Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Journal of Food Composition and Analysis 67 (2018) 135–140

Contents lists available at ScienceDirect

Journal of Food Composition and Analysis


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

Original research article

Effect of cooking on the bioaccessibility of essential elements in different T


varieties of beans (Phaseolus vulgaris L.)
Aline Pereira de Oliveiraa, Bianca dos Santos Oliveira Mateóa, Alexandre Minami Fiorotob,

Pedro Vitoriano de Oliveirab, Juliana Naozukaa,
a
Departamento de Química, Universidade Federal de São Paulo, 275 Prof. Artur Riedel St. 09972270, Diadema, Sao Paulo, Brazil
b
Instituto de Química, Universidade de São Paulo, C.P. 26077, 05513-970, Sao Paulo, SP, Brazil

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

Keywords: The chemical composition of beans (Phaseolus vulgaris L.) provides several health benefits to humans. However,
Phaseolus beans cooking beans before consumption is necessary to reduce the effect of certain toxic and anti-nutritional sub-
Bioaccessibility stances and to increase protein digestibility. However, heat treatment of beans may alter the chemical com-
Ca position and nutrient bioaccessibility. In this study, the effect of cooking on elemental bioaccessibility in seven
Cu
varieties of Phaseolus beans (common, black, fradinho, rajado, bolinha, jalo, and rosinha) were evaluated. In vitro
Fe
S
gastrointestinal digestion of raw and cooked beans was performed to evaluate the bioaccessibility of calcium,
Zn copper, iron, sulfur, and zinc, combined with elemental determination using inductively coupled plasma optical
ICP OES emission spectrometry. The gastrointestinal digestion of raw and cooked grains showed that the cooking de-
Food analysis creased the percentage of soluble iron in the fradinho (29%), rajado (27%), bolinha (20%), and rosinha (7%)
Food composition varieties. The percentages of elemental bioaccessibility for raw or cooked black beans showed that sulfur (93%,
raw and cooked) and copper (88% for raw and 89% for cooked) were the most bioaccessible essential elements;
in contrast, calcium (41%) was the least accessible element in cooked grain.

1. Introduction 78 to 123 mg for raw and cooked beans, respectively, for the common,
black, rosinha, roxo, mulatinho, and rajado varieties. Studies on
Bean (Phaseolus vulgaris L.) is an important component of tropical common and faba beans have shown that after cooking, more than 70%
and subtropical cuisine. It promotes several health benefits (Antunes of the Cu- and Fe-species became insoluble, owing to protein dena-
et al., 1995), being the main source of vegetable protein and the most turation and association with polyphenols and phytate (Naozuka and
important legume in the American and African continent, where animal Oliveira, 2012). In contrast, cooking softened the food matrix and re-
protein is limited because of economic, religious, and cultural reasons leased protein-bound elements, thus facilitating their absorption. In
(Kay, 1979). Varieties of Phaseolus beans are commonly consumed in addition, heating food altered the inherent factors that inhibit mineral
Brazil as it promotes good health, reduces the consumption of lipids and absorption, such as phytates and dietary fiber (Carbonara et al., 2001).
sodium (Geil and Anderson, 1994; Sathe et al., 1984), and presents a Therefore, the influence of cooking on the chemical composition of
good source of vitamins, essential elements (K, Ca, Mg, P, and Fe), beans is evident.
proteins (20–25%), and complex carbohydrates (50–60%) (Ranilla However, the nutritional value of a food depends not only on its
et al., 2009; Saha et al., 2009; Slupski, 2010). Beans possess functional elemental content, but also on its bioavailability. In vivo and in vitro
properties owing to their chemical composition, and they are indicated methods are used to estimate elemental bioavailability and bioacces-
in the dietary manipulation of diseases such as cardiovascular disease, sibility, respectively. Bioaccessibility is defined as the fraction of an
diabetes mellitus, obesity, and cancer (Costa and Rosa, 2008). analyte released from its matrix in the gastrointestinal tract.
However, cooking can alter the chemical composition of beans Subsequently, the element soluble in the gastrointestinal fluid is ab-
(essential elements, proteins, and metalloproteins). The Brazilian Table sorbed during digestion and transformed into metabolically active
of Food Composition (TACO) (2011) lists the chemical composition molecules. Elemental bioaccessibility depends not only on the matrix,
(proteins, lipids, cholesterol, carbohydrates, fiber, and calcium) of dif- but also on the chemical form of the analyte and the analytical model
ferent varieties of Phaseolus. The Ca mass ranged from 17 to 29 mg and used. One of the in vitro methods involves using a gastrointestinal


Corresponding author.
E-mail address: jnaozuka@gmail.com (J. Naozuka).

https://doi.org/10.1016/j.jfca.2018.01.012
Received 2 March 2017; Received in revised form 5 January 2018; Accepted 7 January 2018
Available online 09 January 2018
0889-1575/ © 2018 Elsevier Inc. All rights reserved.
A.P.d. Oliveira et al. Journal of Food Composition and Analysis 67 (2018) 135–140

model, where samples are sequentially exposed to artificial gastric and 2.2. Instrumentation
intestinal juices (Leufroy et al., 2012).
In vitro methods offer an appealing alternative to in vivo studies, as For quantification of Ca, Cu, Fe, S, and Zn, an ICP OES (iCAP 6300
they permit manipulations in the experimental conditions and evalua- Duo, Thermo Fisher Scientific, Cambridge, England) was used, which
tion of different digestion stages, in addition to being more simple, cost- was equipped with axial and radial viewed plasma, a charge-injection
effective, and rapid than in vivo methods (Miller et al., 1981). In vitro device (CID) detector allowing measurements from 166.25 to
models commonly use a simulated digestion system with gastric and 847.00 nm, an Echelle polychromator, and a radiofrequency source of
intestinal fluid. Dialysis with a cellulose membrane is used to evaluate 27.12 MHz. A Meinhard nebulizer combined with a cyclonic spray
nutrient uptake, which simulates the brush border of cell membranes. chamber was used as the sample introduction system. The injector tube
The diffusion of an element across a dialysis membrane during the in- diameter of the torch was 2.0 mm.
testinal stage is considered as a measure of its bioaccessibility. In fact, A shaker table (Q225 M model, Quimis, Diadema, Brazil), was used
these methods are based on simulations of human digestion and allow for solution homogenization and stirring during the extraction proce-
estimation of nutrient bioavailability (US Pharmacopeia 2000). dure. For simulated gastrointestinal digestion, a water bath (Q226M2
Evaluation of the bioaccessibility of essential elements such as Ca, model, Quimis, Diadema, Brazil) was used at 90 rpm for 120 min. A
Cu, Fe, S, and Zn in different varieties of cooked beans, which are ready centrifuge (Q222TM model, Quimis, Diadema, Brazil), was used for
for human consumption, may assist health programs and add greater phase separation.
economic and nutritional value to a food that is widely consumed by The beans were cooked in an electric pressure cooker (Philips Walita
the world population. These elements are essential for human health as Daily Collection, Sao Bernardo do Campo, Brazil). The drying and
they are cofactors for many physiological and metabolic functions milling procedures for the raw and cooked beans were performed in a
(Diplock, 1991). For example, calcium is important for the integrity of drying oven (515 model, FANEM, Guarulhos, Brazil) and a cryogenic
the skeletal structure (O’Dell and Sunde, 1997), copper is required for grinder (MA 775 model, Marconi, Piracicaba, Brazil), respectively.
the oxidizing and reducing activities of certain enzymes (Cozzolino,
2007), iron is essential for energy metabolism and hemoglobin forma- 2.3. Preliminary sample preparation
tion (O’Dell and Sunde, 1997), sulfur is present in several key enzymes
that require free sulfhydryl groups for catalysis, and zinc is the com- The raw beans were cleaned with deionized water and dried in an
ponent of several enzymes involved in biochemical reactions (O’Dell oven at 60 °C to obtain constant mass. The raw beans were milled in the
and Sunde, 1997). cryogenic grinder (particle size value < 100 μm), with 5 min of freezing
The objective of this study was to study the bioaccessibility of Ca, followed by three 2-min cycles of grinding, with 1 min of freezing be-
Cu, Fe, S, and Zn in different varieties of beans after cooking. We used tween each cycle. A part of the raw beans (ca. 20 g) was soaked in
simulated gastrointestinal digestion, dialysis, and elemental determi- deionized water (ca. 200 mL) at room temperature for 24 h. The soaking
nation using inductively coupled plasma optical emission spectrometry water was discarded and the soaked beans were cooked in deionized
(ICP OES) to address this objective. water (1:4 w/v bean: water) for 30 min (Quinteros et al., 2001). The
cooked beans and deionized water were mixed and dried in an oven at
2. Materials and methods 60 °C until a constant mass was obtained. The mixture was ground using
a porcelain mortar and pestle.
2.1. Reagents and samples
2.4. Simulated gastrointestinal digestion assay
All solutions were prepared from analytical reagent grade chemicals
and deionized water (18.2 MΩ cm) obtained from a Milli-Q water pur- The in vitro gastrointestinal digestion procedure using fluids pre-
ification system (Millipore, Burlington, USA) as the solvent. Analytical pared according to the recommendations of the U.S.A. Pharmacopeia
grade nitric acid 65% (w/v), distilled in a quartz sub-boiling still XXIV (2000), was employed to evaluate the bioaccessibility of the es-
(Marconi, Piracicaba, Brazil), was used for constituting reference so- sential elements (Ca, Cu, Fe S, and Zn). The gastric fluid was prepared
lutions and performing simulated gastrointestinal digestion. by dissolving 0.2 g of NaCl and 0.32 g of pepsin in water; 7 mL of HCl
Titrisol standard solutions of 1000 mg L−1 for Ca (CaCl2), Cu 10% (v/v) was subsequently added, and this solution was diluted to
(CuCl2), Fe (FeCl3), S (H2SO4), and Zn (ZnCl2) (Merck, Germany) were 100 mL with deionized water and maintained at pH 1.5. Three grams of
used to prepare the reference analytical solutions after dilution in nitric NaHCO3 was dissolved in 100 mL of highly pure deionized water to
acid 0.1% (v/v) for elemental determination by ICP OES. obtain a 3% (w/v) solution. The intestinal fluid was prepared by dis-
Gastrointestinal fluids were prepared using NaCl (Merck, solving 0.680 g of K2HPO4, 1.0 g pancreatin, 1.25 g bile salts, and
Darmstadt, Germany), pepsin (Sigma Aldrich, St. Louis, USA), HCl 7.7 mL of NaOH (0.2 mol L−1). This solution was diluted to 100 mL and
(Merck, Darmstadt, Germany), NaHCO3 (Reagen, Rio de Janeiro, final pH 6.8.
Brazil), K2HPO4 (Synth, Diadema, Brazil), NaOH (Merck, Synth, For in vitro gastrointestinal digestion, 3.0 mL gastric fluid (pH 1.5)
Germany), pancreatin (Sigma Aldrich, St. Louis, USA), and bile salts was added to 200 mg beans (raw and cooked). The mixture was shaken
(Sigma Aldrich, St. Louis, USA). The pH was adjusted using sodium continuously (100 rpm) in a thermostatic bath at 36 °C for 2 h. To adjust
bicarbonate solution 3% (w/v) (Reagen, Rio de Janeiro, Brazil). the pH to 6.8, 0.4 mL of NaHCO3 3% (w/v) was added. Then, 3.0 mL of
Dialysis was performed using cellulose membranes with 12.4 kDa intestinal fluid (pH 6.5) was added, and the extract solution was shaken
porosity (Sigma Aldrich, St. Louis, USA) in 32-mm tubes. using the same conditions reported in the gastric digestion step. The
Sodium sulfide (Merck, Darmstadt, Germany) 0.3% (w/v) and sul- samples were placed in an ice bath to stop the enzyme activity, and
furic acid (Merck, Darmstadt, Germany) 0.2% (v/v) solutions were used consequently, the simulated digestion was centrifuged to separate the
for treating the membranes. supernatant. These aqueous supernatants (the soluble fraction) were
The different varieties of Phaseolus beans (common, black, rajado, analyzed by ICP OES.
rosinha, bolinha, fradinho, and jalo) were purchased at a local market
in Sao Paulo, Brazil. Six species were of the same brand, whereas jalo 2.5. Dialysis procedure
was of a different brand. The producers assert that the geographic
origin of the beans is São Paulo (rosinha, rajado, and bolinha) and Dialysis was performed to simulate human nutrient absorption. The
Minas Gerais (common, black, fradinho, and jalo). The raw and cooked membranes were treated before dialysis by washing in water for 3–4 h
beans were used in the simulated gastrointestinal digestion assay. for removing glycerol, in Na2S (0.3% w/v) at 80 °C for 1 min, and

136
A.P.d. Oliveira et al. Journal of Food Composition and Analysis 67 (2018) 135–140

storing in H2SO4 (3% v/v). The supernatants (2 mL) were added to the limit of detection (LOD), and limit of quantification (LOQ). The limits of
bags formed from the cellulose membranes, tied with polyethylene detection (LOD) were calculated in accordance with the International
tapes, and immersed in tubes filled with 30 mL water. After 12 h, the Union of Pure and Applied Chemistry (IUPAC) recommendations, using
membrane was carefully opened, and the solution inside the bag (dia- the background equivalent concentration (BEC) and signal-to-back-
lyzed fraction) was collected, and analyzed by ICP OES. ground ratio (SBR) (Mermet and Poussel, 1995; IUPAC, 1978). Thus,
The bioaccessible fraction was determined by subtracting the con- BEC = Crs/SBR; SBR = (Irs − Iblank)/Iblank; LOD = 3 × BEC × RSD/
centration retained in the membrane from the concentration of the 100; where Crs is the concentration of the multi-elemental reference
soluble fraction (Nascimento et al., 2010). The percentages of soluble solution, and Irs and Iblank are the emission intensities for the multi-
(% S) and bioaccessible (% B) fractions were calculated with respect to elemental reference and blank solutions, respectively. The limits of
the corresponding total concentration (Ferreira et al., 2010) and soluble quantification (LOQ) were calculated as 3.3 × LOD (Silva et al., 2007).
elemental concentration, respectively. The values were obtained in μg·g−1, considering a sample mass of
200 mg and a final volume of 6.4 mL.
2.6. Elemental determination in soluble and dialyzed fractions by ICP OES The samples (soluble and dialyzed fractions) were diluted in nitric
acid to resemble the reference analytical solutions. In addition, dilution
Elemental determination was performed after evaluating the ro- decreased the chemical interferences during elemental determination
bustness of the spectrometer conditions. The instrumental conditions by ICP OES. It is worth noting that chemical interferences have been
are shown in Table 1. Calibration was performed using analytical so- evaluated previously (Fioroto et al., 2015).
lutions in HNO3 0.1% (v/v) (Ca, Cu, Fe, S, and Zn: 0.1–60 mg L−1).
Prior to sample analyses by ICP OES, the soluble and bioaccessible 3.2. Determination of Ca, Cu, Fe, S, and Zn in gastrointestinal digestion: the
fractions were centrifuged, and a 3 mL aliquot of the supernatants was effect of cooking
added to 0.5 mL 10% v/v HNO3 and the final volume was made up to
5 mL with deionized water. Bioaccessibility was determined by evaluating the elemental con-
centrations in the soluble and dialyzed fractions. The concentration of
the elements of interest in these fractions was compared to their total
2.7. Statistical analyses
concentration in the different raw and cooked beans; the total con-
centrations were obtained from our previous work (Ferreira et al.,
Statistically significant differences (p < .05) were detected using
2010) and the results are shown in Table 3. For this purpose, the
the two-way analysis of variance (ANOVA) of the Statistica 12.0 System
samples (raw and cooked) were submitted to acid digestion in a closed-
Software (StatSoft, Inc., Tulsa, USA). Differences between means in
vessel microwave oven, using 250 mg sample mass (dried and
significant cases were compared using Tukey's honestly significant
grounded) and the diluted oxidant mixture (3 mL HNO3 + 2 mL
difference (HSD) post hoc test at 5% significance.
H2O2 + 1 mL H2O) (Naozuka et al., 2011). Elemental determination
was performed using ICP OES and flame atomic absorption spectro-
3. Results and discussion metry (FAAS), after the method validation.
The bioaccessibility of Ca, Cu, Fe, S, and Zn in different varieties of
3.1. Method characteristics beans determined using the in vitro method was predicted by analysis of
the gastrointestinal product before (soluble fraction) and after dialysis.
The instrumental parameters for ICP OES analysis are shown in The bioaccessible fractions were determined by subtracting the con-
Table 1. Table 2 lists the characteristic parameters of the analytical centration of the elements in the membrane-bound solution from the
calibration curve, such as linear range, correlation coefficient (R2), concentration in the soluble fraction. The concentrations are shown in
Table 3 as percentages of soluble (% S) and bioaccessible (% B) frac-
Table 1
tions with respect to the corresponding total concentration (Ferreira
Instrumental conditions for Ca, Cu, Fe, S, and Zn determination by ICP OES.
et al., 2010) and soluble elemental concentration, respectively. The
Power 1250 W elemental soluble percentage present in the gastrointestinal fluid before
Nebulizer Meinhard dialysis represents the portion accessible to humans in vivo. The ele-
Spray chamber Cyclonic mental percentage that traversed the membrane during dialysis re-
Plasma gas flow 15 L min−1
Intermediate gas flow 1.0 L min−1
presents the accessible portion of the soluble fraction.
Nebulizer gas flow 0.45 L min−1 Considering the results obtained after analyzing the soluble frac-
Sample introduction 15 L min−1 tions of all the bean varieties (raw and cooked), Fe was found to be the
Analytical wavelengtha (nm) − axial view least accessible, ranging from 11% (black) to 67% (fradinho) in the raw
Ca (II) = 317.9 S(I) = 180.7
beans and 11% (black) to 42% (jalo) in the cooked beans. Previous
Cu(I) = 327.3 Zn(I) = 231.8
Fe(II) = 239.5 studies showed that the bioavailability of Fe to humans from Phaseolus
beans is low (1–2%) because of the presence of phytochemicals such as
a
(I) = atomic line; (II) = ionic line. phytic acid and polyphenols, which are known inhibitors of Fe ab-
sorption (Laparra et al., 2008). In contrast, Zn was the most soluble
element, with solubility percentage ranging from 34 (fradinho) to 161
Table 2
Parameters of the analytical calibration linear range, correlation coefficient (R2), limits of
(black) % and 29 (bolinha) to 146 (rosinha) % for raw and cooked
detection (LOD), and limit of quantification (LOQ) for the elements studied. grains, respectively. The solubilities of Fe and Zn, pH of the intestinal
lumen, dietary factors, and retention time during digestion and ab-
Element Linear R2 LOD LOQ LOD LOQ sorption influence the bioavailability of these elements. In vivo and in
range (soluble) (soluble) (dialyzed) (dialyzed)
vitro studies have reported the negative effects of phytates on Fe and Zn
(mg L−1) (μg g−1) (μg g−1) (μg g−1) (μg g−1)
bioavailability (Luo et al., 2013).
Ca 0.1–60.0 0.9984 1.63 4.93 0.78 2.15 The heating of beans increases protein digestibility from 25 to 60%
Cu 0.1–60.0 0.9983 0.24 0.47 0.11 0.32 (raw) to 85% (cooked), depending on the bean species and the cooking
Fe 0.1–60.0 0.9989 0.01 0.04 0.02 0.07
method (Bressani, 1993). Furthermore, cooking induces desirable sen-
S 0.1–60.0 0.9994 1.25 3.76 0.46 1.39
Zn 0.1–60.0 0.9940 0.02 0.05 0.03 0.10 sory properties in beans, such as sweet taste, cooked bean flavor, and
soft and mushy textures (Ranilla et al., 2009). However, cooking also

137
Table 3
Total concentration and percentages of Ca, Cu, Fe, S, and Zn; soluble (% S) and bioaccessible (% B) in seven varieties of Phaseolus beans (raw and cooked).

Common Black Jalo

Raw Cooked Raw Cooked Raw Cooked


A.P.d. Oliveira et al.

Ca* Total24 0.86 ± 0.20 1.6 ± 0.3 0.76 ± 0.19 2.2 ± 0.3 0.93 ± 0.12 0.96 ± 0.15
Soluble 0.28 ± 0.03a 0.48 ± 0.01a 0.71 ± 0.01a 0.53 ± 0.03a 0.37 ± 0.01a 0.50 ± 0.01a
(%S) (32) (30) (89) (24) (40) (52)
Bioaccessible 0.18 ± 0.01a 0.39 ± 0.01a 0.61 ± 0.02a 0.22 ± 0.02a < LODa*** < LODa***
(%B) (63) (82) (82) (41) (0) (0)
Cu** Total24 9.6 ± 2.6 9.8 ± 1.2 12 ± 6 11 ± 1 14 ± 2 14 ± 1
Soluble 2.7 ± 0.4a,b 4.0 ± 0.3b 5.3 ± 0.1b,c 4.7 ± 0.5b 7.1 ± 0.1c 6.2 ± 0.1c
(%S) (28) (41) (44) (43) (51) (44)
Bioaccessible 2.0 ± 0.2a 3.7 ± 0.1a,b 4.7 ± 0.1a,b 4.2 ± 0.1a,b 5.7 ± 0.3a,c 5.4 ± 0.3a,c
(%B) (73) (91) (88) (89) (80) (87)
Fe** Total24 87 ± 9 81 ± 2 94 ± 3 101 ± 2 84 ± 2 79 ± 2
Soluble 15.7 ± 0.1a 18.6 ± 0.7a 10.3 ± 0.2a 11.1 ± 0.4a 31.1 ± 0.5a 33.2 ± 0.2a
(%S) (18) (23) (11) (11) (37) (42)
Bioaccessible 10.2 ± 0.8a,b 13.4 ± 0.2a,b 8.8 ± 0.6a,b 7.7 ± 0.1a,b 3.1 ± 0.6b 11.6 ± 0.7b
(%B) (65) (72) (85) (69) (10) (35)
S* Total24 1.7 ± 0.1 1.0 ± 0.1 1.8 ± 0.1 2.3 ± 0.1 1.8 ± 0.1 2.1 ± 0.1
Soluble 1.51 ± 0.01b,c 1.84 ± 0.05c 2.29 ± 0.02c 1.5 ± 0.1b 0.65 ± 0.03a 0.50 ± 0.03a
(%S) (89) (184) (127) (63) (36) (24)
Bioaccessible 1.31 ± 0.01b 1.69 ± 0.05b 2.2 ± 0.1c 1.35 ± 0.01b 0.34 ± 0.05a 0.40 ± 0.04a
(%B) (87) (92) (93) (93) (53) (79)
Zn** Total24 33 ± 8 36 ± 4 33 ± 4 35 ± 3 35 ± 3 48 ± 3
Soluble 41 ± 1c 47.2 ± 1c,d 53 ± 2d 36.1 ± 0.6c 56 ± 1d 38 ± 1c
(%S) (124) (131) (161) (103) (160) (79)

138
Bioaccessible 34 ± 2c 39.6 ± 0.2c 39.9 ± 0.4b,c 26 ± 2b.c 22 ± 1a.b < LODa***
(%B) (82) (84) (75) (73) (40) (0)

Fradinho Rajado Bolinha Rosinha

Raw Cooked Raw Cooked Raw Cooked Raw Cooked

*
Ca 0.84 ± 0.16 0.63 ± 0.15 0.96 ± 0.17 0.77 ± 0.14 0.72 ± 0.15 0.77 ± 0.15 0.61 ± 0.13 0.99 ± 0.17
0.45 ± 0.04a 0.33 ± 0.01a 0.44 ± 0.01a 0.49 ± 0.01a 0.51 ± 0.01a 0.58 ± 0.01a 0.57 ± 0.01a 0.55 ± 0.01a
(54) (52) (46) (64) (71) (75) (94) (56)
< LODa*** < LODa*** 0.44 ± 0.01a 0.49 ± 0.04a 0.27 ± 0.01a 0.55 ± 0.01a 0.27 ± 0.03a 0.02 ± 0.03a
(0) (0) (100) (100) (53) (94) (48) (4)
Cu** 3.1 ± 1.8 3.4 ± 0.2 14 ± 3 11 ± 1 9.7 ± 2.9 8.4 ± 0.2 13 ± 5 10 ± 1
2.02 ± 0.3a,b 0.71 ± 0.1a 6.4 ± 0.1c 3.5 ± 0.1b 4.0 ± 0.1b 4.8 ± 0.1b.c 7.2 ± 0.1c 5.3 ± 0.1b.c
(65) (21) (46) (32) (41) (57) (55) (53)
2.0 ± 0.4a,b 0.7 ± 0.4a 4.7 ± 0.5a,c 1.6 ± 0.3a 1.8 ± 0.4a 3.9 ± 0.5a,b 2.3 ± 0.3a 4.7 ± 0.2a,b
(100) (100) (73) (46) (45) (82) (32) (88)
Fe** 61 ± 5 64 ± 1 75 ± 9 68 ± 3 61 ± 4 59 ± 1 56 ± 3 60 ± 1
40.9 ± 0.5a 24.32 ± 0.4b 36.0 ± 0.3a 21.1 ± 0.2b 31.1 ± 0.5a 18.3 ± 0.3b 29.1 ± 0.3a 27.0 ± 0.4a
(67) (38) (48) (21) (51) (31) (52) (45)
20.0 ± 0.6b 14.6 ± 0.8b 20.2 ± 0.7b 12.2 ± 0.4b 16.2 ± 0.8b 5.1 ± 0.4b 11.9 ± 0.5b 17 ± 1b
(49) (60) (56) (58) (52) (28) (41) (63)
Journal of Food Composition and Analysis 67 (2018) 135–140
A.P.d. Oliveira et al. Journal of Food Composition and Analysis 67 (2018) 135–140

causes considerable changes in the composition of numerous chemical


constituents, including amino acids, vitamins, and minerals, as well as

0.51 ± 0.06a

0.45 ± 0.04a
alterations in the bioavailability and bioaccessibility of these nutrients

< LODa***
1.6 ± 0.1

30 ± 1c
(Slupski, 2010). It is noteworthy that thermal treatment of beans im-

38 ± 4
Cooked

proves protein and starch digestibility and raises the nutritive value by
(32)

(89)

(79)

(0)
reducing anti-nutrients such as phytic acid and tannins (Ranilla
et al.2009). Comparison of the concentrations of the nutrient elements
in the soluble fraction of raw and cooked grains showed that cooking
decreased the percentage of soluble Fe in fradinho (29%), rajado (27%),
and bolinha (20%) varieties. Besides losing a part of the soluble iron by
0.70 ± 0.06a
0.9 ± 0.1b

maceration, the decrease in the percentage of soluble Fe after cooking


1.9 ± 0.1

28 ± 1a,b

Values are means ± standard deviation (n = 3). Different superscripts within the same element indicate significant differences (p < .05).
51 ± 1d
Rosinha

35 ± 4

can also be attributed to the changes in elemental interaction with


(146)
Raw

(49)

(75)

(55)

proteins such as albumin, rendering it insoluble after heating


(Fernandes et al., 2010; Oliveira and Naozuka, 2015).
The bioaccessible percentage values show that S (93%, raw and
cooked) and Cu (88% for raw and 89% for cooked) were the most
bioaccessible essential elements; however, Ca (41%) was the least ac-
11.0 ± 0.7a,b
0.65 ± 0.07a

cessible element in the cooked grains. In contrast, the bioaccessibility of


0.9 ± 0.1a

< LODa***
1.7 ± 0.1

Cu, Fe, and Zn decreased after cooking the black beans. For other bean
38 ± 4
Cooked

varieties, the bioaccessibility of most elements increased after cooking,


(50)

(76)

(29)

(0)

especially those of Ca and Cu in bolinha bean.


The bioaccessibility of an element depends not only on the matrix,
but also on the chemical form of the analyte (Leufroy et al., 2012).
Additionally, heating of food alters the bioavailability of macro and
microelements, increases protein digestibility, and reduces the effect of
24.1 ± 0.4b.c

certain anti-nutritional components. These effects improve the ab-


1.5 ± 0.1b

1.1 ± 0.1a

< LODa***
1.7 ± 0.1

36 ± 4
Bolinha

sorption of microelements because protein-bound elements are re-


Raw

(85)

(73)

(67)

leased, and interactions between essential elements and anti-nutrients


(0)

(fibers, polyphenols, oxalate, tannins, and phytates), which render


them nutritionally inaccessible, do not occur (Bertina et al., 2016;
Hemalatha et al., 2007a,b; Sahuquillo et al., 2003).
Contrary to the increase in elemental bioaccessibility by cooking,
0.80 ± 0.06a

studies with cereals, pulses, white beans, chickpeas, and lentils showed
22.1 ± 0.1b
0.7 ± 0.1a

< LODa***
1.7 ± 0.1

that heating did not improve the bioaccessibility of Zn (Hemalatha


35 ± 3
Cooked

et al., 2007a,b; Sahuquillo et al., 2003). In food grains, the differences


(47)

(88)

(63)

(0)

in the extent of elemental dialysis is attributed to the different chemical


forms of these metals (valence states), and differences in the physico-
chemical environment and metal localization in the grains. These ele-
ments are complexed with other constituents, possibly proteins and/or
other food components (Hemalatha et al., 2007a,b).
24.2 ± 0.3b.c

In summary, it is evident that cooking alters elemental solubility


1.3 ± 0.1b

0.9 ± 0.1a

6.0 ± 0.6a
1.7 ± 0.1

and bioaccessibility in different varieties of Phaseolus beans. As ex-


35 ± 5
Rajado

pected, the varieties of raw and cooked beans that presented high
Raw

(76)

(66)

(69)

(25)

bioaccessibility for Cu (jalo, fradinho, rajado, bolinha, and rosinha)


showed a low percentage of Zn bioaccessibility. Heating favors dena-
turation and hydrolysis of proteins, which affects the transition metal
ion chelation capacity of proteins, inter-protein interactions, and the
presence of essential elements, thereby significantly altering elemental
0.82 ± 0.05a,b

bioaccessibility. Heating increased the digestibility and Fe absorption in


a−d
0.7 ± 0.1a

< LODa***

< LODa***
2.1 ± 0.1

common, jalo, fradinho, rajado, and rosinha beans, whereas the


52 ± 3

LOD (μg g−1) = 0.78 (Ca) and 0.03 (Zn).


Cooked

bioaccessibility of Cu decreased in the rajado variety, possibly due to


(39)

(87)

(0)

(0)

the formation of Maillard reaction products (amino sugars) (Pedrosa


and Cozzolino, 2007; Valdés et al., 2011). Furthermore, a second effect
of cooking is the formation of interactions between proteins or elements
with other components present in the beans; for example, Fe forms
insoluble complexes after cooking, which is unavailable for absorption
in humans (Oliveira and Naozuka, 2015). Studies on common beans
1.6 ± 0.2b
1.9 ± 0.1c

showed that heat treatment rendered approximately 80% of the pro-


< LODa***
2.0 ± 0.1

20 ± 1b,c
Fradinho

58 ± 4

teins and more than 70% of Cu and Fe insoluble because of protein


Raw

(93)

(84)

(34)

denaturation and association of the elements with anti-nutrients, such


(0)

as polyphenols and phytates (Hemalatha et al., 2007a,b; Valdés et al.,


Table 3 (continued)

***

2011).
μg/g.

4. Conclusions
**
mg/g;
Zn**
*
S

The bean varieties studied here showed significant differences in the


*

139
A.P.d. Oliveira et al. Journal of Food Composition and Analysis 67 (2018) 135–140

concentration of essential elements that were soluble in the gastro- Hemalatha, S., Platel, K., Srinivasan, K., 2007b. Zinc and iron contents and their bioac-
intestinal fluid, with Fe and Ca being the least bioaccessible elements in cessibility in cereals and pulses consumed in India. Food Chem. 102, 1328–1336.
IUPAC, 1978. Nomenclature, symbols, units and their usage in spectrochemical analysis-II
common, black, jalo, and fradinho beans. Cooking changed the solu- data interpretation Analytical chemistry division. Spectrochim. Acta 33, 241.
bility, and consequently the elemental bioaccessibility, in all varieties Kay, D.E., 1979. Food Legumes. TPI. Crop Digest – Tropical Products Institute, London,
of beans; Zn was inaccessible in jalo, rajado, and rosinha beans after England.
Laparra, J.M., Glahn, R.P., Miller, D.D., 2008. Bioaccessibility of phenols in common
cooking. Heating favored denaturation and protein hydrolysis by al- beans (Phaseolus vulgaris L.) and iron (Fe) availability to caco-2Cells. J. Agric. Food
tering the interactions or chelation capacity of proteins with transition- Chem. 56, 10999–11005.
metal elements, which significantly changed their bioaccessibility. Leufroy, A., Noël, L., Beauchemin, D., Guérin, T., 2012. Bioaccessibility of total arsenic
and arsenic species in seafood as determined by a continuous online leaching method.
Considering the poor bioaccessibility of certain essential elements in Anal. Bioanal. Chem. 402, 2849–2859.
the most important cultivated legume in the world, this study would be Luo, Y.-W., Xie, W.-H., Jin, X.-X., Wang, Q., Zai, X.-M., 2013. Effects of germination and
important for supporting healthy nutritional programs aimed at im- cooking for enhanced in vitro iron, calcium and zinc bioaccessibility from faba bean,
azuki bean and mung bean sprouts. CYTA J. Food 11, 318–323.
proving the mineral content in different varieties of Phaseolus beans.
Mermet, J.M., Poussel, E., 1995. ICP emission spectrometers: analytical figures of merit.
Appl. Spectrosc. 49, 12A–18A.
Compliance with ethics requirements Miller, D.D., Schricker, B.R., Rasmussen, R.R., Campen, D.V.C., 1981. An in vitro method
for estimation of iron availability from meals. Am. J. Clin. Nutr. 34, 2248–2256.
Naozuka, J., Oliveira, P.V., 2012. Cooking effects on iron and proteins content of beans
The authors declare that they have no conflict of interest. (Phaseolus vulgaris L.) by GF AAS and MALDI-TOF MS. J. Braz. Chem. Soc. 23,
This article does not contain any studies with human or animal 156–162.
subjects. Naozuka, J., Vieira, E.C., Nascimento, A.N., Oliveira, P.V., 2011. Elemental analysis of
nuts and seeds by axially viewed ICP OES. Food Chem. 124, 1667–1672.
Nascimento, A.N., Naozuka, J., Oliveira, P.V., 2010. In vitro evaluation of Cu and Fe
Acknowledgments bioavailability in cashew nuts by off-line coupled SEC–UV and SIMAAS. Microchem.
J. 96, 58–63.
O’Dell, B.L., Sunde, R.A., 1997. Handbook of Nutritionally Essential Mineral Elements.
Aline Pereira de Oliveira thanks the Fundação de Amparo à Pesquisa CRC Press, New York,U.S.A.
do Estado de São Paulo/FAPESP (2015/01128-6) for the fellowship Oliveira, A.P., Naozuka, J., 2015. Chemical speciation of iron in different varieties of
provided. Juliana Naozuka thanks the FAPESP for financial support beans (Phaseolus vulgaris L.): cooking effects. J. Braz. Chem. Soc. 26, 2144–2149.
Pedrosa, L.F.C., Cozzolino, S.M.F., 2007. Biodisponibilidade de Nutrientes. Manole, São
(2015/15510-0). Paulo Brazil.
Quinteros, A., Farré, R., Lagarda, M., 2001. Optimization of iron speciation (soluble,
References ferrous and ferric) in beans chickpeas and lentils. J. Food Chem. 75, 365–370.
Ranilla, L.G., Genovese, M.I., Lajolo, F.M., 2009. Effect of different cooking conditions on
phenolic compounds and antioxidant capacity of some selected Brazilian bean
Antunes, P.L., Bilhalava, A.B., Elias, M.C., Soares, G.J., 1995. Valor nutricional de feijão (Phaseolus vulgaris L.) cultivars. J. Agric. Food Chem. 57, 5734–5742.
(Phaseolus vulgaris, L.), cultivares rico 23, carioca: pirata-1 e rosinha-G2. Revista Saha, S., Singh, G., Mahajan, G., Gupta, H.S., 2009. Variability of nutritional and cooking
Brasileira de Agrociência 1, 12–18. quality in bean (Phaseolus vulgaris L.) as a function of genotype. Plant Food Hum.
Bertina, R.L., Maltezc, H.F., Goisd, J.S., Borgesd, D.L.G., Borgesa, G.S.C., Gonzaga, L.V., Nutr. 64, 174–180.
Fetta, R., 2016. Mineral composition and bioaccessibility in Sarcocornia ambigua Sahuquillo, A., Barberá, R., Farré, R., 2003. Bioaccessibility of calcium: iron and zinc from
using ICP-MS. J. Food Comp. Anal. 47, 45–51. three legume samples. Food/Nahrung 47, 438–441.
Bressani, R., 1993. Grain quality of common beans. Food Rev. Int. 9, 237–297. Sathe, S.K., Deshpande, S.S., Salunkhe, D.K., 1984. Dry beans of Phaseolus. A review. Part
Carbonara, M., Grant, G., Aguzzi, A., Pusztai, A., 2001. Investigation of the mechanisms 2. Chemical composition: carbohydrates, fiber, minerals, vitamins and lipids. Crit.
affecting Cu and Fe bioavailability from legumes. Biol. Trace Elem. Res. 84, 181–196. Rev. Food Sci. Nutr. 21, 41–93.
Costa, N.M.B., Rosa, C.O.B., 2008. Alimentos Funcionais: benefícios para a saúde. Metha, Silva, J.C.J., Cadore, S., Nóbrega, J.A., 2007. Dilute-and-shoot procedure for the de-
Viçosa, MG, Brazil. termination of mineral constituents in vinegar samples by axially viewed inductively
Cozzolino, S.M.F., 2007. Biodisponibilidade de Nutrientes. Manole, São Paulo, SP, Brazil. coupled plasma optical emission spectrometry (ICP OES). Food Addit. Contam. 24,
Diplock, A.T., 1991. Antioxidant nutrients and disease prevention: an overview. Am. J. 130–139.
Clin. Nutr. 53, 189S. Slupski, J., 2010. Effect of cooking and sterilisation on the composition of amino acids in
Fernandes, A.C., Nishida, W., Proença, R.P.C., 2010. Influence of soaking on the nutri- immature seeds of flageolet bean (Phaseolus vulgaris L.) cultivars. Food Chem. 121,
tional quality of common beans (Phaseolus vulgaris L.) cooked with or without the 1171–1176.
soaking water: a review. Int. J. Food Sci. Technol. 45, 2209–2218. Tabela Brasileira de Composição de Alimentos (TACO) (Núcleo de estudos e pesquisas em
Ferreira, A.S.T., Naozuka, J., Oliveira, P.V., Kelmer, G., 2010. Effects of the domestic alimentação − UNICAMP-NEPA, 2011), http://www.unicamp.br/nepa/taco/contar/
cooking on elemental chemical composition of beans species (Phaseolus vulgaris L.). J. taco_4_edicao_ampliada_e_revisada.pdf?arquivo=taco_4_versao_ampliada_e_revisada.
Food Process 2014, 1–1-6. pdf. Accessed 30, April 2016.
Fioroto, A.M., Nascimento, A.N., Oliveira, P.V., 2015. In vitro evaluation of Cu, Fe: and U.S.A. Pharmacopeia XXIV & National. Test solutions. The United States Pharmacopeia
Zn bioaccessibility in the presence of babassu mesocarp. J. Agric. Food Chem. 63, and The National Formulary, USP24-NF19; The United States Pharmacopeial
6331–6337. Convention: Rockville, MD, 2000, 2235–2236.
Geil, R.B., Anderson, J.W., 1994. Nutrition and health implications of dry beans: a review. Valdés, S.T., Coelho, C.M.M., Michelluti, D.J., Tramonte, V.L.C.G., 2011. Association of
J. Am. Col. Nutr. 13, 549–558. genotype and preparation methods on the antioxidant activity: and antinutrients in
Hemalatha, S., Platel, K., Srinivasan, K., 2007a. Influence of heat processing on the common beans (Phaseolus vulgaris L.). LWT – Food Sci. Technol. 44, 2104–2111.
bioaccessibility of zinc and iron from cereals and pulses consumed in India. J. Trace
Elem. Med. Biol. 21, 1–7.

140

You might also like