Sunteți pe pagina 1din 7

Food Chemistry 197 (2016) 266–272

Contents lists available at ScienceDirect

Food Chemistry
journal homepage: www.elsevier.com/locate/foodchem

Conventional, ultrasound-assisted, and accelerated-solvent extractions


of anthocyanins from purple sweet potatoes
Zhan Cai a,1, Ziqian Qu a,1, Yu Lan b, Shujuan Zhao a, Xiaohua Ma b, Qiang Wan a, Pu Jing a,⇑, Pingfan Li c,⇑
a
Research Center for Food Safety and Nutrition, Key Lab of Urban Agriculture (South), Bor S. Luh Food Safety Research Center, School of Agriculture & Biology, Shanghai Jiao
Tong University, Shanghai 200240, China
b
Hainan Yedao Group Co. Ltd., Hainan 570105, China
c
Guangdong Industry Technical College, Guangdong 510300, China

a r t i c l e i n f o a b s t r a c t

Article history: Purple sweet potatoes (PSPs) are rich in anthocyanins. In this study, we investigated the extraction
Received 22 July 2015 efficiency of anthocyanins from PSPs using conventional extraction (CE), ultrasound-assisted extraction
Received in revised form 19 October 2015 (UAE), and accelerated-solvent extraction (ASE). Additionally, the effects of these extraction methods
Accepted 24 October 2015
on antioxidant activity and anthocyanin composition of PSP extracts were evaluated. In order of
Available online 10 November 2015
decreasing extraction efficiency, the extraction methods were ASE > UAE > CE for anthocyanins
(218–244 mg/100 g DW) and CE > UAE > ASE for total phenolics (631–955 mg/100 g DW) and
Chemical compounds studied in this article:
flavonoids (28–40 mg/100 g DW). Antioxidant activities of PSP extracts were CE  UAE > ASE for ORAC
Cyanidin 3-sophoroside-5-glucoside
(PubChem CID: 44256732)
(766–1091 mg TE/100 g DW) and ASE > CE  UAE for FRAP (1299–1705 mg TE/100 g DW). Twelve
Peonidin 3-sophoroside-5-glucoside anthocyanins were identified. ASE extracts contained more diacyl anthocyanins and less nonacyl and
(PubChem CID: 44256845) monoacyl anthocyanins than CE and ASE extracts (P < 0.05).
Ó 2015 Elsevier Ltd. All rights reserved.
Keywords:
Extraction efficiency
Anthocyanins
Phenolics
Flavonoids
Antioxidant activity

1. Introduction Rodriguez-Solana, Salgado, Dominguez, & Cortes-Dieguez, 2015).


An effective extraction method maximizes the extraction of
Anthocyanins are flavonoids responsible for the red, purple, target compounds with minimal degradation and is based on
violet, and blue colors of fruits, vegetables, and cereals (Mazza & environmental friendly technologies.
Miniati, 1993). The increasing interest in anthocyanins as food The conventional extraction (CE) of anthocyanins, which is
colorants is attributed to the safety concerns of synthetic food based on the bath stirring extraction method, is performed in
color additives (Giusti & Jing, 2007; Wrolstad & Culver, 2012). acidic solutions to obtain the red stable flavylium cation (Santos,
Purple sweet potatoes (Ipomoea batatas L.), which are rich in Veggi, & Meireles, 2010). Compared to CE, ultrasound-assisted
anthocyanins, have been grown in China as food and food colorants extraction (UAE), which has been used in the extraction of antho-
(Fan, Han, Gu, & Chen, 2008). Anthocyanins from purple sweet cyanins (Lien, Chan, Lai, Huang, & Liao, 2012) and other phenolics
potatoes (PSPs) have bioactive effects, including antioxidant (Teow (Carrera, Ruiz-Rodriguez, Palma, & Barroso, 2012), have several
et al., 2007) and anti-inflammatory activities (Zhang et al., 2009). advantages including reduced processing time and solvent volume.
Additionally, PSP anthocyanins attenuate dimethylnitrosamine- Additionally, UAE has an effective mass transfer and solvent
induced liver injury in rats (Hwang et al., 2011). penetration through the disruption of plant cell walls via acoustical
Extraction is a separation technique that affects the yield, qual- cavitation (Rastogi, 2011). Accelerated-solvent extraction (ASE) is a
ity, and composition of target compounds (Jing & Giusti, 2007; solvent extraction technique that is performed at elevated temper-
atures and pressure, conditions that improve extraction kinetics
and reduce extraction time and solvent volume (Abdel-Aal,
⇑ Corresponding authors.
Akhtar, Rabalski, & Bryan, 2014; Truong, Hu, Thompson, Yencho,
E-mail address: pjing@sjtu.edu.cn (P. Jing).
1
These authors made equivalent contributions to the manuscript.
& Pecota, 2012).

http://dx.doi.org/10.1016/j.foodchem.2015.10.110
0308-8146/Ó 2015 Elsevier Ltd. All rights reserved.
Z. Cai et al. / Food Chemistry 197 (2016) 266–272 267

In this study, we investigated the extraction efficiency of antho- Table 1


cyanins from PSPs using CE, UAE, and ASE. The effects of these Factors and levels of Taguchi orthogonal array designs.

methods on anthocyanin composition and antioxidant activity Design Methods Factor (i) Level (j)
were evaluated. 1 2 3
L9 (34) CE Temperature (°C) 60 70 80
Time (min) 90 120 150
2. Materials and methods Ethanol (%, v/v) 70 80 90
HCl (%, v/v) 0.01 0.05 0.1
2.1. Materials and reagents UAE Temperature (°C) 40 50 60
Time (min) 45 60 75
Ethanol (%, v/v) 80 90 100
Purple sweet potatoes (PSPs) were provided by Hainan Yedao
Power (W) 200 240 280
Ltd. (Hainan, China). Pure standards of gallic acid, quercetin, 6-hy
L4 (23) ASE Temperature (°C) 80 90
droxy-2,5,7,8-tetramethylchromane-2-carboxylic acid (Trolox),
Static time (min) 15 20
fluorescein sodium salt, Folin–Ciocalteu reagent (FC reagent), Static cycle 1 2
2,20 -azobis dihydrochloride (AAPH) and 2,4,6-Tripyridyl-s-triazine
(TPTZ) were purchased from Sigma–Aldrich (Shanghai, China). All
other chemicals and solvents were of the highest commercial grade 2.3. Total monomeric anthocyanins
and were purchased from Anpel (Shanghai, China).
The total monomeric anthocyanins in PSPs extracts using differ-
2.2. Anthocyanins extraction ent extracted methods were determined by the pH differential
method (Giusti & Wrolstad, 2001). An L5S UV–visible spectropho-
2.2.1. Conventional solvent extraction tometer (Shanghai Analytical Instrument, China) was used to read
PSPs were cut into pieces after wash and prepared in a JJ-2 absorbance at 520 and 700 nm. Monomeric anthocyanins content
blender (Jintan, China) into slurry. Ten grams of PSPs slurry were were expressed as cyanidin-3-glucoside equivalents (CGE), using
placed into a 250 mL flask containing 100 mL aqueous ethanol a molecular weight of 449.2 mg/L and the molar absorptivity of
(70%, 80%, and 90%, v/v). The orthogonal array design L9 (34) was 26,900 L cm1 mg1. Cuvettes with 1-cm path length were used.
employed to study effects of temperature (60, 70 and 80 °C), Measurements were performed in triplicates.
extraction time (90, 120, and 150 min), aqueous ethanol (70%,
80%, and 90%, v/v), and HCl concentration (0.1%, v/v) in Table 1. 2.4. Total phenolics
Then, mixtures were centrifuged in L535R (Xiangyi, China) at
4000 rpm at 4 °C for 15 min and the supernatants were collected. Total phenolic content was measured using a modified
The solution was evaporated using a RE-52 rotary evaporator Folin–Ciocalteu method (Waterhouse, 2001). Briefly, a series of
(Yarong, China) and adjusted to 25 mL with corresponding extrac- tubes were prepared with 0.5 mL water or samples and 2.5 mL
tion solutions. The extracts were stored at 18 °C in the dark until Folin–Ciocalteu reagent for 5 min. After that, 2 mL of 75 g/L Na2CO3
analysis. All experiments were conducted in triplicates. solution was added to each test tube and mixed well before
incubation at room temperature for 2 h in dark. Then the
2.2.2. Ultrasound-assisted extraction absorbance of samples and standards was measured at 760 nm in
Ten grams of PSPs slurry were added into a 250-mL conical the L5S UV–visible spectrophotometer after zeroing the
bottles containing 100 mL of 80%, 90%, or 100% (v/v) aqueous spectrophotometer with a water blank. Total phenolic content
ethanol solutions. The bottles were then capped and put into a was calculated as gallic acid equivalents (GAE) based on a gallic acid
THC ultrasound extraction (Ji Ning Tianhua Ultrasonic Electronics, standard curve. Disposable cuvettes of 1-cm path length were used.
China). Four three-level factors including temperatures (40, 50, and Each sample was evaluated using three replications.
60 °C), duration (40, 50, and 60 min), ethanol concentration (80%,
90%, and 100%, v/v) and ultrasonic output power (200, 240, and 2.5. Total flavonoids
280 W) were studied via the Taguchi orthogonal array design L9
(34) in Table 1. Aqueous ethanol solution was acidified by adding The total flavonoids content of each PSP extracts was deter-
0.1% (v/v) HCl for all UAE experiments. Mixtures were centrifuged mined using a slightly modified method described previously
at 4000 rpm at 4 °C for 15 min after extraction. The supernatants (Meda, Lamien, Romito, Millogo, & Nacoulma, 2005). Briefly, 5 mL
were collected and adjusted to 25 mL after evaporation. All exper- of 2% aluminum trichloride (AlCl3) in methanol was mixed with
iments were conducted in triplicates. The extracts were stored at the same volume of a sample solution and let stand for 10 min.
18 °C in the dark until analysis. Then, the absorbance of samples or standards was measured at
415 nm in the L5S UV–visible spectrophotometer against a blank
2.2.3. Accelerated-solvent extraction sample consisting of a 5 mL methanol solution without AlCl3. The
The accelerated-solvent extractions was carried out with an ASE total flavonoids content was calculated using a standard curve
350 Accelerated Solvent Extractor system (Dionex, Sunnyvale, with quercetin (0–50 mg/L) as the standard. The mean of three
USA). Ten grams of PSPs slurry were added into a 100-mL Zirco- readings was used and expressed as mg of quercetin equivalents
nium extraction cell (Dionex, Sunnyvale, USA). The extraction cells (QE) per 100 g of PSPs.
were arranged in the sample carousel and extracted in a Taguchi L4
(23) design including three two-level factors in Table 1. Factors and 2.6. Antioxidant activity assays
levels have been chosen as a result of considering the cost of
conducting experiments and single factorial experimental results. 2.6.1. Oxygen radical absorbance capacity (ORAC) assay
The 80% (v/v) aqueous ethanol containing 0.1% (v/v) HCl was used The oxygen radical absorbance capacity of the extracts from
as the extraction solvent for all experiments. The supernatants PSPs was investigated according to previously reported procedure
were collected and adjusted to 25 mL after evaporation. Three (Jing et al., 2014) with a slight modification in an Infinite F200 Pro
trails were conducted for all experiments. The extracts were stored microplate reader (Tecan, Männedorf, Switzerland). Samples and
at 18 °C in the dark until analysis. Trolox standards were prepared with ethanol. All other reagents
268 Z. Cai et al. / Food Chemistry 197 (2016) 266–272

were prepared in 75 mmol/L phosphate buffer (pH 7.4). Briefly, column volumes of 0.01% HCl–methanol. The methanol was
each well in 96-well plate contained 25 lL sample or ethanol for removed by rotary evaporation at 40 °C and the residue was taken
blank and 150 lL, 4  106 mmol/L fluorescein solution. The plate up to about 1 ml with deionized water. The samples were stored at
with cover was incubated for 30 min in 37 °C, and then 25 lL, 18 °C for LC–MS and HPLC analysis.
0.153 mol/L AAPH were added to each well to start reaction. The The determination of high-resolution masses of anthocyanins in
fluorescence was recorded every minute for 120 min (excitation/ extracts was carried out in a Waters Acquity UPLC system (Waters
emission: 485/535 nm) at 37 °C. Trolox equivalents were calcu- Corporation, Milford, MA, USA) coupled to the Waters Synapt
lated using the relative area under the curve for samples compared HDMS Q-TOF MS detector. Five microliter injected on a COSMOSIL
to a Trolox standard curve prepared under the same experimental C18-MS-II column (150  3.0 mm, 5 lm), and the mobile phase
conditions. Results were expressed as Trolox equivalents (TE) per consisted of solvents A (formic acid: water = 1:100, v/v) and B
100 g of PSPs. All experiments were performed in triplicates. (formic acid: acetonitrile = 1:100, v/v), using the following
gradient: 5% B between 0 and 0.5 min, 5–20% B between 0.5 and
2.6.2. Ferric reducing antioxidant capacity (FRAP) assay 6 min, 20–35% B between 6 and 8.5 min, 35–50% B between 8.5
The FRAP method was carried out based on the reduction of and 9.5 min, 50–100% B between 9.5 and 11 min. Peak detection
Fe3+-TPTZ to a blue colored Fe2+-TPTZ with modification (Jing was carried out online by electrospray ionization in the positive
et al., 2012). Briefly, a portion of 10 mmol/L TPTZ was mixed with mode. The applied electrospray/ion optics parameters were
the same volume of 20 mmol/L FeCl3  6H2O, and 10 times higher set as follows: capillary voltage, 3.0 kV; sampling cone, 35 V;
volume of 300 mmol/L acetate buffer (pH 3.6). The mixture was collision energy, 4 eV; source temperature, 100 °C; desolvation
incubated at 37 °C for 30 min. Then, 3 mL of FRAP reagent, temperature, 300 °C; desolvation gas, 500 L/h. Spectra were
100 lL of sample or standards and 300 lL of distilled water were collected using full ion scan mode over the mass-to-charge (m/z)
added to the test tubes and incubated. The absorbance was range 200–2000 au. Scan time, 0.3 s; inter scan time, 0.02 s.
measured at 593 nm in the L5S UV–visible spectrophotometer. Quantitative analysis of anthocyanins were conducted using the
Trolox was used as standard for comparison and adequate dilution Finnigan Surveyor Plus system (Thermo Scientific, USA). Separation
of sample was performed. The results were reported as Trolox was achieved by reverse phase elution on a Shim-pack VP-ODS
equivalents (TE) per 100 g of PSPs. All experiments were conducted column (4.6 mm  250 mm, 5 lm, Shimadzu, Kyoto, Japan). The
for three times. chromatographic conditions were: flow rate 1 mL/min, sample
injection volume of 10 lL and mobile phase A (formic acid/water,
2.7. Qualitative and quantitative analysis of PSP anthocyanin-rich 1:100, v/v) and mobile phase B (formic acid/acetonitrile, 1:100,
extracts v/v). A gradient program was used as follows: 0–5 min, 15% B;
5–30 min, 15–20% B; 30–45 min, 20–40% B; 45–50 min, 40% B.
Anthocyanin-rich extracts (1 mL) were semi-purified using the Spectral information over the wavelength range of 200–800 nm
method of our previous study (Jing et al., 2012). About 1 mL of was collected. And the detection wavelength was set at 530 nm.
anthocyanins extract was loaded onto a C18 Sep-Pak solid cartridge Each extract was analyzed in triplicates.
(ANPEL, Shanghai, China), which was preconditioned with 2
column volumes of methanol and 3 column volumes of 0.01% 2.8. Experimental design and statistical analysis
HCl–water (v/v). Sugars and other polar compounds were removed
with 3-column volumes of 0.01% HCl–water, and anthocyanins The Taguchi orthogonal array design was applied to determine
and other phenolics were bound to the C18 cartridge. Finally, effects of extraction variables and their main effects. Four three-
anthocyanins were recovered from the cartridge with three level or three two-level factors were selected based on results of

Table 2
Total anthocyanin content of purple sweet potatoes using conventional extraction method (CE) at various levels of temperature, time, ethanol concentration and HCl ratio in
solvent.

Run # Temperature (°C) Time (min) Ethanol (%, v/v) HCl concentration (%, v/v) Monomeric anthocyaninsa (mg/100 g DW)
1 60 90 70 0.01 200.92 ± 15.09 cd
2 60 120 80 0.05 208.04 ± 7.24 cd
3 60 150 90 0.1 204.06 ± 6.40 cd
4 70 90 80 0.1 215.29 ± 6.88d
5 70 120 90 0.01 140.06 ± 9.78a
6 70 150 70 0.05 162.52 ± 1.81ab
7 80 90 90 0.05 179.79 ± 16.54bc
8 80 120 70 0.1 155.88 ± 5.55ab
9 80 150 80 0.01 165.66 ± 14.49ab
ANOVA analysis
SS 7267.250 4443.882 3036.794 2411.389
Df 2 2 2 2
F 33.313 20.371 13.921 11.054
P value 0.000⁄⁄ 0.000⁄⁄ 0.000⁄⁄ 0.001⁄⁄
Range analysis
Ki1 204.34 198.67 173.11 168.88
Ki2 172.62 167.99 196.33 183.45
Ki3 167.11 177.41 174.64 191.74
Rj 37.23 30.67 23.22 22.86

Kij shows the average response of each factor at different levels where i represent a factor and j a level. Rj = Ki.max–Ki.min for each factor.
SS: sum of squares.
Df: degree of freedom.
a
Means within a column for amount of monomeric anthocyanins followed by the same letter are not significantly different at P < 0.05 (Tukey HSD test).
Z. Cai et al. / Food Chemistry 197 (2016) 266–272 269

their single factorial assays and listed as Taguchi L9 (34) or L4 (23) Table 4
orthogonal design in Table 1, respectively. Three trails have been Yields of monomeric anthocyanins from PSPs extracts using accelerated-solvent
extraction (ASE) at two-level temperature, static time and static cycle.
conducted for each experiment. Statistical analysis was performed
with an analysis of variance (ANOVA) and a range analysis. Run # Temperature Static time Static Monomeric anthocyaninsa
The univariate ANOVA in General Linear Model was applied to (°C) (min) cycle (mg/100 g DW)

determine main effects using SPSS (version16.0, SPSS Inc., Chicago, 1 80 20 2 185.48 ± 2.99a
IL, USA). The Tukey HSD test was used to identify differences in 2 80 15 1 142.46 ± 1.83a
3 90 20 1 197.14 ± 8.33b
means. Tests were conducted in triplicate determinations with 4 90 15 2 252.34 ± 10.59c
data reported as mean ± standard deviation.
ANOVA analysis
Range analysis was used to confirm the effect of each factor and SS 11078.979 111.264 7235.376
determine the optimal level of different factors. The range analysis Df 1 1 1
was applied to parameter optimization. The average response for F 228.638 2.296 149.317
each factor was computed at each level, and labeled as Kij, where P value 0.000⁄⁄ 0.168 0.000⁄⁄
i represent a factor and j a level. The Rj represented the difference Range analysis
of Ki.max and Ki.min was calculated, where Ki.max and Ki.min were the Ki1 163.97 198.90 169.80
Ki2 224.74 191.31 218.91
largest and smallest values among the Kij for each factor, respec-
Rj 60.77 7.59 49.11
tively. According to the largest donating rule (Rao, Kumar,
Prakasham, & Hobbs, 2008), the largest Rj value indicates the most Kij shows the average response of each factor at different levels where i represent a
factor and j a level. Rj = Ki.max–Ki.min for each factor.
significant influence to the monomeric anthocyanins yields. The
SS: sum of squares.
optimal operation parameters were determined when the highest Df: degree of freedom.
monomeric anthocyanin yield among Ki1, Ki2, and Ki3 for each fac- a
Means within a column for amount of monomeric anthocyanins followed by the
tor was clearly distinguished. same letter are not significantly different at P < 0.05 (Tukey HSD test).

Anthocyanin yield from CE ranged from 140.06 to


3. Results and discussion
215.29 mg/100 g DW (Table 2). In CE, the extraction parameter con-
tributions on anthocyanin yield were temperature > time > ethanol
3.1. Anthocyanin extraction efficiency from PSP
concentration > HCl concentration (P < 0.001). Additionally, range
analysis was applied for extraction parameter optimization. Based
This study evaluated anthocyanin extraction efficiency from
on the Ri values, the significance of the extraction parameters in
PSPs using CE, UAE, and ASE. In CE, aqueous ethanol was used as
the range analysis was consistent with the ANOVA results;
a solvent to assess the effects of temperature, time, and HCl con-
therefore, temperature was the most significant factor, while HCl
centration on the anthocyanin extraction efficiency from PSPs.
concentration was the least significant one. The theoretical
Tables 2–4 show the Taguchi’s orthogonal array results, which
maximum anthocyanin yield at 60 °C, 90-min extraction time,
were subjected to ANOVA to determine the significance of main
80% (v/v) ethanol, and 0.1% HCl (v/v) had the highest Kij values
effects, followed by range analysis for optimizing the extraction
among all factor levels.
parameters of each method.
Anthocyanin yield from UAE ranged from 171.94 to
214.92 mg/100 g DW (Table 3). Based on the ANOVA results,
Table 3
Yields of monomeric anthocyanins from PSPs extracts using ultrasound-assisted
temperature, ethanol concentration, and power were significant
extraction method (UAE) at various levels of temperature, time, ethanol concentration
and power.

Run # Temperature Time Ethanol Power Monomeric Table 5


(°C) (min) (%, v/v) (w) anthocyaninsa Total amounts of anthocyanins, phenolics, and flavonoids, and antioxidant activities
(mg/100 g DW) using CE, UAE, and ASE at optimized conditions.
1 40 45 80 200 183.53 ± 10.99ab
Extraction methods
2 50 60 80 240 198.26 ± 10.62bcd
3 60 75 80 280 182.56 ± 4.59ab CE UAE ASE
4 60 60 90 200 214.92 ± 11.59d
Optimized parameters
5 40 75 90 240 190.53 ± 1.33abc
Temperature (°C) 60 50 90
6 50 45 90 280 204.30 ± 11.83bcd
Time/static time 90 45 15
7 50 75 100 200 209.37 ± 13.04 cd
(min)
8 60 45 100 240 205.63 ± 1.33bcd
Ethanol (%, v/v) 80 90 80
9 40 60 100 280 171.94 ± 1.57a
HCl (%, v/v) 0.1 0.1 0.1
ANOVA analysis Power (W) n.a. 200 n.a.
SS 2562.036 65.877 1030.588 1283.768 Static cycle n.a. n.a. 2
Df 2 2 2 2
Extractive efficiency
F 16.354 0.421 6.579 8.195
Monomeric 217.58 ± 2.90a 229.41 ± 4.59b 244.07 ± 11.84c
P value 0.000⁄⁄ 0.663 0.007⁄⁄ 0.003⁄⁄
anthocyaninsa
Range analysis Total phenolics b 955.05 ± 20.90a 769.65 ± 1.50b 630.67 ± 6.54c
Ki1 182.00 197.82 188.12 202.61 Total flavonoids c 40.54 ± 0.35a 32.90 ± 0.56b 27.81 ± 0.96c
Ki2 203.98 195.04 203.25 198.14 ORAC d 1091.06 ± 60.24a 1036.94 ± 110.80a 766.13 ± 114.24b
Ki3 201.04 194.15 195.65 186.27 FRAP d 1299.43 ± 2.32a 1303.14 ± 1.22a 1704.76 ± 6.38b
Rj 21.98 3.67 15.13 16.34
Values are represented as means ± SD (n = 3); within each row, means with the
Kij shows the average response of each factor at different levels where i represent a different letter are significantly different (P 6 0.05).
a
factor and j a level. Rj = Ki.max–Ki.min for each factor. Total monomeric anthocyanins was calculated as cyanidin-3-glucoside equiv-
SS: sum of squares. alents (mg CGE/100 g DW).
b
Df: degree of freedom. Total phenolics were calculated as gallic acid equivalents (mg GAE/100 g DW).
a c
Means within a column for amount of monomeric anthocyanins followed by the Total flavonoids was calculated as quercetin equivalents (mg QE/100 g DW).
d
same letter are not significantly different at P < 0.05 (Tukey HSD test). Antioxidant activities were calculated as trolox equivalents (mg TE/100 g DW).
270 Z. Cai et al. / Food Chemistry 197 (2016) 266–272

(P < 0.05), whereas time was not significant (P = 0.663). Addition- static cycle. The theoretical maximum anthocyanin yield at 90 °C,
ally, the ANOVA results revealed that the extraction parameter two static cycles, and 15-min static time had the highest Kij values
contributions on the response (i.e., anthocyanin yield) were among all factor levels.
temperature > power > ethanol concentration. The range analysis
results were consistent with the ANOVA results. The theoretical 3.2. Bioactive compound extraction efficiency from PSP
maximum anthocyanin yield at 50 °C, 45-min extraction time,
90% ethanol (v/v), and 200 W had the highest Kij values among CE, UAE, and ASE were performed under optimized extraction
all factor levels. parameters to assess the extraction efficiency of anthocyanins,
Anthocyanin yield from ASE ranged from 142.46 to total phenolics, and total flavonoids from PSPs (Table 5). Antho-
252.34 mg/100 g DW (Table 4). Based on the ANOVA results, the cyanin yield from the extraction methods was 217.58–244.07 mg
main effects were temperature and static cycle (P < 0.05), whereas cyanidin-3-glucoside equivalents (CGE)/100 g DW; an anthocyanin
static time was not significant (P = 0.168). Furthermore, tempera- yield of 0–663 mg/100 g DW (or 0–210 mg/100 g FW) has been
ture had a more significant contribution on anthocyanin yield than reported in different PSP genotypes using pressurized liquid

Table 6
Qualitative and quantitative analyses of anthocyanins in purple sweet potatoes extracts using different extraction methods.

Peak No. kmax M/Z Anthocyanins Peak area% of anthocyaninsA


CE UAE ASE
2 523 773, 449, 287 Cy-3-soph-5-glc 4.4 ± 0.0a 4.0 ± 0.1a 2.0 ± 0.0b
3 521 787, 463, 301 Pn-3-soph-5-glc 9.0 ± 0.3a 10.1 ± 0.1a 5.9 ± 0.1b
Total amount of nonacyl anthocyanins 13.4 ± 0.3a 14.1 ± 0.2a 7.9 ± 0.1b
4 523 893, 731, 449, 287 Cy-3-p-hydroxybenzoyl soph-5-glc 4.1 ± 0.0a 4.3 ± 0.1a 4.2 ± 0.1a
5 530 907, 745, 463, 301 Pn-3-p-hydroxybenzoyl soph-5-glc 1.0 ± 0.0a 0.8 ± 0.0a 0.5 ± 0.0b
6 523 949, 787, 449, 287 Cy-3-feruloyl soph-5-glc 2.9 ± 0.0a 1.2 ± 0.0b 0.8 ± 0.0c
7 519 935, 773, 449, 287 Cy-3-caffeoyl soph-5-glc 14.0 ± 0.0a 13.4 ± 0.0a 11.2 ± 0.1a
8 520 963, 801, 463, 301 Pn-3-feruloyl soph-5-glc 6.2 ± 0.1a 6.5 ± 0.1a 3.7 ± 0.0b
10 520 949, 787, 463, 301 Pn-3-caffeoyl soph-5-glc 1.7 ± 0.1a 1.9 ± 0.0a 2.4 ± 0.1b
Total amount of monoacyl anthocyanins 29.9 ± 0.2a 28.1 ± 0.1a 22.8 ± 0.3b
9 521 1055, 893, 449, 287 Cy-3-caffeoyl-p-hydroxybenzoyl soph-5-glc 9.7 ± 0.0a 10.6 ± 0.1a 15.8 ± 0.0b
11 523 1111, 949, 449, 287 Cy-3-caffeoyl-feruloyl soph-5-glc 1.3 ± 0.0a 1.4 ± 0.0a 1.4 ± 0.0a
12 521 1069, 907, 463, 301 Pn-3-caffeoyl-p-hydroxybenzoyl soph-5-glc 28.1 ± 0.0a 28.9 ± 0.0a 24.8 ± 0.0a
13 520 1125, 963, 463, 301 Pn-3-caffeoyl-feruloyl soph-5-glc 17.6 ± 0.0a 16.9 ± 0.0a 27.2 ± 0.0b
Total amount of diacyl anthocyanins 56.7 ± 0.0a 57.8 ± 0.1a 86.8 ± 0.0b

Abbreviation: cy, cyanidin; soph, sophoroside; glc, glucoside; pn, peonidin. Values are represented as means ± SD (n = 3). Within each line, means with the same letter are not
significantly different (P 6 0.05).
A
Calculation was based on percentage of peak areas in HPLC chromatogram.
Z. Cai et al. / Food Chemistry 197 (2016) 266–272 271

extraction (Truong et al., 2010). In this study, the concentration of 4. Conclusions


anthocyanins extracted from PSPs was higher than that reported in
previous studies (Montilla, Hillebrand, & Winterhalter, 2011; Extraction methods affect the yield, purity, and composition of
Rodriguez-Saona, Millar, & Trumble, 1998; Teow et al., 2007): anthocyanins. This study was aimed to compare CE, UAE, and ASE
32.2–53.1 mg CGE/100 g FW (or 107.32–176.98 mg CGE/100 g as to the effectiveness of extraction of anthocyanins from purple
DW; Teow et al., 2007) and 6.5–29.1 mg/100 g FW (Montilla sweet potatoes. While requiring high temp, ASE was a pressured-
et al., 2011). liquid and short-time extraction method and thus favored the
Total phenolic and flavonoid yield was 630.67–955.05 mg gallic extraction of anthocyanins that are heat labile from purple sweet
acid equivalents (GAE)/100 g and 27.81–40.54 mg quercetin equiv- potatoes. CE and UAE are long-time extraction methods at rela-
alents (QE)/100 g based on dry mass, respectively (Table 5). In this tively low temperature and resulted in low anthocyanin yields
study, the concentration of total phenolics was higher than that with high impurities of other phenolics in this study. As far as
obtained by Teow et al. (2007) by the CE method (21.2–42.2 mg the aspect of anthocyanin composition is concerned, ASE could
GAE/100 g FW, or 70.66–140.65 mg GAE/100 g DW). Differences extract more diacyl anthocyanins and less nonacyl and monoacyl
in bioactive components might be attributed to differences in anthocyanins than CE or UEA, suggesting anthocyanins in ASE
PSP cultivars or extraction assays. extracts were more stable than those extracted by CE or UAE. These
Interestingly, anthocyanin yield followed the order ASE > UAE > methods also affected antioxidant activities of PSPs extracts. When
CE, which was opposite to the total phenolic and flavonoid yield applied to food matrix as antioxidant agents, PSPs anthocyanin-
(CE > UAE > ASE; P < 0.05). A short-time extraction method at high rich extracts via the three methods demonstrated different activi-
temperatures (e.g., ASE) might favor the extraction of anthocyanins ties depending on the mechanisms involved. CE and UAE extracts
which are heat-labile, whereas long-time extraction methods were more efficient in breaking radical chain activity while ASE
at relatively low temperatures (e.g., CE and UAE) contribute to would be more suitable as an antioxidant via the mechanism of
low anthocyanin yields. Other thermo-stable phenolics could be ferric reduction. The anthocyanin extraction method from PSP
extracted from PSPs with long extraction durations. should take into consideration the extraction efficiency of bioactive
compounds, antioxidant activities, and types of anthocyanins.
3.3. Antioxidant activity of PSP extracts
Acknowledgements
Extraction conditions affect antioxidant activities of plant com-
pounds (Michiels, Kevers, Pincemail, Defraigne, & Dommes, 2012). This study was founded by National Nature Science Foundation
The ORAC value of CE extracts was 1091.06 mg trolox equivalents of China (Grant No. 31371756) and Hainan Provincial department
(TE)/100 g DW, similar to that of UAE extracts (1036.94 mg of Science and Technology (ZDXM2014008).
TE/100 g DW), but higher than that of ASE extracts (766.13 mg
TE/100 g DW; P = 0.01; Table 5). The ORAC values obtained in this References
study were higher than those obtained by Teow et al. (2007), who
reported 68.08–83.35 mg TE/100 g FW (or 226.71–277.56 mg Abdel-Aal, E. S. M., Akhtar, H., Rabalski, I., & Bryan, M. (2014). Accelerated,
TE/100 g DW) for sweet potato cultivars containing 70% water. microwave-assisted, and conventional solvent extraction methods affect
anthocyanin composition from colored grains. Journal of Food Science, 79(2),
However, the FRAP value of ASE extracts was 1704.76 mg C138–C146.
TE/100 g DW, which was significantly higher than that of CE and Carrera, C., Ruiz-Rodriguez, A., Palma, M., & Barroso, C. G. (2012). Ultrasound
UAE extracts (P < 0.01). PSP anthocyanins had higher FRAP values, assisted extraction of phenolic compounds from grapes. Analytica Chimica Acta,
732, 100–104.
whereas other flavonoids or phenolics had higher ORAC values. Fan, G. J., Han, Y. B., Gu, Z. X., & Chen, D. M. (2008). Optimizing conditions for
anthocyanins extraction from purple sweet potato using response surface
3.4. PSP anthocyanin composition methodology (RSM). LWT-Food Science and Technology, 41(1), 155–160.
Giusti, M. M., & Jing, P. (2007). Natural pigments of berry fruits: functionality and
application. In Y. Zhao (Ed.), Berry fruit: Value-added products for health
PSP extracts were subjected to HPLC, UV–visible, and LC–MS. A promotion (1st ed., pp. 105–146). Florida: Taylor & Francis.
total of 13 peaks of PSP anthocyanin-rich extracts had a maximum Giusti, M. M., & Wrolstad, R. E. (2003). Acylated anthocyanins from edible sources
and their applications in food systems. Biochemical Engineering Journal, 14,
absorbance at 519–530 nm. On the other hand, 27 anthocyanin
217–225.
peaks have been detected in purple-fleshed sweet potatoes grown Giusti, M. M., & Wrolstad, R. E. (2001). Characterization and measurement of
in Korea (Lee, Park, Choi, & Jung, 2013). Only 12 anthocyanins anthocyanins by UV–visible spectroscopy. In R. E. Wrolstad, T. E. Acree, E. A.
(Table 6) were identified by LC–MS based on fragmentation Decker, M. H. Penner, D. S. Reid, S. J. Schwartz, C. F. Shoemaker, D. Smith, & P.
Sporns (Eds.), Current protocols in food analytical chemistry (1st ed.,
patterns of individual peaks and past studies (Kim et al., 2012; pp. F1.2.1–F1.2.13). NY: John Wiley & Sons Inc.. Vol. 1.
Lee et al., 2013). Peonidin-3-sophoroside-5-glucoside and cyani Hwang, Y. P., Choi, J. H., Yun, H. J., Han, E. H., Kim, H. G., Kim, J. Y., ... Jeong, H. G.
din-3-sophoroside-5-glucoside represented 7.9–13.4% of the total (2011). Anthocyanins from purple sweet potato attenuate
dimethylnitrosamine-induced liver injury in rats by inducing Nrf2-mediated
anthocyanins and were identified as the basic structure of other antioxidant enzymes and reducing COX-2 and iNOS expression. Food and
acylated anthocyanins. Some acylated anthocyanins with one or Chemical Toxicology, 49(1), 93–99.
two p-hydroxybenzoic, ferulic and/or caffeic acid were monoacyl Jing, P., & Giusti, M. M. (2007). Effects of extraction conditions on improving the
yield and quality of an anthocyanin-rich purple corn (Zea mays L.) color extract.
(22.8–29.9%) and diacyl anthocyanins (56.7–86.8%), respectively. Journal of Food Science, 72(7), C363–C368.
The extraction methods affected the peak area percentage of Jing, P., Ye, T., Shi, H. M., Sheng, Y., Slavin, M., Gao, B. Y., ... Yu, L. L. (2012).
anthocyanins (Table 6). In the ASE extracts, the fraction of nonacyl, Antioxidant properties and phytochemical composition of China-grown
pomegranate seeds. Food Chemistry, 132(3), 1457–1464.
monoacyl, and diacyl anthocyanins in the 12 anthocyanin peaks Jing, P., Zhao, S. J., Ruan, S. Y., Sui, Z. Q., Chen, L. H., Jiang, L. L., & Qian, B. J. (2014).
were 7.9%, 22.8%, and 86.8% (peak area percentage), respectively. Quantitative studies on structure-ORAC relationships of anthocyanins from
ASE extracts contained more diacyl anthocyanins and less nonacyl eggplant and radish using 3D-QSAR. Food Chemistry, 145, 365–371.
Jing, P., Zhao, S. J., Ruan, S. Y., Xie, Z. H., Dong, Y., & Yu, L. L. (2012). Anthocyanin and
and monoacyl anthocyanins than CE or UEA extracts (P < 0.05),
glucosinolate occurrences in the roots of Chinese red radish (Raphanus sativus
which could be attributed to the thermal degradation effects of L.), and their stability to heat and pH. Food Chemistry, 133(4), 1569–1576.
CE and UAE on acyl anthocyanins (Abdel-Aal et al., 2014). ASE Kim, H. W., Kim, J. B., Cho, S. M., Chung, M. N., Lee, Y. M., Chu, S. M., ... Lee, D. J.
extracts were rich in acylated anthocyanins, which are more stable (2012). Anthocyanin changes in the Korean purple-fleshed sweet potato,
Shinzami, as affected by steaming and baking. Food Chemistry, 130(4), 966–972.
than those present in CE and UAE extracts (Giusti & Wrolstad, Lee, M. J., Park, J. S., Choi, D. S., & Jung, M. Y. (2013). Characterization and
2003). quantitation of anthocyanins in purple-fleshed sweet potatoes cultivated in
272 Z. Cai et al. / Food Chemistry 197 (2016) 266–272

Korea by HPLC-DAD and HPLC-ESI-QTOF-MS/MS. Journal of Agricultural and Food techniques for volatile (GC–MS and GC/FID) and phenolic compounds (HPLC-
Chemistry, 61(12), 3148–3158. ESI/MS/MS) from lamiaceae species. Phytochemical Analysis, 26(1), 61–71.
Lien, C. Y., Chan, C. F., Lai, Y. C., Huang, C. L., & Liao, W. C. (2012). Ultrasound-assisted Santos, D. T., Veggi, P. C., & Meireles, M. A. A. (2010). Extraction of antioxidant
anthocyanin extraction of purple sweet potato variety TNG73, Ipomoea batatas, compounds from Jabuticaba (Myrciaria cauliflora) skins: Yield, composition and
L.. Separation Science and Technology, 47(8), 1241–1247. economical evaluation. Journal of Food Engineering, 101(1), 23–31.
Mazza, G., & Miniati, E. (1993). Anthocyanins in fruits, vegetables and grains London. Teow, C. C., Truong, V. D., McFeeters, R. F., Thompson, R. L., Pecota, K. V., & Yencho, G.
CRC Press. C. (2007). Antioxidant activities, phenolic and beta-carotene contents of
Meda, A., Lamien, C. E., Romito, M., Millogo, J., & Nacoulma, O. G. (2005). sweet potato genotypes with varying flesh colours. Food Chemistry, 103(3),
Determination of the total phenolic, flavonoid and proline contents in Burkina 829–838.
Fasan honey, as well as their radical scavenging activity. Food Chemistry, 91(3), Truong, V. D., Deighton, N., Thompson, R. T., McFeeters, R. F., Dean, L. O., Pecota, K.
571–577. V., & Yencho, G. C. (2010). Characterization of anthocyanins and anthocyanidins
Michiels, J. A., Kevers, C., Pincemail, J., Defraigne, J. O., & Dommes, J. (2012). in purple-fleshed sweet potatoes by HPLC-DAD/ESI-MS/MS. Journal of
Extraction conditions can greatly influence antioxidant capacity assays in plant Agricultural and Food Chemistry, 58(1), 404–410.
food matrices. Food Chemistry, 130(4), 986–993. Truong, V. D., Hu, Z., Thompson, R. L., Yencho, G. C., & Pecota, K. V. (2012).
Montilla, E. C., Hillebrand, S., & Winterhalter, P. (2011). Anthocyanins in purple Pressurized liquid extraction and quantification of anthocyanins in
sweet potato (Ipomoea batatas L.) varieties. In J. A. T. D. Silva (Ed.), Fruit vegetable purple-fleshed sweet potato genotypes. Journal of Food Composition and
and cereal science and biotechnology (pp. 19–24). Global Science Books. Analysis, 26(1–2), 96–103.
Rao, R. S., Kumar, C. G., Prakasham, R. S., & Hobbs, P. J. (2008). The Taguchi Waterhouse, A. L. (2001). Determination of total phenolics. In R. E. Wrolstad, T. E.
methodology as a statistical tool for biotechnological applications: A critical Acree, H. An, E. A. Decker, M. H. Penner, D. S. Reid, S. J. Schwartz, C. F.
appraisal. Biotechnology Journal, 3, 510–523. Shoemaker, & P. Sporns (Eds.), Current protocols in food analytical chemistry (1st
Rastogi, N. K. (2011). Opportunities and challenges in application of ultrasound in ed., pp. I1.1.1–11.11.12). NY: John Wiley & Sons Inc..
food processing. Critical Reviews in Food Science and Nutrition, 51(8), 705–722. Wrolstad, R. E., & Culver, C. A. (2012). Alternatives to those artificial FD&C food
Rodriguez-Saona, C., Millar, J. G., & Trumble, J. T. (1998). Isolation, identification, colorants. Annual Review of Food Science and Technology, 3(1), 59–77.
and biological activity of isopersin, a new compound from avocado idioblast oil Zhang, Z. F., Fan, S. H., Zheng, Y. L., Lu, J., Wu, D. M., Shan, Q., & Hu, B. (2009). Purple
cells. Journal of Natural Products, 61(9), 1168–1170. sweet potato color attenuates oxidative stress and inflammatory response
Rodriguez-Solana, R., Salgado, J. M., Dominguez, J. M., & Cortes-Dieguez, S. (2015). induced by D-galactose in mouse liver. Food and Chemical Toxicology, 47(2),
Comparison of soxhlet, accelerated solvent and supercritical fluid extraction 496–501.

S-ar putea să vă placă și