You are on page 1of 10

Food Research International xxx (xxxx) xxx–xxx

Contents lists available at ScienceDirect

Food Research International


journal homepage: www.elsevier.com/locate/foodres

Evaluation of bioactive compounds potential and antioxidant activity of


brown, green and red propolis from Brazilian northeast region
Julianna Karla Santana Andradea, Marina Denadaia,⁎, Christean Santos de Oliveiraa,
Maria Lucia Nunesb, Narendra Naraina
a
Laboratory of Flavor and Chromatographic Analysis, PROCTA, Federal University of Sergipe, 49100-000 São Cristóvão, SE, Brazil
b
Department of Food Technology, Federal University of Ceara, CEP 60020-180 Fortaleza, Brazil

A R T I C L E I N F O A B S T R A C T

Keywords: The aim of the present study was to determine the contents of bioactive compounds present in brown, green and
Propolis red species of propolis cultivated in the Brazilian northeast states of Alagoas and Sergipe. The contents of
Bioactive compounds phenolic compounds, flavonoids and antioxidant activity (DPPH, ABTS+, FRAP, ORAC) were determined.
Flavonoids Identification and quantification of phenolic and flavonoid compounds were performed by using UHPLC-QqQ-
Antioxidant activity
MS/MS system. The results revealed high contents of total phenolics and flavonoids. Among the three species,
UHPLC-QqQ-MS/MS
the antioxidant potential had higher capacity in the red propolis. The presence of some of bioactive compounds
viz. acacetin, artepellin C, eriodictyol, gallic acid, isorhamnetin, protocatechuic acid, vanillin and vanillic acid in
Brazilian red propolis is reported for the first time in this work. Positive correlation between total phenolics
versus the FRAP and ORAC methods was established which led to conclusion that antioxidant activity of propolis
is mainly due to its phenolic compounds.

1. Introduction The chemical composition of propolis varies according to the geo-


graphic region, climate, environmental conditions and collecting season
Propolis is a resinous substance, dark in color and it is collected by (López et al., 2014; Sawaya, Barbosa da Silva Cunha, & Marcucci,
honeybees, mainly from leaf and flower buds, stems and cracks in the bark 2011). Although phenolic compounds are the most abundant in pro-
of many species of trees. This material is transported to the hive and mixed polis, yet > 300 compounds have been identified in different species,
with beeswax, producing a strongly adhesive substance, which for cen- such as phenolic acids, flavonoids including flavones, flavanones, fla-
turies has been used worldwide in traditional medicine (Daugsch, Moraes, vonols and chalcones, terpenes, aromatic aldehydes, alcohols, fatty
Fort, & Park, 2008; Pellati, Orlandini, Pinetti, & Benvenuti, 2011). acids, stilbenes, steroids, amino acids, lignans and sugars (Akyol et al.,
In South America, Brazil is well known for its green propolis, produced 2013; da Silva Frozza et al., 2013; Righi et al., 2011).
by Apis mellifera, which collect resins mainly from a native shrub Baccharis Raw propolis cannot be used as feedstock and it must be purified.
dracunculifolia (López, Schmidt, Eberlin, & Sawaya, 2014). However, due This process should remove the waxy material and preserve the fraction
to the large Brazilian biodiversity, there are 13 types of propolis, which of polyphenols, which are considered to contribute most to the curative
include less common brown and red propolis and these are classified based effects than the other constituents of propolis. The most popular tech-
on the place of production where these are found (Sawaya et al., 2004). nique for the production of propolis extracts is ethanol extraction due to
Since propolis is a natural product characterizing for several biological and the fact that active substances of propolis are more easily soluble in
pharmacological properties, it has attracted interest of researchers in the etanol (Pietta, Gardana, & Pietta, 2002). According to other authors,
last decades. The therapeutic properties of propolis are well known in extraction using the hydroalcoholic solvent (ethanol) have been de-
popular medicine, due its antiseptic, antitumoral (Franchi et al., 2012; scribed as the most suitable means for the extraction of biologically
Frozza et al., 2017), antiinflammatory (Bufalo et al., 2013; Franchi et al., active phenolic components from propolis materials (Cottica et al.,
2012), immunomodulatory (Bufalo et al., 2013), antioxidant (Franchi 2011; Frozza et al., 2017; López et al., 2014; Pellati et al., 2011). Sun,
et al., 2012), antibacterial, antimicrobial activities (Bufalo et al., 2013; Wu, Wang, and Zhang (2015) stated in his study that ethanol/water
Franchi et al., 2012; Sforcin & Bankova, 2011; Szliszka et al., 2013; Viuda- solvents have a significant effect on the phenolic composition and an-
Martos, Ruiz-Navajas, Fernandez-Lopez, & Perez-Alvarez, 2008). tioxidant properties of the propolis extracts.


Corresponding author.
E-mail address: marrydenadai@gmail.com (M. Denadai).

http://dx.doi.org/10.1016/j.foodres.2017.08.066
Received 11 July 2017; Received in revised form 29 August 2017; Accepted 30 August 2017
0963-9969/ © 2017 Published by Elsevier Ltd.

Please cite this article as: Andrade, J.K.S., Food Research International (2017), http://dx.doi.org/10.1016/j.foodres.2017.08.066
J.K.S. Andrade et al. Food Research International xxx (xxxx) xxx–xxx

The compounds daidzein, biochanin A (López et al., 2014; Silva The samples of brown, green and red propolis were collected for a
et al., 2015), pinocembrin and quercetin (Daugsch et al., 2008; de period of 7 days in the afternoon time at about 16 h. Three batches of
Mendonça et al., 2015) are biomarkers of Brazilian red propolis. Besides propolis sample weighing about 50 g each were collected to carry out
these compounds, the red propolis is known to possess chemical sub- the analysis. Raw propolis were ground using a mill (IKA, Brazil, A11
stances which were not found in other varieties of propolis, such as basic) and the powder was later stored in glass containers which were
vestitol and neovestitol, C-glycoside, liquiritigenin, isoliquiritigenin, maintained at 0 °C.
formononetin, and medicarpin (Bueno-Silva et al., 2013; da Silva
Frozza et al., 2013; López et al., 2014; Piccinelli et al., 2011). However, 2.3. Preparation of propolis extracts
the principal constituents of Brazilian green propolis are caffeic acid, p-
coumaric acid, ferulic acid, naringenin, kaempferol, isorhamnetin, sa- Raw propolis sample (2 g) was extracted with 15 mL of ethanol:-
kurametin, pinocembrin, kaempferide and artepellin C (Szliszka et al., water (70:30, v/v) in an ultrasound bath apparatus (3L Alpha Plus) at
2013). 35 °C for 60 min. The extracts were centrifuged at 400g for 10 min,
In Brazil, green and red propolis species are the most studied using a centrifuge (Eppendorf, 5810R). The supernatants were col-
(Frozza et al., 2017; Hatano et al., 2012; Silva et al., 2015), although lected, dried at 35 °C in a drying oven, and resuspended in 2 mL of
there are some reports in literature on other species, like brown pro- ethanol:water (70:30, v/v). Finally, samples were filtered through
polis (Bittencourt et al., 2015). However, there are no reports on the 0.2 μm cellulose filters (Millipore, Bedford, MA, USA) and injected in
presence and comparative data on phenolic compounds and antioxidant the UHPLC-QqQ-MS/MS system.
activity among the various varieties of Brazilian propolis. Thus the in-
vestigation on the chemical composition of different species and their 2.4. Determination of total phenolics content
comparative evaluation becomes important.
In recent years, the use of Ultra-high Performance Liquid The total content of phenolic compounds was determined by using
Chromatographic system coupled with Tandem Mass Spectrometry Folin-Ciocalteu phenol reagent (Shetty, Curtis, Levin, Witkowsky, & Ang,
(UHPLC-QqQ-MS/MS) is becoming the most commonly employed 1995). One milliliter of the ethanolic extract was transferred to test tubes
method for determination of phenolic compounds, due to the system's and 1 mL of 95% ethanol solution, 5 mL of distilled water and 0.5 mL of
high sensitivity, selectivity and high-throughput capability. Thus the Folin-Ciocalteu reagent (1N) were added, followed by homogenization in
objective of this study was to use UHPLC-QqQ-MS/MS system to identify vórtex. Later, 1 mL of sodium carbonate solution 5% (w/v) was added. The
and quantify phenolic compounds and to evaluate their antioxidant ac- test tubes were kept in dark for 60 min, and then homogenized in vórtex.
tivity in brown, red and green propolis species grown in the states of The absorbance was measured at a wavelength of 725 nm against a blank
Alagoas and Sergipe, pertaining to the Northeast region in Brazil. consisting of 95% ethanol solution, using a spectrophotometer (Molecular
Devices, Sunnyvale, CA, USA; SpectraMax M2). For the quantification of
2. Materials and methods these extracts, a calibration curve was constructed from the analysis of
different concentrations of gallic acid varying from 0.0008–0.1 mg/mL, and
2.1. Analytical standards and reagents its data on absorbances based on calibration equation: y = 10.571x
+ 0.0111 (r2 = 0.9958). Results were expressed in terms of milligrams of
All the organic solvents employed were of HPLC grade. Water used gallic acid equivalent (GAE) per g of fresh sample weight.
for the mobile phase was purified through a Milli-Q system (Millipore,
São Paulo, Brazil; Direct-Q® 3UV). The solvents acetonitrile and formic 2.5. Determination of total flavonoids contents
acid used were of HPLC grade 98% of purity obtained from Sigma
Aldrich and Fluka Analytica (St Louis, MO, USA). Apigenin (C15H10O5), The total flavonoids content was determined according to the
acacetin (C16H12O5), artepellin C (C19H24O3), biochanin A (C16H12O5), method proposed by Meda, Lamien, Romito, Millogo, and Nacoulma
cinnamic acid (C9H8O2), α-cyano-4-hydroxycinnamic acid (C10H7O3N), (2005). Five hundred microliters of extract were transferred to test tube
caffeic acid (C9H8O4), ferulic acid (C10H10O4), caffeic acid phenyl ester and 0.5 mL of a 20 mg·mL− 1 methanolic solution of aluminum chloride
(CAPE) (C17H16O4), (+)-catechin (C15H14O6), chrysin (C15H10O4), (ALCL3) was added. Samples were homogenized on a vortex and left in
epicatechin (C15H14O6), eriodictyol (C15H12O6), ethyl gallate the dark for 30 min. The spectrophotometer (Molecular Devices, Sun-
(C9H10O5), gallic acid (C7H6O5), isorhamnetin (C16H12O7), kaempferide nyvale, CA, USA; SpectraMax M2) was set at the wavelength of 415 nm
(C16H12O6), kaempferol (C15H10O6), luteolin (C15H10O6), narigenin and absorbance reading measured. The calibration curve was con-
(C15H12O5), p-coumaric acid (C9H8O3), protocatechuic acid (C7H6O4), structed from different concentrations of quercetin varying from
pinocembrin (C15H12O4), quercetin-3-glucoside (C21H20O12), chrolo- 0.0008–0.1 mg/mL, and its data on absorbances based on calibration
genic acid (C16H18O9), rutin (C27H30O16), vanillin (C8H8O3) and va- equation: y = 21.874x − 0.0047 (r2 = 0.9998). The results were ex-
nillic acid (C8H8O4) were purchased from Sigma-Aldrich (Saint Louis, pressed in terms of milligrams of quercetin per g of propolis.
MO, USA). The reagents and standards: etanol; aluminum chloride;
sodium carbonate, potassium phosphate buffer, sodium citrate, ferrous 2.6. Determination of antioxidant activity
sulfate; Folin-Ciocalteu phenol reagent; fluorescein; 6-hydroxy-2,5,7,8-
tetramethylchromo-2-carboxylic acid (Trolox); 2,2-diphenyl-1-pi- 2.6.1. DPPH assay
crylhydrazyl radical (DPPH%); 2,2′-azino-bis (3-ethylbenzthiazoline) 6- The antioxidant capacity was determined by the DPPH radical (2,2-
sulfonic acid (ABTS+); 2,2-Azobis (2-Amidino-Propane) dichloride diphenyl-1-picrylhydrazyl) scavenging method recommended by Kwon,
(AAPH); and FRAP reagent were obtained from Sigma Aldrich and Vattem, and Shetty (2006) with slight modifications. A 250 μL aliquot
Fluka Analytica (St Louis, MO, USA). of extract was mixed with 1.25 mL of DPPH. After 5 min, the absor-
bance was read at 517 nm using a spectrophotometer (Molecular De-
2.2. Propolis samples vices, Sunnyvale, CA, USA; SpectraMax M2). The readings were com-
pared with the controls, containing 95% ethanol instead of extract. The
Brown, red and green raw propolis samples were collected from percentage inhibition was calculated by Eq. (1). Different concentra-
apiaries located at Marechal Deodoro, Alagoas state, Brazil, at co- tions of Trolox varying from 0 to 0.0014 mmol Trolox/mL were used to
ordinates 9° 45′34.454″ S, 35° 50′24.986″ W in January 2016 while the construct the calibration curve based on calibration equation:
raw red propolis was collected from apiary, located in Brejo Grande, y = − 421.1x + 0.7193 (r2 = 0.9958), and expressed the results in
Sergipe, Brazil, at coordinates 10° 25′28″ S, 36° 27′44″ W, in July 2016. terms of μmol Trolox per g of propolis.

2
J.K.S. Andrade et al. Food Research International xxx (xxxx) xxx–xxx

% Inhibition =
Abscontrol − Absextract
× 100 200 °C, gas flow: 12 L·min− 1, nebulizer: 20 psi, sheat gas temp: 400 °C,
Abscontrol (1) sheat gas flow: 11 L·min− 1, capillary voltage: 3500 V, nozzle: 500 V,
dwell time: 9.8, acceleration cell voltage: 5 V.
where: Abscontrol = absorbance of the control sample and Ab-
The standard calibration curve was prepared using eight data
sextract = absorbance of the extract. Result was expressed as percentage
points, covering the concentration ranges (20–1000 ng·mL− 1). All
inhibition.
analyses were performed in triplicate.

2.6.2. ABTS+ assay


2.8. Statistical analysis
Free radical ABTS+ was captured according to the methodology
proposed by Nenadis, Wang, Tsimidou, and Zhang (2004), with slight
The test results were evaluated by the method of analysis of var-
modifications. An aliquot of 30 μL of extract was transferred to test
iance (ANOVA) with comparison of means by Tukey test at 95% con-
tubes with 3.0 mL of the ABTS+ radical and homogenized in vortex.
fidence level with the help of SAS version 9.0 software. The correlation
After 6 min, the absorbance value was read in the spectrophotometer
between results (Total phenolics, flavonoids, DPPH, ABTS*, FRAP and
(Molecular Devices, Sunnyvale, CA, USA; SpectraMax M2) at 734 nm.
ORAC) was determined from the Pearson correlation coefficient, which
Different concentrations of Trolox varying from 0 to 0.0018 mmol
quantifies the linear association between two quantitative variables.
Trolox/mL were used to construct the calibration curve based on cali-
This correlation was determined using Statistica 7.0 software.
bration equation: y = − 247.07x + 0.6706 (r2 = 0.996). Results were
expressed in terms of μmol Trolox per g of propolis.
3. Results and discussion

2.6.3. FRAP assay


3.1. Total phenolic and flavonoid contents
The antioxidant activity through the iron reduction method (FRAP)
was determined according to the methodology proposed by Thaipong,
Phenolic and flavonoid compounds are the main components re-
Boonprakob, Crosby, Cisneros-Zevallos, and Hawkins Byrne (2006). For sponsible for the functional property of propolis. Table 1 shows sig-
the analysis, an extract aliquot of 90 μL was transferred into test tubes,
nificant difference by Turkey's test (p < 0.05) for total phenolic and
with 270 μL of distilled water, mixed with 2.7 mL of the FRAP reagent, flavonoids composition among brown, green and red varieties. Red
homogenized on a vortex and kept at 37 °C in a water bath. The ab-
propolis had 91.3 mg GAE·g− 1 of total phenolics, which was the highest
sorbance value was read in a spectrophotometer (Molecular Devices, among all varieties. Regarding flavonoids content, green propolis
Sunnyvale, CA, USA; SpectraMax M2) at 595 nm after 30 min. A cali-
showed the highest amount (59.45 mg quercetin·g− 1 of propolis) when
bration curve, ranging from 0 to 0.00064 mmol Trolox/mL, was con-
compared with red and brown.
structed based on calibration equation: y = 1740.5x + 0.1223
Some studies with Brazilian propolis have reported the phenolic
(r2 = 0.998). The results were expressed in terms of μmol Trolox per g
concentrations ranging from 48 to 87 mg GAE·g− 1 for propolis of states
of propolis.
of Paraná (Cottica et al., 2011) and 49–100 mg GAE·g− 1 for the etha-
nolic extracts of green propolis from the state of São Paulo
2.6.4. ORAC assay (Mello & Hubinger, 2012), although the results vary depending on the
The absorption capacity of the oxygen radical (ORAC) was analyzed collecting region and species. Bittencourt et al. (2015) reported with
according to the methodology proposed by Albarici (2009), with slight respect to the content of total phenolics the value of 48.24 ± 1.09 mg
modifications. 1.50 mL of fluorescein working solution (63 mmol·L− 1) GAE·g− 1 of dry weight for ethanolic extract of Brazilian brown propolis
and 0.75 mL of sample were added, and tubes were maintained at 37 °C and 185.52 ± 1.09 mg GAE·g− 1 of dry weight for ethanolic extract of
for 15 min. Later, 0.75 mL of AAPH working solution was added. The Brazilian green propolis.
experiment was performed in a 96-well microplate and 350 μL of the Regarding the flavonoid content, the red propolis (31.48 mg of
sample was added. The antioxidant activity was monitored at every 5 min quercetin·g− 1 of propolis) exhibited a similar result was found by Righi
by measuring fluorescence for 1000 min in a spectrophotometer (Mole- et al. (2011) for red propolis obtained from Maceio, state of Alagoas, in
cular Devices, USA; SpectraMax M2) using an excitation filter of 485 nm northeastern Brazil (32.9 mg of quercetin·g− 1 of propolis). According
and emission of 520 nm. Different concentrations of Trolox varying from to these authors, flavonoid contents of Brazilian red propolis were
5.0–90.0 μmol Trolox/mL were used to construct the calibration curve found to vary from 27 and 43 mg·g− 1. Due to the geographical origin of
based on equation: y = 0.4339x + 0.5079 (r2 = 0.9984). The results the samples, there is diversity in polyphenol content of propolis, which
were expressed in terms of μmol Trolox per g of propolis. affects the biological activity and pharmacological effects among the
species (Falcao et al., 2010; Gómez-Caravaca, Gómez-Romero, Arráez-
2.7. Identification and quantification of phenolic and flavonoid compounds Román, Segura-Carretero, & Fernández-Gutiérrez, 2006). Cottica et al.
using UHPLC-QqQ-MS/MS system (2011) obtained values of total flavonoids between 10 and 26 mg
quercetin·g− 1 of Brazilian propolis extract from Paraná. Hatano et al.
Identification and quantification of phenolic and flavonoid com- (2012), obtained the flavonoid content of 51.9 ± 2.4 mg·g− 1 of
pounds present in propolis was performed using an Agilent UHPLC-MS/ ethanolic extract of Brazilian green propolis from the state of Minas
MS system of the Triple Quadrupole type (Agilent 6490 Triple Quad LC- Gerais. Wang et al. (2016) obtained values of 53.0 ± 0.22 mg
MS/MS), with electrospray ionization in positive mode, except for
caffeic acid phenethyl ester (CAPE) which was in negative mode, in Table 1
SRM (Selected Reaction Monitoring) experiments. The column used was Total phenolic compounds and flavonoids contents (mean ± standard deviation; n = 9)
Ascents Express F5 (150 × 2.1 mm, 2.7 μm particle; Sigma Aldrich) in brown, green and red propolis.

operating at a flow rate of 0.2 ml·min− 1 and a temperature of 40 °C.


Propolis Total phenolic compounds (mg Total flavonoids content (mg
The mobile phases used were ultrapure water + 0.1% formic acid GAE·g− 1 of propolis) quercetin·g− 1 of propolis)
(Phase A) and acetonitrile + 0.1% formic acid (Phase B). The injection
volume was 2 μL. The gradient elution was as follows: 0.0–1.0 min, Brown 55.74 ± 0.48c 30.89 ± 0.20c
Green 90.55 ± 1.52b 59.45 ± 0.82a
15% B; 7.0–9.0 min, 25% B; 13.0–16.0 min, 30%; 17.0 min, 35%;
Red 91.32 ± 0.49a 31.48 ± 0.56b
21.0–23.0 min, 40% B; 25.0 min, 45% B; 28.0 min, 50% B; 30.0 min,
15% B. ⁎
Means followed by same lower-case letters in the same column do not differ statistically
The parameters of the mass spectrometer were: gas temperature was from each other at the 5% significance level (p < 0.05).

3
J.K.S. Andrade et al. Food Research International xxx (xxxx) xxx–xxx

quercetin·g− 1 of ethanol extract from Brazilian propolis of Olimpia of These compounds presented the higher concentration in red propolis sam-
states São Paulo. ples, when compared with the other species.
It is observed that the green and brown Brazilian propolis of the Acacetin, artepellin C, eriodictyol, gallic acid, isorhamnetin, pro-
state of Alagoas and the Brazilian red propolis of the state of Sergipe, tocatechuic acid, vanillin and vanillic acid were identified in red pro-
obtained different concentrations of phenolics and flavonoids among polis analyzed in this work. Fig. 2 illustrates the SRM chromatograms of
themselves and when compared to the propolis of the varieties (brown, the quantification transitions of these phenolic compounds in red pro-
green and red) from other Brazilian states. This can be explained due to polis extract. The presence of these compounds in Brazilian red propolis
the fact that the chemical composition of propolis can vary significantly is reported for the first time in this work.
depending on the different geographic regions, climatic conditions, the The distinct chemical composition of red propolis plays an im-
vegetation growing around hives, as well as the time of collection and portant role in the treatment of human diseases, such as dental caries,
these differences can influence the biological activity (López et al., candidiasis, cancer, skin wound and oxidative stress-related conditions
2014; Piccinelli et al., 2011; Sawaya et al., 2011). (Freires, de Alencar, & Rosalen, 2016). Biochanin A and daidzein, which
are well known biomarkers of Brazilian red propolis, were also found in
3.2. Identification and determination of phenolic and flavonoids compounds the samples. Daidzein has significant antiinflammatory potential, and
by UHPLC-QqQ-MS/MS previous studies demonstrated that administration of this compound
significantly reduces the number of neutrophils, release antiin-
Some phenolic compounds cannot be separated with good efficiency flammatory cytokines, and other benefits as well (Franchin et al.,
in a single LC-MS system. Tandem mass spectrometric technique is a 2017). Additionally, the presence of a 5-hydroxil group in the chemical
powerful tool to provide better information and the characteristic structure of biochanin A appears to enhance inhibitory activity of some
spectra profile of the target compounds. The SRM mode of UHPLC-QqQ- enzymatic systems, including testosterone 5α-reductase, which con-
MS/MS is a highly sensitive technique used to determine phenolic verts testosterone to dihydrotestosterone, as reported by Bae et al.
compounds based on the monitoring of specified molecular ion and (2012). The inhibition of this enzyme represents an important phar-
their respective ion transitions, which avoid interference of overlapping macological approach against androgen dependent diseases, such
LC peaks (Sherwood et al., 2009; Wu et al., 2016). prostate hyperplasia.
For quantification of target compounds, a calibration curve was In general, the results obtained in this study exhibit more quantities of
prepared by analyzing at different concentrations (20–1000 ng/mL). polyphenolic compounds in ethanolic extracts from green propolis, when
Regression analysis produced r2 values above 0.99 and maximum de- compared with the red and brown extracts. These results are in agreement
tection limit (MDL) ranged from 0.02 to 1.12 ng·mL− 1 (Table 2). with total phenolic and flavonoid results, which demonstrate that green
In order to obtain maximum sensitivity for identification of all propolis exhibited the highest amount of these class of compounds. In an
compounds, two SRM transitions were optimized for each compound. earlier study reported by (Barbaric et al., 2011), the chemical composition
The first transitions were used for quantification, and the second ones of ethanolic extracts from propolis was found to have lower levels of the
for confirmatory purposes. The data on precursor and product ions, phenolic compounds viz. ferulic acid (0.0358 to 0.3022 mg·g− 1 of pro-
with associated collision energy for fragmentation and the retention polis), p-coumaric acid (0.00441 to 0.1562 mg·g− 1 of propolis), pino-
times are presented in Table 2. The target compounds were identified in cembrin (0.2487 to 7.6651 mg·g− 1 of propolis), apigenin (0.0019 to
accordance with the SRM chromatogram of the two selected transitions, 0.0189 mg·g− 1 of propolis) and kaempferol (0.0697 to 0.2931 mg·g− 1 of
and their corresponding retention times. The chromatographic separa- propolis). When determining the phenolic compounds in propolis from
tion of the 28 target compounds was optimized in order to get the best China, Sun et al. (2015) found low quantities of p-coumaric acid
resolution, efficiency and selectivity, particularly for enantiomeric se- (1.40 mg·g− 1), cinnamic acid (0.10 mg·g− 1), kaempferol (0.59 mg·g− 1)
paration of isomers, in case of (+)-catechin and epicatechin, which and quercetin (0.10 mg·g− 1). Vargas-Sánchez et al. (2014), in a previous
exhibit the same molar mass and the same patterns of fragmentation, study with Mexican propolis, found similar results for kaempferol
consequently resulting in identical product ions. After testing several (1.45 mg·g− 1) when comparing with the green propolis of the present
conventional reverse phase stationary phases, the Ascentis Express F5 work. In the same study, pinocembrin (1.65 mg·g− 1) and luteolin
column provided a proper resolution of these isomeric catechins. (0.87 mg·g− 1) also exhibited similar results for brown and red propolis,
The quantification of individual compounds was calculated using respectively.
the peak areas of the identified compounds relative to the peak areas of Daidzein, formononetine and biochanin A were identified by Silva
the corresponding analytical standard. The data on contents of flavo- et al. (2015) in red propolis from Brazilian northeast region. These
noids and phenolic acids found in propolis samples are presented in compounds are biomarkers of chemical composition for red propolis
Table 3. The chromatographic profile of the three propolis species is species (Franchi et al., 2012; López et al., 2014; Piccinelli et al., 2011).
show in Fig. 1. Frozza et al. (2017) found low quantities of biochanin A (0.27 mg·g− 1)
The results indicate that the 3 propolis species (brown, green and red) in Brazilian red propolis from the state of Alagoas. Other phenolic
had similar bioactive profile composition, but these were quantitatively compounds, such as isoliquiritigenina, pinocembrina and quercetin
different. The flavonoids artepellin C, kaempferide, kaempferol, pinocem- were also identified in red propolis samples (Daugsch et al., 2008; de
brim, and the phenolic acids such as p-coumaric, chrologenic and caffeic Mendonça et al., 2015).
acids were the most abundant compounds found in brown and green pro- Among the several target polyphenolic compounds found in the
polis. However, kaempferide was the majoritary compound (6.04 mg·g− 1) different propolis species (brown, green and red) from Brazil, artepillin
in green propolis followed by p-coumaric acid (5.34 mg·g− 1), artepellin C C, kaempferide, p-coumaric acid, luteolin, chrologenic acid, kaemp-
(4.80 mg·g− 1), chrologenic acid (2.81 mg·g− 1), kaempferol (1.48 mg·g− 1) ferol, caffeic acid and pinocembrin were in majority. These compounds
and caffeic acid (1.06 mg·g− 1). Brown propolis contained higher quantity have been reported in the literature as beneficial for human health.
(3.72 mg·g− 1) of artepellin C followed by kaempferide (3.48 mg·g− 1), p- Some authors have attributed antibacterial, antitumoral and antiin-
coumaric acid (2.64 mg·g− 1), chrologenic acid (1.76 mg·g− 1) and pino- flammatory activities for artepillin C (Hata et al., 2012; Ikeda et al.,
cembrim (1.66 mg·g− 1). Artepellin C was identified in all 3 propolis species 2011; Paulino et al., 2008) and p-coumaric acid (Punithavathi, Prince,
analyzed in this work. As expected, artepellin C concentration was higher in Kumar, & Selvakumari, 2011), including other therapeutic properties.
green propolis, due to its characteristic composition in this specie (Hata Chrologenic acid has demonstrated good antioxidant activity in in vitro
et al., 2012; Ikeda et al., 2011). Major compounds found in red propolis studies (Farooqui & Farooqui, 2012), preventing the DNA oxidation
were luteolin (1.75 mg·g− 1), naringenin (0.96 mg·g− 1), kaempferol resulting in the prevention of many chronical diseases. Caffeic acid,
(0.59 mg·g− 1), pinocembrin (0.41 mg·g− 1) and biochanin A (0.39 mg·g− 1). luteolin and kaempferol have been studied by many authors (Bufalo

4
J.K.S. Andrade et al. Food Research International xxx (xxxx) xxx–xxx

Table 2
Analytical MS/MS parameters retention time, linearity parameters, correlation coefficients and detection limits for phenolic and flavonoid compounds.

Compound Molecular ESI polarity Precursor ion Fragments Collision Retention Calibration equationa Correlation MDL
formula m/z energy (eV) time (min) coefficient (r2) (ng·mL− 1)

Acacetin C16H12O5 + 285 242b 25 24.2 y = 6587 + 69212x 0.998 0.022


270
Apigenin C15H10O5 + 271 153b 25 17.7 y = 11116 + 2146x − 0.4294x2 0.999 0.077
119
Artepellin C C19H24O3 + 301 245b 15 29.3 y = 12979 + 1680x + 0.7351x 2
0.995 0.017
189
Caffeic acid C9H8O4 + 181 135b 25 4.4 y = 433.1 − 558.4 0.998 0.400
163 15
Biochanin A C16H12O5 + 285 253b 25 24.5 y = 8429.9 + 5661.8x − 1.45x2 0.997 0.130
152
Caffeic acid phenethyl C17H16O4 − 283 161b 25 23.3 y = 121.8 − 565.1x 0.999 0.150
ester (CAPE) 179 15
Chrysin C15H10O4 + 255 153b 25 25.6 y = 3297 + 29158x 0.998 1.29
209
(+)-Catechin C15H14O6 + 291 139b 15 3.1 y = 1460 − 2538x 0.998 0.116
123
Epicatechin C15H14O6 + 291 139b 15 3.6 y = 1965 − 28235x 0.997 2.39
123
Daidzein C15H10O4 + 255 199b 25 11.8 y = 164591 + 8489x − 2,96x 2
0.997 0.229
137
Ethyl gallate C9H10O5 + 199 127b 15 5.6 y = 976.8 − 6209x 0.998 0.103
153
Eriodictyol C15H12O6 + 289 153b 30 13.1 y = 2707 − 15399x 0.998 1.61
145 25
Chrologenic acid C16H18O9 + 355 163b 15 3.16 y = 10531 − 57618x 0.996 0.149
145 25
Gallic acid C7H6O5 + 171 109b 15 2.17 y = 171 − 2135x 0.995 16.5
125
Cinnamic acid C9H8O2 + 149 103b 25 12.5 y = 1139 − 10095x 0.998 0.749
131 15
α-Cyano-4- C10H7O3N + 190 172b 15 8.48 y = 242 + 1725x 0.997 9.33
hydroxycinnamic 144
acid
Isorhamnetin C16H12O7 + 317 302b 25 18.6 y = − 6352 + 1254x + 0.18x2 0.995 2.03
229 35
Luteolin C15H10O6 + 287 153b 35 15.0 y = 2600 + 50848x 0.998 1.51
135 35
Kaempferide C16H12O6 + 301 258b 35 24.3 y = 1296 − 6160x 0.997 0.013
286 25
Kaempferol C15H10O6 + 287 153b 35 17.6 y = 1193 − 6547x 0.997 1.80
121 25
Narigenin C15H12O5 + 273 153b 35 16.0 y = 3146 − 7237x 0.998 0.040
119
p-Coumaric acid C9H8O3 + 165 147b 15 6.51 y = 971.3 − 6661x 0.997 1.12
91 35
Protocatechuic acid C7H6O4 + 155 65b 15 2.80 y = 414.6 − 2985x 0.997 0.270
137 35
Pinocembrin C15H12O4 + 257 153b 15 23.1 y = − 68899 + 4477x 0.998 0.070
131 25
Ferulic acid C10H10O4 + 195 177b 15 7.47 y = 2811 − 28884x 0.997 0.192
134 20
Quercetin-3-glucoside C21H20O12 + 465 303b 15 7.24 y = 1541 − 7238x 0.998 0.926
85 35
Rutin C27H30O16 + 611 303b 35 6.46 y = 840.7 − 3036x 0.997 0.066
465 15
Vanillin C8H8O3 + 153 93b 15 6.51 y = 9927 − 72786x 0.998 0.054
65 35
Vanillic acid C8H8O4 + 169 93b 15 4.57 y = 20015 − 3631x + 0.4320x2 0.996 0.221
65 35

MDL - Method Detection Limits.


a
Y is the value of the peak area. X is the concentration of the standard compound.
b
Quantitation transition.

et al., 2013; Chen et al., 2008; Filomeni et al., 2012) against different flavonoids, such kaempferide, and prenylated phenylpropanoids
cancer cell lines and these compounds, inhibit significantly the pro- (Falcao et al., 2010; Gómez-Caravaca et al., 2006). These observations
liferation of tumoral cells. are in agreement with our findings, which exhibited that kaempferide
The main characteristic of chemical composition of propolis from was the major compound found in green and brown propolis species.
tropical locals, particularly in Brazil, is the presence of non-typical The phenolic contents in propolis may vary depending on the origin of

5
J.K.S. Andrade et al. Food Research International xxx (xxxx) xxx–xxx

Table 3 Bărnuţiu (2011) observed lower DPPH values ranging from


Phenolic and flavonoid compounds contents (mg·g− 1) in brown, green and red propolis. 0.29 ± 0.10 to 1.24 ± 0.01 mmol Trolox·g− 1 propolis of Transylva-
nian propolis extracts. Kalogeropoulos, Konteles, Troullidou,
Compounds Content (mg·g− 1 of propolis)
Mourtzinos, and Karathanos (2009) found DPPH values for propolis
Brown Green Red from Greece ranging from 0.33 to 1.11 mmol Trolox·g− 1 propolis.
Vargas-Sánchez et al. (2014), when analyzing raw propolis from
Acacetin 0.08 ± 0.011 0.11 ± 0.01 0.08 ± 0.00
Mexico, found a percentage inhibition of the DPPH radical of 64.8%,
α-Cyano-4-hydroxycinnamic acid < LQ < LQ < LQ
Apigenin 0.03 ± 0.01 0.01 ± 0.00 0.01 ± 0.00 which is lower than found in this work.
Artepellin C 3.72 ± 0.00 4.80 ± 0.03 0.01 ± 0.00 The results of ORAC exhibited higher antioxidant responses for
Biochanin A 0.01 ± 0.00 0.01 ± 0.00 0.39 ± 0.02 green and red propolis (6734 and 6665 μmol Trolox·g− 1 propolis, re-
Caffeic acid 0.59 ± 0.06 1.06 ± 0.04 < LQ spectively) when compared with brown propolis (5343 μmol
Caffeic acid phenethyl ester 0.09 ± 0.01 0.29 ± 0.02 < LQ
Trolox·g− 1 propolis). However, no significant difference (p < 0.05)
(CAPE)
± Catechin 0.01 ± 0.00 0.01 ± 0.00 < LQ was observed by Tukey test. The results of ORAC tests on propolis, and
Chlorogenic acid 1.76 ± 0.03 2.81 ± 0.01 0.01 ± 0.00 the antioxidant responses may vary due the differences in species and
Chrysin 0.14 ± 0.04 0.07 ± 0.01 < LQ their geographic localization (Hatano et al., 2012; Silva et al., 2011;
Cinnamic acid 0.10 ± 0.01 0.10 ± 0.00 < LQ
Sun et al., 2015).
Daidzein 0.001 ± 0.00 < LQ 0.12 ± 0.00
Epicatechin 0.01 ± 0.00 0.01 ± 0.00 < LQ
Eriodictyol 0.01 ± 0.00 0.01 ± 0.00 0.01 ± 0.00 3.4. Pearson correlation coefficients
Gallic acid 0.04 ± 0.00 0.04 ± 0.00 0.01 ± 0.00
Isorhamnetin 0.40 ± 0.03 0.42 ± 0.01 0.01 ± 0.00 The mechanism used to measure the antioxidant capacity is dif-
Kaempferide 3.48 ± 0.03 6.04 ± 0.08 < LQ
ferent in each method, which may cause slight differences between the
Kaempferol 0.61 ± 0.19 1.48 ± 0.00 0.59 ± 0.00
Luteolin 0.04 ± 0.02 0.01 ± 0.00 1.75 ± 0.01 results. Thus, all methods were compared using the Pearson correlation
Narigenin 0.15 ± 0.03 0.25 ± 0.03 0.96 ± 0.03 coefficient (r) between the variables (Total Phenolic contents,
p-Coumaric acid 2.64 ± 0.22 5.34 ± 0.28 0.01 ± 0.00 Flavonoids, DPPH, ABTS*, FRAP and ORAC), which quantifies the
Pinocembrin 1.66 ± 0.15 0.93 ± 0.08 0.41 ± 0.02
linear association between two quantitative variables. This correlation
Protocatechuic acid 0.18 ± 0.01 0.11 ± 0.00 0.05 ± 0.00
Quercetin-3-glucoside 0.24 ± 0.06 0.45 ± 0.06 < LQ
data are presented in Table 5.
Rutin 0.13 ± 0.02 0.16 ± 0.02 0.05 ± 0.00 A very strong positive correlation was observed between total
Trans-isoferulic acid 0.06 ± 0.00 0.15 ± 0.02 < LQ phenolics and the FRAP and ORAC methods (r = 0.9888 and
Vanillin 0.04 ± 0.01 0.02 ± 0.00 0.01 ± 0.00 r = 0.9980, respectively), which suggests the relation of the phenolic
Vanillic acid 0.02 ± 0.00 0.01 ± 0.00 0.01 ± 0.00
compounds with the antioxidant activity. As for the DPPH method,

LQ: limit of quantification. there was a strong positive correlation (r = 0.9492) with the FRAP
method and a very strong and positive correlation between the DPPH
and ABTS* methods (r = 0.9744). The antioxidant activity evaluated
by the ABTS method showed a strong positive correlation (r = 0.8542)
the samples, and its differences may affect the biological activity and with the FRAP assay. The ORAC method presented a very strong cor-
pharmacological effects. relation (r = 0.9774) with the iron reduction potential (FRAP).
The correlation observed between the parameters shows statistically
that the flavonoids are not responsible for the antioxidant activity, since
3.3. Antioxidant activity the Pearson correlation coefficient was 0.0526 and − 0.1730 for the
DPPH and ABTS* assays, respectively and of 0.3643 and 0.5529 for the
The Table 4 presents the results of DPPH, ABTS+, FRAP and ORAC FRAP and ORAC assays, respectively. This fact suggests that the anti-
activities of different species of propolis. All the extracts possessed oxidant activity of brown, green and red Brazilian propolis is related to
potent and diverse antioxidant activity with significant variations other phenolic compounds, and not directly related to the class of fla-
(p < 0.05). Red propolis exhibited the higher antioxidant potential in vonoids.
DPPH, ABTS+ and FRAP methods, followed by green and brown spe- According to Cottica et al. (2011), a negative correlation (− 0.9082)
cies. was observed between DPPH and total flavonoids content in a Brazilian
The antioxidant activity of three propolis when compared with the propolis collected in Maringá in the state of Paraná. In addition, the
results of propolis reported in other studies, show wide difference in authors also showed that the FRAP value may not be related to the
their antioxidant potential. FRAP antioxidant assay exhibited the values flavonoid content, since the correlation coefficient was only 0.4915.
ranging from 471 to 633 mmol Trolox·g− 1propolis. Turan et al. (2015) This fact corroborated with the results presented in this study. Similar
reported FRAP value of 311.0 ± 2.5 mg Trolox·g− 1 in extracts of results were also reported by Can, Yildiz, Şahin, Asadov, and Kolayli
propolis cultivated in Turkey, which was lower than the result herein (2015), showing a very strong and positive correlation (0.98) between
obtained. the FRAP values and the total phenolic compounds content in the
With regard to DPPH assay, inhibition percentage varied from 86 to propolis samples, which suggests that the total antioxidant activity of
90.7%. Red propolis exhibited the higher inhibition (90.72%) of DPPH the propolis is due to their phenolic compounds.
radical, followed by the green (88.53%) and brown (86.06%) propolis.
Skaba et al. (2013) found low values of DPPH (1230.07 ± 135.55 4. Conclusions
μmol Trolox·g− 1) and ABTS%+ (1223.06 ± 137.40 μmol Trolox·g− 1)
in ethanolic extract of Brazilian green propolis from the state of Minas Ethanolic extracts of Brazilian propolis from different species
Gerais, Brazil. In other study, Bonvehí and Gutiérrez (2011) obtained (brown, green and red) presented high flavonoid and phenolic com-
values of ABTS%+ between 560 and 1430 μmol Trolox·g− 1 in ethanolic pounds contents. In addition, the results obtained by DPPH, ABTS,
extract of propolis from Spain. Mihai, Mărghitaş, Dezmirean, and FRAP and ORAC assays revealed high antioxidant activity, indicating

6
J.K.S. Andrade et al. Food Research International xxx (xxxx) xxx–xxx

Fig. 1. Total ion chromatogram (TIC) of ethanolic extracts of Brazilian propolis (brown, green and red) in the SRM mode. (For interpretation of the references to color in this figure
legend, the reader is referred to the web version of this article.)

that propolis is a promising source of biologically active polyphenols. The results on total phenolic compounds and flavonoids in all 3 propolis
The UHPLC-QqQ-MS/MS system provided good selectivity and effi- species (brown, green and red) obtained from the Alagoas and Sergipe
ciency, in the determination of 28 phenolic compounds with good states of Brazilian northeastern region were correlated with their anti-
precision and accuracy. Comparing the 3 different propolis species oxidant activities (DPPH, ABTS, FRAP and ORAC). A very strong positive
analyzed, brown and green propolis displayed higher amounts of correlation between total phenolics and the FRAP and ORAC methods was
bioactive compounds than the red species, which reveals the differences established which led to conclusion that antioxidant activity of propolis is
in the chemical composition of the different species. mainly due to its phenolic components composition.

7
J.K.S. Andrade et al. Food Research International xxx (xxxx) xxx–xxx

Fig. 2. SRM chromatograms of ethanolic extract of red propolis according to quantification transitions of acacetin, artepellin C, gallic acid, eriodictyol, isorhamnetin, protocatechuic acid,
vanillin and vanillic acid.

Table 4
Antioxidant activity (mean ± standard deviation; n = 9) of brown, green and red propolis.

Propolis Antioxidant activity (μmol Trolox·g− 1 propolis)

DPPH ABTS+ FRAP ORAC

Brown 4431.00 ± 45.11bB 1868.45 ± 131.39cC 471.51 ± 37.79bD 5343.84 ± 114.94bA


Green 4554.35 ± 80.20aB 2214.96 ± 20.61bC 604.20 ± 20.72aD 6734.87 ± 124.56aA
Red 4663.80 ± 68.01aB 2913.55 ± 95.26aC 633.18 ± 40.20aD 6665.75 ± 58.78aA


Means followed by same lower-case letters in the same column do not differ statistically from each other at the 5% significance level (p < 0.05).
⁎⁎
Means followed by same capital letters in the same row do not differ statistically from each other at the 5% significance level (p < 0.05).

Abbreviations
Table 5
Pearson correlation coefficients (r) of total phenolic compounds, flavonoids and anti-
oxidant activity (DPPH, ABTS*, FRAP and ORAC).
ABTS%+ 2,2-Azino-bis (3-ethylbenzthiazoline-6-sulfonic acid)
ABTS radical cation scavenging activity
Parameters Total Flavonoids DPPH ABTS⁎ FRAP ORAC AAPH 2,2-Azobis (2-Amidino-Propane) dichloride
phenolic DPPH radical scavenging activity
Total phenolic 1 0.4994 0.8914 0.7669 0.9888⁎ 0.9980⁎⁎
DPPH% 2,2-diphenyl-1-picryhydrazyl radical
Flavonoids 1 0.0526 −0.1730 0.3643 0.5529 FRAP ferric reducing antioxidant power
DPPH 1 0.9744⁎ 0.9492 0.8612 GAE gallic acid equivalent
ABTS 1 0.8542⁎ 0.7250 Trolox 6-hydroxy-2,5,7,8-tetramethylchroman-2-carboxylic acid
FRAP 1 0.9774⁎
ALCL3 methanolic solution of aluminum chloride
ORAC 1
ORAC oxygen radical absorbance capacity

Significant at 5% probability (p < 0.05).
⁎⁎
Significant at 1% probability (p < 0.01).

8
J.K.S. Andrade et al. Food Research International xxx (xxxx) xxx–xxx

Acknowledgment novel anti-inflammatory drugs. European Journal of Medicinal Chemistry. http://dx.


doi.org/10.1016/j.ejmech.2017.06.050 (In press).
Freires, I. A., de Alencar, S. M., & Rosalen, P. L. A. (2016). A pharmacological perspective
Special thanks are also extended to Apicola Almar ltda, Maceio, on the use of Brazilian Red Propolis and its isolated compounds against human dis-
Alagoas, Brazil for supply of the green and brown propolis and to Dr. eases. European Journal of Medicinal Chemistry, 110, 267–279.
Frozza, C. O.d. S., Santos, D. A., Rufatto, L. C., Minetto, L., Scariot, F. J., Echeverrigaray,
Juliana Cordeiro Cardoso of the University Tiradentes, Aracaju, Brazil S., ... Henriques, J. A. P. (2017). Antitumor activity of Brazilian red propolis fractions
for supplying the red propolis samples. against Hep-2 cancer cell line. Biomedicine & Pharmacotherapy, 91, 951–963.
Gómez-Caravaca, A. M., Gómez-Romero, M., Arráez-Román, D., Segura-Carretero, A., &
Fernández-Gutiérrez, A. (2006). Advances in the analysis of phenolic compounds in
Funding sources products derived from bees. Journal of Pharmaceutical and Biomedical Analysis, 41(4),
1220–1234.
All authors gratefully acknowledge financial support received from Hata, T., Tazawa, S., Ohta, S., Rhyu, M.-R., Misaka, T., & Ichihara, K. (2012). Artepillin C,
a major ingredient of Brazilian propolis, induces a pungent taste by activating TRPA1
CNPq, Brazil vide research project ‘Instituto Nacional de Ciência e
channels. PloS One, 7(11), 48072.
Tecnologia de Frutos Tropicais’ (Project 465335/2014-4) in developing Hatano, A., Nonaka, T., Yoshino, M., Ahn, M.-R., Tazawa, S., Araki, Y., & Kumazawa, S.
this work. Authors (Julianna Andrade, Marina Denadai) acknowledge (2012). Antioxidant activity and phenolic constituents of red propolis from
and thank CAPES (Ministry of Education, Brazil) for their fellowships. Shandong, China. Food Science and Technology Research, 18(4), 577–584.
Ikeda, R., Yanagisawa, M., Takahashi, N., Kawada, T., Kumazawa, S., Yamaotsu, N., ...
Tsuda, T. (2011). Brazilian propolis-derived components inhibit TNF-a-mediated
Notes down regulation of adiponectin expression via different mechanisms in 3T3-L1 adi-
pocytes. Biochimica et Biophysica Acta (BBA) - General Subjects, 1810(7), 695–703.
Kalogeropoulos, N., Konteles, S. J., Troullidou, E., Mourtzinos, I., & Karathanos, V. T.
The authors declare that there is no competing financial interest. (2009). Chemical composition, antioxidant activity and antimicrobial properties of
propolis extracts from Greece and Cyprus. Food Chemistry, 116, 452–461.
References Kwon, Y.-I. I., Vattem, D. A., & Shetty, K. (2006). Evaluation of clonal herbs of Laminaceae
species against diabetes and hypertension. Asia Pacific Journal of Clinical Nutrition,
15(1), 107–118.
Akyol, S., Ozturk, G., Ginis, Z., Armutcu, F., Yigitoglu, M. R., & Akyol, O. (2013). In vivo López, B. G.-C., Schmidt, E. M., Eberlin, M. N., & Sawaya, A. C. H. F. (2014).
and in vitro antineoplastic actions of caffeic acid phenethyl ester (CAPE): therapeutic Phytochemical markers of different types of red propolis. Food Chemistry, 146,
perspectives. Nutrition and Cancer, 65(4), 515–526. 174–180.
Albarici, T. R. (2009). Protocolo 4 - Atividade antioxidante pelo método orac. In T. R. Meda, A., Lamien, C. E., Romito, M., Millogo, J., & Nacoulma, O. G. (2005).
Albarici, D. M. de Freitas, & J. D. C. Pessoa (Eds.). Protocolos de análise para polpa de Determination of the total phenolic, flavonoid and proline contents in Burkina Fasan
açaí: um guia prático de consulta (pp. 33–38). Embrapa Instrumentação Agropecuária. honey, as well as their radical scavenging activity. Food Chemistry, 91(3), 571–577.
Bae, M., Woo, M., Kusuma, I. W., Arung, E. T., Yang, C. H., & Kim, Y. U. (2012). Inhibitory Mello, B. C. B. S., & Hubinger, M. D. (2012). Antioxidant activity and polyphenol contents
effects of isoflavonoids on rat prostate testosterone 5α-reductase. Journal of in Brazilian green propolis extracts prepared with the use of ethanol and water as
Acupuncture and Meridian Studies, 5(6), 319–322. solvents in different pH values. International Journal of Food Science and Technology,
Barbaric, M., Miskovic´, K., Bojic, M., Loncar, M. B., Smolcic-Bubalo, A., Debeljak, Z., & 47(12), 2510–2518.
Medic-Saric, M. (2011). Chemical composition of the ethanolic propolis extracts and Mihai, C. M., Mărghitaş, L. A., Dezmirean, D. S., & Bărnuţiu, L. (2011). Correlation be-
its effect on HeLa cells. Journal of Ethnopharmacology, 135(3), 772–778. tween polyphenolic profile and antioxidant activity of propolis from Transylvania.
Bittencourt, M. L. F., Ribeiro, P. R., Franco, R. L. P., Hilhorst, H. W. M., de Castro, R. D., & Animal Science and Biotechnologies, 44(2), 100–103.
Fernandez, L. G. (2015). Metabolite profiling, antioxidant and antibacterial activities Nenadis, N., Wang, L. F., Tsimidou, M., & Zhang, H. Y. (2004). Estimation of scavenging
of Brazilian propolis: Use of correlation and multivariate analyses to identify po- activity of phenolic compounds using the ABTS(*+) assay. Journal of Agricultural and
tential bioactive compounds. Food Research International, 76, 449–457. Food Chemistry, 52, 4669–4674.
Bonvehí, J. S., & Gutiérrez, A. L. (2011). Antioxidant activity and total phenolics of Paulino, N., Abreu, S. R. L., Uto, Y., Koyama, D., Nagasawa, H., Hori, H., ... Bretz, W. A.
propolis from the Basque Country (Northeastern Spain). Journal of the American Oil (2008). Anti-inflammatory effects of a bioavailable compound, Artepillin C, in
Chemists' Society, 88, 1387–1395. Brazilian propolis. European Journal of Pharmacology, 587, 296–301.
Bueno-Silva, B., Alencar, S. M., Koo, H., Ikegaki, M., Silva, G. V. J., Napimoga, M. H., & Pellati, F., Orlandini, G., Pinetti, D., & Benvenuti, S. (2011). HPLC-DAD and HPLC-ESI-
Rosalen, P. L. (2013). Anti-inflammatory and antimicrobial evaluation of neo vestitol MS/MS methods for metabolite profiling of propolis extracts. Journal of
and vestitol isolated from Brazilian red propolis. Journal of Agricultural and Food Pharmaceutical and Biomedical Analysis, 55(5), 934–948.
Chemistry, 61(19), 4546–4550. Piccinelli, A. L., Lotti, C., Campone, L., Cuesta-Rubio, O., Campo Fernandez, M., &
Bufalo, M. C., Ferreira, I., Costa, G., Francisco, V., Liberal, J., Cruz, M. T., ... Sforcin, J. M. Rastrelli, L. (2011). Cuban and Brazilian red propolis: Botanical origin and com-
(2013). Propolis and its constituent caffeic acid suppress LPS-stimulated pro-in- parative analysis by high-performance liquid chromatography-photodiode array de-
flammatory response by blocking NF-kappaB and MAPK activation in macrophages. tection/electrospray ionization tandem mass spectrometry. Journal of Agricultural and
Journal of Ethnopharmacology, 149, 84–92. Food Chemistry, 59, 6484–6491.
Can, Z., Yildiz, O., Şahin, H., Asadov, A., & Kolayli, S. (2015). Phenolic profile and an- Pietta, P. G., Gardana, C., & Pietta, A. M. (2002). Analytical methods for quality control of
tioxidant potential of propolis from Azerbaijan. Mellifera, 15(1), 16–28. propolis. Fitoterapia, 74, S7–S20.
Chen, H.-Q., Jin, Z.-Y., Wang, X.-J., Xu, X.-M., Deng, L., & Zhao, J.-W. (2008). Luteolin Punithavathi, V. R., Prince, P. S. M., Kumar, R., & Selvakumari, J. (2011).
protects dopaminergic neurons from inflammation-induced injury through inhibition Antihyperglycaemic, antilipid peroxidative and antioxidant effects of gallic acid on
of microglial activation. Neuroscience Letters, 448(2), 175–179. streptozotocin induced diabetic Wistar rats. European Journal of Pharmacology,
Cottica, S. M., Sawaya, A. C. H. F., Eberlin, M. N., Franco, S. L., Zeoula, L. M., & 650(1), 465–471.
Visentainer, J. V. (2011). Antioxidant activity and composition of propolis obtained Righi, A. A., Alves, T. R., Negri, G., Marques, L. M., Breyer, H., & Salatino, A. (2011).
by different methods of extraction. Journal of the Brazilian Chemical Society, 22, Brazilian red propolis: Unreported substances, antioxidant and antimicrobial activ-
929–935. ities. Journal of the Science of Food and Agriculture, 91(13), 2363–2370.
Daugsch, A., Moraes, C. S., Fort, P., & Park, Y. K. (2008). Brazilian red propolis - Chemical Sawaya, A. C., Tomazela, D. M., Cunha, I. B., Bankova, V. S., Marcucci, M. C., Custodio, A.
composition and botanical origin. Evidence-based Complementary and Alternative R., & Eberlin, M. N. (2004). Electrospray ionization mass spectrometry fingerprinting
Medicine, 5(4), 435–441. of propolis. Analyst, 129(8), 739–744.
de Mendonça, I. C. G., Porto, I. C. C.d. M., do Nascimento, T. G., de Souza, N. S., Oliveira, Sawaya, A. C. H. F., Barbosa da Silva Cunha, I., & Marcucci, M. C. (2011). Analytical
J. M.d. S., Arruda, R. E.d. S., ... Barreto, F. S. (2015). Brazilian red propolis: phyto- methods applied to diverse types of Brazilian propolis. Chemistry Central Journal,
chemical screening, antioxidant activity and effect against cancer cells. BMC 5, 27.
Complementary and Alternative Medicine, 15, 357. Sforcin, J. M., & Bankova, V. (2011). Propolis: Is there a potential for the development of
Falcao, S. I., Vilas-Boas, M., Estevinho, L. M., Barros, C., Domingues, M. R. M., & Cardoso, new drugs? Journal of Ethnopharmacology, 133(2), 253–260.
S. M. (2010). Phenolic characterization of Northeast Portuguese propolis: Usual and Sherwood, C. A., Eastham, A., Lee, L. W., Risler, J., Mirzaei, H., Falkner, J. A., & Martin,
unusual compounds. Analytical and Bioanalytical Chemistry, 396, 887–897. D. B. (2009). Rapid optimization of MRM-MS instrument parameters by subtle al-
Farooqui, T., & Farooqui, A. A. (2012). Beneficial effects of propolis on human health and teration of precursor and product m/z targets. Journal of Proteome Research, 8(7),
neurological diseases. Frontiers in Bioscience (Elite Edition), 4, 779–793. 3746–3751.
Filomeni, G., Graziani, I., De Zio, D., Dini, L., Centonze, D., Rotilio, G., & Ciriolo, M. R. Shetty, K., Curtis, O. F., Levin, R. E., Witkowsky, R., & Ang, W. (1995). Prevention of
(2012). Neuroprotection of kaempferol by autophagy in models of rotenone-mediated vitrification associated with in vitro shoot culture of oregano. (Origanum vulgare) by
acute toxicity: Possible implications for Parkinson's disease. Neurobiology of Aging, Pseudomonas spp. Journal of Plant Physiology, 147(3), 447–451.
33(4), 767–785. da Silva Frozza, C. O., Garcia, C. S. C., Gambato, G., de Souza, M. D. O., Salvador, M.,
Franchi, G. C., Moraes, C. S., Toreti, V. C., Daugsch, A., Nowill, A. E., & Park, Y. K. (2012). Moura, S., ... Roesch-Ely, M. (2013). Chemical characterization, antioxidant and
Comparison of effects of the ethanolic extracts of Brazilian propolis on human leu- cytotoxic activities of Brazilian red propolis. Food and Chemical Toxicology, 52,
kemic cells as assessed with the MTT Assay. Evidence-based Complementary and 137–142.
Alternative Medicine918956 (6). Silva, R. O., Andrade, V. M., Rêgo, E. S. B., Dória, G. A. A., Lima, B. S., Silva, F. A., ...
Franchin, M., Freires, I. A., Lazarini, J. G., Nani, B. D., da Cunha, M. G., Colón, D. F., ... Gomes, M. Z. (2015). Acute and sub-acute oral toxicity of Brazilian red propolis in
Rosalen, P. L. (2017). The use of Brazilian propolis for discovery and development of rats. Journal of Ethnopharmacology, 170, 66–71.

9
J.K.S. Andrade et al. Food Research International xxx (xxxx) xxx–xxx

Silva, V., Genta, G., Moller, M. N., Masner, M., Thomson, L., Romero, N., ... Denicola, A. Turan, I., Demir, S., Misir, S., Kilinc, K., Mentese, A., Aliyazicioglu, Y., & Deger, O. (2015).
(2011). Antioxidant activity of Uruguayan propolis. In vitro and cellular assays. Cytotoxic effect of Turkish propolis on liver, colon, breast, cervix and prostate cancer
Journal of Agricultural and Food Chemistry, 59, 6430–6437. cell lines. Tropical Journal of Pharmaceutical Research, 14(5), 777–782.
Skaba, D., Morawiec, T., Tanasiewicz, M., Mertas, A., Bobela, E., Szliszka, E., ... Król, W. Vargas-Sánchez, R. D., Torrescano-Urrutia, G. R., Acedo-Félix, E., Carvajal-Millán, E.,
(2013). Influence of the toothpaste with Brazilian ethanol extract propolis on the oral González-Córdova, A. F., Vallejo-Galland, B., ... Sánchez-Escalante, A. (2014).
cavity health. Evidence-based Complementary and Alternative Medicine215391 (12 Antioxidant and antimicrobial activity of commercial propolis extract in beef patties.
pages). Journal of Food Science, 79(8), C1499–C1504.
Sun, C., Wu, Z., Wang, Z., & Zhang, H. (2015). Effect of ethanol/water solvents on phe- Viuda-Martos, M., Ruiz-Navajas, Y., Fernandez-Lopez, J., & Perez-Alvarez, J. A. (2008).
nolic profiles and antioxidant properties of Beijing propolis extracts. Evidence-based Functional properties of honey, propolis, and royal jelly. Journal of Food Science,
Complementary and Alternative Medicine, 2015, 9. 73(9), R117–R124.
Szliszka, E., Kucharska, A. Z., Sokół-Łętowska, A., Mertas, A., Czuba, Z. P., & Król, W. Wang, X., Sankarapandian, K., Cheng, Y., Woo, S. O., Kwon, H. W., Perumalsamy, H., &
(2013). Chemical composition and anti-inflammatory effect of ethanolic extract of Ahn, Y.-J. (2016). Relationship between total phenolic contents and biological
Brazilian green propolis on activated J774A.1 macrophages. Evidence-based properties of propolis from 20 different regions in South Korea. BMC Complementary
Complementary and Alternative Medicine976415. and Alternative Medicine, 16–65.
Thaipong, K., Boonprakob, U., Crosby, K., Cisneros-Zevallos, L., & Hawkins Byrne, D. Wu, Y., Jiang, X., Zhang, S., Dai, X., Liu, Y., Tan, H., ... Xia, T. (2016). Quantification of
(2006). Comparison of ABTS, DPPH, FRAP, and ORAC assays for estimating anti- flavonol glycosides in Camellia sinensis by MRM mode of UPLC-QQQ-MS/MS. Journal
oxidant activity from guava fruit extracts. Journal of Food Composition and Analysis, of Chromatography B, 1017–1018, 10–17.
19(2006), 669–675.

10

You might also like