You are on page 1of 8

Article

pubs.acs.org/JAFC

Anthocyanins Contents, Profiles, and Color Characteristics of Red


Cabbage Extracts from Different Cultivars and Maturity Stages
Neda Ahmadiani,† Rebecca J. Robbins,§ Thomas M. Collins,§ and M. Monica Giusti*,†

Department of Food Science and Technology, The Ohio State University, 2015 Fyffe Road, Columbus, Ohio 43210, United States
§
Analytical and Applied Sciences Group, Mars Inc., 800 High Street, Hackettstown, New Jersey 07840, United States

ABSTRACT: Red cabbage (Brassica oleracea L.) is an excellent source of food colorant. This study aimed to evaluate the
anthocyanin pigment contents and profiles from seven red cabbage cultivars at two maturity stages (8 weeks apart) and evaluate
their color characteristics and behavior under acidic and neutral pH. Anthocyanin concentrations ranged from 1111 to 1780 mg
Cy3G/100 g DM and did not increase with time. Cultivar and maturation affected pigment profile. Some varieties accumulated
≥30% of diacylated pigments, and proportions of monoacylated pigments decreased with time. Extracts from selected varieties at
first harvesting time produced colors similar (λmax = 520 nm and ΔE = 6.1−8.8) to FD&C Red No. 3 at pH 3.5. At pH 7, extracts
from the second harvest with s higher proportion of diacylation produced λmax ≃ 610 nm, similar to FD&C Blue No. 2. Cultivar
selection and maturation affected color and stability of red cabbage extracts at different pH values.
KEYWORDS: anthocyanins, red cabbage (Brassica oleracea L.), cultivar, harvesting time, color

■ INTRODUCTION
Anthocyanins are water-soluble pigments with potential
Knowing the anthocyanin composition of red cabbage cultivar
and maturation time would help us to select the cultivar and
maturation that could provide the desired characteristics for a
application in coloring of different food products.1,2 Colorants
specific application.


made of these pigments are currently manufactured for food
use from horticultural crops and processing wastes.3 Fruit and
MATERIALS AND METHODS
vegetable juices containing anthocyanins such as concentrated
red cabbage, black carrot, purple sweet potato, radish, bilberry, Plant Materials. Seven red cabbage cultivars, Cairo, Kosaro,
Integro, Buscaro, Azurro, Primero, and Bandolero (three heads from
and elderberry are being used as approved food color additives each cultivar), at two maturity stages (harvested 13 and 21 weeks after
in most countries.2,4 In addition, anthocyanins are proven to be transplanting) were donated by Bejo Seeds Inc. (Geneva, NY, USA).
good antioxidant compounds due to their effective free radical Cabbages were grown side-by-side during the summer season. Samples
scavenging properties and have shown numerous potential were shipped immediately after harvest and refrigerated until analyzed
health benefits in in vitro and vivo studies.5−9 (within a week). The water content in each sample was determined by
Red cabbage (Brassica oleracea L.) is an edible source with placing 4−5 g of sample in a mechanical convection incubator
high content and high potential yield per unit area of (Precision Scientific, Buffalo, NY, USA) at 37 °C for 2 days to dry.
anthocyanins.10 Red cabbage anthocyanin extract is known to Extraction and Purification. Cabbage heads were sliced, and ≃30
g was frozen with liquid nitrogen and kept frozen until analyzed the
have considerable amounts of mono- or diacylated cyanidin
following day. The frozen materials were ground using a stainless steel
anthocyanins.11,12 Type and acylation of anthocyanins are two Waring Commercial Blender (New Hartford, CT, USA) coupled with
important factors that determine their color characteristics at a 0.95 L container.27 The acetone/chloroform extraction procedure
certain pH values.13−17 Due to its anthocyanin compositions, was adopted from that of Giusti and Wrolstad.28 Frozen plant powder
red cabbage anthocyanin extracts can exhibit a wide spectrum was mixed with 30 mL of acetone. The mixture was filtered through a
of color, ranging from orange through red to purple and blue Whatman no. 1 filter (Whatman Inc., Florham, NJ, USA), and the
based upon the pH of the environment.18 Acylation of residual cake was washed with 70% aqueous acetone acidified with
anthocyanins also influences their antioxidant properties and 0.1% formic acid (≃250 mL) until the powder was white and the
filtrate was clear. The filtrates were combined, transferred to a
stability in the food matrix.15,19,20 Diacylated anthocyanins are
separatory funnel, and mixed with 1 volume of chloroform. The phases
linked to higher antioxidant activity compared to the other non- were allowed to separate for 4−5 h. The aqueous phase was collected,
and monoacylated ones.12 Anthocyanin pigments with higher and the residual acetone was evaporated using a Büchi rotavapor
number of acylation have also shown good stabilities to light (Brinkmann Instruments, Inc., Westbury, NY, USA). The aqueous
and processing temperature.20 extract was purified using a Sep-Pak C18 cartridge (Waters Corp.,
There are several known intrinsic and extrinsic factors that Milford, MA, USA). The cartridge was activated with methanol and
affect anthocyanin content and composition in plants.21−23 washed with acidified water before the sample was loaded. The
Plants cultivar and maturation time are among the essential cartridge was further washed with acidified water (0.1% formic acid),
factors that can influence the phytochemical content, including
anthocyanins.24−26 The objectives of this study were to evaluate Received: January 15, 2014
the anthocyanins content and profile from different red cabbage Revised: July 2, 2014
cultivars at two maturity stages and evaluate their color Accepted: July 3, 2014
characteristics and behavior under acidic and neutral pH. Published: July 3, 2014

© 2014 American Chemical Society 7524 dx.doi.org/10.1021/jf501991q | J. Agric. Food Chem. 2014, 62, 7524−7531
Journal of Agricultural and Food Chemistry Article

Figure 1. HPLC chromatograms of two representative red cabbage anthocyanin extracts: second-harvested Cairo (top) and first-harvested Integro
(bottom) red cabbage at 510−540 nm. Refer to Table 1 for major peak identifications. HPLC conditions: solvent A, 4.5% formic acid in LC-MS
grade water; solvent B, LC-MS acetonitrile; gradient, 0−50 min, 0−30% B. The major peaks (1−8) were found in all other red cabbage samples.

and the anthocyanins were recovered with acidified methanol (0.1% monitored at 510−540 nm. Peak areas at this region were then
formic acid). The methanol was removed using the rotavapor, and the integrated and normalized. The proportion of the total peak area of
volume was taken to 25 mL with acidified water (0.1% formic acid). each individual anthocyanin was calculated and reported as percentage
Anthocyanins Content. The total monomeric anthocyanin was of total peak area at 510−540 nm.
determined by using the pH differential method according to Giusti For MS analyses, a 0.2 mL/min volume was diverted into the MS
and Wrolstad.29 The extract was diluted using pH 1 (0.025 M and ionized under positive ion condition using an electrospray probe.
potassium chloride) and pH 4.5 (0.4 M sodium acetated) buffers with Data were monitored using total ion scan (SCAN) (m/z 200−1200)
a dilution factor of 100. The solutions were allowed to equilibrate for and selected ion monitoring at m/z 271 (pelargonidin), m/z 287
15 min in the dark. Absorbance was read on 1 cm path length cuvettes (cyanidin), m/z 301 (peonidin), m/z 303 (delphinidin), m/z 317
at 520 and 700 nm using a Shimadzu UV−visible spectrophotometer (petunidin), and m/z 331 (malvidin).
(Shimadzu, Columbia, MD, USA). The total monomeric anthocyanin Color and Spectrophotometric Analyses. A ColorQuest XE
was calculated on the basis of the dry matter (DM) and fresh matter colorimeter (HunterLab, Hunter Associates Laboratories Inc., Reston,
(FM) and reported as milligrams of cyanidin-3-glucoside (Cy3G) per VA, USA) was used to measure the color characteristics (Hunter CIE
100 g of sample using the equation LCh) of the samples. The equipment was set for transmittance with
specular included, and D65/10° was used for the measurements.
total monomeric anthocyanin (mg/L) Samples were placed in a 1 cm path length plastic cuvette and CIE L*,
a*, b*, chroma (c*), and hue angle (h°) were measured.
= [((A520 − A 700)pH1 − (A520 − A 700)pH4.5 ) × DF To measure the color in acidic condition, the extract was mixed (in
× 1000 × MW]/(ε × P) triplicate) with distilled (DI) water (1:20 v/v). The pH of the
solutions were measured after 30 min of equilibration and was 3.5. To
where DF is the dilution factor, MW is the molecular weight (449.2 for measure the color in neutral conditions, the extract was diluted with
Cy3G), ε is the molar absorptivity coefficient (26900 cm−1 mg−1 for water (1:2 v/v, in triplicate). The diluted solutions were then mixed
Cy3G), and P is the cuvette path length. with 0.1 M potassium phosphate buffer (pH 7) with a dilution factor
Alkaline Hydrolysis of Anthocyanins. Alkaline hydrolysis (DF) of 20. The maximum absorbances (λmax) in the visible range of
(saponification) was adopted from the method of Giusti and the neutral solutions were recorded using a Shimadzu UV−visible
Wrolstad.28 Purified red cabbage anthocyanin extract (10 mL) was spectrophotometer 2450.
mixed with 10 mL of 10% KOH in a capped test tube and set aside for FD&C Red No. 3, FD&C No. 40, and FD&C Blue No. 2 (Noveon
15 min at room temperature in the dark. The solution was neutralized Hilton Davis, Inc., Cincinnati, OH, USA) were also dissolved in DI
using 2 N HCl until the color turned pink. The neutralized sample was water to concentrations that most closely matched the lightness (L*)
then purified using a Sep-Pak C18 cartridge (Waters Corp., Milford, and chroma (c*) of the samples.
MA, USA) and prepared for HPLC analysis. Statistical Analysis. Seven different cultivars at two maturity
Chromatographic Analysis. A Shimadzu Prominence reverse stages (three heads each) for a total of 42 samples were analyzed using
phase high-pressure LC-MS was coupled to an SPD-M20-A photo- principal component analysis (PCA). PCAs for total monomeric
diode array and a single-quadrupole electrospray ionization (ESI) mass anthocyanin, nonacylated, monoacylated, and diacylated pigments
spectrophotometer (Shimadzu Scientific, Inc.). For data analysis, LC- (based on the proportional peak areas shown in Table 2) were
MS Solution software was used (Shimadzu Scientific, Inc.). The performed. Autoscaling was used to normalize each variable before the
column was a 100 × 4.5 mm Kinetex PFP 2.6 μm (Phenomenex Inc., analysis.
Torrance, CA, USA). The solvents were phase A, 4.5% formic acid in To compare the anthocyanin contents and profile, normality of the
LC-MS grade water, and phase B, LC-MS acetonitrile (Fisher variables were first checked using the Kolmogorov−Smirnov test (α =
Scientific Inc., Fair Lawn, NJ, USA), and the gradient was 0−50 0.05). Analysis of variance (ANOVA) was then used to analyze the
min, 0−30% B. Injection volume was 20 μL. Spectral data were total monomeric anthocyanin and the proportion of each group of
obtained from 250 to 700 nm, and elution of anthocyanins was pigments separately using the following model: Yijk = μ + vi + tj + vtij +

7525 dx.doi.org/10.1021/jf501991q | J. Agric. Food Chem. 2014, 62, 7524−7531


Journal of Agricultural and Food Chemistry Article

αk + εijk, where Y is the individual variable, μ is the grand mean, vi is


the cultivar effect, tj is the harvesting time effect, vtij is the interaction
between the main factors, αk is the heads effect defined as random
factor, and εijk is the random error of the model. When a significant
difference was obtained (P value < 0.05), the Tukey means-
comparison test was used to compare each pair of means.
All statistical analyses were done on the basis of at least three
independent replicate samples from each individual head. Results were
analyzed by using Minitab 16 statistical software (Minitab Inc.).

■ RESULTS AND DISCUSSION


HPLC-PDA-MS and Identification of Major Pigments.
According to the HPLC-PDA data obtained at 510−540 nm,
up to 23 peaks were observed. In previous studies, up to 36
different anthocyanins have been detected in red cabbage.29−31
Figure 1 shows examples of anthocyanin profiles (Cairo and
Integro extracts) obtained by HPLC-PDA. The eight major
peaks, representing ≃90% of the total anthocyanins and Figure 2. Correlation between the first two principal components and
common to all seven cultivars at both maturity stages, were the variables as well as score plot with respect to cultivars and
selected for further identification and analyses. The λmax at the harvesting time for seven different red cabbage cultivars: a, total
monomeric anthocyanin; b, diacylated pigments; c, monoacylated
UV and vis ranges, molecular ions, and fragments along with
pigments; d, nonacylated pigments. AZ, Azurro; Ba, Bandolero; Bu,
tentative identification of each peak are presented in Table 1. Buscaro; Ca, Cairo; In, Integro; Ko, Kosaro; Pr, Primero. Symbols in
The pigments were identified using HPLC-PDA and HPLC- black and gray represent harvest times after 13 or 21 weeks,
MS and compared with data reported in the literature.11,12 respectively.

Table 1. PDA Absorbance and MS Data for Red Cabbage


Anthocyaninsa According to this analysis, the samples were clearly separated
diagonally on the basis of their maturation time. With
RTb λvisc λacyld maturation the percentage of diacylation increased for most
peak (min) (nm) (nm) M+ e identification
samples, so for most of the second-harvested samples PC1 is
1 12.13 513 773 (287) Cy-3diG-5G negative. Also, because the early mature samples (week 13) had
2 16.95 528 334 979 (287) Cy-3diG-5G + sinapicf slightly higher amounts of total monomeric anthocyanin, PC2
3 27.41 523 313 919 (287) Cy-3diG-5G + p- tend to be more positive (Figure 2).
coumaric
4 28.33 523 326 949 (287) Cy-3diG-5G + ferulic
PCA is a proper way to obtain relevant information from the
5 28.85 524 329 979 (287) Cy-3diG-5G + sinapic
original variables into fewer new latent variables (PCs).
6 30.53 536 319 1125 (287) Cy-3diG-5G + ferulic
According to our results, this analysis was helpful in
and ferulic classification of the samples at two maturity stages. Most
7 31.55 536 330 1155 (287) Cy-3diG-5G + sinapic samples with different maturity times were separated on the
and ferulic basis of the first two principal components. Azurro and Primero
8 32.31 536 331 1185 (287) Cy-3diG-5G + sinapic cultivars, however, were exceptions because their pigment
and sinapic
a
profiles were not significantly changed with maturation.
m/z 287 was the major fragment in all eight peaks. Cy-3diG-5G, Anthocyanin Content. The average anthocyanin contents
cyanidin-3-diglucoside-5-glucoside. bRetention time. cλ vis−max. dλ of for the 13- and 21-week-harvested plants were ≃1442 and 1269
acylation. eMass ion. fTentative identification.
mg Cy3G/100 g DM respectively, and these values for the fresh
matter were ≃150 and 145 mg Cy3G/100 g FM, respectively.
As shown in Table 1, m/z 287 was the fragment in all eight Piccaglia et al. also reported the anthocyanin content of three
anthocyanins indicating that cyanidin derivatives were the red cabbage cultivars in Italy to be >1000 mg/100 g DM.10 The
major aglycon as reported in previous studies.12,30,32−34 All of anthocyanin content for the fresh weight red cabbage reported
the pigments were nonacylated, monoacylated, and diacylated by Oregon State University database and Wu et al. were,
derivatives of cyanidin-3-diglucoside-5-glucoside (Cy-3diG- however, 25 and 322 ± 40.8 mg/100 g FW, respectively.35,36
5G), which was also confirmed by saponification (results are According to our findings, cultivar made a difference in
not shown). The acylating groups were aromatic acids: sinapic, anthocyanin contents at both maturity stages.
ferulic, and p-coumaric acids (Table 1). As shown in Table 2, Buscaro and Integro harvested after 13
Anthocyanin Content and Proportion of the Pig- weeks had the highest anthocyanin contents. Anthocyanin
ments. Principal Component Analysis of Red Cabbage contents, however, did not change significantly from the first to
Anthocyanin Extracts. The analysis of the objects (i.e., red the second harvest except for the Buscaro (DM) cultivar, which
cabbage cultivars at two maturity stages) is performed visually showed a lower anthocyanin content when plants were left
using the scores plot (Figure 2), where the objects are longer on the ground (Table 2). Accumulation of anthocyanins
represented as a function of the principal components (PCs). can be explained by developmental factors, which could be
As shown in Figure 2, PC1 correlated positively with different in different varieties.23
monoacylated pigments as opposed to the diacylated pigments. Proportion of Major Pigments. Although red cabbage is
PC2, on the other hand, was more affected by the nonacylated known to have more than 20 different anthocyanins, relative
pigments and the total monomeric anthocyanin. PC1 and PC2 proportions of 8 of them (Figure 1 and Table 1) represented
extracted 83.7 and 12.8% of the total variances, respectively. ≃90% of the total anthocyanins and were measured and
7526 dx.doi.org/10.1021/jf501991q | J. Agric. Food Chem. 2014, 62, 7524−7531
Journal of Agricultural and Food Chemistry Article

Table 2. Anthocyanin Contents (Total Monomeric Anthocyanin) and Percentage of Major Pigments (Percent Total Peak Area
at 510−540 nm) in Seven Red Cabbage Cultivars at Two Different Harvesting Timesa
total monomeric
anthocyanin
(mg Cy3G/100 g)
cultivar harvest time (week) DMb FMc nonacylated pigmentsd (%) monoacylated pigmentse (%) diacylated pigmentsf (%)
Primero 13 1111 e 109 gh 21.24 bc 67.63 a 4.56 g
21 1026 e 104 h 26.84 a 63.71 ab 4.09 g

Integro 13 1660 ab 185 a 18.17 cde 65.22 ab 7.23 fg


21 1637 ab 188 a 25.59 ab 53.32 cde 12.28 ef

Azurro 13 1392 bcd 144 bcdef 15.85 def 69.33 a 9.19 fg


21 1217 de 137 cdef 19.33 cde 65.01 ab 9.21 fg

Kosaro 13 1247 cde 128 efgh 15.51 ef 55.43 cd 11.42 f


21 1001 e 115 fgh 20.25 cd 46.74 ef 19.41 cd

Cairo 13 1389 bcd 153 bcde 20.08 cde 58.23 bc 12.87 ef


21 1256 cde 168 ab 28.5 a 43.96 fg 17.74 de

Bandolero 13 1517 abc 165 abcd 13.1 f 48.67 def 24.54 bc


21 1512 abc 165 abcd 19.07 cde 36.75 gh 29.1 b

Buscaro 13 1780 a 170 abcd 13.36 f 52.1 cde 23.59 bcd


21 1236 de 137 defg 17.04 cdef 35.48 h 35.45 a
a
Different letters in the same column indicate significant differences (p < 0.05). bDry matter. cFresh matter. dPeak 1 (Table 1). ePeaks 2, 3, 4, and 5
(Table 1). fPeak 6, 7, and 8 (Table 1).

Table 3. CIE L*a*b*, Chroma (c*), Hue (h°), and λmax of Seven Red Cabbage Cultivar Extracts at Two Different Harvesting
Times at pH 3.5 and Their Comparison with Two FD&C Synthetic Red Dyes
cultivar HTa (week) L* a* b* c* h° λmax (nm) ΔE1b ΔE2c
Primero 13 76.2 ± 0.8 45.3 ± 0.6 −3.5 ± 0.7 47.4 ± 3.5 359.2 ± 4.4 520 ± 1.2 15.9 ± 2 18 ± 0.7
21 77.1 ± 2 42.1 ± 4 −2.7 ± 0.7 42.2 ± 4 356.3 ± 0.9 520 ± 1.4 18.5 ± 4.2 17.9 ± 1

Integro 13 67.4 ± 2.6 59.5 ± 1.7 −0.5 ± 0.3 59.6 ± 4.6 357.5 ± 2.6 520 ± 0.8 8.8 ± 3.1 23.4 ± 1.3
21 66.2 ± 1.8 61 ± 2.5 −1.9 ± 1.2 61 ± 2.5 358.2 ± 1.2 520 ± 1.1 9.3 ± 2.2 24.8 ± 1.4

Azurro 13 69.6 ± 3 56.1 ± 1 −2.8 ± 2 56.6 ± 5.3 357.7 ± 2.7 520 ± 0.6 6.8 ± 2.9 20.7 ± 2
21 71.2 ± 2.9 53.2 ± 4.7 −2.9 ± 1.4 53.3 ± 4.6 356.7 ± 1.9 520 ± 0.5 9.1 ± 2.2 20.2 ± 1.4

Kosaro 13 69 ± 3.2 57 ± 3.5 −5.9 ± 1.2 57.3 ± 5 354.1 ± 2.6 520 ± 0.3 6.1 ± 2.7 23.7 ± 3.7
21 69.5 ± 1.5 55.5 ± 2.4 −7.1 ± 0.9 56 ± 2.4 352.7 ± 1 520 ± 0.8 7 ± 0.8 24.8 ± 1.6

Cairo 13 68.9 ± 3.5 57.1 ± 1.6 −4.5 ± 0.6 55.9 ± 3.6 357 ± 1.4 520 ± 1.4 6.6 ± 3.2 23 ± 1.3
21 64.9 ± 1.5 63 ± 2 −4.3 ± 0.4 63.2 ± 2 356.1 ± 0.5 520 ± 0.6 10.1 ± 2 28.1 ± 1.6

Bandolero 13 62.5 ± 1.6 67.3 ± 4.3 −4.7 ± 1.3 68.4 ± 1.5 356.8 ± 4.4 520 ± 0.8 14.8 ± 3 32 ± 1.5
21 60.7 ± 1.8 68.2 ± 2.1 −4.8 ± 1.5 68.4 ± 2 356 ± 1.4 530 ± 1.2 15.9 ± 2.7 33.6 ± 1.3

Buscaro 13 62.9 ± 4.1 66.9 ± 2.3 −3.7 ± 1 64.5 ± 1.2 357.9 ± 1.4 520 ± 0.9 12.8 ± 3 29.9 ± 4.4
21 61.7 ± 4.2 66 ± 4.5 −6.4 ± 2.9 66.4 ± 4.2 354.4 ± 2.8 530 ± 1.5 14.2 ± 5.7 33 ± 3

FD&C Red No. 3 74.2 59.99 −6.1 60.3 354.19 520


FD&C Red No. 40 75.64 45.11 14.48 47.38 17.79 500
a
Harvesting time. bColor difference with FD&C Red No. 3. cColor difference with FD&C Red No. 40.

compared among cultivars and maturation stages as they are According to our results, nonacylated pigments represented
more likely to affect the color and stability of the extract. The an average of ≃19.6% (±4.5) of the major pigments (Table 2).
Charron et al. also found the percentage of nonacylated
pigments were also grouped into nonacylated, monoacylated, pigment in red cabbage to be 21.3%.31 Among the major
and diacylated. pigments identified, the amounts of mono- and diacylated
7527 dx.doi.org/10.1021/jf501991q | J. Agric. Food Chem. 2014, 62, 7524−7531
Journal of Agricultural and Food Chemistry Article

anthocyanins were on average 54.4 (±12.6) and 15.8% (±8.7), Bandolero and Buscaro extracts as opposed to Primero. The
respectively (Table 2). Red cabbage has been identified as a wavelengths of maximum absorbances (λmax) of the samples
highly acylated anthocyanin source according to previous were similar to that of Red No. 3 (≃520 nm) except for late-
research.30,31 harvested Bandolero and Buscaro cultivars (λmax ≃ 530 nm).
Before the percentages of each group of pigments were Generally, samples showed more similar color characteristics to
compared using ANOVA, the normality test was done to verify FD&C Red No. 3 than to FD&C Red No. 40 synthetic dyes
the validity of the test, and none of the variables had a p value (ΔE1 < ΔE2). Under acidic pH, maturation of the plants did
smaller than 0.05. not seem to significantly affect the color of the solutions. Early-
Cultivar had a significant (p value < 0.05) impact on pigment harvested Azurro, Kosaro, Cairo, and Integro at the tested
profile. The proportions of monoacylated and diacylated concentrations were the samples that produced colors most
pigments were the major differences found among the cultivars. similar to FD&C Red No. 3 with similar λmax and ΔE between
Primero and Azurro had the highest proportions of 6.1 and 8.8 (Table 3).
monoacylated pigments and the lowest proportions of Stability in Neutral pH. The spectral characteristics of the
diacylated pigments, whereas Buscaro and Bandolero varieties samples at pH 7 were monitored over 72 h of refrigerated
had the lowest proportions of monoacyltaed pigments and the storage. Figure 3 shows the changes in λmax for three
highest proportion of diacylated pigments. The effect of
maturity on pigment profile was largely dependent on the
cultivar for monoacylated and diacylated pigments (p value <
0.05); this interaction was not observed for the nonacylated
pigments.
As shown in Table 2, the proportion of nonacylated
pigments increased significantly in the 21-week-mature plants
except for Azurro and Buscaro cultivars, which did not show a
significant increase. More mature Primero, Cairo, and Integro
had highest amounts of nonacylated pigments.
The proportions of monoacylated anthocyanins were,
however, decreased with longer maturation time, except for
Primero and Azurro cultivars, which did not show a significant
decrease on monoacylated pigments. The proportion of Figure 3. Change in λmax for the second-harvested Buscaro, Bandolero,
monoacylated pigments was highest (≃67%) for cultivar and Primero at pH 7 over 72 h of refrigeration. Most color changes
happened during the first 6 h of storage.
Primero, Integro, and Azurro first harvests as compared to
the other cultivars (≃54%) harvested at the same time (Table
2). representative samples. Solutions colored with Buscaro and
The proportion of diacylated pigments varied remarkably Bandolero red cabbage extracts showed the highest variations in
among cultivars. They increased significantly in Kosaro and the λmax, whereas solutions colored with Primero extract
Buscaro due to maturation. Buscaro and Bandolero, harvested showed the least changes over the 72 h. Most of the changes
after 21 weeks, had the highest (≃30%) proportion of observed happened during the first 5−6 h (Figure 3). Table 4
diacylated pigments, whereas Primero (≃4%) had the lowest shows the λmax of all seven varieties 30 min after the extracts
at both maturity stages (Table 2). were mixed with buffer pH 7 and 6 h afterward (refrigerated
Because the plants were grown side by side, the main reason storage). The λmax for samples containing significantly higher
for the profile differences could arise from intrinsic factors such amounts of diacylated anthocyanins (Buscaro and Bandolero)
as plant genetics and enzymes and their activities throughout at this pH seemed to be the highest (≃600 nm) compared to
the maturation, which influence the anthocyanin synthesis the other samples. Torskangerpoll and Andersen also
within the plant. Variation in anthocyanin structures can be investigated the absorbance and color change in cyanidin 3-
correlated with alteration of single genes, which influence the (2″-(2‴-sinapoylglucosyl)-6″-sinapoylglucoside)-5-glucoside (a
enzymatic step of anthocyanin synthesis pathways.37 The diacylated pigment isolated from red cabbage) at different pH
reason for accumulation of certain anthocyanins in certain values. They also found out that at pH 7.2 the pigment had λmax
cultivars could be due to the difference in the activities of of ≃605 nm.38 A bathochromic effect of up to 7 nm was
different genes. Also, during the maturation, the activity of the observed for most samples after 6 h of refrigerated storage.
genes controlling the synthesis of monoacylated pigments More mature Buscaro and Bandolero varieties produced a λmax
could have slowed, whereas those responsible for the synthesis (≃610 nm) similar to that of FD&C Blue No. 2. For other
of nonacylated anthocyanins or for the addition of a second samples such as Primero and Integro, however, minute
acyl group may have remained active. bathochromic effects were observed (Table 4).
Color Characteristics and Stability. Color in Acidic pH. Anthocyanins tend to have lower stabilities at neutral to
Table 3 shows the color characteristics of solutions colored alkaline pH values.14,39 After mixing with buffer pH 7 followed
with red cabbage anthocyanin extracts from the different by 6 h of refrigeration, the color degradation was highest
cultivars at pH 3.5 and their color comparisons to synthetic (≃53%) and lowest (≃19%) for the second-harvested Primero
colorants FD&C Red No. 3 and No. 40. The lightness of the and Buscaro, respectively (Table 4). Higher stabilities of the
samples was approximately between 60 and 77 (Table 3). The Buscaro and Bandolero cultivars at the tested pH values could
synthetic colorant solutions were also adjusted to have a similar be explained by the larger number of diacylated anthocyanins,
lightness. Extracts from all seven cultivars produced a deep pink which have higher stabilities due to the pigments intramolecular
color at pH 3.5 with a hue angle close to 350°. Chroma values and/or intermolecular copigmentation and self-association
of the samples were higher for solutions prepared with reactions.15 Torskangerpoll and Andersen also demonstrated
7528 dx.doi.org/10.1021/jf501991q | J. Agric. Food Chem. 2014, 62, 7524−7531
Journal of Agricultural and Food Chemistry Article

Table 4. Maximum Absorbance (λmax) of Red Cabbage that diacylation of anthocyainins increases their color
Extracts at Two Different Harvesting Times Measured after stabilities.38
30 min and 6 h of Refrigeration Storage in Buffer pH 7 The color characteristics of pH 7 buffer solutions colored
Compared to FD&C Synthetic Blue No. 2 and Stabilities with the different red cabbage extracts after 6 h of refrigeration
(Percent Degradation) during This Timea compared to FD&C Blue No. 2 are shown in Table 5. The
lightness of the samples ranged between ≃60 and 85. Chroma
λmax (nm)
values were higher for Bandolero and Buscaro cultivars at the
cultivar HTb (weeks) 30 min 6h % degradation tested pH value (Table 5). The synthetic FD&C Blue No. 2
Primero 13 591.4 ± 0.6 591.6 ± 0.8 50.1 ± 5.5 was diluted until the lightness was similar to that of the
21 591.3 ± 0.4 591.9 ± 1 53.3 ± 8.9 evaluated samples. ΔE values of the samples when compared to
this synthetic dye were between 17.6 and 26.4 (Table 5). The
Integro 13 591.4 ± 1.1 592.1 ± 0.1 38.9 ± 5.4 color difference with FD&C Blue No. 2 at the tested
21 598.6 ± 2.3 599.7 ± 2 44 ± 9.9 concentration for the late-harvested Kosaro sample was the
smallest; however, the λmax values for the second-harvested
Azurro 13 590.8 ± 0.8 593.5 ± 0.5 41.5 ± 4.9 Buscaro and Bandolero cultivars were closer to this value for
21 595.9 ± 0.7 597.8 ± 2.1 49.3 ± 7.9 FD&C Blue No. 2 (Tables 4 and 5). Acylation of anthocyanins
with aromatic acids (e.g., cinnamic acid) has proven to increase
Kosaro 13 596.3 ± 1 599.9 ± 1.4 31 ± 3.3
the λmax and shift the hue angle to purple color under acidic
21 602 ± 0.3 604.8 ± 0.8 39.5 ± 5.2
conditions.16,17 As shown in Table 1, late-harvested Buscaro
and Bandolero cultivars, with the highest percentages of
Cairo 13 593.6 ± 0.2 595.3 ± 4.2 42.1 ± 8.4
diacylated pigments, also exhibited the highest λmax under
21 599.7 ± 2.4 602.9 ± 0.5 36.7 ± 8.2
neutral conditions.
In conclusion, anthocyanin content varied among red
Bandolero 13 601.8 ± 1.2 609.4 ± 1.2 27.1 ± 4
21 599.7 ± 2.3 610.1 ± 1.6 22.5 ± 3.8
cabbage cultivars, and leaving the cabbages in the ground for
additional time did not increase the pigment content. The
Buscaro 13 600.1 ± 0.7 606.7 ± 1.9 28.6 ± 2.3
pigment profiles changed among the cultivars and maturation
21 601.8 ± 0.4 610.1 ± 1.8 19.1 ± 3.8 stages. For most cultivars, the amount of nonacylated and
diacylated pigments increased as opposed to the monoacylated
FD&C Blue No. 2 610 610 pigments. Cultivars with lower proportions of diacylated
pigments better reproduced the color of FD&C Red No. 3
a
Percent degradation was calculated by dividing the absorbance at λmax
after 6 h by the absorbance at λmax after 30 min in buffer pH 7 × 100. under acidic conditions, whereas cultivars with higher
b
Harvesting time. proportions of diacylated pigments better matched the colors
of FD&C Blue No. 2 at neutral pH. Pigment profile also
affected the color stability at neutral pH. For future studies,

Table 5. CIE L*a*b*, Chroma (c*), and Hue (h°) of Seven Red Cabbage Extracts at Two Different Harvesting Times after 6 h
of Refrigeration Storage at pH 7 Compared to FD&C Blue No. 2
cultivar HTa (weeks) L* a* b* c* h° ΔEb
Primero 13 83.6 ± 0.6 2.6 ± 0.2 −9.9 ± 0.6 10.2 ± 2.4 284.6 ± 1.5 25.7 ± 0.5
21 84.8 ± 2 2.1 ± 0.6 −8.7 ± 2.2 9 ± 2.2 284.5 ± 5.1 26.4 ± 2

Integro 13 70.5 ± 2.9 1.7 ± 2.2 −22.4 ± 2.8 22.5 ± 2.7 274.7 ± 5.8 21.4 ± 1
21 69.9 ± 3.2 −0.2 ± 0.6 −23.5 ± 2.6 23.5 ± 2.6 269.5 ± 1.3 19.8 ± 0.7

Azurro 13 77.7 ± 1.5 1.4 ± 0.2 −15.8 ± 1.5 15.9 ± 1.4 275.3 ± 1.3 21.2 ± 1.3
21 78.5 ± 3.7 1.2 ± 0.4 −15.2 ± 3.1 15.3 ± 3.1 274.7 ± 1.9 21.5 ± 1.3

Kosaro 13 72.1 ± 1.5 −1.3 ± 0.7 −22.1 ± 1.9 22.1 ± 2.4 266.7 ± 1.3 17.9 ± 2.3
21 71.9 ± 2.5 −1.8 ± 1.8 −22.5 ± 2 22.6 ± 2.2 265.6 ± 4.2 17.6 ± 1.1

Cairo 13 69.3 ± 4.4 −0.5 ± 1.7 −24.7 ± 4.2 24.7 ± 4.2 269.2 ± 3.5 20.2 ± 1.1
21 66.2 ± 3.4 −2.3 ± 0.6 −27.9 ± 3 28 ± 3 265.3 ± 0.8 20.2 ± 1.8

Bandolero 13 60.8 ± 1.8 −4.9 ± 0.7 −31.9 ± 2.7 32.3 ± 4.3 261.2 ± 1.4 22.9 ± 1.2
21 59.6 ± 3.1 −5.7 ± 0.6 −33.2 ± 2.6 33.7 ± 2.5 260.2 ± 1.6 23.9 ± 3.6

Buscaro 13 63.4 ± 3 −3.8 ± 1.5 −29.9 ± 2.9 30.2 ± 2.1 262.9 ± 2.9 21.2 ± 1.6
21 63.4 ± 1.4 −5.4 ± 0.2 −30.6 ± 0.8 31 ± 0.8 260 ± 0.7 20.3 ± 1.4

FD&C Blue No. 2 78.06 −17.95 −24.28 30.19 233.52

a
Harvesting time. bColor difference with FD&C Blue No. 2.

7529 dx.doi.org/10.1021/jf501991q | J. Agric. Food Chem. 2014, 62, 7524−7531


Journal of Agricultural and Food Chemistry Article

however, year-to-year variability of these cultivars should be (15) Giusti, M. M.; Wrolstad, R. E. Acylated anthocyanins from
investigated. edible sources and their applications in food systems. Biochem. Eng. J.


2003, 14, 217−225.
AUTHOR INFORMATION (16) Stintzing, F. C.; Stintzing, A. S.; Carle, R.; Frei, B.; Wrolstad, R.
E. Color and antioxidant properties of cyanidin-based anthocyanin
Corresponding Author
pigments. J. Agric. Food Chem. 2002, 50, 6172−6181.
*(M.M.G.) Phone: (614) 247-8016. Fax: (614) 292-0218. E- (17) Giusti, M. M.; Rodriguez-Saona, L. E.; Wrolstad, R. E. Molar
mail: giusti.6@osu.edu. absorptivity and color characteristics of acylated and non-acylated
Funding pelargonidin-based anthocyanins. J. Agric. Food Chem. 1999, 47, 4631−
We are thankful to MARS Global Chocolate, Hackettstown, NJ, 4637.
USA, for providing funding for the project. (18) Walkowiak-Tomczak, D.; Czapski, J. Colour changes of a
Notes preparation from red cabbage during storage in a model system. Food
Chem. 2007, 104, 709−714.
The authors declare no competing financial interest.


(19) Tamura, H.; Yamagami, A. Antioxidative activity of mono-
acylated anthocyanins isolated from Muscat Bailey A grape. J. Agric.
ACKNOWLEDGMENTS Food Chem. 1994, 42, 1612.
We thank Ken McCammon and Bejo Seeds, Inc., for providing (20) Dyrby, M.; Westergaard, N.; Stapelfeldt, H. Light and heat
the plant materials. Also, we especially thank Marçal Plans sensitivity of red cabbage extract in soft drink model systems. Food
Pujolras from Universitat Politècnica de Catalunya for his help Chem. 2001, 72, 431−437.
with data statistical analysis. (21) Chalker-Scott, L. Environmental significance of anthocyanins in


plant stress responses. Photochem. Photobiol. 1999, 70, 1−9.
ABBREVIATIONS USED (22) Connor, A. M.; Luby, J. J.; Tong, C. B. S.; Finn, C. E.; Hancock,
J. F. Genotypic and environmental variation in antioxidant activity,
DM, dry matter; FM, fresh matter; AZ, Azurro; Ba, Bandolero; total phenolic content, and anthocyanin content among blueberry
Bu, Buscaro; Ca, Cairo; In, Integro; Ko, Kosaro; Pr, Primero


cultivars. J. Am. Soc. Hortic. Sci. 2002, 127, 89−97.
(23) Awad, M. A.; de Jager, A.; van der Plas, L. H. W.; van der Krol,
REFERENCES A. R. Flavonoid and chlorogenic acid changes in skin of ‘Elstar’ and
(1) Giusti, M. M.; Schwartz, S.; Elbe, H. V. Colorants. In Fennema’s ‘Jonagold’ apples during development and ripening. Sci. Hortic.−
Food Chemistry, 4th ed.; Damodaran, S., Parkin, K., Fennema, O. R., Amsterdam 2001, 90, 69−83.
Eds.; CRC Press/Taylor & Francis: Boca Raton, FL, USA, 2008; pp (24) Solomon, A.; Golubowicz, S.; Yablowicz, Z.; Grossman, S.;
571−638. Bergman, M.; Gottlieb, H. E.; Altman, A.; Kerem, Z.; Flaishman, M. A.
(2) Henry, B. S. Natural food colours. In Natural Food Colorants, 2nd Antioxidant activities and anthocyanin content of fresh fruits of
ed.; George, A. F., Hendry, J. D., Eds.; Blackie Chapman & Hall: common fig (Ficus carica L.). J. Agric. Food Chem. 2006, 54, 7717−
London, UK, 1996; pp 40−79. 7723.
(3) Wrolstad, R. E.; Smith, D. E. Color analysis. In Food Analysis, 4th (25) Fawole, O. A.; Opara, U. L. Effects of maturity status on
ed.; Nielsen, S. S., Ed.; Springer Science+Business Media: Berlin, biochemical content, polyphenol composition and antioxidant capacity
Germany, 2010; pp 573−586. of pomegranate fruit arils (cv. ‘Bhagwa’). S. Afr. J. Bot. 2013, 85, 23−
(4) Socaciu, C. Food Colorants: Chemical and Functional Properties; 31.
Taylor & Francis: Boca Raton, FL, USA, 2008. (26) Josuttis, M.; Verrall, S.; Stewart, D.; Krueger, E.; McDougall, G.
(5) Wrolstad, R. E. Symposium 12: Interaction of natural colors with J. Genetic and environmental effects on tannin composition in
other ingredients − anthocyanin pigments − bioactivity and coloring strawberry (Fragaria × ananassa) cultivars grown in different
properties. J. Food Sci. 2004, 69, C419−C421. European locations. J. Agric. Food Chem. 2013, 61, 790−800.
(6) Kong, J.-M.; Chia, L.-S.; Goh, N.-K.; Chia, T.-F.; Brouillard, R. (27) Wrolstad, R. E.; Durst, R. W.; Giusti, M. M.; Rodriguez-Saona,
Analysis and biological activities of anthocyanins. Phytochemistry 2003, L. E. Analysis of anthocyanins in nutraceuticals. In Quality Management
64, 923−933. of Nutraceuticals; American Chemical Society: Washington, DC, USA,
(7) Heins, A.; Stockmann, H.; Schwarz, K. Designing “anthocyanin-
2001; Vol. 803, pp 42−62.
tailored” food composition. In Biologically-Active Phytochemicals in
(28) Giusti, M. M.; Wrolstad, R. E. Characterization of red radish
Food; Pfannhauser, W., Fenwick, G. R., Khokhar, S., Eds.; Springer:
anthocyanins. J. Food Sci. 1996, 61, 322−326.
Berlin, Germany, 2001; pp 378−381.
(29) Arapitsas, P.; Sjoberg, P. J. R.; Turner, C. Characterisation of
(8) He, J. A.; Giusti, M. M. Anthocyanins: natural colorants with
anthocyanins in red cabbage using high resolution liquid chromatog-
health-promoting properties. Doyle, M. P., Klaenhammer, T. R., Eds.
Annu. Rev. Food Sci. Technol. 2010, 1, 163−186. raphy coupled with photodiode array detection and electrospray
(9) Ghosh, D.; Konishi, T. Anthocyanins and anthocyanin-rich ionization-linear ion trap mass spectrometry. Food Chem. 2008, 109,
extracts: role in diabetes and eye function. Asia Pac. J. Clin. Nutr. 2007, 219−226.
16, 200−208. (30) Wu, X. L.; Prior, R. L. Identification and characterization of
(10) Piccaglia, R.; Marotti, M.; Baldoni, G. Factors influencing anthocyanins by high-performance liquid chromatography-electrospray
anthocyanin content in red cabbage (Brassica oleracea var capitata L f ionization-tandem mass spectrometry in common foods in the United
rubra (L) Thell). J. Sci. Food Agric. 2002, 82, 1504−1509. States: vegetables, nuts, and grains. J. Agric. Food Chem. 2005, 53,
(11) McDougall, G. J.; Fyffe, S.; Dobson, P.; Stewart, D. 3101−3113.
Anthocyanins from red cabbage − stability to simulated gastro- (31) Charron, C. S.; Clevidence, B. A.; Britz, S. J.; Novotny, J. A.
intestinal digestion. Phytochemistry 2007, 68, 1285−1294. Effect of dose size on bioavailability of acylated and nonacylated
(12) Wiczkowski, W.; Szawara-Nowak, D.; Topolska, J. Red cabbage anthocyanins from red cabbage (Brassica oleracea L. var. capitata). J.
anthocyanins: profile, isolation, identification, and antioxidant activity. Agric. Food Chem. 2007, 55, 5354−5362.
Food Res. Int. 2013, 51, 303−309. (32) Park, S.; Arasu, M. V.; Lee, M. K.; Chun, J. H.; Seo, J. M.; Lee, S.
(13) Mazza, G.; Miniati, E. Anthocyanins in fruits, vegetables, and W.; Al-Dhabi, N. A.; Kim, S. J. Quantification of glucosinolates,
grains; CRC Press: Boca Raton, FL, USA, 1993. anthocyanins, free amino acids, and vitamin C in inbred lines of
(14) Cabrita, L.; Fossen, T.; Andersen, O. M. Colour and stability of cabbage (Brassica oleracea L.). Food Chem. 2014, 145, 77−85.
the six common anthocyanidin 3-glucosides in aqueous solutions. Food (33) Scalzo, R. L.; Genna, A.; Branca, F.; Chedin, M.; Chassaigne, H.
Chem. 2000, 68, 101−107. Anthocyanin composition of cauliflower (Brassica oleracea L. var.

7530 dx.doi.org/10.1021/jf501991q | J. Agric. Food Chem. 2014, 62, 7524−7531


Journal of Agricultural and Food Chemistry Article

botrytis) and cabbage (B. oleracea L. var. capitata) and its stability in
relation to thermal treatments. Food Chem. 2008, 107, 136−144.
(34) Sun, J. H.; Xiao, Z. L.; Lin, L. Z.; Lester, G. E.; Wang, Q.;
Harnly, J. M.; Chen, P. Profiling polyphenols in five brassica species
microgreens by UHPLC-PDA-ESI/HRMSn. J. Agric. Food Chem.
2013, 61, 10960−10970.
(35) Wu, X. L.; Beecher, G. R.; Holden, J. M.; Haytowitz, D. B.;
Gebhardt, S. E.; Prior, R. L. Concentrations of anthocyanins in
common foods in the United States and estimation of normal
consumption. J. Agric. Food Chem. 2006, 54, 4069−4075.
(36) Linus Pauling Institute: Micronutrient Research for Optimum
Health, Oregon State University; http://lpi.oregonstate.edu/
infocenter/phytochemicals/flavonoids/flavtab2.html (accessed July
15, 2013).
(37) Holton, T. A.; Cornish, E. C. Genetics and biochemistry of
anthocyanin biosynthesis. Plant Cell 1995, 7, 1071−1083.
(38) Torskangerpoll, K.; Andersen, Ø. M. Colour stability of
anthocyanins in aqueous solutions at various pH values. Food Chem.
2005, 89, 427−440.
(39) Fossen, T.; Cabrita, L.; Andersen, O. M. Colour and stability of
pure anthocyanins influenced by pH including the alkaline region.
Food Chem. 1998, 63, 435−440.

7531 dx.doi.org/10.1021/jf501991q | J. Agric. Food Chem. 2014, 62, 7524−7531

You might also like