Acessibilidade / Reportar erro

Simplified process of extraction of polyphenols from agroindustrial grape waste

Abstract

The extraction, and stability of polyphenols from the grape residue were studied. The extractions were performed following the Box-Behnken design, and the surface response methodology was used to model the extraction of total anthocyanins (TA), flavonols (TF), and phenolics (TP). The extraction was optimized simultaneously by the desirability function. The degradation kinetics of monomeric anthocyanins, and the increase in polymeric color were modeled under refrigeration conditions. The extraction with a temperature of 60 °C, solid:liquid ratio of 1/25 g/mL, and time of 80 min, maximized the recovery of TA (30.96 mg/100 g), TF (73.34 mg/100 g), and TP (856.78 mg EAG/100 g). The degradation of monomeric anthocyanins, and the increase in polymeric color followed a kinetic first-order reaction, with reaction rates (k) of 4.10 x 10-3 days-1, and 3.46 x 10-3 days-1, respectively. The half-life (t1/2) of anthocyanins was 169 days. The ethanol-citric acid solution allowed polyphenols to be efficiently extracted from the grape residue, and had a positive effect on the stability of anthocyanins.

Keywords:
Box-Behnken design; phytochemicals; grape residue; solid-liquid extraction

1 Introduction

One of the industries historically, and economically important in many countries of the world is the processing of grapes, as they are obtained from various products such as wine, raisins, juices, jellies, among others. The grape fruit is recognized for its nutritional properties and beneficial to health due to its bioactive compounds (Dhekney, 2016Dhekney, S. A. (2016). Grape. In B. Caballero, P. M. Finglas, & F. Toldrá (Eds.), Encyclopedia of food and health (pp. 261-265). London: Elsevier. http://dx.doi.org/10.1016/B978-0-12-384947-2.00360-3.
http://dx.doi.org/10.1016/B978-0-12-3849...
).

The wine, and grape juice industries generate abundant waste, which causes additional costs for its elimination (Devesa-Rey et al., 2011Devesa-Rey, R., Vecino, X., Varela-Alende, J. L., Barral, M. T., Cruz, J. M., & Moldes, A. B. (2011). Valorization of winery waste vs. the costs of not recycling. Waste Management (New York, N.Y.), 31(11), 2327-2335. http://dx.doi.org/10.1016/j.wasman.2011.06.001. PMid:21752623.
http://dx.doi.org/10.1016/j.wasman.2011....
; Mammadova et al., 2020Mammadova, S. M., Fataliyev, H. K., Gadimova, N. S., Aliyeva, G. R., Tagiyev, A. T., & Baloglanova, K. V. (2020). Production of functional products using grape processing residuals. Food Science and Technology, 2061, 1-7. http://dx.doi.org/10.1590/fst.30419.
http://dx.doi.org/10.1590/fst.30419...
). These residues are approximately 30% by weight of the fruit used in processing, consisting of seeds, husks, and stalks (Teixeira et al., 2014Teixeira, A., Baenas, N., Dominguez-Perles, R., Barros, A., Rosa, E., Moreno, D. A., & Garcia-Viguera, C. (2014). Natural bioactive compounds from winery by-products as health promoters: a review. International Journal of Molecular Sciences, 15(9), 15638-15678. http://dx.doi.org/10.3390/ijms150915638. PMid:25192288.
http://dx.doi.org/10.3390/ijms150915638...
). Many investigations point out that residues derived from fruit processing present phytochemicals (Morais et al., 2015Morais, D. R., Rotta, E. M., Sargi, S. C., Schmidt, E. M., Bonafe, E. G., Eberlin, M. N., Sawaya, A. C. H. F., & Visentainer, J. V. (2015). Antioxidant activity, phenolics and UPLC-ESI(-)-MS of extracts from different tropical fruits parts and processed peels. Food Research International, 77, 392-399. http://dx.doi.org/10.1016/j.foodres.2015.08.036.
http://dx.doi.org/10.1016/j.foodres.2015...
), with antioxidant properties (O’Shea et al., 2012O’Shea, N., Arendt, E. K., & Gallagher, E. (2012). Dietary fibre and phytochemical characteristics of fruit and vegetable by-products and their recent applications as novel ingredients in food products. Innovative Food Science & Emerging Technologies, 16, 1-10. http://dx.doi.org/10.1016/j.ifset.2012.06.002.
http://dx.doi.org/10.1016/j.ifset.2012.0...
; Ribeiro et al., 2018Ribeiro, T. P., Lima, M. A. C., Alves, R. E., Gonçalves, A. L. S., & Souza, A. P. C. (2018). Chemical characterization of winemaking byproducts from grape varieties cultivated in Vale do São Francisco, Brazil. Food Science and Technology, 38(4), 577-583. http://dx.doi.org/10.1590/fst.01116.
http://dx.doi.org/10.1590/fst.01116...
), of which phenolic compounds have been reported frequently (Bataglion et al., 2015Bataglion, G. A., Da Silva, F. M. A., Eberlin, M. N., & Koolen, H. H. F. (2015). Determination of the phenolic composition from Brazilian tropical fruits by UHPLC-MS/MS. Food Chemistry, 180, 280-287. http://dx.doi.org/10.1016/j.foodchem.2015.02.059. PMid:25766829.
http://dx.doi.org/10.1016/j.foodchem.201...
).

The grape residue is a potential source of phenolic compounds (Goula et al., 2016Goula, A. M., Thymiatis, K., & Kaderides, K. (2016). Valorization of grape pomace: drying behavior and ultrasound extraction of phenolics. Food and Bioproducts Processing, 100, 132-144. http://dx.doi.org/10.1016/j.fbp.2016.06.016.
http://dx.doi.org/10.1016/j.fbp.2016.06....
), of this group, stands out the anthocyanins for their ability to confer color, and functional properties, which can be used in the elaboration of functional foods (Aguilera et al., 2016Aguilera, Y., Mojica, L., Rebollo-Hernanz, M., Berhow, M., De Mejía, E. G., & Martín-Cabrejas, M. A. (2016). Black bean coats: new source of anthocyanins stabilized by B-cyclodextrin copigmentation in a sport beverage. Food Chemistry, 212, 561-570. http://dx.doi.org/10.1016/j.foodchem.2016.06.022. PMid:27374568.
http://dx.doi.org/10.1016/j.foodchem.201...
). Recently, powdered grape residues were used in the ice cream formulation (Vital et al., 2018Vital, A. C. P., Santos, N. W., Matumoto-Pintro, P. T., da Silva Scapim, M. R., & Madrona, G. S. (2018). Ice cream supplemented with grape juice residue as a source of antioxidants. International Journal of Dairy Technology, 71(1), 183-189. http://dx.doi.org/10.1111/1471-0307.12412.
http://dx.doi.org/10.1111/1471-0307.1241...
), and yogurt (Mammadova et al., 2020Mammadova, S. M., Fataliyev, H. K., Gadimova, N. S., Aliyeva, G. R., Tagiyev, A. T., & Baloglanova, K. V. (2020). Production of functional products using grape processing residuals. Food Science and Technology, 2061, 1-7. http://dx.doi.org/10.1590/fst.30419.
http://dx.doi.org/10.1590/fst.30419...
), improving their functional properties. Grape seed extracts were used to enrich milk for yogurt processing, observing an alteration of the fermentation time, and the yogurt quality attributes (Alwazeer et al., 2020Alwazeer, D., Bulut, M., & Tunçtürk, Y. (2020). Fortification of milk with plant extracts modifies the acidification and reducing capacities of yoghurt bacteria. International Journal of Dairy Technology, 73(1), 117-125. http://dx.doi.org/10.1111/1471-0307.12643.
http://dx.doi.org/10.1111/1471-0307.1264...
). Although the results are promising, the use of HCl in the obtaining bioactive extracts represents a potential risk to human health.

The conventional solid-liquid extraction method is alternatively used to obtain phytochemicals. Extraction occurs as result of diffusion of the compound of interest to the solvent, this phenomenon is produced by the affinity and selectivity of the solvent used (Takeuchi et al., 2009Takeuchi, T. M., Pereira, C. G., Braga, M. E. M., Maróstica, M. R., Leal, P. F., & Meireles, M. A. A. (2009). Low-pressure solvent extraction (solid–liquid extraction, microwave assisted, and ultrasound assisted) from condimentary plants. In M. A. A. Meireles (Ed.), Extracting bioactive compounds for food products: theory and applications (pp. 140-142). London: CRC Press, Taylor & Francis Group.). Solvents such as ethanol, methanol, and acetone are often used in the extraction of phytochemicals, these solvents are acidified with hydrochloric acid in order to improve the stability of anthocyanins (Lees & Francis, 1972Lees, D. H., & Francis, F. J. (1972). Standardization of Pigment Analyses in Cranberries. HortScience, 7, 83-84.; Rodriguez-Saona & Wrolstad, 2001Rodriguez-Saona, L. E., & Wrolstad, R. E. (2001). Extraction, isolation, and purification of anthocyanins. Current Protocols in Food Analytical Chemistry, 1, F1.1.1-F1.1.11. https://doi.org/10.1002/0471142913.faf0101s00.
https://doi.org/10.1002/0471142913.faf01...
). The concern about the use of toxic solvents in the extraction of phytochemicals was approached by Pedro et al., (2016)Pedro, A. C., Granato, D., & Rosso, N. D. (2016). Extraction of anthocyanins and polyphenols from black rice (Oryza sativa L.) by modeling and assessing their reversibility and stability. Food Chemistry, 191, 12-20. http://dx.doi.org/10.1016/j.foodchem.2015.02.045. PMid:26258696.
http://dx.doi.org/10.1016/j.foodchem.201...
, observed good results in the extraction of phytochemicals with ethanol acidified with citric acid. Some factors that significantly influenced the extraction were temperature, solvent concentration, solid-liquid ratio, and time (Li et al., 2012Li, Y., Han, L., Ma, R., Xu, X., Zhao, C., Wang, Z., Chen, F., & Hu, X. (2012). Effect of energy density and citric acid concentration on anthocyanins yield and solution temperature of grape peel in microwave-Assisted extraction process. Journal of Food Engineering, 109(2), 274-280. http://dx.doi.org/10.1016/j.jfoodeng.2011.09.021.
http://dx.doi.org/10.1016/j.jfoodeng.201...
; Pedro et al., 2016Pedro, A. C., Granato, D., & Rosso, N. D. (2016). Extraction of anthocyanins and polyphenols from black rice (Oryza sativa L.) by modeling and assessing their reversibility and stability. Food Chemistry, 191, 12-20. http://dx.doi.org/10.1016/j.foodchem.2015.02.045. PMid:26258696.
http://dx.doi.org/10.1016/j.foodchem.201...
). It was observed that in the extraction of anthocyanins other compounds are also extracted (Pedro et al., 2016Pedro, A. C., Granato, D., & Rosso, N. D. (2016). Extraction of anthocyanins and polyphenols from black rice (Oryza sativa L.) by modeling and assessing their reversibility and stability. Food Chemistry, 191, 12-20. http://dx.doi.org/10.1016/j.foodchem.2015.02.045. PMid:26258696.
http://dx.doi.org/10.1016/j.foodchem.201...
).

Due to the current interest in the consumption of foods with functional properties, it is necessary to develop efficient processes, using non-toxic chemicals reagents for the safe extraction of phytochemicals. For this reason, the present research aimed to optimize the extraction of polyphenols from grape residue from the juice industry, aiming their use in food processing, and to evaluate their stability during storage.

2 Materials and methods

2.1 Sample

A total of 25 kg of grape residue from cv. Isabel (Vitis labrusca) was supplied by a pulp processing industry in the municipality of Goiana, Pernambuco, Brazil. The residue was homogenized and divided into sample units of 800 g each and stored at -18 ± 1 °C.

2.2 Chemical reagents

The reagents used in the extraction were: absolute ethanol 99.9% (Merck KGaA, Emsure, Germany) and citric acid 99.5% (Química Moderna, Brazil). The reagents used in the quantitative analyzes were: hydrochloric acid, potassium chloride, sodium acetate, chloroform and acetone (Fmaia, Brazil); Folin-Ciocalteu's reagent (Merck, Germany); Gallic acid 98% (Vetec Química fina Ltda, Brazil); Sodium carbonate (Sigma-Aldrich, Brazil) and potassium metabisulphite 96% (K2S2O5) (Dinâmica, Química Contemporânea Ltda, Brazil).

2.3 Preparation of the residue for extraction

The residue was oven dried (MARCONI, MA035, Brazil) with air circulation at 40 °C for 18 h. The dried residue was ground in a knife mill with cooling (TECNAL, TE-631/2, Brazil) for 1 min at 7,000 rpm. The flour was sieved sequentially with two stainless steel sieves (Bertel, Caieiras, Brazil) of 42 and 80 mesh. The residue with a particle size between 355-180 μm was selected for the experiments. This residue was vacuum packed at 98.66 kPa in a vacuum sealer (SELOVAC, 200B, Brazil) and stored protected from light at -18±1°C until the experiments were run.

2.4 Characterization of the wet and dry residue

Moisture content

The moisture of the residue was determined by infrared radiation before and after drying, an infrared apparatus (Marte, ID50, Brazil) was used with a constant temperature of 105 °C.

Quantification of total anthocyanins, total flavonols and total phenolics

Total anthocyanins (TA) and total flavonols (TF) were quantified following the methodology of Lees & Francis (1972)Lees, D. H., & Francis, F. J. (1972). Standardization of Pigment Analyses in Cranberries. HortScience, 7, 83-84., Using 3 g of the wet residue and 2 g of the dried residue for extraction. Absorbance readings were performed on a UV-Visible spectrophotometer (SHIMADZU, UV-1650PC, Japan) at 535 and 374 nm, for TA and TF, respectively. Calculations were performed with Equations 1 and 2, TA were expressed in mg of cyanidin-3-glycoside per 100 g of sample and TF in mg equivalent in quercetin per 100 g of sample.

T A = A b s 535 n m x D F 98.2 (1)
T F = A b s 374 n m x D F 76.5 (2)

Where: DF – Dilution factor; Abs – Absorbance; TA - Total anthocyanins (mg/100g); TF - Total flavonols (mg/100g)

Total phenolics (TP) were quantified from the extracts obtained to quantify TA and TF, following the methodology of Wettasinghe & Shahidi (1999)Wettasinghe, M., & Shahidi, F. (1999). Evening primrose meal: a source of natural antioxidants and scavenger of hydrogen peroxide and oxygen-derived free radicals. Journal of Agricultural and Food Chemistry, 47(5), 1801-1812. http://dx.doi.org/10.1021/jf9810416. PMid:10552455.
http://dx.doi.org/10.1021/jf9810416...
. The absorbance was read at 725 nm on the UV-Visible Spectrophotometer. TP was calculated with a standard curve constructed with gallic acid and the results were expressed in mg equivalent in gallic acid (EAG) per 100 g of sample.

2.5 Experimental design for extractions

The effect of temperature (X1), solid:liquid ratio (X2) and time (X3) on TA, TF and TP extraction were investigated. The tests were performed according to the Box-Behnken design adjusted to the three independent variables, with total of 15 assays including three replicates at the center point (Table 1). The levels of the variables were adjusted according to the experiences of Pedro et al. (2016)Pedro, A. C., Granato, D., & Rosso, N. D. (2016). Extraction of anthocyanins and polyphenols from black rice (Oryza sativa L.) by modeling and assessing their reversibility and stability. Food Chemistry, 191, 12-20. http://dx.doi.org/10.1016/j.foodchem.2015.02.045. PMid:26258696.
http://dx.doi.org/10.1016/j.foodchem.201...
, as well as the extraction assays. The extractions were performed in a rotary evaporator (Heidolph, Laborota 4000), with agitation of 90±2 rpm at atmospheric pressure according to the experimental design. 2 g of the dry residue was mixed with the extraction solvent (ethanol acidified with 1.5 mol/L citric acid solution in a ratio of 80/20 vol/vol). After completion of the extraction, the sample-solvent mixture was filtered and the volume of the filtrate was checked to 100 mL with the extraction solvent. The extracts were stored at -18 ± 1 °C in amber glass vials. The TA, TF and TP were quantified by reading the absorbance of the extracts after 24 h, as described by Lees & Francis (1972)Lees, D. H., & Francis, F. J. (1972). Standardization of Pigment Analyses in Cranberries. HortScience, 7, 83-84..

Table 1
Box-Behnken design with experimental and fitted data for the extraction of polyphenols from grape residue cv. Isabel.

2.6 Elaboration and characterization of the optimized extract concentrate

The optimized concentrate extract (OCE) was prepared following the flowchart as shown in Figure 1. For each extraction was used 20 g of the dried residue. The parameters of each operation were established in the extraction assays and after the optimization of the extraction. The concentration was added to the process for the purpose of removing the ethanol.

Figure 1
Flowchart of the process of elaboration of optimization extract concentrate of grape residue.

The OCE was analyzed to quantify the monomeric anthocyanins in mg/L of Malvidine-3,5-diglucoside (Toaldo et al., 2013Toaldo, I. M., Fogolari, O., Pimentel, G. C., de Gois, J. S., Borges, D. L. G., Caliari, V., & Bordignon-Luiz, M. (2013). Effect of grape seeds on the polyphenol bioactive content and elemental composition by ICP-MS of grape juices from Vitis labrusca L. Lebensmittel-Wissenschaft + Technologie, 53(1), 1-8. http://dx.doi.org/10.1016/j.lwt.2013.02.028.
http://dx.doi.org/10.1016/j.lwt.2013.02....
) by the pH-differential method (Lee et al., 2005Lee, J., Durst, R. W., & Wrolstad, R. E. (2005). Determination of total monomeric anthocyanin pigment content of fruit juices, beverages, natural colorants, and wines by the pH differential method: collaborative study. Journal of AOAC International, 88(5), 1269-1278. http://dx.doi.org/10.1093/jaoac/88.5.1269. PMid:16385975.
http://dx.doi.org/10.1093/jaoac/88.5.126...
). TP and TF were quantified by reading the absorbance according to the methodology described by Lees & Francis (1972)Lees, D. H., & Francis, F. J. (1972). Standardization of Pigment Analyses in Cranberries. HortScience, 7, 83-84.. The pH was measured with the aid of a pH meter, TECNAL, Tec-3MP, Brazil (Association of Official Analytical Chemistry, 2002Association of Official Analytical Chemistry – AOAC. (2002). Official methods of analysis of the Association of Official Analytical Chemists (16th ed.). Gaithersburg: AOAC.). Solids soluble in °Brix were determined by reading on an automatic refractometer at 25 °C (REICHERT, r2i300, USA). The water activity was determined by the direct method at 25 °C in an Aqualab (4T analyzer, DECAGON DEVICES, Brazil). The color was characterized using the CIELAB parameters (L* a* b*) as described by McGuire (1992)McGuire, R. G. (1992). Reporting of objective color measurements. HortScience, 27(12), 1254-1255. http://dx.doi.org/10.21273/HORTSCI.27.12.1254.
http://dx.doi.org/10.21273/HORTSCI.27.12...
.

Percentage of polymeric color of optimized extract concentrate

The percentage of the polymeric color was determined according to the Giusti, & Wrolstad (2001)Giusti, M. M., & Wrolstad, R. E. (2001). Characterization and measerument of anthocyanins by UV-visible spectroscopy. Current Protocols in Food Analytical Chemistry, 1, F1.2.1-F1.2.13. https://doi.org/10.1002/0471142913.faf0102s00.
https://doi.org/10.1002/0471142913.faf01...
, with Equation 3. The diluted OCE (1 part OCE and 4 parts distilled water) was used for the determinations.

P o l y m e r i c c o l o r % = P o l y m e r i c c o l o r C o l o r d e n s i t y x 100 (3)

The polymeric color and color density were calculated with Equation 4, using absorbance readings of the OCE diluted with bleaching treatment and without bleaching, respectively.

A 420 n m A 700 n m + A 520 n m A 700 n m x D F (4)

Where: A – Absorbance; DF – Dilution factor

2.7. Stability study of anthocyanins

The OCE was packed in 20 mL amber vials and stored at 4 °C in a refrigerated incubator (BOD, TECNAL, TE-371, Brazil). The degradation of the monomeric anthocyanins was monitored by the pH-differential method (Lee et al., 2005Lee, J., Durst, R. W., & Wrolstad, R. E. (2005). Determination of total monomeric anthocyanin pigment content of fruit juices, beverages, natural colorants, and wines by the pH differential method: collaborative study. Journal of AOAC International, 88(5), 1269-1278. http://dx.doi.org/10.1093/jaoac/88.5.1269. PMid:16385975.
http://dx.doi.org/10.1093/jaoac/88.5.126...
) and the polymeric color by the method described in item 2.6.1. The determinations were performed initially every 3 days and after 21 days every 5 days for a period of 83 days. The degradation kinetics of the monomeric anthocyanins and the increase of the polymeric color were determined using the first order reaction model as proposed by Sharma et al., (2016)Sharma, R. J., Gupta, R. C., Singh, S., Bansal, A. K., & Singh, I. P. (2016). Stability of anthocyanins- and anthocyanidins-enriched extracts, and formulations of fruit pulp of Eugenia jambolana (‘jamun’). Food Chemistry, 190, 808-817. http://dx.doi.org/10.1016/j.foodchem.2015.06.029. PMid:26213042.
http://dx.doi.org/10.1016/j.foodchem.201...
, using Equation 5.

C = C 0 e ± k * t (5)

Where: t – Time; k – The first-order kinetic rate constant; C0 – Concentrations of monomeric anthocyanins (mg.L-1) and polymeric color (%) at time zero; C – Concentrations of monomeric anthocyanins (mg.L-1) and polymeric color (%) at time t.

The half-life time of the monomeric anthocyanins was calculated with Equation 6.

t 1 2 = ln 2 k (6)

Where: k – The first-order kinetic rate constant; t12 – Half-life time

2.8. Statistical analysis

The data of the assays for optimization of the extraction were submitted to analysis of variances (ANOVA) and the differences among the means by Tukey test (p < 0.05). The normality and homoscedasticity of the ANOVA residues were obtained by the Anderson-Darling and Breush-Pagan tests, respectively. The analysis of multiple linear regression by response surface methodology (RSM) was performed using the Equation 7 model:

Y i = β 0 + i = 1 k β i X i + i = 1 k β i i X i 2 + i > j k β i j X i X j (7)

Where the response function (Yi) is composed of linear, quadratic and interactive components. The constant β0 denotes the intercept of the model; Βi, βii and βij represent the coefficients of the linear, quadratic and iterative components of the model, respectively. Xi and Xj are the independent variables, and k represents the number of factors that were investigated. The quality of the fit of the experimental data to the model was evaluated by the test of lack of fit, coefficient of ordinary and adjusted regression. The assumption of normality of residuals was verified by Anderson-Darling test. Simultaneous optimization of TA, TF and TP extraction was performed using the desirability function (Derringer & Suich, 1980Derringer, G., & Suich, R. (1980). Simultaneous optimisation of several response variables. Journal of Quality Technology, 12(4), 214-219. http://dx.doi.org/10.1080/00224065.1980.11980968.
http://dx.doi.org/10.1080/00224065.1980....
).

The degradation kinetics of monomeric anthocyanins and the increase in polymeric color of OCE were adjusted by non-linear regression. The normality of the residues and the ordinary determination coefficient were used to evaluate the quality of the adjustment. Statistical analyzes were performed with the aid of MATLAB® R2010a 7.10.0.499 software (MathWorks, USA).

3 Results and discussion

3.1 Characterization of the wet and dry residue

The moisture content of the wet residue was 51.39±1.53 g/100g wet basis (w.b.), a near content was observed in the grape residue of vinification by Minjares-Fuentes et al. (2014)Minjares-Fuentes, R., Femenia, A., Garau, M. C., Meza-Velázquez, J. A., Simal, S., & Rosselló, C. (2014). Ultrasound-assisted extraction of pectins from grape pomace using citric acid: A response surface methodology approach. Carbohydrate Polymers, 106(1), 179-189. http://dx.doi.org/10.1016/j.carbpol.2014.02.013. PMid:24721067.
http://dx.doi.org/10.1016/j.carbpol.2014...
. The dried residue had a moisture content of 5.33±0.44 g/100g w.b., a similar value was reported by Ribeiro et al. (2015)Ribeiro, L. F., Ribani, R. H., Francisco, T. M. G., Soares, A. A., Pontarolo, R., & Haminiuk, C. W. I. (2015). Profile of bioactive compounds from grape pomace (Vitis vinifera and Vitis labrusca) by spectrophotometric, chromatographic and spectral analyses. Journal of Chromatography. B, Analytical Technologies in the Biomedical and Life Sciences, 1007, 72-80. http://dx.doi.org/10.1016/j.jchromb.2015.11.005. PMid:26590878.
http://dx.doi.org/10.1016/j.jchromb.2015...
. The TA content of the residue before and after drying was 11.51±0.95 mg/100g and 28.32±0.34 mg/100g, respectively. The Higher value was observed by Liazid et al. (2011)Liazid, A., Guerrero, R. F., Cantos, E., Palma, M., & Barroso, C. G. (2011). Microwave assisted extraction of anthocyanins from grape skins. Food Chemistry, 124(3), 1238-1243. http://dx.doi.org/10.1016/j.foodchem.2010.07.053.
http://dx.doi.org/10.1016/j.foodchem.201...
in the grape skin of cv. Tintilla Rota (V. vinifera), reporting a content of 154.59 mg/100g. The TA content observed in this work is probably due to the low anthocyanin content of the cv. Isabel (Yamamoto et al., 2015Yamamoto, L. Y., de Assis, A. M., Roberto, S. R., Bovolenta, Y. R., Nixdorf, S. L., García-Romero, E., Gómez-Alonso, S., & Hermosín-Gutiérrez, I. (2015). Application of abscisic acid (S-ABA) to cv. Isabel grapes (V. vinifera × V. labrusca) for color improvement: effects on color, phenolic composition and antioxidant capacity of their grape juice. Food Research International, 77, 572-583. http://dx.doi.org/10.1016/j.foodres.2015.10.019.
http://dx.doi.org/10.1016/j.foodres.2015...
), and/or the culture conditions (Zhou et al., 2020Zhou, S. H., Guo, R. R., Wei, R. F., Liu, J. B., Yu, H., Shi, X. F., Zhang, Y., Xie, T. L., & Cheng, G. (2020). Effects of bagging or the combination of umbrella and bag treatments on anthocyanin accumulation in the berry skin of ‘Kyoho’ (Vitis labruscana) grape. Food Science and Technology, 40(2), 394-400. http://dx.doi.org/10.1590/fst.41218.
http://dx.doi.org/10.1590/fst.41218...
; Sun et al., 2019Sun, X., Liu, L., Ma, T., Yu, J., Huang, W., Fang, Y., & Zhan, J. (2019). Effect of high Cu2+ stress on fermentation performance and copper biosorption of Saccharomyces cerevisiae during wine fermentation. Food Science and Technology, 39(1), 19-26. http://dx.doi.org/10.1590/1678-457x.24217.
http://dx.doi.org/10.1590/1678-457x.2421...
).

The TP content of the wet and dry residue was 273.20 ± 18.27 mg EAG/100 g and 804.26 ± 44.53 mg EAG/100 g, respectively. The phenolic content of the grape residue is influenced by the grape cultivar and the processing conditions (Abe et al., 2007Abe, L. T., Mota, R. V., Lajolo, F. M., & Genovese, M. I. (2007). Compostos fenólicos e capacidade antioxidante de cultivares de uvas Vitis labrusca L. e Vitis vinifera L. Food Science and Technology, 27(2), 394-400. http://dx.doi.org/10.1590/S0101-20612007000200032.
http://dx.doi.org/10.1590/S0101-20612007...
). This fact was observed in the skin of grapes of several cultivars derived from ten different vinification processes (Harsha et al., 2013Harsha, P. S. C., Gardana, C., Simonetti, P., Spigno, G., & Lavelli, V. (2013). Characterization of phenolics, in vitro reducing capacity and anti-glycation activity of red grape skins recovered from winemaking by-products. Bioresource Technology, 140, 263-268. http://dx.doi.org/10.1016/j.biortech.2013.04.092. PMid:23707914.
http://dx.doi.org/10.1016/j.biortech.201...
).

The TF content in the wet and dry residue was 21.65 ± 0.64 mg/100 g and 58.95 ± 0.48 mg/100 g, respectively. Harsha et al. (2013)Harsha, P. S. C., Gardana, C., Simonetti, P., Spigno, G., & Lavelli, V. (2013). Characterization of phenolics, in vitro reducing capacity and anti-glycation activity of red grape skins recovered from winemaking by-products. Bioresource Technology, 140, 263-268. http://dx.doi.org/10.1016/j.biortech.2013.04.092. PMid:23707914.
http://dx.doi.org/10.1016/j.biortech.201...
reported similar contents in grape skins derived from various vinification processes. In a similar study, it was reported close levels in Cabernet Franc and Sauvignon grapes skins derived from vinification (Barcia et al., 2014Barcia, M. T., Pertuzatti, P. B., Rodrigues, D., Gómez-Alonso, S., Hermosín-Gutiérrez, I., & Godoy, H. T. (2014). Occurrence of low molecular weight phenolics in Vitis vinifera red grape cultivars and their winemaking by-products from São Paulo (Brazil). Food Research International, 62, 500-513. http://dx.doi.org/10.1016/j.foodres.2014.03.051.
http://dx.doi.org/10.1016/j.foodres.2014...
).

3.2 Surface responses

The experimental data from the extraction assays are presented in Table 1. The normality and homoscedasticity of ANOVA residues and multiple linear regression analysis were verified by the Anderson-Darling and Breush-Pagan test (p > 0.05), respectively. Norman (2010)Norman, G. (2010). Likert scales, levels of measurement and the “laws” of statistics. Advances in Health Sciences Education : Theory and Practice, 15(5), 625-632. http://dx.doi.org/10.1007/s10459-010-9222-y. PMid:20146096.
http://dx.doi.org/10.1007/s10459-010-922...
, say that parametric statistics are robust when data are not normal. However, obtaining efficient estimates is linked to adherence to the ANOVA assumptions for the residues.

Significant statistical differences (p < 0.05) were found between the essays and the means of the essays, by ANOVA and the Tukey test, respectively. The regression results are shown in Table 2, the quadratic model was significant (p < 0.05) for TA, TP and TF. Likewise, there was no lack of fit significant (p > 0.05) for the three data groups. The adjusted models for TA, TP and TF explain 93%, 94% and 94%, respectively, the variation of the response variables.

Table 2
Estimated regression coefficients for the model quadratic polynomial and the ANOVA of the fitted model.

Response surface for extraction of total anthocyanins

In Table 2, it is observed that the linear components of temperature (X1), solid:liquid ratio (X2), time (X3), the interaction between the two (X1X2) and the intercept were significant and positive. This indicates that these components increase the value of the TA content. The temperature influenced significantly the extraction of anthocyanins from residues of Tulipa gesneriana L. petals (Arici et al., 2016Arici, M., Karasu, S., Baslar, M., Toker, O. S., Sagdic, O., & Karaagacli, M. (2016). Tulip petal as a novel natural food colorant source: extraction optimization and stability studies. Industrial Crops and Products, 91, 215-222. http://dx.doi.org/10.1016/j.indcrop.2016.07.003.
http://dx.doi.org/10.1016/j.indcrop.2016...
) and in Nitraria tangutorun seeds (Sang et al., 2017Sang, J., Sang, J., Ma, Q., Hou, X., & Li, C. Q. (2017). Extraction optimization and identification of anthocyanins from Nitraria tangutorun Bobr. seed meal and establishment of a green analytical method of anthocyanins. Food Chemistry, 218, 386-395. http://dx.doi.org/10.1016/j.foodchem.2016.09.093. PMid:27719925.
http://dx.doi.org/10.1016/j.foodchem.201...
).

Response surface for extraction of total phenolics

The coefficients of the linear components of the temperature (X1), time (X3), of the interaction between the temperature with time (X1X3) and the intercept were significant and positive (Table 2). The temperature had a significant effect on the microwave-assisted extraction of phenolic compounds in grape skins (Cvjetko Bubalo et al., 2016Cvjetko Bubalo, M., Ćurko, N., Tomašević, M., Kovačević Ganić, K., & Radojčić Redovniković, I. (2016). Green extraction of grape skin phenolics by using deep eutectic solvents. Food Chemistry, 200, 159-166. http://dx.doi.org/10.1016/j.foodchem.2016.01.040. PMid:26830574.
http://dx.doi.org/10.1016/j.foodchem.201...
). Extraction of phenolic compounds from grape residue was significantly influenced by temperature and solid:liquid ratio (Pinelo et al., 2005Pinelo, M., Rubilar, M., Jerez, M., Sineiro, J., & Núñez, M. J. (2005). Effect of solvent, temperature, and solvent-to-solid ratio on the total phenolic content and antiradical activity of extracts from different components of grape pomace. Journal of Agricultural and Food Chemistry, 53(6), 2111-2117. http://dx.doi.org/10.1021/jf0488110. PMid:15769143.
http://dx.doi.org/10.1021/jf0488110...
).

Response surface for extraction of total flavonols

The coefficients of the linear components of temperature (X1), time (X3) and intercept were significant (p<0.05) and positive indicating that these components tend to increase TF extraction. The time had a significant influence on the extraction of flavonols from citrus flowers (González-Centeno et al., 2014González-Centeno, M. R., Knoerzer, K., Sabarez, H., Simal, S., Rosselló, C., & Femenia, A. (2014). Effect of acoustic frequency and power density on the aqueous ultrasonic-assisted extraction of grape pomace (Vitis vinifera L.) - A response surface approach. Ultrasonics Sonochemistry, 21(6), 2176-2184. http://dx.doi.org/10.1016/j.ultsonch.2014.01.021. PMid:24548543.
http://dx.doi.org/10.1016/j.ultsonch.201...
)

3.3 Extraction optimization

Equations 8, 9 and 10, with significant components determined by RSM, were used for optimization. Extractions of TA, TF and TP were optimized simultaneously by the desirability function. The individual desirability (d1, d2 and d3) was calculated for TA, TP and TF by unilateral transformation. The global desirability (D) was obtained by the geometric mean of the individual desirability’s.

T A = 20.66 3.94 X 1 1.94 X 2 1.79 X 3 + 2.61 X 1 X 2 (8)
T F = 63.00 + 7.65 X 1 + 2.68 X 3 (9)
T P = 682.49 75.14 X 1 32.55 X 3 + 66.59 X 1 X 3 (10)

Where: TA – Total anthocyanins; TF – Total flavonols; TP – Total phenolics; X1 – Temperature (°C); X2 – Solid/liquid rate (g/mL); X3 – Time (min)

The conditions that maximize the three response variables obtained for a D = 0.77, was the temperature of 60 °C, solid:liquid ratio of 1/25 g/mL and 80 min. Under these conditions it is possible to extract 30.96 mg/100g of TA; 73.34 mg/100g TF and 856.78 mg EAG/100g TP. Similar optimized extraction temperature was observed in the extraction of anthocyanins from grape skins by Li et al. (2012)Li, Y., Han, L., Ma, R., Xu, X., Zhao, C., Wang, Z., Chen, F., & Hu, X. (2012). Effect of energy density and citric acid concentration on anthocyanins yield and solution temperature of grape peel in microwave-Assisted extraction process. Journal of Food Engineering, 109(2), 274-280. http://dx.doi.org/10.1016/j.jfoodeng.2011.09.021.
http://dx.doi.org/10.1016/j.jfoodeng.201...
. Different optimized conditions were observed in the extraction of anthocyanins from black rice (Pedro et al., 2016Pedro, A. C., Granato, D., & Rosso, N. D. (2016). Extraction of anthocyanins and polyphenols from black rice (Oryza sativa L.) by modeling and assessing their reversibility and stability. Food Chemistry, 191, 12-20. http://dx.doi.org/10.1016/j.foodchem.2015.02.045. PMid:26258696.
http://dx.doi.org/10.1016/j.foodchem.201...
).

The predicted value for TP extraction was higher than that reported by González-Centeno et al. (2014)González-Centeno, M. R., Knoerzer, K., Sabarez, H., Simal, S., Rosselló, C., & Femenia, A. (2014). Effect of acoustic frequency and power density on the aqueous ultrasonic-assisted extraction of grape pomace (Vitis vinifera L.) - A response surface approach. Ultrasonics Sonochemistry, 21(6), 2176-2184. http://dx.doi.org/10.1016/j.ultsonch.2014.01.021. PMid:24548543.
http://dx.doi.org/10.1016/j.ultsonch.201...
, on ultrasound extraction of grape residue. In another study, higher near total phenolics content of 957 mg EAG/100 g was reported in the grape residue (Goula et al., 2016Goula, A. M., Thymiatis, K., & Kaderides, K. (2016). Valorization of grape pomace: drying behavior and ultrasound extraction of phenolics. Food and Bioproducts Processing, 100, 132-144. http://dx.doi.org/10.1016/j.fbp.2016.06.016.
http://dx.doi.org/10.1016/j.fbp.2016.06....
). The value predicted for the extraction of total flavonols was higher than reported by González-Centeno et al (2014)González-Centeno, M. R., Knoerzer, K., Sabarez, H., Simal, S., Rosselló, C., & Femenia, A. (2014). Effect of acoustic frequency and power density on the aqueous ultrasonic-assisted extraction of grape pomace (Vitis vinifera L.) - A response surface approach. Ultrasonics Sonochemistry, 21(6), 2176-2184. http://dx.doi.org/10.1016/j.ultsonch.2014.01.021. PMid:24548543.
http://dx.doi.org/10.1016/j.ultsonch.201...
, on ultrasonic extraction from grape residue.

3.4 Elaboration and characterization of the optimized extract concentrate

Approximately 200 mL of OCE was obtained for each liter of alcoholic extract. Table 3 shows the results of the characterization. Most soluble solids in the extract are assumed to be the citric acid used in the acidification of the extraction solvent. Water activity indicates that there is water available for chemical and enzymatic reactions, which could accelerate the degradation of phytochemicals (Schwartz et al., 2010Schwartz, S. J., von Elbe, J. H., & Giusti, M. M. (2010). Colorants. In S. Damodaran, K. L. Parkin, & O. R. Fennema (Eds.), Food chemistry (4th ed., pp. 571-632). London: CRC Press, Taylor & Francis Group.). The pH of the extract was similar to that reported as convenient to maintain the stability of anthocyanins (Sui et al., 2014Sui, X., Dong, X., & Zhou, W. (2014). Combined effect of pH and high temperature on the stability and antioxidant capacity of two anthocyanins in aqueous solution. Food Chemistry, 163, 163-170. http://dx.doi.org/10.1016/j.foodchem.2014.04.075. PMid:24912712.
http://dx.doi.org/10.1016/j.foodchem.201...
).

Table 3
Result of the characterization of the optimized extract concentrate.

The positive and superior value of parameter a* compared to b* indicates that OCE has a predominantly red color, with luminosity (L *) of 24.32. The similar color was observed in wines without aging by Avizcuri et al. (2016)Avizcuri, J.-M., Sáenz-Navajas, M.-P., Echávarri, J.-F., Ferreira, V., & Fernández-Zurbano, P. (2016). Evaluation of the impact of initial red wine composition on changes in color and anthocyanin content during bottle storage. Food Chemistry, 213, 123-134. http://dx.doi.org/10.1016/j.foodchem.2016.06.050. PMid:27451163.
http://dx.doi.org/10.1016/j.foodchem.201...
. The percentage of OCE polymeric color indicates the presence of degraded and/or polymerized anthocyanins, probably generated in the grape processing in the industry (Kirca, & Cemeroglu, 2003Kirca, A., & Cemeroglu, B. (2003). Degradation kinetics of anthocyanins in blood orange juice and concentrate. Food Chemistry, 81(4), 583-587. http://dx.doi.org/10.1016/S0308-8146(02)00500-9.
http://dx.doi.org/10.1016/S0308-8146(02)...
). Extracts with composition and similar characteristics obtained in this work are being applied in yogurts (Chouchouli et al., 2013Chouchouli, V., Kalogeropoulos, N., Konteles, S. J., Karvela, E., Makris, D. P., & Karathanos, V. T. (2013). Fortification of yoghurts with grape (Vitis vinifera) seed extracts. Lebensmittel-Wissenschaft + Technologie, 53(2), 522-529. http://dx.doi.org/10.1016/j.lwt.2013.03.008.
http://dx.doi.org/10.1016/j.lwt.2013.03....
).

3.5 Stability of anthocyanins

Table 4 shows the mean values of the monomeric anthocyanin content and the percentage of polymeric color determined periodically during 83 days of storage at 4 °C. In Table 5, the kinetic parameters for the degradation of monomeric anthocyanins and the increase of the polymeric color are observed. A 50% lower reaction rate for anthocyanin degradation was observed in extracts of Hibiscus sabdariffa, during storage at 4 °C (Sinela et al., 2017Sinela, A., Rawat, N., Mertz, C., Achir, N., Fulcrand, H., & Dornier, M. (2017). Anthocyanins degradation during storage of Hibiscus sabdariffa extract and evolution of its degradation products. Food Chemistry, 214, 234-241. http://dx.doi.org/10.1016/j.foodchem.2016.07.071. PMid:27507471.
http://dx.doi.org/10.1016/j.foodchem.201...
). Cissé et al. (2012)Cissé, M., Bohuon, P., Sambe, F., Kane, C., Sakho, M., & Dornier, M. (2012). Aqueous extraction of anthocyanins from Hibiscus sabdariffa: Experimental kinetics and modeling. Journal of Food Engineering, 109(1), 16-21. http://dx.doi.org/10.1016/j.jfoodeng.2011.10.012.
http://dx.doi.org/10.1016/j.jfoodeng.201...
reported a degradation rate of anthocyanins in Hibiscus sabdariffa extract, similar to that found in this present study. They also reported that the increase in polymer color is directly related to the degradation of anthocyanins.

Table 4
Variation of monomeric anthocyanins content and percentage of polymeric color in the concentrated extract stored at 4 °C
Table 5
Kinetic parameters of the degradation of monomeric anthocyanins and generation of polymeric color of the concentrated extract in storage at 4 °C.

4 Conclusions

The optimization of the polyphenol extraction of the grape residue with ethanol acidified with citric acid showed good results. It was observed that temperature and time significantly influenced the extraction of the three phytochemicals. While the solid:liquid ratio was only significant for the extraction of anthocyanins. The quadratic model used to adjust the experimental data was significant, although only linear and interaction components were significant. The obtained extract is rich in polyphenols and from the toxicological point of view is safe, due to the use of only ethanol and citric acid in the extraction, both considered nontoxic. Therefore, the extract can be used in the preparation of food safely. The degradation of monomeric anthocyanins and the increase of polymeric color followed a first order kinetics. The degradation rate of anthocyanins was similar to that reported in other studies, therefore, it can be inferred that the ethanol acidified with the citric acid used in the extraction of polyphenols has a positive effect on the stability of anthocyanins under refrigeration.

Acknowledgements

This study was financed in part by the Coordenação de Aperfeiçoamento de Pessoal de Nível Superior - Brasil (CAPES) - Finance Code 001. Authors are also acknowledged with the National Institute of Science and Technology of Tropical Fruits and CNPq for the financial support; to the Organization of American States (OAS) and to the Grupo Coimbra das Universidades Brasileiras (GCUB). The authors are also grateful to the Company “Agroindustria Frutnaã” for providing the agro industrial waste of grape cv. Isabel used is this research.

  • Pratical Application: Extraction of antioxidant compounds from agroindustrial residues of grapes for food use.

References

  • Abe, L. T., Mota, R. V., Lajolo, F. M., & Genovese, M. I. (2007). Compostos fenólicos e capacidade antioxidante de cultivares de uvas Vitis labrusca L. e Vitis vinifera L. Food Science and Technology, 27(2), 394-400. http://dx.doi.org/10.1590/S0101-20612007000200032
    » http://dx.doi.org/10.1590/S0101-20612007000200032
  • Aguilera, Y., Mojica, L., Rebollo-Hernanz, M., Berhow, M., De Mejía, E. G., & Martín-Cabrejas, M. A. (2016). Black bean coats: new source of anthocyanins stabilized by B-cyclodextrin copigmentation in a sport beverage. Food Chemistry, 212, 561-570. http://dx.doi.org/10.1016/j.foodchem.2016.06.022 PMid:27374568.
    » http://dx.doi.org/10.1016/j.foodchem.2016.06.022
  • Alwazeer, D., Bulut, M., & Tunçtürk, Y. (2020). Fortification of milk with plant extracts modifies the acidification and reducing capacities of yoghurt bacteria. International Journal of Dairy Technology, 73(1), 117-125. http://dx.doi.org/10.1111/1471-0307.12643
    » http://dx.doi.org/10.1111/1471-0307.12643
  • Arici, M., Karasu, S., Baslar, M., Toker, O. S., Sagdic, O., & Karaagacli, M. (2016). Tulip petal as a novel natural food colorant source: extraction optimization and stability studies. Industrial Crops and Products, 91, 215-222. http://dx.doi.org/10.1016/j.indcrop.2016.07.003
    » http://dx.doi.org/10.1016/j.indcrop.2016.07.003
  • Association of Official Analytical Chemistry – AOAC. (2002). Official methods of analysis of the Association of Official Analytical Chemists (16th ed.). Gaithersburg: AOAC.
  • Avizcuri, J.-M., Sáenz-Navajas, M.-P., Echávarri, J.-F., Ferreira, V., & Fernández-Zurbano, P. (2016). Evaluation of the impact of initial red wine composition on changes in color and anthocyanin content during bottle storage. Food Chemistry, 213, 123-134. http://dx.doi.org/10.1016/j.foodchem.2016.06.050 PMid:27451163.
    » http://dx.doi.org/10.1016/j.foodchem.2016.06.050
  • Barcia, M. T., Pertuzatti, P. B., Rodrigues, D., Gómez-Alonso, S., Hermosín-Gutiérrez, I., & Godoy, H. T. (2014). Occurrence of low molecular weight phenolics in Vitis vinifera red grape cultivars and their winemaking by-products from São Paulo (Brazil). Food Research International, 62, 500-513. http://dx.doi.org/10.1016/j.foodres.2014.03.051
    » http://dx.doi.org/10.1016/j.foodres.2014.03.051
  • Bataglion, G. A., Da Silva, F. M. A., Eberlin, M. N., & Koolen, H. H. F. (2015). Determination of the phenolic composition from Brazilian tropical fruits by UHPLC-MS/MS. Food Chemistry, 180, 280-287. http://dx.doi.org/10.1016/j.foodchem.2015.02.059 PMid:25766829.
    » http://dx.doi.org/10.1016/j.foodchem.2015.02.059
  • Chouchouli, V., Kalogeropoulos, N., Konteles, S. J., Karvela, E., Makris, D. P., & Karathanos, V. T. (2013). Fortification of yoghurts with grape (Vitis vinifera) seed extracts. Lebensmittel-Wissenschaft + Technologie, 53(2), 522-529. http://dx.doi.org/10.1016/j.lwt.2013.03.008
    » http://dx.doi.org/10.1016/j.lwt.2013.03.008
  • Cissé, M., Bohuon, P., Sambe, F., Kane, C., Sakho, M., & Dornier, M. (2012). Aqueous extraction of anthocyanins from Hibiscus sabdariffa: Experimental kinetics and modeling. Journal of Food Engineering, 109(1), 16-21. http://dx.doi.org/10.1016/j.jfoodeng.2011.10.012
    » http://dx.doi.org/10.1016/j.jfoodeng.2011.10.012
  • Cvjetko Bubalo, M., Ćurko, N., Tomašević, M., Kovačević Ganić, K., & Radojčić Redovniković, I. (2016). Green extraction of grape skin phenolics by using deep eutectic solvents. Food Chemistry, 200, 159-166. http://dx.doi.org/10.1016/j.foodchem.2016.01.040 PMid:26830574.
    » http://dx.doi.org/10.1016/j.foodchem.2016.01.040
  • Derringer, G., & Suich, R. (1980). Simultaneous optimisation of several response variables. Journal of Quality Technology, 12(4), 214-219. http://dx.doi.org/10.1080/00224065.1980.11980968
    » http://dx.doi.org/10.1080/00224065.1980.11980968
  • Devesa-Rey, R., Vecino, X., Varela-Alende, J. L., Barral, M. T., Cruz, J. M., & Moldes, A. B. (2011). Valorization of winery waste vs. the costs of not recycling. Waste Management (New York, N.Y.), 31(11), 2327-2335. http://dx.doi.org/10.1016/j.wasman.2011.06.001 PMid:21752623.
    » http://dx.doi.org/10.1016/j.wasman.2011.06.001
  • Dhekney, S. A. (2016). Grape. In B. Caballero, P. M. Finglas, & F. Toldrá (Eds.), Encyclopedia of food and health (pp. 261-265). London: Elsevier. http://dx.doi.org/10.1016/B978-0-12-384947-2.00360-3
    » http://dx.doi.org/10.1016/B978-0-12-384947-2.00360-3
  • Giusti, M. M., & Wrolstad, R. E. (2001). Characterization and measerument of anthocyanins by UV-visible spectroscopy. Current Protocols in Food Analytical Chemistry, 1, F1.2.1-F1.2.13. https://doi.org/10.1002/0471142913.faf0102s00
    » https://doi.org/10.1002/0471142913.faf0102s00
  • González-Centeno, M. R., Knoerzer, K., Sabarez, H., Simal, S., Rosselló, C., & Femenia, A. (2014). Effect of acoustic frequency and power density on the aqueous ultrasonic-assisted extraction of grape pomace (Vitis vinifera L.) - A response surface approach. Ultrasonics Sonochemistry, 21(6), 2176-2184. http://dx.doi.org/10.1016/j.ultsonch.2014.01.021 PMid:24548543.
    » http://dx.doi.org/10.1016/j.ultsonch.2014.01.021
  • Goula, A. M., Thymiatis, K., & Kaderides, K. (2016). Valorization of grape pomace: drying behavior and ultrasound extraction of phenolics. Food and Bioproducts Processing, 100, 132-144. http://dx.doi.org/10.1016/j.fbp.2016.06.016
    » http://dx.doi.org/10.1016/j.fbp.2016.06.016
  • Harsha, P. S. C., Gardana, C., Simonetti, P., Spigno, G., & Lavelli, V. (2013). Characterization of phenolics, in vitro reducing capacity and anti-glycation activity of red grape skins recovered from winemaking by-products. Bioresource Technology, 140, 263-268. http://dx.doi.org/10.1016/j.biortech.2013.04.092 PMid:23707914.
    » http://dx.doi.org/10.1016/j.biortech.2013.04.092
  • Kirca, A., & Cemeroglu, B. (2003). Degradation kinetics of anthocyanins in blood orange juice and concentrate. Food Chemistry, 81(4), 583-587. http://dx.doi.org/10.1016/S0308-8146(02)00500-9
    » http://dx.doi.org/10.1016/S0308-8146(02)00500-9
  • Lee, J., Durst, R. W., & Wrolstad, R. E. (2005). Determination of total monomeric anthocyanin pigment content of fruit juices, beverages, natural colorants, and wines by the pH differential method: collaborative study. Journal of AOAC International, 88(5), 1269-1278. http://dx.doi.org/10.1093/jaoac/88.5.1269 PMid:16385975.
    » http://dx.doi.org/10.1093/jaoac/88.5.1269
  • Lees, D. H., & Francis, F. J. (1972). Standardization of Pigment Analyses in Cranberries. HortScience, 7, 83-84.
  • Li, Y., Han, L., Ma, R., Xu, X., Zhao, C., Wang, Z., Chen, F., & Hu, X. (2012). Effect of energy density and citric acid concentration on anthocyanins yield and solution temperature of grape peel in microwave-Assisted extraction process. Journal of Food Engineering, 109(2), 274-280. http://dx.doi.org/10.1016/j.jfoodeng.2011.09.021
    » http://dx.doi.org/10.1016/j.jfoodeng.2011.09.021
  • Liazid, A., Guerrero, R. F., Cantos, E., Palma, M., & Barroso, C. G. (2011). Microwave assisted extraction of anthocyanins from grape skins. Food Chemistry, 124(3), 1238-1243. http://dx.doi.org/10.1016/j.foodchem.2010.07.053
    » http://dx.doi.org/10.1016/j.foodchem.2010.07.053
  • Mammadova, S. M., Fataliyev, H. K., Gadimova, N. S., Aliyeva, G. R., Tagiyev, A. T., & Baloglanova, K. V. (2020). Production of functional products using grape processing residuals. Food Science and Technology, 2061, 1-7. http://dx.doi.org/10.1590/fst.30419
    » http://dx.doi.org/10.1590/fst.30419
  • McGuire, R. G. (1992). Reporting of objective color measurements. HortScience, 27(12), 1254-1255. http://dx.doi.org/10.21273/HORTSCI.27.12.1254
    » http://dx.doi.org/10.21273/HORTSCI.27.12.1254
  • Minjares-Fuentes, R., Femenia, A., Garau, M. C., Meza-Velázquez, J. A., Simal, S., & Rosselló, C. (2014). Ultrasound-assisted extraction of pectins from grape pomace using citric acid: A response surface methodology approach. Carbohydrate Polymers, 106(1), 179-189. http://dx.doi.org/10.1016/j.carbpol.2014.02.013 PMid:24721067.
    » http://dx.doi.org/10.1016/j.carbpol.2014.02.013
  • Morais, D. R., Rotta, E. M., Sargi, S. C., Schmidt, E. M., Bonafe, E. G., Eberlin, M. N., Sawaya, A. C. H. F., & Visentainer, J. V. (2015). Antioxidant activity, phenolics and UPLC-ESI(-)-MS of extracts from different tropical fruits parts and processed peels. Food Research International, 77, 392-399. http://dx.doi.org/10.1016/j.foodres.2015.08.036
    » http://dx.doi.org/10.1016/j.foodres.2015.08.036
  • Norman, G. (2010). Likert scales, levels of measurement and the “laws” of statistics. Advances in Health Sciences Education : Theory and Practice, 15(5), 625-632. http://dx.doi.org/10.1007/s10459-010-9222-y PMid:20146096.
    » http://dx.doi.org/10.1007/s10459-010-9222-y
  • O’Shea, N., Arendt, E. K., & Gallagher, E. (2012). Dietary fibre and phytochemical characteristics of fruit and vegetable by-products and their recent applications as novel ingredients in food products. Innovative Food Science & Emerging Technologies, 16, 1-10. http://dx.doi.org/10.1016/j.ifset.2012.06.002
    » http://dx.doi.org/10.1016/j.ifset.2012.06.002
  • Pedro, A. C., Granato, D., & Rosso, N. D. (2016). Extraction of anthocyanins and polyphenols from black rice (Oryza sativa L.) by modeling and assessing their reversibility and stability. Food Chemistry, 191, 12-20. http://dx.doi.org/10.1016/j.foodchem.2015.02.045 PMid:26258696.
    » http://dx.doi.org/10.1016/j.foodchem.2015.02.045
  • Pinelo, M., Rubilar, M., Jerez, M., Sineiro, J., & Núñez, M. J. (2005). Effect of solvent, temperature, and solvent-to-solid ratio on the total phenolic content and antiradical activity of extracts from different components of grape pomace. Journal of Agricultural and Food Chemistry, 53(6), 2111-2117. http://dx.doi.org/10.1021/jf0488110 PMid:15769143.
    » http://dx.doi.org/10.1021/jf0488110
  • Ribeiro, L. F., Ribani, R. H., Francisco, T. M. G., Soares, A. A., Pontarolo, R., & Haminiuk, C. W. I. (2015). Profile of bioactive compounds from grape pomace (Vitis vinifera and Vitis labrusca) by spectrophotometric, chromatographic and spectral analyses. Journal of Chromatography. B, Analytical Technologies in the Biomedical and Life Sciences, 1007, 72-80. http://dx.doi.org/10.1016/j.jchromb.2015.11.005 PMid:26590878.
    » http://dx.doi.org/10.1016/j.jchromb.2015.11.005
  • Ribeiro, T. P., Lima, M. A. C., Alves, R. E., Gonçalves, A. L. S., & Souza, A. P. C. (2018). Chemical characterization of winemaking byproducts from grape varieties cultivated in Vale do São Francisco, Brazil. Food Science and Technology, 38(4), 577-583. http://dx.doi.org/10.1590/fst.01116
    » http://dx.doi.org/10.1590/fst.01116
  • Rodriguez-Saona, L. E., & Wrolstad, R. E. (2001). Extraction, isolation, and purification of anthocyanins. Current Protocols in Food Analytical Chemistry, 1, F1.1.1-F1.1.11. https://doi.org/10.1002/0471142913.faf0101s00
    » https://doi.org/10.1002/0471142913.faf0101s00
  • Sang, J., Sang, J., Ma, Q., Hou, X., & Li, C. Q. (2017). Extraction optimization and identification of anthocyanins from Nitraria tangutorun Bobr. seed meal and establishment of a green analytical method of anthocyanins. Food Chemistry, 218, 386-395. http://dx.doi.org/10.1016/j.foodchem.2016.09.093 PMid:27719925.
    » http://dx.doi.org/10.1016/j.foodchem.2016.09.093
  • Schwartz, S. J., von Elbe, J. H., & Giusti, M. M. (2010). Colorants. In S. Damodaran, K. L. Parkin, & O. R. Fennema (Eds.), Food chemistry (4th ed., pp. 571-632). London: CRC Press, Taylor & Francis Group.
  • Sharma, R. J., Gupta, R. C., Singh, S., Bansal, A. K., & Singh, I. P. (2016). Stability of anthocyanins- and anthocyanidins-enriched extracts, and formulations of fruit pulp of Eugenia jambolana (‘jamun’). Food Chemistry, 190, 808-817. http://dx.doi.org/10.1016/j.foodchem.2015.06.029 PMid:26213042.
    » http://dx.doi.org/10.1016/j.foodchem.2015.06.029
  • Sinela, A., Rawat, N., Mertz, C., Achir, N., Fulcrand, H., & Dornier, M. (2017). Anthocyanins degradation during storage of Hibiscus sabdariffa extract and evolution of its degradation products. Food Chemistry, 214, 234-241. http://dx.doi.org/10.1016/j.foodchem.2016.07.071 PMid:27507471.
    » http://dx.doi.org/10.1016/j.foodchem.2016.07.071
  • Sui, X., Dong, X., & Zhou, W. (2014). Combined effect of pH and high temperature on the stability and antioxidant capacity of two anthocyanins in aqueous solution. Food Chemistry, 163, 163-170. http://dx.doi.org/10.1016/j.foodchem.2014.04.075 PMid:24912712.
    » http://dx.doi.org/10.1016/j.foodchem.2014.04.075
  • Sun, X., Liu, L., Ma, T., Yu, J., Huang, W., Fang, Y., & Zhan, J. (2019). Effect of high Cu2+ stress on fermentation performance and copper biosorption of Saccharomyces cerevisiae during wine fermentation. Food Science and Technology, 39(1), 19-26. http://dx.doi.org/10.1590/1678-457x.24217
    » http://dx.doi.org/10.1590/1678-457x.24217
  • Takeuchi, T. M., Pereira, C. G., Braga, M. E. M., Maróstica, M. R., Leal, P. F., & Meireles, M. A. A. (2009). Low-pressure solvent extraction (solid–liquid extraction, microwave assisted, and ultrasound assisted) from condimentary plants. In M. A. A. Meireles (Ed.), Extracting bioactive compounds for food products: theory and applications (pp. 140-142). London: CRC Press, Taylor & Francis Group.
  • Teixeira, A., Baenas, N., Dominguez-Perles, R., Barros, A., Rosa, E., Moreno, D. A., & Garcia-Viguera, C. (2014). Natural bioactive compounds from winery by-products as health promoters: a review. International Journal of Molecular Sciences, 15(9), 15638-15678. http://dx.doi.org/10.3390/ijms150915638 PMid:25192288.
    » http://dx.doi.org/10.3390/ijms150915638
  • Toaldo, I. M., Fogolari, O., Pimentel, G. C., de Gois, J. S., Borges, D. L. G., Caliari, V., & Bordignon-Luiz, M. (2013). Effect of grape seeds on the polyphenol bioactive content and elemental composition by ICP-MS of grape juices from Vitis labrusca L. Lebensmittel-Wissenschaft + Technologie, 53(1), 1-8. http://dx.doi.org/10.1016/j.lwt.2013.02.028
    » http://dx.doi.org/10.1016/j.lwt.2013.02.028
  • Vital, A. C. P., Santos, N. W., Matumoto-Pintro, P. T., da Silva Scapim, M. R., & Madrona, G. S. (2018). Ice cream supplemented with grape juice residue as a source of antioxidants. International Journal of Dairy Technology, 71(1), 183-189. http://dx.doi.org/10.1111/1471-0307.12412
    » http://dx.doi.org/10.1111/1471-0307.12412
  • Wettasinghe, M., & Shahidi, F. (1999). Evening primrose meal: a source of natural antioxidants and scavenger of hydrogen peroxide and oxygen-derived free radicals. Journal of Agricultural and Food Chemistry, 47(5), 1801-1812. http://dx.doi.org/10.1021/jf9810416 PMid:10552455.
    » http://dx.doi.org/10.1021/jf9810416
  • Yamamoto, L. Y., de Assis, A. M., Roberto, S. R., Bovolenta, Y. R., Nixdorf, S. L., García-Romero, E., Gómez-Alonso, S., & Hermosín-Gutiérrez, I. (2015). Application of abscisic acid (S-ABA) to cv. Isabel grapes (V. vinifera × V. labrusca) for color improvement: effects on color, phenolic composition and antioxidant capacity of their grape juice. Food Research International, 77, 572-583. http://dx.doi.org/10.1016/j.foodres.2015.10.019
    » http://dx.doi.org/10.1016/j.foodres.2015.10.019
  • Zhou, S. H., Guo, R. R., Wei, R. F., Liu, J. B., Yu, H., Shi, X. F., Zhang, Y., Xie, T. L., & Cheng, G. (2020). Effects of bagging or the combination of umbrella and bag treatments on anthocyanin accumulation in the berry skin of ‘Kyoho’ (Vitis labruscana) grape. Food Science and Technology, 40(2), 394-400. http://dx.doi.org/10.1590/fst.41218
    » http://dx.doi.org/10.1590/fst.41218

Publication Dates

  • Publication in this collection
    18 Dec 2020
  • Date of issue
    2021

History

  • Received
    27 July 2020
  • Accepted
    24 Sept 2020
Sociedade Brasileira de Ciência e Tecnologia de Alimentos Av. Brasil, 2880, Caixa Postal 271, 13001-970 Campinas SP - Brazil, Tel.: +55 19 3241.5793, Tel./Fax.: +55 19 3241.0527 - Campinas - SP - Brazil
E-mail: revista@sbcta.org.br