyna na poziom skladników hematologicznych i biochemicznych krwi indyków.
Annales Universitatis Mariae Curie-Sklodowska, Lublin-Polonia
2006, 24, 357–363.
[19] Sharma MC, Joshi C, Pathak NN, Kaur H. Copper status and enzyme, hormone, vitamin and immune function in heifers.
Res. Vet. Sci. 2005, 79, 113–123.
[20] Lin W, Stayton I, Huang Y, Zhou XD, Ma Y. Cytotoxicity and cell membrane depolarization induced by aluminum oxide nanoparticles in human lung epithelial cells A549. Toxicol. Environ.
Chem. 2008, 90, 983–996.
[21] Birben E, Sahiner UM, Sackesen C, Erzurum S, Kalayci O.
Oxidative stress and antioxidant defense. World Allergy Organ.
J. 2012, 5, 9–19.
[22] Wu XZ, Zhang TT, Guo JG, Liu Z, Yang FH, Gao XH. Copper bioavailability, blood parameters, and nutrient balance in mink.
J. Anim. Sci. 2015, 93, 176–184.
[23] Mroczek-Sosnowska N, Batorska M, Lukasiewicz M, Wnuk A,
Sawosz E, Jaworski S, Niemiec J. Effect of nanoparticles of copper and copper sulfate administered in ovo on hematological and biochemical blood markers of broiler chickens. Annals of
Warsaw University of Life Sciences-SGGW. Anim. Sci. 2013, 52,
141–149.
[24] McDowell LR, Mineral In Animal And Human Nutrition,
Academic Press Inc.: San Diego, New York, Boston, London,
Sydney, Tokyo, Toronto, 1992, pp. 524.
[25] Chesters J.K. Trace element-gene interactions. Nutr. Rev. 1992,
50, 217–223.
[26] Karimi A, Sadeghi G, Vaziry A. The effect of copper in excess of the requirement during the starter period on subsequent performance of broiler chicks. J. Appl. Poult. Res. 2011, 20,
203–209.
[27] EFSA Panel on Additives and Products or Substances used in
Animal Feed (FEEDAP). Revision of the currently authorised maximum copper content in complete feed. EFSA J. 2016, 14,
4563.
[28] Kim JW, Kim JH, Shin JE, Kil DY. Relative bioavailability of copper in tribasic copper chloride to copper in copper sulfate for laying hens based on egg yolk and feather copper concentrations.
Poult. Sci. 2016, 95, 1591–1597.
[29] Kincaid RL. Assessment of trace mineral status of ruminants: a review. Proc. Am. Soc. Anim. Sci. 1999, 77, 1–10.
[30] National Research Council Nutrient Requirements of Poultry,
1994 9th Rev Ed. NAS-NRC, Washington DC.
[31] Leeson S. Copper metabolism and dietary needs. World Poult.
Sci. 2009, 65, 353–365.
[32] Kim JW, Kil DY. Determination of relative bioavailability of copper in tribasic copper chloride to copper in copper sulfate for broiler chickens based on liver and feather copper concentrations. Anim. Feed Sci. Technol. 2015, 210, 138–143.
[33] Ammerman CB, Baker DH, Lewis AJ. Bioavailability of Nutrients for Animals: Amino Acids, Minerals, and Vitamins, Academic
Press: San Diego, CA, 1995.
[34] Luo XG, Ji F, Lin YX, Steward FA, Lu L, Liu B, Yu SX. Effects of dietary supplementation with copper sulfate or tribasic copper chloride on broiler performance, relative copper bioavailability, and oxidation stability of vitamin E in feed. Poult. Sci. 2005,
84, 888–893.
[35] Marchetti M, Ashmead HD, Tossani N, Marchetti S, Ashmead
SD. Comparison of the rates of vitamin degradation when mixed with metal sulphates or metal amino acid chelates.
J. Food Compost. Anal. 2000, 13, 875–884.
[36] Pesti GM, Bakalli RI. Studies on the feeding of cupric sulfate pentahydrate and cupric citrate to broiler chickens. Poult. Sci.
1996, 75, 1086–1091.
[37] Ledoux DR, Henry PR, Ammerman CB, Rao PV, Miles RD. Estimation of the relative bioavailability of inorganic copper sources for chicks using tissue uptake of copper. J. Anim. Sci. 1991, 69,
215–222.
[38] Liu Z, Bryant MM, Roland DA Sr. Layer performance and phytase retention as influenced by copper sulfate pentahydrate and tribasic copper chloride. J. Appl. Poult. Res. 2005, 14, 499–505.
[39] Miles RD, O’keefe SF, Henry PR, Ammerman CB, Luo XG. The effect of dietary supplementation with copper sulfate or tribasic copper chloride on broiler performance, relative copper bioavailability, and dietary prooxidant activity. Poult. Sci. 1998,
77, 416–425.
[40] Arias VJ, Koutsos EA. Effects of copper source and level on intestinal physiology and growth of broiler chickens. Poult. Sci.
2006, 85, 999–1007.
[41] Świątkiewicz S, Arczewska-Włosek A, Józefiak D. The efficacy of organic minerals in poultry nutrition: review and implications of recent studies. World Poult. Sci. J. 2014, 70, 475–486.
[42] Leeson S, Summers JD. Commercial Poultry Nutrition
Publication, University Books: Guelph, Canada, 2005.
[43] Zhao J, Allee G, Gerlemann G, Ma L, Gracia MI, Parker D,
Vazquez-Anon M, Harrell RJ. Effects of a chelated copper as growth promoter on performance and carcass traits in pigs.
Asian-Australas. J. Anim. Sci. 2014, 27, 965–973.
[44] Banks KM, Thompson KL, Rush JK, Applegate TJ. Effects of copper source on phosphorus retention in broiler chicks and laying hens. Poult. Sci. 2004, 83, 990–996.
[45] Mondal MK, Das TK, Biswas P, Samanta CC, Bairagi B.
Influence of dietary inorganic and organic copper salt and level of soybean oil on plasma lipids, metabolites and mineral balance of broiler chickens. Anim. Feed Sci. Technol. 2007, 139,
212–233.
[46] Miles RD, Henry PR, Sampath VC, Shivarzad M, Comer CW.
Relative bioavailability of novel amino acid chelates of manganese and copper for chicks. J. Appl. Poult. Res. 2003, 12,
417–423.
[47] Guo R, Henry PR, Holwerda RA, Cao J, Littell RC, Miles RD,
Ammerman CB. Chemical characteristics and relative bioavailability of supplemental organic copper sources for poultry.
Anim. Sci. 2001, 79, 1132–1141.
[48] Bao YM, Choct M, Iji PA, Bruerton K. Effect of organically complexed copper, iron, manganese, and zinc on broiler performance, mineral excretion, and accumulation in tissues. J. Appl.
Poult. Res. 2007, 16, 448–455.
[49] Jegede AV, Oduguwa OO, Bamgbose AM, Fanimo AO, Nollet L.
Growth response, blood characteristics and copper accumulation in organs of broilers fed on diets supplemented with organic and inorganic dietary copper sources. Br. Poult. Sci.
2011, 52, 133–139.
[50] Kwiecień M, Winiarska-Mieczan A, Piedra JV, Bujanowicz-Haraś
B, Chałabis-Mazurek A. Effects of copper glycine chelate on liver and faecal mineral concentrations, and blood parameters in broilers. Agric. Food Sci. 2015, 24, 92–103.
[51] Liu S, Lu L, Li S, Xie J, Zhang L, Wamg R, Luo X. Copper in organic proteinate or inorganic sulfate form is equally bioavailable for broiler chicks fed a conventional corn-soybean meal diet. Biol. Trace Elem. Res. 2012, 147, 142–148.
[52] Das TK, Mondal MK, Biswas P, Bairagi B, Samanta CC. Influence of level of dietary inorganic and organic copper and energy level on the performance and nutrient utilization of broiler chickens. Asian-Australas. J. Anim. Sci. 2010, 23, 82.
[53] Shamsudeen P, Shrivastava HP. Biointeraction of chelated and inorganic copper with aflatoxin on growth performance of broiler chicken. Int. J. Vet. Sci. 2013, 2, 106–110.
[54] Kim JS, Adamcakova-Dodd A, O’Shaughnessy PT, Grassian VH,
Thorne PS. Effects of copper nanoparticle exposure on host defense in a murine pulmonary infection model. Part. Fibre
Toxicol. 2011, 8, 29.
[55] Kwiecień M, Winiarska-Mieczan A, Zawiślak K, Sroka S. Effect of copper glycinate chelate on biomechanical, morphometric and chemical properties of chicken femur. Ann. Anim. Sci.
2014, 14, 127–139.
[56] Jegede AV, Oduguwa OO, Oso AO, Fafiolu AO, Idowu OMO,
Nollet L. Growth performance, blood characteristics and plasma lipids of growing pullet fed dietary concentrations of organic and inorganic copper sources. Livest. Sci. 2012,145, 298–302.
[57] Attia YA, Qota EM, Zeweil HS, Bovera F, Abd Al-Hamid AE,
Sahledom MD. Effect of different dietary concentrations of inorganic and organic copper on growth performance and lipid metabolism of White Pekin male ducks. Br. Poult. Sci. 2012, 53,
77–88.
[58] Kulkarni RC, Shrivastava HP, Mandal AB. Effect of iron supplementation on serum trace mineral status and cholesterol in broiler chickens. Indian J. Anim. Res. 2013, 47, 89–90.
[59] Rahman ZU, Besbasi F, Afan AM, Bengali EA, Zendah MI, Hilmy
M, Mukhtar MR, Jaspal SAS, Aslam N. Effects of copper supplement on haematological profiles and broiler meat composition.
Int. J. Agric. Biol. 2001, 1560–8530, 203–205.
[60] Yang W, Wang J, Liu L. Effect of high dietary copper on somatostatin and growth hormone-releasing hormone levels in the hypothalamic of growing pigs. Biol. Trace Elem. Res. 2011, 143,
893–900.
[61] Zhu D, Yu B, Ju C, Mei S, Chen D. Effect of high dietary copper on the expression of hypothalamic appetite regulators in weanling pigs. J. Anim. Feed Sci. 2011, 20, 60–70.
[62] Zhou W, Kornegay ET, Lindemann MD, Swinkels JW, Welten MK,
Wong EA. Stimulation of growth by intravenous injection of copper in weanling pigs. J. Anim. Sci. 1994, 72, 2395–2403.
[63] Prashanth L, Kattapagari K, Chitturi R, Baddam VR, Prasad
L. A review on role of essential trace elements in health and disease. NTR Univ. Health Sci. 2015, 4, 75–85.
[64] Wang ZL. Characterizing the structure and properties of individual wire-like nanoentities. Adv. Mater. 2000, 12, 1295–1298.
[65] Din MI, Rehan R. Synthesis, characterization, and applications of copper nanoparticles. Anal. Lett. 2017, 50, 50–62.
[66] Rosi NL, Mirkin CA. Nanostructures in biodiagnostics. Chem.
Rev. 2005, 105, 1547–1562.
[67] Volokitin Y, Sinzig J, De Jongh L, Schmid G, Vargaftik M, Moiseevi I. Quantum size effects in the thermodynamic properties of metallic nanoparticles. Nature 1996, 384, 621–623.
[68] Sasaki K, Matsubara K, Kawamura S, Saito K, Yagi M, Norimatsu W, Sasai R, Yui T. Synthesis of copper nanoparticles within the interlayer space of titania nanosheet transparent films. J. Mater. Chem. C 2016, 4, 1476–1481.
[69] Lee Y, Choi J-R, Lee KJ, Stott NE, Kim D. Large-scale synthesis of copper nanoparticles by chemically controlled reduction for applications of inkjet-printed electronics. Nanotechnology
2008, 19, 415604.
[70] Ruparelia JP, Chatterjee AK, Duttagupta SP, Mukherji S. Strain specificity in antimicrobial activity of silver and copper nanoparticles. Acta Biomater. 2008, 4, 707–716.
[71] Rajendran D, Thulasi A, Jash S, Selvaraju S, Rao SBN. Synthesis and application of nano minerals in livestock industry. In
Animal Nutrition & Reproductive Physiology (Recent Concepts).
Sampath, KT, Ghosh, J, Bhatta, R, Eds., Satish Serial Publishing
House: Delhi 2013, 517–530.
[72] Male KB, Hrapovic S, Liu Y, Wang D, Luong JH. Electrochemical detection of carbohydrates using copper nanoparticles and carbon nanotubes. Anal. Chem. Acta 2004, 516, 35–41.
[73] Chan GH, Zhao J, Hicks EM, Schatz GC, Van Duyne RP. Plasmonic properties of copper nanoparticles fabricated by nanosphere lithography. Nano Lett. 2007, 7, 1947–1952.
[74] Chen S, Sommers JM. Alkanethiolate-protected copper nanoparticles: spectroscopy, electrochemistry, and solid-state morphological evolution. J. Physiol. Chem. B 2001, 105, 8816.
[75] Shankar S, Rhim J-W. Effect of copper salts and reducing agents on characteristics and antimicrobial activity of copper nanoparticles. Mater. Lett. 2014, 132, 307–311.
[76] Yallappa S, Manjanna J, Sindhe MA, Satyanarayan ND, Pramod
SN, Nagaraja K. Microwave assisted rapid synthesis and biological evaluation of stable copper nanoparticles using T. arjuna bark extract. Spectrochim. Acta Part A: Mol. Biomol.
Spectrosc. 2013, 110, 108–115.
[77] Valodkar M, Rathore PS, Jadeja RN, Thounaojam M, Devkar RV,
Thakore S. Cytotoxicity evaluation and antimicrobial studies of starch capped water soluble copper nanoparticles. J. Hazard.
Mater. 2012, 201–202, 244–249.
[78] Buzea C, Pacheco II, Robbie K. Nanomaterials and nanoparticles: sources and toxicity. Biointerphases 2007, 2, 17–71.
[79] Pradhan, S, Hedberg, J, Blomberg, E, Wold, S, Odnevall Wallinder, I. Effect of sonication on particle dispersion, administered dose and metal release of non-functionalized, non-inert metal nanoparticles. J. Nanopart. Res. 2016, 18, 285.
[80] Taurozzi JS, Hackley VA, Wiesner MR. Ultrasonic dispersion of nanoparticles for environmental, health and safety assessment – issues and recommendations. Nanotoxicology 2010, 5,
711–729.
[81] Powers KW, Palazuelos M, Moudgil BM, Roberts SM. Characterization of the size, shape, and state of dispersion of nanoparticles for toxicological studies. Nanotoxicology 2007, 1, 42–51.
[82] Singh N, Manshian B, Jenkins GJS. NanoGenotoxicology: the
DNA damaging potential of engineered nanomaterials. Biomaterials 2009, 30–24, 3891–3914.
[83] Elder A, Vidyasagar S, DeLouise L. Physicochemical factors that affect metal and metal oxide nanoparticle passage across epithelial barriers. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2009, 1, 434–450.
[84] Chandra S, Kumar A, Tomar PK. Synthesis and characterization of copper nanoparticles by reducing agent. J. Saudi Chem. Soc.
2014, 18, 149–153.
[85] Soomro RA, Sherazi STH, Sirajuddin MN, Shah MR, Kalwar NH,
Hallam KR, Shah A. Synthesis of air stable copper nanoparticles and their use in catalysis. Adv. Mater. Lett. 2013, 5, 191–198.
[86] Meng H, Chen Z, Xing GM. Ultrahigh reactivity provokes nanotoxicity: explanation of oral toxicity of nano-copper particles.
Toxicol. Lett. 2007, 175, 102–110.
[87] Cho WS, Duffin R, Poland CA. Differential pro-inflammatory effects of metal oxide nanoparticles and their soluble ions in vitro and in vivo; zinc and copper nanoparticles, but not their ions, recruit eosinophils to the lungs. Nanotoxicology 2012, 6,
22–35.
[88] Nel A, Xia T, Madler L, Li N. Toxic potential of materials at the nanolevel. Science 2006, 311, 622–627.
[89] Aillon KL, Xie Y, El-Gendy N, Berkland CJ, Forrest ML. Effects of nanomaterials physicochemical properties on in vivo toxicity.
Adv. Drug Deliv. Rev. 2009, 61, 457–466.
[90] Zhu M, Nie G, Meng H, Xia T, Nel A, Zhao Y. Physicochemical properties determine nanomaterial cellular uptake, transport, and fate. Acc. Chem. Res. 2013, 46, 622–631.
[91] Jackelen A-ML, Jungbauer M, Glavee GN. Nanoscale materials synthesis. 1. Solvent effects on hydridoborate reduction of copper ions. Langmuir 1999, 15, 2322–2326.
[92] Xia X, Xie C, Cai S, Yang Z, Yang X. Corrosion characteristics of copper microparticles and copper nanoparticles in distilled water. Corros. Sci. 2006, 48, 3924–3932.
[93] Shi M, Kwon HS, Peng Z, Elder A, Yang H. Effects of surface chemistry on the generation of reactive oxygen species by copper nanoparticles. ACS Nano 2012, 6, 2157–2164.
[94] Cedervall T, Lynch I, Lindman S, Berggard T, Thulin E, Nilsson
H, Dawson KA, Linse S. Understanding the nanoparticleprotein corona using methods to quantify exchange rates and affinities of proteins for nanoparticles. Proc. Natl. Acad. Sci.
2007, 104, 2050–2055.
[95] Baroli B, Ennas MG, Loffredo F, Isola M, Pinna R, Lopez-Quintela MA. Penetration of metallic nanoparticles in human fullthickness skin. J. Investig. Dermatol. 2007, 127, 1701–1712.
[96] Zahr AS, Davis CA, Pishko MV. Macrophage uptake of coreshell nanoparticles surface modified with poly (ethylene glycol). Langmuir 2006, 22, 8178–8185.
[97] Hoshino A, Fujioka K, Oku T. Physicochemical properties and cellular toxicity of nanocrystal quantum dots depend on their surface modification. Nano Lett. 2004, 4, 2163–2169.
[98] Pietroiusti A, Massimiani M, Fenoglio I. Low doses of pristine and oxidized single-wall carbon nanotubes affect mammalian embryonic development. ACS Nano 2011, 5, 4624–4633.
[99] Georgieva JV, Kalicharan D, Couraud P. Surface characteristics of nanoparticles determine their intracellular fate in and processing by human blood-brain barrier endothelial cells in vitro. Mol. Ther. 2011, 19, 318–325.
[100] Mamonova IA, Matasov MD, Babushkina IV, Losev OE, Chebotareva YG, Gladkova EV, Borodulina YV. Study of physical properties and biological activity of copper nanoparticles.
Nanotechnol. Russ. 2013, 8, 303–308.
[101] Rakhmetova AA, Alekseeva TP, Bogoslovskaya OA, Leipunskii
IO, Ol’khovskaya IP, Zhigach AN, Glushchenko NN. Wound-healing properties of copper nanoparticles as a function of physicochemical parameters. Nantechnol. Russ. 2010, 5, 271–276.
[102] Huang J, Chen C, He N, Hong J, Lu Y, Qingbiao L, Shao W, Sun
D, Wang XH, Wang Y, Yiang X. Biosynthesis of silver and gold nanoparticles by novel sundried Cinnamomum camphora leaf.
Nanotechnology 2007, 18, 105–106.
[103] Ingle A, Gade A, Pierrat S, Sonnichsen C, Rai MK. Mycosynthesis of silver nanoparticles using the fungus Fusarium acuminatum and its activity against some human pathogenic bacteria. Curr. Nanosci. 2008, 4, 141–144.
[104] Lewinski N, Colvin V, Drezek R, Cytotoxicity of nanoparticles.
Small 2008, 4, 26–49.
[105] Nel AE, Mädler L, Velegol D, Xia TE, Hoek MV, Somasundaran
P, Klaessig F, Castranova V, Thompson M. Understanding biophysicochemical interactions at the nano–bio interface. Nat.
Mater. 2009, 8, 543–557.
[106] Zhao F, Zhao Y, Liu Y, Chang XL, Chen CY, Zhao YL. Cellular uptake, intracellular trafficking, and cytotoxicity of nanomaterials. Small 2011, 7, 1322–1337.
[107] Civardi C, Schubert M, Fey A, Wick P, Schwarze FW. Micronized copper wood preservatives: efficacy of ion, nano, and bulk copper against the brown rot fungus Rhodonia placenta. PLoS
One 2015, 10, e0142578.
[108] Zaboli K, Aliarabi H, Bahari AA, Abbasalipourkabir R. Role of dietary nano-zinc oxide on growth performance and blood levels of mineral: a study on in Iranian Angora (Markhoz) goat kids. J. Pharm. Health Sci. 2013, 2, 19–26.
[109] Hillyer JF, Albrecht RM. Gastrointestinal persorption and tissue distribution of differently sized colloidal gold nanoparticles. J. Pharm. Sci. 2001, 90, 1927–1936.
[110] Mroczek-Sosnowska N, Sawosz E, Vadalasetty K, Łukasiewicz
M, Niemiec J, Wierzbicki M, Kutwin M, Jaworski S, Chwalibog A. Nanoparticles of copper stimulate angiogenesis at systemic and molecular level. Int. J. Mol. Sci. 2015a, 16,
4838–4849.
[111] Al-Bairuty GA, Boyle D, Henry TB, Handy RD. Sublethal effects of copper sulphate compared to copper nanoparticles in rainbow trout (Oncorhynchus mykiss) at low pH: physiology and metal accumulation. Aquat. Toxicol. 2016, 174,
188–198.
[112] Hoet PH, Brüske-Hohlfeld I, Salata OV. Nanoparticles–known and unknown health risks. J. Nanobiotechnol. 2004, 2, 12.
[113] O’ Hagan DT. The intestinal uptake of particles and the implications for drug and antigen delivery. J. Anat. 1996, 189,
477–482.
[114] Hillery AM, Jani PU, Florence AT. Comparative, quantitative study of lymphoid and nonlymphoid uptake of 60 nm polystyrene particles. J. Drug Target. 1994, 2, 151–156.
[115] Food Safety Authority of Ireland. The Relevance for Food
Safety of Applications of Nanotechnology in the Food and
Feed Industries Abbey Court, Lower Abbey Street, Dublin
2008, 1.
[116] Gangadoo S, Stanley D, Hughes RJ, Moore RJ, Chapman J.
Nanoparticles in feed: Progress and prospects in poultry research. Trends Food Sci. Technol. 2016, 58, 115–126.
[117] Sadauskas E, Wallin H, Stoltenberg M, Vogel U, Doering P,
Larsen A. Kupffer cells are central in the removal of nanoparticles from the organism. Part. Fibre Toxicol. 2007, 4, 1.
[118] Jiang X, Rocker C, Hafner M, Brandholt S, Dorlich RM,
Nienhaus GU. Endo- and exocytosis of zwitterionic quantum dot nanoparticles by live HeLa cells. ACS Nano 2010, 4,
6787–6797.
[119] Gangadoo S, Taylor-Robinson AW, Chapman J. From replacement to regeneration: are bio-nanomaterials the emerging prospect for therapy of defective joints and bones. J. Biotechnol. Biomater. 2015, 5, 2.
[120] Alber F, Dokudovskaya S, Veenhoff LM, Zhang WH, Kipper J,
Devos D, Suprapto A, Karni-Schmidt O, Williams R, Chait BT,
Sali A, Rout MP. The molecular architecture of the nuclear pore complex. Nature 2007, 450, 695–701.
[121] Jani P, Halbert GW, Langridge J, Florence AT. Nanoparticle uptake by the rat gastrointestinal mucosa: quantitation and particle size dependency. J. Pharm. Pharmacol. 1990, 42, 821–826.
[122] Oberdörster G, Oberdörster E, Oberdörster J. Nanotoxicology: an emerging discipline evolving from studies of ultrafine particles. Environ. Health Perspect. 2005, 113, 823–839.
[123] Thulasi A, Rajendran D, Jash S, Selvaraju S, Jose VL, Velusamy
S, Mathivanan S. Nanobiotechnology of Animal Nutrition,
Sampath, KT, Ghosh, J, Bhatta, R, Eds., Satish Serial Publishing House: New Delhi Chapter 24, 2013, pp. 499–515.
[124] Fröhlich E, Roblegg E. Models for oral uptake of nanoparticles in consumer products. Toxicology 2012, 291, 10–17.
[125] Mroczek-Sosnowska N, Lukasiewicz M, Wnuk A, Sawosz E,
Niemiec J. Effect of copper nanoparticles and copper sulfate administered in ovo on copper content in breast muscle, liver and spleen of broiler chickens. Annals of Warsaw University of
Life Sciences-SGGW. Anim. Sci. 2014, 53, 135–142.
[126] Mroczek-Sosnowska N, Łukasiewicz M, Wnuk A, Sawosz E,
Niemiec J, Skot A, Jaworski S, Chwalibog A. In ovo administration of copper nanoparticles and copper sulfate positively influences chicken performance: effect of Cu on chicken performance. J. Sci. Food Agric. 2015b, 96, 3058–3062.
[127] El Basuini MF, El-Hais AM, Dawood MAO, Abou-Zeid AE-S,
EL-Damrawy SZ, Khalafalla MME-S, Koshio S, Ishikawa
M, Dossou S. Effect of different levels of dietary copper nanoparticles and copper sulfate on growth performance, blood biochemical profiles, antioxidant status and immune response of red sea bream (Pagrus major). Aquaculture 2016,
455, 32–40.
[128] Lien TF. Nanosize of copper sulfate and effects on growth, copper availability, and excretion of pigs. Int. J. Livest. Res.
2009, 1, 30–36.
[129] Refaie AM, Ghazal MN, Easa FM, Barakat SA, GE Y, WH E.
Nano-copper as a new growth promoter in the diet of growing
New Zealand white rabbits. Egypt Journal of Rabbit Science
2015, 25, 39–57.
[130] Wang C, Wang MQ, Ye SS, Tao WJ, Du YJ. Effects of copperloaded chitosan nanoparticles on growth and immunity in broilers. Poult. Sci. 2011, 90, 2223–2228.
[131] Miroshnikov S, Yausheva E, Sizova E, Miroshnikova E. Comparative assessment of effect of copper nano- and microparticles in chicken. Orient. J. Chem. 2015, 31, 2327–2336.
[132] Kumar P, Biswas A, Bharti VK, Srivastava RB. Effects of dietary copper supplementation on performance and blood biochemical parameters in broiler chickens at cold desert region in
India. J. Vet. Sci. Photon 2013, 114, 166–172.
[133] Canli E.G, Canli M. Effects of aluminum, copper, and titanium nanoparticles on some blood parameters in Wistar rats. Turk.
J. Zool. 2017, 41, 259–266.
[134] Mommsen TP, Vijayan MM, Moon TW. Cortisol in teleosts: dynamics, mechanisms of action, and metabolic regulation.
Rev. Fish Biol. Fish. 1999, 9, 211–268.
[135] Zahedi M, Ghalehkandi JG, Ebrahimnezhad Y, Emami F. Effects of different levels of copper sulfate on blood biochemical traits in Japanese quail (Coturnix coturnix japonica), Int. J.
Biosci. (IJB) 2013, 3, 221–226.
[136] Payvastegan S, Farhoomand P, Delfani N. Growth performance, organ weights and, blood parameters of broilers fed diets containing graded levels of dietary canola meal and supplemental copper. J. Poult. Sci. 2013, 50,
354–363.
[137] Kim S, Chao PY, Allen KGD. Inhibition of elevated hepatic glutathione abolishes copper deficiency cholesterolemia. FASEB
J. 1992, 6, 2467–2471
[138] Bakalli RI, Pesti GM, Ragland WL, Konjufca V. Dietary copper in excess of nutritional requirement reduces plasma and breast muscle cholesterol in chickens. Poult. Sci. 1995, 74,
360–365
[139] Yang YX, Guo J, Yoon SY, Jin Z, Choi JY, Piao XS, Kim BW, Ohh
SJ, Wang MH, Chae BJ. Early energy and protein reduction: effects on growth, blood profiles and expression of genes related to protein and fat metabolism in broilers. Br. Poult.
Sci. 2009, 50, 218–227.
[140] Fukawa K, Nishimura N, Irino O, Nitta, K. Experimental studies on anti-tumor effects of lysozyme-II: effect of lysozyme on immunopotentiation. Gan To Kagaku Ryoho 1982, 9,
1832–1837.
[141] Dmoch M, Polonis A. Wplyw tiokompleksu miedziowego na wybrane wskalniki hematologiczne, biochemiczne i zawartosc skladników mineralnych we krwi kurczat brojlerów. Acta Sci.
Pol., Zootechnica 2007, 6, 11–18.
[142] Mullally AM, Vogelsang GB, Moliterno AR. Warted sheep and premature infants the role of trace metals in hematopoiesis.
Blood Rev. 2004, 18, 227–234.
[143] Maryam G, Samaneh Z. The effect of copper oxide nanoparticles as feed additive on some the blood proteins of broiler chickens. Mol. Biol. Res. Commun. 2014, 3, 144–145.
[144] Yazdankhah S, Rudi K, Bernhoft A. Zinc and copper in animal feed – development of resistance and co-resistance to antimicrobial agents in bacteria of animal origin. Microb. Ecol.
Health Dis. 2014, 25, 25862.
[145] Ventola CL. The antibiotic resistance crisis: part 1: causes and threats. Pharm. Ther. 2015, 40, 277–283.
[146] DeAlba-Montero I, Guajardo-Pacheco J, Morales-Sánchez E,
Araujo-Martínez R, Loredo-Becerra GM, Martínez-Castañón
G-A, Ruiz F, Compeán Jasso ME. Antimicrobial properties of copper nanoparticles and amino acid chelated copper nanoparticles produced by using a soya extract. Bioinorg. Chem.
Appl. 2017, 2017, 1–6.
[147] Sánchez-Sanhueza G, Fuentes-Rodríguez D, Bello-Toledo H.
Nanopartículas de xobre como potencial agente antimicrobiano en la desinfección de canales radiculares: revisión sistemática. Int. J. Odontostomatol. 2016, 10, 547–554.
[148] Rudramurthy G, Swamy M, Sinniah U, Ghasemzadeh, A.
Nanoparticles: alternatives against drug-resistant pathogenic microbes. Molecules 2016, 21, 836.
[149] Chang Y-N, Zhang M, Xia L, Zhang J, Xing G. The toxic effects and mechanisms of CuO and ZnO nanoparticles. Materials
2012, 5, 2850–2871.
[150] Amro NA, Kotra P, WaduMesthrige K, Bulychev A, Mobashery
S, Liu GY. High-resolution atomic force microscopy studies of the Escherichia coli outer membrane: structural basis for permeability. Langmuir 2000, 16, 2789–2796.
[151] Azam A, Ahmed A, Oves M. Size-dependent antimicrobial properties of CuO nanoparticles against Gram-positive andnegative bacterial strains. Int. J. Nanomed. 2012, 7, 3527–
3535.
[152] Fang J, Lyon DY, Wiesner MR, Dong J, Alvarez. Effect of a fullerene water suspension on bacterial phospholipids and membrane phase behavior. Environ. Sci. Technol. 2007, 41,
2636–2642.
[153] Applerot G, Lellouche J, Lipovsky A, Nitzan Y, Lubart R,
Gedanken A, Banin E. Understanding the antibacterial mechanism of CuO nanoparticles: revealing the route of induced oxidative stress. Small 2012, 8, 3326–3337.
[154] Sondi I, Salopek-Sondi B. Silver nanoparticles as antimicrobial agent: a case study on E. coli as a model for Gram-negative bacteria. J. Colloid Interface Sci. 2004, 275, 177–182.
[155] Kim B-E, Nevitt T, Thiele DJ. Mechanisms for copper acquisition, distribution and regulation. Nat. Chem. Biol. 2008, 4,
176–185.
[156] Chatterjee AK, Chakraborty R, Basu T. Mechanism of antibacterial activity of copper nanoparticles. Nanotechnology 2014,
25, 135101.
[157] Das R, Gang S, Nath SS, Bhattacharjee R. Linoleic acid capped copper nanoparticles for antibacterial activity. J. Bionanosci.
2010, 4, 82–86.[158] Duran N, Marcato PD, De Conti R, Alves OL, Costa FTM, Brocchi
M. Potential use of silver nanoparticles on pathogenic bacteria, their toxicity, and possible mechanisms of action. J. Braz.
Chem. Soc. 2010, 21, 949–959.
[159] Prabhu BM, Ali SF, Murdock RC, Hussain SM, Srivatsan M.
Copper nanoparticles exert size and concentration dependent toxicity on somatosensory neurons of rat. Nanotoxicology
2010, 4, 150–160.
[160] Diaz-Visurrage J, Gutierrez C, Plessing C, Garcia A. Metal nanostructures as antimicrobial agents. Formatex, Badajoz,
Spain 2011, 257, 210–218.
[161] Patra JK, Baek K-H. Antibacterial activity and synergistic antibacterial potential of biosynthesized silver nanoparticles against foodborne pathogenic bacteria along with its anticandidal and antioxidant effects. Front. Microbiol. 2017, 8, 167.
[162] Theivasanthi T, Alagar M. Studies of copper nanoparticles effects on micro-organisms. 2011. arXiv:1110.1372
[physics].https://arxiv.org/abs/1110.1372
[163] Hajipour MJ, Fromm KM, Akbar Ashkarran A, Jimenez de
Aberasturi D, e Larramendi IR, Rojo T, Serpooshan V, Parak
WJ, Mahmoudi, M. Antibacterial properties of nanoparticles.
Trends Biotechnol. 2012, 30, 499–511.
[164] Khurana C, Andhariya N, Chudasama B, Vala AK, Pandey OP.
Influence of antibiotic adsorption on biocidal activities of silver nanoparticles. IET Nanobiotechnol. 2016, 10, 69–74.
[165] Sportelli M, Picca R, Ronco R, Bonerba E, Tantillo G, Pollini
M, Sannino A, Valentini A, Cataldi T, Cioffi N. Investigation of industrial polyurethane foams modified with antimicrobial copper nanoparticles. Materials 2016, 9, 544.
[166] Ballo MKS, Rtimi S, Pulgarin C, Hopf N, Berthet A, Kiwi J, Moreillon P, Entenza JM, Bizzini A. In vitro and in vivo effectiveness of an innovative silver-copper nanoparticle coating of catheters to prevent methicillin-resistant Staphylococcus aureus infection.
Antimicrob. Agents Chemother. 2016, 60, 5349–5356.
[167] Gunawan C, Teoh WY, Marquis CP, Amal R. Cytotoxic origin of copper (II) oxide nanoparticles: comparative studies with micron-sized particles, leachate, and metal salt. ACS Nano
2011, 5, 7214–7225.
[168] Rajasekaran P, Santra S. Hydrothermally treated chitosan hydrogel loaded with copper and zinc particles as a potential micronutrient-based antimicrobial feed additive. Front. Vet.
Sci. 2015, 2, 62.
[169] Højberg O, Canibe N, Poulsen HD, Hedemann MS, Jensen
BB. Influence of dietary zinc oxide and copper sulfate on the gastrointestinal ecosystem in newly weaned piglets. Appl.
Environ. Microbiol. 2005, 71, 2267–2277.
[170] Carlson D, Poulsen HD, Sehested J. Influence of weaning and effect of post weaning dietary zinc and copper on electrophysiological response to glucose, theophylline and 5-HT in piglet small intestinal mucosa. Comp. Biochem. Physiol. A
Mol. Integr. Physiol. 2004, 137, 757–765.
[171] Seiler C, Berendonk TU. Heavy metal driven co-selection of antibiotic resistance in soil and water bodies impacted by agriculture and aquaculture. Front. Microbiol. 2012, 3, 399.
[172] Baker-Austin C, Wright MS, Stepanauskas R, McArthur JV. Coselection of antibiotic and metal resistance. Trends Microbiol.
2006, 14, 176–182.
[173] Schrand AM, Rahman MF, Hussain SM, Schlager JJ, Smith DA,
Syed AF. Metal-based nanoparticles and their toxicity assessment. WIREs Nanomed. Nanobiotechnol. 2010, 2, 554–568.
[174] Alzahrani E, Ahmed RA. Synthesis of copper nanoparticles with various sizes and shapes: application as a superior nonenzymatic sensor and antibacterial agent. Int. J. Electrochem.
Sci. 2016, 11, 4712–4723.
[175] Shobha G, Moses V, Anand, S. Biological synthesis of copper nanoparticles and its impact. Int. J. Pharm. Sci. Invent. 2014,
3, 6–28.
[176] Ramyadevi J, Jeyasubramanian K, Marikani A, Rajakumar G,
Rahuman AA. Synthesis and antimicrobial activity of copper nanoparticles. Mater. Lett. 2012, 71, 114–116.
[177] Yoon KY, Byeon JH, Park JH, Hwang J. Susceptibility constants of Escherichia coli and Bacillus subtilis to silver and copper nanoparticles Sci. Total Environ. 2007, 373, 572–575.
[178] Xu JF, Ji W, Shen ZX, Tang SH, Ye XR, Jia DZ, Xin XQ. Preparation and characterization of CuO nanocrystals. J. Solid State Chem.
2009, 147, 516–519.
[179] Raffi M, Mehrwan S, Bhatti T, Akhter J, Hameed A, Yawar W, ul Hasan, M. Investigations into the antibacterial behavior of copper nanoparticles against Escherichia coli. Ann. Microbiol.
2010, 60, 75–80.
[180] Lee S, Chung H, Kim S, Lee I. The genotoxic effect of ZnO and
CuO nanoparticles on early growth of buckwheat, Fagopyrum esculentum. Water Air Soil Pollut. 2013, 88, 1971–1977.
[181] Pramanik A, Laha D, Bhattacharya D, Pramanik P, Karmakar
P. A novel study of antibacterial activity of copper iodide nanoparticle mediated by DNA and membrane damage. Colloids
Surf. B Biointerfaces 2012, 96, 50–55.
[182] Huh AJ, Kwon YJ. “Nanoantibiotics”: a new paradigm for treating infectious diseases using nanomaterials in the antibiotics resistant era. J. Control Release. 2011, 156, 128–145.
[183] Tamayo LA, Zapata PA, Rabagliati FM, Azócar MI, Muñoz LA,
Zhou X, Thompson GE, Páez MA. Antibacterial and noncytotoxic effect of nanocomposites based in polyethylene and copper nanoparticles. J. Mater. Sci. Mater. Med. 2015,
26, 129.
[184] Chakraborty R, Sarkar RK, Chatterjee AK, Manju U,
Chattopadhyay AP, Basu T. A simple, fast and cost-effective method of synthesis of cupric oxide nanoparticle with promising antibacterial potency: unraveling the biological and chemical modes of action. Biochim. Biophys. Acta (BBA) –
General Subjects 2015, 1850, 845–856.
[185] Wu B, Huang R, Sahu M, Feng X, Biswas P, Tang YJ. Bacterial responses to Cu doped TiO2
nanoparticles. Sci. Total Environ.
2010, 408, 1755–1758.
[186] Gajjar P, Pettee B, Britt DW, Huang W, Johnson PW, Anderson AJ. Antimicrobial activities of commercial nanoparticles against an environmental soil microbe, Pseudomonas putida
KT2440. J. Biol. Eng. 2009, 3, 9.
[187] Kaur G, Singh P, Mehta SK, Kumar S, Dilbaghi N, Chaudhary
GR. A facile route for the synthesis of Co, Ni and Cu metallic nanoparticles with potential antimicrobial activity using novel metallosurfactants. Appl. Surf. Sci. 2017, 404, 254–262.
[188] Du BD, Phu DV, Quoc LA, Hien NQ. Synthesis and investigation of antimicrobial activity of Cu2
O nanoparticles/zeolite.
J. Nanopart. 2017, 2017, 1–6. doi: 10.1155/2017/705684.
[189] Dobrovolný K, Ulbrich P, Švecová M, Rimpelová S, Malinčík
J, Kohout M, Svoboda J, Bartůněk V. Copper nanoparticles in glycerol-polyvinyl alcohol matrix: in situ preparation, stabilisation and antimicrobial activity. J. Alloys Compds 2017, 697,
147–155.

[190] Morsi RE, Alsabagh AM, Nasr SA, Zaki MM. Multifunctional nanocomposites of chitosan, silver nanoparticles, copper nanoparticles and carbon nanotubes for water treatment: antimicrobial characteristics. Int. J. Biol. Macromol. 2017, 97,
264–269.
[191] Woźniak-Budych MJ, Przysiecka Ł, Langer K, Peplińska B,
Jarek M, Wiesner M, Nowaczyk G, Jurga S. Green synthesis of rifampicin-loaded copper nanoparticles with enhanced antimicrobial activity. J. Mater. Sci. Mater. Med. 2017, 28, 42.
[192] Villanueva ME, Diez AM, González JA, Pérez CJ, Orrego M,
Piehl L, Teves S, Copello GJ. Antimicrobial activity of starch hydrogel incorporated with copper nanoparticles. ACS Appl.
Mater. Interfaces 2016, 8, 16280–16288.
[193] Abboud Y, Saffaj T, Chagraoui A, El Bouari A, Brouzi K, Tanane
O, Ihssane B. Biosynthesis, characterization and antimicrobial activity of copper oxide nanoparticles (CONPs) produced using brown alga extract (Bifurcaria bifurcata). Appl. Nanosci.
2013, 4, 571–576.
[194] Usman M, El Zowalaty M, Ibrahim NA, Salama M, Shameli K,
Zainuddin N. Synthesis, characterization, and antimicrobial properties of copper nanoparticles. Int. J. Nanomed. 2013, 8,
4467–4479.
[195] Shannahan JH, Brown JM. Engineered nanomaterial exposure and the risk of allergic disease. Curr. Opin. Allergy Clin. Immunol. 2014, 14, 95–99.
[196] Moghimi SM, Hunter AC. Capture of stealth nanoparticles by the body’s defences. Crit. Rev. Ther. Drug Carrier Syst. 2001,
18, 527–550.
[197] Geiser M, Rothen-Rutishauser B, Kapp N, Schürch S, Kreyling
W, Schulz H, Semmler M, Hof VI, Heyder J, Gehr P. Ultrafine particles cross cellular membranes by nonphagocytic mechanisms in lungs and in cultured cells. Environ. Health Perspect.
2005, 113, 1555–1560.
[198] Manolova V, Flace A, Bauer M, Schwarz K, Saudan P, Bachmann
MF. Nanoparticles target distinct dendritic cell populations according to their size. Eur. J. Immunol. 2008, 38, 1404–1413.
[199] Chono S, Tanino T, Seki T, Morimoto K. Influence of particle size on drug delivery to rat alveolar macrophages following pulmonary administration of ciprofloxacin incorporated into liposomes. J. Drug Target. 2006, 14, 557–566.
[200] Borm PJ, Kreyling W. Toxicological hazards of inhaled nanoparticles potential implications for drug delivery. J. Nanosci.
Nanotechnol. 2004, 4, 521–531.
[201] Li, N, Xia, T. and Nel, A. E. The role of oxidative stress in ambient particulate matter induced lung diseases and its implications in the toxicity of engineered nanoparticles. Free Radic.
Biol. Med. 2008, 44, 1689–1699.
[202] Pettibone JM, Adamcakova-Dodd A, Thorne PS,
O’Shaughnessy PT, Weydert JA, Grassian VH. Inflammatory response of mice following inhalation exposure to iron and copper nanoparticles. Nanotoxicology 2008, 2, 189–204.
[203] Pineda L, Sawosz E, Vadalasetty KP, Chwalibog A. Effect of copper nanoparticles on metabolic rate and development of chicken embryos. Anim. Feed Sci. Technol. 2013, 186,
125–129.
[204] Wang T, Long X, Cheng Y, Liu Z, Yan S. A comparison effect of copper nanoparticles versus copper sulphate on juvenile
Epinephelus coioides: growth parameters, digestive enzymes, body composition, and histology as biomarkers. Int. J.
Genomics 2015, 2015, 1–10.
[205] Manke A, Wang L, Rojanasakul Y. Mechanisms of nanoparticleinduced oxidative stress and toxicity. Biomed Res. Int. 2013,
2013, e942916.
[206] Zhang, L, Bai, R, Liu, Y, Meng, L, Li, B, Wang, L, Xu, L, Le
Guyader, L, Chen, C. The dose-dependent toxicological effects and potential perturbation on the neurotransmitter secretion in brain following intranasal instillation of copper nanoparticles. Nanotoxicology 2012, 6, 562–575.
[207] Wang T, Long X, Liu Z, Cheng Y, Yan S. Effect of copper nanoparticles and copper sulphate on oxidation stress, cell apoptosis and immune responses in the intestines of juvenile
Epinephelus coioides. Fish Shellfish Immunol. 2014, 44,
674–682.
[208] Wang T, Long X, Chen X, Liu Y, Liu Z, Han S, Yan S. Integrated transcriptome, proteome and physiology analysis of
Epinephelus coioides after exposure to copper nanoparticles or copper sulfate. Nanotoxicology 2017, 11, 236–246.
[209] Hua J, Vijver MG, Ahmad F, Richardson MK, Peijnenburg
WJ. Toxicity of different-sized copper nano- and submicron particles and their shed copper ions to zebrafish embryos.
Environ. Toxicol. Chem. 2014, 33, 1774–1782.
[210] Hedberg J, Karlsson HL, Hedberg Y, Blomberg E, Odnevall
Wallinder I. The importance of extracellular speciation and corrosion of copper nanoparticles on lung cell membrane integrity. Colloids Surf. B Biointerfaces 2016, 141, 291–300.
[211] Song L, Connolly M, Fernández-Cruz ML, Vijver MG,
Fernández M, Conde E, de Snoo GR, Peijnenburg WJ, Navas
JM. Species-specific toxicity of copper nanoparticles among mammalian and piscine cell lines. Nanotoxicology 2014, 8,
383–393.
[212] Sadiq R, Khan QM, Mobeen A, Hashmat AJ. In vitro toxicological assessment of iron oxide, aluminium oxide and copper nanoparticles in prokaryotic and eukaryotic cell types. Drug
Chem. Toxicol. 2015, 38, 152–161.
[213] Palza H, Galarce N, Bejarano J, Beltran M, Caviedes P. Effect of copper nanoparticles on the cell viability of polymer composites. Int. J. Polym. Mater. Polym. Biomater. 2017, 66, 462–468.
[214] Ganesan S, Anaimalai Thirumurthi N, Raghunath A, Vijayakumar S, Perumal E. Acute and sub-lethal exposure to copper oxide nanoparticles causes oxidative stress and teratogenicity in zebrafish embryos: oxidative stress and teratogenicity of
CuO-NPs. J. Appl. Toxicol. 2016, 36, 554–567.
[215] Sun Y, Zhang G, He Z, Wang Y, Cui J. Effects of copper oxide nanoparticles on developing zebrafish embryos and larvae.
Int. J. Nanomed. 2016, 11, 905–918.
[216] Liao M, Liu H. Gene expression profiling of nephrotoxicity from copper nanoparticles in rats after repeated oral administration. Environ. Toxicol. Pharmacol. 2012, 34,
67–80.
[217] Doudi M, Setorki M. Acute effect of nano-copper on liver tissue and function in rat. Nanomed. J. 2014, 1, 331–338.
[218] Ostaszewska T, Chojnacki M, Kamaszewski M,
Sawosz-Chwalibóg E. Histopathological effects of silver and copper nanoparticles on the epidermis, gills, and liver of Siberian sturgeon. Environ. Sci. Pollut. Res. 2016, 23,
1621–1633.
[219] Lee I-C, Kim J-C, Ko J-W, Park S-H, Lim J-O, Shin I-S, Moon C,
Kim S-H, Her J-D. Comparative toxicity and biodistribution of copper nanoparticles and cupric ions in rats. Int. J. Nanomed.
2016, 11, 2883–2900.
[220] Chen Z, Meng H, Xing G, Chen C, Zhao Y, Jia G, Wang T, Yuan
H, Ye C, Zhao F, Chai Z, Zhu C, Fang X, Ma B, Wan L. Acute toxicological effects of copper nanoparticles in vivo. Toxicol.
Lett. 2006, 163, 109–120.
[221] Hill EK, Li J. Current and future prospects for nanotechnology in animal production. J. Anim. Sci. Biotechnol. 2017, 8, 26.
[222] Prescott JF, Baggot JD. Antimicrobial Therapy in Veterinary
Medicine, 2nd ed., Iowa State University Press: Ames, IA,
1993, pp. 564–565.
[223] Mahler GJ, Esch MB, Tako E, Southard TL, Archer SD, Glahn RP,
Shuler ML. Oral exposure to polystyrene nanoparticles affects iron absorption. Nat. Nanotechnol. 2012, 7, 264–271.
[224] Selim NA, Radwan NL, Youssef SF, Eldin TAS, Elwafa SA. Effect of inclusion inorganic, organic or nano selenium forms in broiler diets on: 1-growth performance, carcass and meat characteristics. Int. J. Poult. Sci. 2015, 14, 135–143.
[225] Bagheri M, Golchin-Gelehdooni S, Mohamadi M, Tabidian A.
Comparative effects of nano, mineral and organic selenium on growth performance, immunity responses and total antioxidant activity in broiler chickens. Int. J. Biol. Pharm. Allied Sci. (IJBPAS) 2015, 4, 583–595.
[226] Boostani A, Sadeghi AA, Mousavi SN, Chamani M, Kashan N.
The effects of organic, inorganic, and nano-selenium on blood attributes in broiler chickens exposed to oxidative stress. Acta
Sci. Vet. 2015, 43, 1–6.
[227] Milani NC, Sbardella M, Ikeda NY, Arno A, Mascarenhas BC,
Miyada VS. Dietary zinc oxide nanoparticles as growth promoter for weanling pigs. Anim. Feed Sci. Technol. 2017, 227, 13–23.
[228] Mishra A, Swain RK, Mishra SK, Panda N, Sethy K. Growth performance and serum biochemical parameters as affected by nano zinc supplementation in layer chicks. Indian J. Anim.
Nutr. 2014, 31, 384–388.
[229] Swain PS, Rao SBN, Rajendran D, Dominic G, Selvaraju S.
Nano zinc, an alternative to conventional zinc as animal feed supplement: a review. Anim. Nutr. 2016, 2, 134–141.
[230] Sirirat N, Lu J-J, Tsung-Yu Hung A, Chen S-Y, Lien T-F. Effects different levels of nanoparticles chromium picolinate supplementation on growth performance, mineral retention, and immune responses in broiler chickens. J. Agric. Sci. 2012,
4, 48.
[231] Pineda L, Chwalibog A, Sawosz E, Lauridsen C, Engberg R,
Elnif J. Effect of silver nanoparticles on growth performance, metabolism and microbial profile of broiler chickens. Arch.
Anim. Nutr. 2012, 66, 416–429.
[232] Fondevila M, Herrer R, Casallas MC, Abecia L, Ducha JJ. Silver nanoparticles as a potential antimicrobial additive for weaned pigs. Anim. Feed Sci. Technol. 2009, 150, 259–269.
[233] Sawosz F, Pineda L, Hotowy A, Hyttel P, Sawosz E, Szmidt M,
Niemiec T, Chwalibog A. Nano-nutrition of chicken embryos.
The effect of silver nanoparticles and glutamine on molecular responses, and the morphology of pectoral muscle. Comp.
Biochem. Physiol. A 2012, 161, 315–319.
[234] Bhanja S, Hotowy A, Mehra M, Sawosz E, Pineda L, Vadalasetty K, Kurantowicz N, Chwalibog A. In ovo administration of silver nanoparticles and/or amino acids influence metabolism and immune gene expression in chicken embryos. Int. J. Mol.
Sci. 2015, 16, 9484–9503.
[235] Ravikumar S, Gokulakrishnan R. The inhibitory effect of metal oxide nanoparticles against poultry pathogens. Int. J. Pharm.
Sci. Drug Res. 2012, 4, 157–159.
[236] Mroczek-Sosnowska N, Łukasiewicz M, Adamek D, Kamaszewski M, Niemiec J, Wnuk-Gnich A, Scott A, Chwalibog A,
Sawosz E. Effect of copper nanoparticles administered in ovo on the activity of proliferating cells and on the resistance of femoral bones in broiler chickens. Arch. Anim. Nutr. 2017, 71,
327–332.
[237] Zheng SM, Guo LX, Zhan X. Copper silicate nanoparticles: effects of intestinal microflora, nitrogen metabolism and ammonia emission from excreta of yellow-feathered broilers.
Chin. J. Anim. Nutr. 2013, 8, 1837–1844.
[238] Nguyen QK, Nguyen DD, Nguyen VK, Nguyen KT, Nguyen HC,
Tran XT, Nguyen HC, Phung DT. Impact of biogenic nanoscale metals Fe, Cu, Zn and Se on reproductive LV chickens. Adv.
Nat. Sci. Nanosci. Nanotechnol. 2015, 6, 35017.
[239] Minglei S, Zheng L, Xiaoye G, Xiu’an Z. Copper silicate nanoparticles: effects of intestinal microflora, nitrogen metabolism and ammonia emission from excreta of yellow-feathered broilers. Nutr. Feed Sci. Technol. 2013, 25, 1843–1850.
[240] Muralisankar T, Saravana Bhavan P, Radhakrishnan S,
Seenivasan C, Srinivasan V. The effect of copper nanoparticles supplementation on freshwater prawn Macrobrachium rosenbergii post larvae. J. Trace Elem. Med. Biol. 2016, 34,
39–49.
[241] El Basuini MF, El-Hais AM, Dawood MAO, Abou-Zeid AE-S,
EL-Damrawy SZ, Khalafalla MME-S, Koshio S, Ishikawa M, Dossou S. Effects of dietary copper nanoparticles and vitamin C supplementations on growth performance, immune response and stress resistance of red sea bream. Pagrus major. Aquac.
Nutr. 2017, 1–12. doi:10.1111/anu.12508
[242] Han, X.-Y, Du, W.-L, Fan, C.-L, Xu, Z.-R. Changes in composition a metabolism of caecal microbiota in rats fed diets supplemented with copper-loaded chitosan nanoparticles: CSN changes composition and metabolism of caecal microbiota in rats. J. Anim. Physiol. Anim. Nutr. 2010, 94, e138–e144.
[243] Han X-Y, Du W-L, Huang, Q-C, Xu Z-R, Wang Y-Z. Changes in small intestinal morphology and digestive enzyme activity with oral administration of copper-loaded chitosan nanoparticles in rats. Biol. Trace Elem. Res. 2012, 145, 355–360.
[244] Sarvestani S, Rezvani MR, Zamiri MJ, Shekarforoush S, Atashi
H, Mosleh N. The effect of nanocopper and mannan oligosaccharide supplementation on nutrient digestibility and performance in broiler chickens. J. Vet. Res. 2016, 71, 153–161.
[245] Ognik K, Stępniowska A, Cholewińska E, Kozłowski K. The effect of administration of copper nanoparticles to chickens in drinking water on estimated intestinal absorption of iron, zinc, and calcium. Poult. Sci. 2016, 95, 2045–2051.
[246] Wang M-Q, Du Y-J, Wang C, Tao W-J, He Y-D, Li H. Effects of copper-loaded chitosan nanoparticles on intestinal microflora and morphology in weaned piglets. Biol. Trace Elem. Res.
2012, 149, 184–189.
[247] Sawosz E, Binek M, Grodzik M, Zielińska M, Sysa P, Szmidt
M, Niemiec T, Chwalibog A. Influence of hydrocolloidal silver nanoparticles on gastrointestinal microflora and morphology of enterocytes of quails. Arch. Anim. Nutr. 2010, 61, 444–451.
[248] Tomaszewska E, Muszyński S, Ognik K, Dobrowolski P,
Kwiecień M, Juśkiewicz J, Chocyk D, Świetlicki M, Blicharski T,
Gładyszewska B. Comparison of the effect of dietary copper nanoparticles with copper (II) salt on bone geometric and structural parameters as well as material characteristics in a rat model. J. Trace Elem. Med. Biol. 2017, 42, 103–110.

Related topics:
Authors:
Abdullah Scott
Recommend
Comment
Share
Profile picture
Would you like to discuss another topic? Create a new post to engage with experts in the community.
Featured users in Poultry Industry
Lorena Ramos
Lorena Ramos
Cargill
United States
Kendra Waldbusser
Kendra Waldbusser
Pilgrim´s
United States
Phillip Smith
Phillip Smith
Tyson
Tyson
United States
Join Engormix and be part of the largest agribusiness social network in the world.