Acessibilidade / Reportar erro

Scale-up in the synthesis of nanoparticles for encapsulation of agroindustrial active principles

Escalonamento na síntese de nanopartículas para encapsulamento de princípios ativos agroindustriais

ABSTRACT

Given the potentiality of nanoparticles (NP) to load substances as active principle of drugs, cosmetics and food, there is interest in increasing its production. Particularly in the agroindustrial area, the byproducts are source of functional compounds that must be stabilized by encapsulation, for example, to allow its application. Considering the amount of generated byproducts, it is necessary to understand the encapsulation synthesis in a high scale dimension. In this way, the active compounds vitamin C (from the byproduct of the acerola processing) and polyphenols (from the byproduct of grape processing) were nanoencapsulated into chitosan NP at three reactional volumes: 0.1; 1.0 and 10.0 dm3. The NP were characterized for yield, particle size, morphology, encapsulation efficiency and release profile. In general, the increase in scale did not influence these parameters. It is found a close similarity for NP size value between the 0.1 and 10.0 dm3 reactional volumes. For example, for the blank control, the size was 155±6 and 227±118 nm respectively for 0.1 and 10.0 dm3 reactional volumes. Similarly for the encapsulated acerola byproduct extract (373±44 and 400±83 nm) and ascorbic acid (279±29 and 217±29 nm). For the encapsulated skin grape extract, the size decreased sharply from 1040 to 308 nm. The yield per volume ratio was about 1.3 mg cm-3. Additional analysis for NP with encapsulated skin grape extract, by Scanning Electronic Microscopy, showed uniformly distributed spherical structures and the release profile was similar for all reactional volumes. Thus, the system is suitable for scale-up for NP production.

Index terms:
Vitamin C; polyphenols; byproduct; chitosan

RESUMO

Dada a potencialidade das nanopartículas (NP) de serem empregadas como vetores para administração de drogas, cosméticos e alimentos, tem-se procurado aumentar a sua produção. Particularmente no setor agroindustrial, os subprodutos são fontes de compostos ativos que precisam ser estabilizados para garantir sua aplicação. No entanto, dado o volume de subproduto gerado, é preciso estudar o encapsulamento em escalas maiores. Assim, visando determinar a influência do escalonamento nas propriedades da NP, a vitamina C (extraída do resíduo do processamento de acerola) e polifenóis (obtido do resíduo do processamento de uva) foram nanoencapsulados em NP de quitosana a partir de três volumes reacionais diferentes: 0,1; 1,0 e 10,0 dm3. As NP foram caracterizadas em relação ao rendimento, tamanho de partícula, morfologia e perfil de liberação. De forma geral, o aumento de escala não influenciou os parâmetros estudados. Assim, os valores de tamanho das NP entre os volumes reacionais de 0,1 e 10,0 dm3 foram muito similares. Por exemplo, para o controle (não encapsulado), o tamanho foi 155±6 e 227±118 nm, respectivamente, para os volumes reacionais de 0,1 e 10,0 dm3. O mesmo padrão foi observado para o extrato do subproduto de acerola encapsulado (373±44 e 400±83 nm) e o ácido ascórbico (279±29 e 217±29 nm). Para o extrato da casca de uva encapsulado, houve um decréscimo de tamanho, variando de 1040 a 308 nm. A razão rendimento por volume ficou em torno de 1,3 mg cm-3. Análises adicionais para NP com extrato da casca de uva encapsulado, por Microscopia Eletrônica de Varredura, mostrou estruturas esféricas uniformemente distribuídas e o perfil de liberação foi independente do volume reacional. Desta forma, o sistema é adequado para o escalonamento.

Termos para indexação:
Vitamina C; polifenóis; subprodutos; quitosana

INTRODUCTION

The nanoencapsulation is a useful procedure to apply controlled release of active compounds with potential application in the pharmacological (Mora-Huertas; Fessi; Elaissari, 2010MORA-HUERTAS, C. E.; FESSI, H.; ELAISSARI, A. Polymer-based nanocapsules for drug delivery. International Journal of Pharmaceutics, 385(1-2):113-142, 2010.), food (Fathi; Martín; McClements, 2014FATHI, M.; MARTÍN, Á.; MCCLEMENTS, D. J. Nanoencapsulation of food ingredients using carbohydrate based delivery systems. Trends in Food Science & Technology, 39(1):18-39, 2014.) and agricultural (Badang; Chakraborty, 2019BADANG, M. C. N.; CHAKRABORTY, S. Carbohydrate polymers as controlled release devices for pesticides. Journal of Carbohydrate Chemistry, 38(1):67-85, 2019.) areas. Several matrices can be used to achieve the encapsulation and to deliver the active substance. Among the matrices, those based on biodegradable material is attractive due to apply “Generally Recognized As Safe” (GRAS) materials (Kumari; Yadav; Yadav, 2010KUMARI, A.; YADAV, S. K.; YADAV, S. C. Biodegradable polymeric nanoparticles based drug delivery systems. Colloids and Surfaces B: Biointerfaces , 75(1):1-18, 2010.). This designation is fundamental mainly for food and pharmaceutical applications (Santos et al., 2016SANTOS, P. P. et al. Biodegradable polymers as wall materials to the synthesis of bioactive compound nanocapsules. Trends in Food Science & Technology , 53:23-33, 2016.). Chitosan, as an example of such material, is a special polysaccharide that has been widely used to encapsulate and deliver several classes of active compounds such as vitamins (Britto et al., 2014BRITTO, D. et al. Entrapment characteristics of hydrosoluble vitamins loaded into chitosan and N,N,N-trimethy chitosan nanoparticles. Macromolecular Research, 22(12):1261-1267, 2014.; Britto et al., 2012BRITTO, D. et al. N,N,N-trimethyl chitosan nanoparticles as a vitamin carrier system. Food Hydrocolloids, 27(2):487-493, 2012.) and essential oil (Feyzioglu; Tornuk, 2016FEYZIOGLU, G. C.; TORNUK, F. Development of chitosan nanoparticles loaded with summer savory (Satureja hortensis L.) essential oil for antimicrobial and antioxidant delivery applications. LWT - Food Science and Technology, 70:104-110, 2016.). Regarding the methods, the ionic gelation is largely studied due to its easiness and mild condition of synthesis (Gan et al., 2005GAN, Q. et al. Modulation of surface charge, particle size and morphological properties of chitosan-TPP nanoparticles intended for gene delivery. Colloids and Surfaces B, 44(2-3): 65-73, 2005.). This procedure is useful for polysaccharide polyelectrolyte such as chitosan that allows interaction with charged and polar species to form stable nanohydrogel. In the case of chitosan, that is a cationic polyelectrolyte, the most used ionic gelation inductor is the sodium tripolyphosphate (TPP), with five negative charged groups (Britto et al., 2014BRITTO, D. et al. Entrapment characteristics of hydrosoluble vitamins loaded into chitosan and N,N,N-trimethy chitosan nanoparticles. Macromolecular Research, 22(12):1261-1267, 2014.; Britto et al., 2012BRITTO, D. et al. N,N,N-trimethyl chitosan nanoparticles as a vitamin carrier system. Food Hydrocolloids, 27(2):487-493, 2012.).

However, many factors can influence the chitosan-TPP system in the encapsulation of actives compounds. The chitosan concentration, the ratio chitosan to TPP, pH, kind of acid used to solubilize the chitosan and the solubility of the encapsulated compound play important role in the nanoparticle structure, mainly in its size (Gan et al., 2005GAN, Q. et al. Modulation of surface charge, particle size and morphological properties of chitosan-TPP nanoparticles intended for gene delivery. Colloids and Surfaces B, 44(2-3): 65-73, 2005.) and load capacity/efficiency (Britto et al., 2012BRITTO, D. et al. N,N,N-trimethyl chitosan nanoparticles as a vitamin carrier system. Food Hydrocolloids, 27(2):487-493, 2012.; Gan; Wang, 2007GAN, Q.; WANG, T. Chitosan nanoparticle as protein delivery carrier: Systematic examination of fabrication conditions for efficient loading and release. Colloids and Surfaces B: Biointerfaces, 59(1):24-34, 2007.). In an attempt to better understand the attribution of such parameters, studies have been conducted to evaluate the influence of the chitosan molecular weight, the ratio of crosslinking agent TPP and the pH on the property of synthetized NP (Gan et al., 2005GAN, Q. et al. Modulation of surface charge, particle size and morphological properties of chitosan-TPP nanoparticles intended for gene delivery. Colloids and Surfaces B, 44(2-3): 65-73, 2005.). Concerning the NP sizes, it increases steadily with chitosan concentration, equally for low, medium or high molecular weight chitosan. Small-sized NP can be achieved by reacting a low concentration of low or medium molecular weight chitosan. Moreover, by increasing the ratio of chitosan to TPP, the NP size also increases. Finally, low pH favored small-sized NP (Gan et al., 2005GAN, Q. et al. Modulation of surface charge, particle size and morphological properties of chitosan-TPP nanoparticles intended for gene delivery. Colloids and Surfaces B, 44(2-3): 65-73, 2005.; Gokce et al., 2014GOKCE, Y. et al. Ultrasonication of chitosan nanoparticle suspension: Influence on particle size. Colloids and Surfaces A: Physicochemical and Engineering Aspects, 462:75-81, 2014.).

Several efforts have been driven to get the better condition for encapsulation by chitosan-TPP ionic gelation, e.g., through factorial design (Kiilll et al., 2017KIILLL, C. P. et al. Synthesis and factorial design applied to a novel chitosan/sodium polyphosphate nanoparticles via ionotropic gelation as an RGD delivery system. Carbohydrate Polymers, 157:1695-1702, 2017.) or by systematic statistical Design of Experiments (Sreekumar et al., 2018SREEKUMAR, S. et al. Parameters influencing the size of chitosan-TPP nano- and microparticles. Scientific Reports, 8(4695):1-11, 2018.). However, the large-scale production of NP based on TPP has not been explored. While, for other NP systems, such studies were led, for example, for silver-chitosan NP (Wongpreecha et al., 2018WONGPREECHA, J. et al. One-pot, large-scale green synthesis of silver nanoparticles-chitosan with enhanced antibacterial activity and low cytotoxicity. Carbohydrate Polymers , 199: 641-648, 2018.) and chitosan nanofibers (Wang et al., 2016WANG, L. et al. Needleless electrospinning for scaled-up production of ultrafine chitosan hybrid nanofibers used for air filtration. RSC Advances, 6(107):105988-105995, 2016.). In this way, it is necessary to study such synthesis in large scale, aiming its potential application in the industrial segment. Regarding this, the vitamins are important active compounds to be encapsulated as an ingredient in premixes and feed for fish diets, mainly that hydrosoluble ones (Britto et al., 2016BRITTO, D. et al. Analysis of thermal and aqueous suspension stabilities of chitosan based nanoencapsulated vitamins. Química Nova, 39(9):1126-1130, 2016.). Other important substances are the phenolic compounds as an active antifungal ingredient in edible coatings to extend the shelflife of perishable fruits and vegetables (Friedman, 2014FRIEDMAN, M. Antibacterial, antiviral, and antifungal properties of wines and winery byproducts in relation to their flavonoid content, Journal of Agricultural and Food Chemistry, 62(26):6025-6042, 2014.).

Active substances such as vitamin and phenolic compound are found in agroindustrial byproduct (Valduga et al., 2008VALDUGA, E. et al. Extração, secagem por atomização e microencapsulamento de antocianinas do bagaço da uva “Isabel” (Vitis labrusca). Ciência e Agrotecnologia, 32(5):1568-1574, 2008.). Thus, it is important to evaluate the possibility of direct nanoencapsualtion of such substance from the aqueous extract. This can be carried out, for example, from acerola byproduct, once a previous study showed that in the processed residue still has a significant amount of residual vitamin C (Nascimento et al., 2019NASCIMENTO, J. A. A. et al. Stability of nanocomposite edible films based on polysaccharides and vitamin C from agroindustrial residue. Materials Research, 22(3): e20190057, 2019.). Likewise, for winery byproducts, many phenolic compounds remain in the residue (Melo et al., 2015MELO, P. S. et al. Winery by-products: Extraction optimization, phenolic composition and cytotoxic evaluation to act as a new source of scavenging of reactive oxygen species. Food Chemistry , 181:160-169, 2015.).

In this way, aiming the synthesis of such NP in a scale compatible with the generated volume of byproducts, some NP characteristics via ionic gelation with TPP were investigated in different scale of production. For this, aqueous extract from the byproduct of acerola juice production (rich in vitamin C) and aqueous/ethanol extract from the byproduct of wine industries (rich in polyphenols) were used as models.

MATERIAL AND METHODS

Material

Sodium tripolyphosphate (TPP) and medium molecular weight chitosan (Chi), 80% deacetylated, were purchased from Aldrich Chemical Company Inc. (USA) and used as supplied. Others chemical as acetic acid (HAc) (Vetec®, 99.7%); metaphosphoric acid P.A. (Vetec®); hydrochloric acid (37% HCl, Alphatec®); oxalic acid dihydrate P.A. (Vetec®, 99.5-102.5%), L-(+)-ascorbic acid P.A. (AA) (Vetec®) and ethanol P.A. (Sciavicco®, 99.5%) were used as commercially purchased.

The byproduct from acerola processing (Malpighia emarginata D.C., ‘Sertaneja’ variety) was supplied by Niagro-Nichirei do Brasil Agrícola Ltda. and the byproduct from winery activity (Vitis Vinifera, ‘Egiodolla’ variety) was supplied by ViniBrasil, both located in Petrolina, PE, Brazil.

Extraction of vitamin C from acerola byproduct

The basic methodology for the extraction of vitamin C from acerola byproduct consisted in weighing 5.0 g of the residue in a Falcon tube and subjecting it to extraction with 50.0 cm3 chitosan aqueous solution (3.0 mg cm-3) in HAc 0.5% (v/v). After manual stirring for dispersion of the residue, the mixture was centrifuged at 12,857 x g (centrifuge Eppendorf, model 5804 R, rotor F-34-6-38, 10,000 rpm) for 10 minutes and 8 °C for separation of the solid residue and obtaining the supernatant rich in vitamin C. This supernatant was then subjected to NP synthesis and encapsulation, as described in following.

Extraction of polyphenols from wine byproduct

The methodology for extraction of polyphenols was based in a hydro-alcoholic extraction. For this, at first, the seeds were separated from the wine byproduct and 200 g of grape skin was extracted with water-ethanol, as described in the literature (Moura et al., 2011MOURA, R. S. et al. Addition of açaí (Euterpe oleracea) to cigarettes has a protective effect against emphysema in mice. Food and Chemical Toxicology, 49(4):855-863, 2011.). After the filtration, rota-evaporation and lyophilization, a dry material was obtained and stored in the dark at -20 ºC until the utilization.

Synthesis and scale-up of chitosan nanoparticle

The basic procedure for chitosan NP synthesis via ionic gelation consisted in to react 50.0 cm3 of chitosan solution (3.0 mg cm-3) dissolved in HAc (0.5%, v/v) with 50.0 cm3 TPP solution (1.2 mg cm-3) to get 0.1 dm3 final volume (Britto et al., 2014BRITTO, D. et al. Entrapment characteristics of hydrosoluble vitamins loaded into chitosan and N,N,N-trimethy chitosan nanoparticles. Macromolecular Research, 22(12):1261-1267, 2014.). In this case, the pH of the reactant solutions of chitosan and TPP was respectively 2.93 and 9.58. After the synthesis, the pH of the NP suspension was 4.4. The TPP solution was added to the chitosan solution at an approximate speed of 1.0 cm3 s-1 under magnetic stirring (~ 500 rpm). This NP suspension was taken as a blank sample and called NPblk. The same procedure was repeated, accounting, however, for final volumes of 1.0 dm3 and 10.0 dm3. Some adjustment was necessary to implement such scale-up, mainly for synthesis at 10.0 dm3. The first was substituting the magnetic stirring for a mechanical one, maintaining the rotation at 500 rpm. It was necessary also to increase the TPP speed addition to near 100.0 cm3 s-1.

The encapsulation of vitamin C from the acerola byproduct followed the same procedure, substituting the initial chitosan solution by the supernatant rich in vitamin C. This NP suspension was denominated NPace. For comparison, a control system, containing 30.0 mg AA added to 50.0 cm3 chitosan solution in HAc, was also prepared and encapsulated in the same way. This NP suspension was denominated NPAA.

The polyphenol extract was encapsulated following the same process, with the previous addition of 80.0 mg of dry extract to the TPP solution. In this case, the TPP concentration was 1.0 mg cm-3. The NP suspension was called NPgra.

Characterization of nanoparticles suspension

The NP suspensions were analyzed for particle size and zeta potential by Dynamic Light Scattering (DLS) technique (Malvern Instruments®, model ZS Zen 3600). The NP suspension was filled into the disposable cuvette without dilution.

The Encapsulation Efficiency (EE) was determined by UV-visible spectrophotometry (ThermoScientific® spectrophotometer, model MultiSkan GO). First, the NP suspension was centrifuged at 35,392 x g (centrifuge Beckman Coulter, model Avanti J26-XP, rotor JA 25.15, 20,000 rpm) for 20 minutes at 8 °C and the supernatant quantified for vitamin C or phenolic compound. For this, calibrations curves were set with appropriated concentrations of AA in metaphosphoric acid (1.0% w/v) or freeze-dried skin grape extract in ethanol-water solution (1:1). The maximum absorbance in the UV-Visible spectrum (200-400 nm range) was 240 nm for AA and 278 nm for grape extract. The resultant calibration curve fit was y = 0.0131 + 56.2476x (R2 = 0.9995) for ascorbic acid and y = -0.1878 + 8.7375x (R2 = 0.9738) for grape extract. The EE was calculated according to the following equation:

EE = total initial substance non encapsulated substance total initial substance × 100 (1)

The precipitate resulted from the centrifugation was submitted to lyophilization for yield calculation and expressed as a ratio of recovered mass per reactional volume.

Given the high EE value for NP suspension with encapsulated skin grape extract, it was chosen for morphologic and release profile additional analyses. The morphologic analysis was done by Scanning Electron Microscopy (SEM) (Tescan® equipment, model Vega 3). For this, the samples were previously centrifuged at 35,392 x g (centrifuge Beckman Coulter, model Avanti J26-XP, rotor JA 25.15, 20,000 rpm) for 20 minutes at 8 °C and stabilized in a sucrose solution at 5% before lyophilization (cryogenic treatment). Following, the dry sample was washed with deionized water under ultrasonic bath several times. After that, the suspension was deposited in a glass sheet and coated with a gold layer.

For the release profile, 10.0 g of the centrifuged precipitated NP was suspended in 60% ethanol-phosphate buffer solution under orbital agitation at 120 rpm and 25 ºC. At pre-determinate time intervals, aliquots of the suspension were taken, centrifuged and the supernatant quantified by UV-Visible spectroscopy, as described above. The experiment was done in duplicate. The normalization was done by the Origin® Lab SR0 software, v. 8.0725, applying the [0.1] method.

RESULTS AND DISCUSSION

Encapsulation efficiency and yield in the scale-up

The systems did not show significant variations for supernatant concentration, encapsulation efficiency and yield as function of the reactional volume (Table 1). However, for NP suspensions with encapsulated ascorbic acid and acerola extract, there was a slight decrease in EE according to the increase of the reactional volume. On the other hand, for NPgra there was no trend in the EE variation. Comparing EE values for NPace and NPgra, it is seen a great difference, in which the former showed lower EE values.

Table 1:
Values for supernatant concentration (SC, mg cm-3), encapsulation efficiency (EE, %) and yield ratio (YR, mg cm-3), as a ratio of recovered mass per reactional volume, for nanoparticles suspensions in the three reactional volumes.

The variation observed for the parameters SC, EE and YR are in agreement with discussed before and is due mainly to the solubility and solvatation of each encapsulated compound in the reactional solvent (Britto et al., 2012BRITTO, D. et al. N,N,N-trimethyl chitosan nanoparticles as a vitamin carrier system. Food Hydrocolloids, 27(2):487-493, 2012.). Another important data in the characterization of NP is its yield expressed in terms of total mass recovered after centrifugation and separation from the supernatant. According to the results, the yield ratio did not show significant variations for any of the reactional volumes (Table 1). These results indicate the stability of the system through the scale-up in the range 0.1 to 10.0 dm3.

Nanoparticle size and zeta potential in the scale-up

The size of NP measured by DLS showed variations in function of the reactional volume and the presence of encapsulated substance (Table 2). For the blank system (without active compound), the NP size had the lowest value, considering the reactional volume of 0.1 dm3. For 1.0 dm3, it increased but remained the same value for 10.0 dm3. This size distribution is in accordance with reported early (Gan et al., 2005GAN, Q. et al. Modulation of surface charge, particle size and morphological properties of chitosan-TPP nanoparticles intended for gene delivery. Colloids and Surfaces B, 44(2-3): 65-73, 2005.; Britto et al., 2012BRITTO, D. et al. N,N,N-trimethyl chitosan nanoparticles as a vitamin carrier system. Food Hydrocolloids, 27(2):487-493, 2012.; Neves et al., 2014NEVES, A. L. P. et al. Factorial design as tool in chitosan nanoparticles development by ionic gelation technique. Colloids and Surfaces A: Physicochemical and Engineering Aspects , 445:34-39, 2014.). This result attests the feasibility of the methods for scale-up, considering the adaptations for large scale that was made. In the same way, for the NP suspension with encapsulated ascorbic acid, the variation in size was not significant, mainly for the reactional volumes of 0.1 dm3 and 10.0 dm3. Generally, comparing the blank control with the encapsulated ones, there is an increase in the NP size. This was observed before and it is related to the increase of crosslinking density among the species (Gan; Wang, 2007GAN, Q.; WANG, T. Chitosan nanoparticle as protein delivery carrier: Systematic examination of fabrication conditions for efficient loading and release. Colloids and Surfaces B: Biointerfaces, 59(1):24-34, 2007.; Britto et al., 2012BRITTO, D. et al. N,N,N-trimethyl chitosan nanoparticles as a vitamin carrier system. Food Hydrocolloids, 27(2):487-493, 2012.). This find confirms the huge influence caused on the NP architecture by the addition of third species (encapsulated substance) in comparison with the dual system (polymer and crosslinker only) (Britto et al., 2012BRITTO, D. et al. N,N,N-trimethyl chitosan nanoparticles as a vitamin carrier system. Food Hydrocolloids, 27(2):487-493, 2012.; Geçer et al., 2010GEÇER, A. et al. Trimethyl chitosan nanoparticles enhances dissolution of the poorly water-soluble drug Candesartan Cilexetil. Macromolecular Research , 18(10):986-991, 2010.). For skin grape extract, there was a sharp downward trend in size for larger reactional volume. Finally, for acerola extract, the first and last reactional volumes were similar regarding NP size (Nascimento et al., 2019NASCIMENTO, J. A. A. et al. Stability of nanocomposite edible films based on polysaccharides and vitamin C from agroindustrial residue. Materials Research, 22(3): e20190057, 2019.). According to the results, it is found a close similarity for NP size distribution between the first and the last reactional volumes studied. In this way, the ionic gelation process with chitosan and TPP is viable for scale-up.

Table 2:
Values for nanoparticle size (nm) and zeta potential (ZP, mV) for nanoparticles suspensions in the three reactional volumes.

Regarding the zeta potential value, there was a slight decrease for NPblk mainly for the highest reactional volume (Table 2). However, for NPAA, NPgra and NPace, there is no specific trend, except a high overall value for NPgra, mainly at 0.1 dm3 and 1.0 dm3, in comparison with NPblk. The zeta potential gives a measurement of the stability of the system regarding the net surface charge. When chitosan and TPP are mixed in dilute acid solution, they spontaneously form nanocomplexes with an overall positive surface charge (Gan et al., 2005GAN, Q. et al. Modulation of surface charge, particle size and morphological properties of chitosan-TPP nanoparticles intended for gene delivery. Colloids and Surfaces B, 44(2-3): 65-73, 2005.). In this way, depending on the structure of the complex polymer-encapsulated substance-TPP, its value can either increase or decrease. The general high values for zeta potential observed for the encapsulated systems confirm the stable character of these nanoparticle suspensions in all the reactional volumes.

Nanoparticles morphology

The SEM analysis showed regular and spherical NP distribution for NPgra obtained at 0.1 dm3 (Figure 1a). However, it is also seen large structure marked as “A”, resulted from the aggregate formation. On the other hand, for NPgra obtained at 1.0 dm3, the NP distribution is very irregular (Figure 1b). Selecting randomly 10 structures in Figure 1a and calculating its diameters, it was found an average of 431 nm (Table 3). This value is inferior that calculated by the DLS technique. The main reason for this is that the SEM technique is related to the dry solid-state in comparison with the hydrodynamical one by the DLS technique.

Figure 1:
Scanning Electron Microscopy pictures for the nanoparticles suspension with encapsulated skin grape extract deposited on a glass sheet for sample synthetized at a) 0.1 dm3 and b) 1.0 dm3.

Table 3:
Size distribution for nanoparticles suspension with encapsulated skin grape extract synthetized at 0.1 dm3, according to Scanning Electron Microscopy measurement in Figure 1.

The morphological analysis also confirmed the size variation between the reactional volumes of 0.1 and 1.0 dm3. Despite the irregularity in the shape distribution (Figure 1b), in general, the size distribution for the sample synthetized at 1.0 dm3 is inferior that observed for the sample synthetized at 0.1 dm3 (Figure 1a) that is accordance with found in Table 2.

The isolation process by centrifugation and lyophilization generated some “defects”, resulting in agglomeration in the isolated hydrogel (“A” structures in Figure 1a). Although, in general, the cryogenic treatment preserves the NP structure (Shahgaldian et al., 2003SHAHGALDIAN, P. et al. A study of the freeze-drying conditions of calixarene based solid lipid nanoparticles. European Journal of Pharmaceutics and Biopharmaceutics, 55(2):181-184, 2003.). However, some physical-chemistry variation in the isolation process, e.g., aggregation after centrifugation, dispersion by ultrasound, can influence the regularity and distribution of the NP.

Release profile

The release profile for the sample NPgra showed differences in the three reactional volumes regarding the absolute concentration (Figure 2a). In a period of 24 hours, the sample NPgra, synthetized at the lowest reactional volume, exhibited an accumulated final concentration of 0.3 mg cm3; for NPgra synthetized at intermediate volume, the accumulated final concentration was 2.0 mg cm3 and, finally, for that synthetized at the highest volume, the accumulated final concentration was 3.4 mg cm3. On the other hand, for the normalized concentration (Figure 2b), the release profile was similar in all cases.

Figure 2:
Release profile for the nanoparticles suspension with encapsulated skin grape extract synthetized in the three reactional volumes: 0.1 dm3 (▲), 1.0 dm3 (●) and 10.0 dm3 (■) relative to a) absolute concentration and b) normalized concentration.

The difference found in the accumulated final concentration in function of the reactional volume may be related to the number of centrifugation cycles necessary to isolate the NP. For high reactional volume, it was necessary more centrifugation cycles to precipitate the NP, getting a more concentrated hydrogel. However, the close similarity in the normalized plot (Figure 2b) indicates the absence of difference in the release profile. The profile is typical of an initial burst release, reaching a plateau of constant release after near 5 hours. Such time is strictly dependent on the several factors, e.g., solvent composition, pH, molecular weight of the encapsulated compound, NP characteristics and others (Britto et al., 2012BRITTO, D. et al. N,N,N-trimethyl chitosan nanoparticles as a vitamin carrier system. Food Hydrocolloids, 27(2):487-493, 2012.). In a recent study with the polyphenol epigallocatechin gallate encapsulated in chitosan-zein NP, the plateau was reached in different time interval whether the release condition was 50% or 95% of ethanol (Liang et al., 2017LIANG, J. et al. Encapsulation of epigallocatechin gallate in zein/chitosan nanoparticles for controlled applications in food systems. Food Chemistry, 231:19-24, 2017.). On the other hand, for another polyphenol gallic acid encapsulated in poly(lactic-co-glycolic acid) nanoparticles, the plateau was reached at near 10 hours (Alves; Mainardes; Khalil, 2016ALVES, A. C. S.; MAINARDES, R. M.; KHALIL, N. M. Nanoencapsulation of gallic acid and evaluation of its cytotoxicity and antioxidant activity. Materials Science and Engineering: C, 60:126-134, 2016.). Particularly, for the polyphenol extracted from winery byproduct, it is a very complex substance, having many different compounds with specific physical-chemistry properties and, therefore, different release profile.

CONCLUSIONS

The major characteristic of the nanoparticle, that is its size, showed a close similarity for the lowest and highest reactional volumes. Further, for skin grape extract, the nanoparticle size decreased constantly in function of increasing reactional volume. This attests that the ionic gelation system with chitosan and TPP is suitable for scaling-up the nanoencapsulation of agroindustrial byproduct extract from acerola and grape. However, due to its high dependency on the kind of substance to be encapsulated, the suitability for other kinds of active substance must be studied. All other properties of the system, e.g., yielding, encapsulation efficiency, morphology and release profile showed minimal variation according to the reactional volume, attesting its suitability for scaling-up.

REFERENCES

  • ALVES, A. C. S.; MAINARDES, R. M.; KHALIL, N. M. Nanoencapsulation of gallic acid and evaluation of its cytotoxicity and antioxidant activity. Materials Science and Engineering: C, 60:126-134, 2016.
  • BADANG, M. C. N.; CHAKRABORTY, S. Carbohydrate polymers as controlled release devices for pesticides. Journal of Carbohydrate Chemistry, 38(1):67-85, 2019.
  • BRITTO, D. et al. N,N,N-trimethyl chitosan nanoparticles as a vitamin carrier system. Food Hydrocolloids, 27(2):487-493, 2012.
  • BRITTO, D. et al. Entrapment characteristics of hydrosoluble vitamins loaded into chitosan and N,N,N-trimethy chitosan nanoparticles. Macromolecular Research, 22(12):1261-1267, 2014.
  • BRITTO, D. et al. Analysis of thermal and aqueous suspension stabilities of chitosan based nanoencapsulated vitamins. Química Nova, 39(9):1126-1130, 2016.
  • FATHI, M.; MARTÍN, Á.; MCCLEMENTS, D. J. Nanoencapsulation of food ingredients using carbohydrate based delivery systems. Trends in Food Science & Technology, 39(1):18-39, 2014.
  • FEYZIOGLU, G. C.; TORNUK, F. Development of chitosan nanoparticles loaded with summer savory (Satureja hortensis L.) essential oil for antimicrobial and antioxidant delivery applications. LWT - Food Science and Technology, 70:104-110, 2016.
  • FRIEDMAN, M. Antibacterial, antiviral, and antifungal properties of wines and winery byproducts in relation to their flavonoid content, Journal of Agricultural and Food Chemistry, 62(26):6025-6042, 2014.
  • GAN, Q.; WANG, T. Chitosan nanoparticle as protein delivery carrier: Systematic examination of fabrication conditions for efficient loading and release. Colloids and Surfaces B: Biointerfaces, 59(1):24-34, 2007.
  • GAN, Q. et al. Modulation of surface charge, particle size and morphological properties of chitosan-TPP nanoparticles intended for gene delivery. Colloids and Surfaces B, 44(2-3): 65-73, 2005.
  • GEÇER, A. et al. Trimethyl chitosan nanoparticles enhances dissolution of the poorly water-soluble drug Candesartan Cilexetil. Macromolecular Research , 18(10):986-991, 2010.
  • GOKCE, Y. et al. Ultrasonication of chitosan nanoparticle suspension: Influence on particle size. Colloids and Surfaces A: Physicochemical and Engineering Aspects, 462:75-81, 2014.
  • KIILLL, C. P. et al. Synthesis and factorial design applied to a novel chitosan/sodium polyphosphate nanoparticles via ionotropic gelation as an RGD delivery system. Carbohydrate Polymers, 157:1695-1702, 2017.
  • KUMARI, A.; YADAV, S. K.; YADAV, S. C. Biodegradable polymeric nanoparticles based drug delivery systems. Colloids and Surfaces B: Biointerfaces , 75(1):1-18, 2010.
  • LIANG, J. et al. Encapsulation of epigallocatechin gallate in zein/chitosan nanoparticles for controlled applications in food systems. Food Chemistry, 231:19-24, 2017.
  • MELO, P. S. et al. Winery by-products: Extraction optimization, phenolic composition and cytotoxic evaluation to act as a new source of scavenging of reactive oxygen species. Food Chemistry , 181:160-169, 2015.
  • MORA-HUERTAS, C. E.; FESSI, H.; ELAISSARI, A. Polymer-based nanocapsules for drug delivery. International Journal of Pharmaceutics, 385(1-2):113-142, 2010.
  • MOURA, R. S. et al. Addition of açaí (Euterpe oleracea) to cigarettes has a protective effect against emphysema in mice. Food and Chemical Toxicology, 49(4):855-863, 2011.
  • NASCIMENTO, J. A. A. et al. Stability of nanocomposite edible films based on polysaccharides and vitamin C from agroindustrial residue. Materials Research, 22(3): e20190057, 2019.
  • NEVES, A. L. P. et al. Factorial design as tool in chitosan nanoparticles development by ionic gelation technique. Colloids and Surfaces A: Physicochemical and Engineering Aspects , 445:34-39, 2014.
  • SANTOS, P. P. et al. Biodegradable polymers as wall materials to the synthesis of bioactive compound nanocapsules. Trends in Food Science & Technology , 53:23-33, 2016.
  • SHAHGALDIAN, P. et al. A study of the freeze-drying conditions of calixarene based solid lipid nanoparticles. European Journal of Pharmaceutics and Biopharmaceutics, 55(2):181-184, 2003.
  • SREEKUMAR, S. et al. Parameters influencing the size of chitosan-TPP nano- and microparticles. Scientific Reports, 8(4695):1-11, 2018.
  • VALDUGA, E. et al. Extração, secagem por atomização e microencapsulamento de antocianinas do bagaço da uva “Isabel” (Vitis labrusca). Ciência e Agrotecnologia, 32(5):1568-1574, 2008.
  • WANG, L. et al. Needleless electrospinning for scaled-up production of ultrafine chitosan hybrid nanofibers used for air filtration. RSC Advances, 6(107):105988-105995, 2016.
  • WONGPREECHA, J. et al. One-pot, large-scale green synthesis of silver nanoparticles-chitosan with enhanced antibacterial activity and low cytotoxicity. Carbohydrate Polymers , 199: 641-648, 2018.

Publication Dates

  • Publication in this collection
    06 Mar 2020
  • Date of issue
    2019

History

  • Received
    07 Oct 2019
  • Accepted
    19 Dec 2019
Editora da Universidade Federal de Lavras Editora da UFLA, Caixa Postal 3037 - 37200-900 - Lavras - MG - Brasil, Telefone: 35 3829-1115 - Lavras - MG - Brazil
E-mail: revista.ca.editora@ufla.br