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

Food Chemistry 114 (2009) 10631068

Contents lists available at ScienceDirect

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

The impact of dehydration process on antinutrients and protein digestibility


of some legume ours
Mara A. Martn-Cabrejas a,*, Yolanda Aguilera a, Mercedes M. Pedrosa b, Carmen Cuadrado b,
Teresa Hernndez c, Soledad Daz c, Rosa M. Esteban a
a

Departamento de Qumica Agrcola, Facultad de Ciencias, Universidad Autnoma de Madrid (UAM), 28049 Madrid, Spain
Departamento de Tecnologa de Alimentos, SGIT-INIA, Ctra Corua km 7.5, 28040 Madrid, Spain
c
Instituto de Fermentaciones Industriales, CSIC, Juan de la Cierva, 3, 28006 Madrid, Spain
b

a r t i c l e

i n f o

Article history:
Received 28 May 2008
Received in revised form 1 October 2008
Accepted 30 October 2008

Keywords:
Dehydration process
Antinutrients
In vitro protein digestibility
Legume ours

a b s t r a c t
Dehydrated foods are specially designed for patients with mastication or/and deglutition problems. This
study has assessed the effects of soaking, cooking and industrial dehydration treatments on antinutrient
factors and also on protein digestibility in legume ours (chickpea, lentil and bean). A general decline of
phytic acid was observed during dehydration, being the most accentuated in case of lentil (44%), followed
by white beans and pink-mottled cream beans. Beans were the legumes that showed the highest levels of
enzyme inhibitors and lectins, however processing such as cooking and dehydration signicantly reduced
(p < 0.05) their levels further to negligible concentrations. The dehydration did not cause further effects
than ordinary cooking in reduction of the concentration of polyphenolic compounds of ours. However, a
higher increase of in vitro protein digestibility (IVPD) was produced by dehydration in all legumes from
12% to 15%. Thus, dehydrated legume ours could be considered ready-to-use for special meals to specic
populations.
2008 Elsevier Ltd. All rights reserved.

1. Introduction
Nowadays, a great interest of food industry is the challenge of
designing, developing, and commercialising special foods for every
population group such as infant, youth, adults, pregnant women,
sportsmen, and elder people (Mota & Empis, 2000). In this sense,
texture-modied foods are specially designed to patients with
chewing problems and swallowing dysfunction which are common among elderly people and affect perception, food choice
and the ability to eat (Rothenborg et al., 2007). Texture-modied
foods present the aspect, taste and similar characteristics to the
traditional foods that allow the accurate composition, energy
and nutrients to be known in every prepared dish. Dehydrated
foods are considered texture-modied foods which may be rehydrated with the appearance and texture of canned or conventionally prepared foods (Nickels, 2006; Vega-Mercado, Gngora-Nieto,
& Barbosa-Cnovas, 2001). Hence, dehydration of foods is a controlled effect to preserve the structure or create a new one that
serves for functional purposes (Aguilera, Chiralt, & Fito, 2003). In
addition, these foods present the following advantages: maximum
microbiology security (low food manipulation), soft and homogenous texture, prolonged preservation time, lesser production of

* Corresponding author. Tel.: +34 91 4978678; fax: +34 91 4973830.


E-mail address: maria.martin@uam.es (M.A. Martn-Cabrejas).
0308-8146/$ - see front matter 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.foodchem.2008.10.070

food by-products, easy use and storage food (reduction of volume


and weight).
One suitable food to be dehydrated are legumes, they are valuable sources of dietary protein to complement cereals, roots and
tubers. Legume starch is more slowly digested than cereal starch
and its ingestion produces less abrupt changes in plasma glucose
and insulin. Legume seeds are also valuable sources of dietary ber,
vitamins and minerals including folate, thiamine and riboavin
(Messina, 1999). Thus, they are important components of a healthy
diet. However, their nutritional quality is limited by the presence
of heat labile and heat-stable antinutritional factors (ANF) that exhibit undesirable physiological effects (Pusztai, Bardocz, & MartnCabrejas, 2004). The ANFs are structurally different compounds
broadly divided into two categories: proteins (such as lectins and
protease inhibitors) and others such as phytate, tannins or proanthocyanidins, oligosaccharides, saponins and alkaloids. The presence, distribution and negative impact of the ingestion of ANFs in
grain legumes have extensively been reported (Grant, 1991).
In general, raw legumes contain far higher levels of ANFs than
their processed forms hence processing is necessary before the
incorporation of these grains into food or animal diets (Hajs &
Osagie, 2004). Few studies about the industrial process of dehydration after soaking and cooking treatments have been carried out
(Martn-Cabrejas et al., 2006) in order to know the nutritional
improvement and hence, to obtain legume ours with high nutritive-value.

1064

M.A. Martn-Cabrejas et al. / Food Chemistry 114 (2009) 10631068

In the present study, the effect of soaking, cooking and dehydration treatment has been assessed for eliminating ANFs and
improvement of the protein digestibility from legumes. The
obtained dehydrated legume ours could be considered readyto-use for special meals to specic groups of populations with
mastication and/ or swallowing problems.

2. Materials and methods

1964). The a-amylase inhibitor content of seed extracts was determined by the starch/iodine procedure of Piergiovanni (1992). Suitable aliquots of the seed extracts were mixed with a-amylase
solution and incubated at 40 C for 30 min to allow formation of
the inhibitor-enzyme complex and the amount of remaining aamylase was then determined. The enzyme inhibitor contents of
seed extracts were calculated by comparison of the amount of sample or inhibitor required to cause 50% inhibition of enzyme activity
and were expressed as mg-equivalent of inhibitor standard g 1 seed
meal. For inhibitor levels, triplicate assays were conducted.

2.1. Samples
2.5. Lectins
Seeds of chickpea (Castellano and Sinaloa varieties), lentil and
two cultivars of beans (white bean and pink-mottled cream bean)
were used in the present study. They were obtained from the agrifood industry Vegenat SA (Badajoz, Spain). From each cultivar
there were three batches of 250 g of raw and processed samples.
The seeds were freeze-dried and were milled to our and passed
through a 250 lm sieve.
2.2. Processing conditions
Legumes were subjected to an industrial dehydration process
carried out in Vegenat SA. The processing consisted of the following steps: raw material was soaked in tap water (1:10 w/v) for 16 h
at 20 C. After draining the soaking water, the soaked legumes
were cooked by boiling for 70 min in case of chickpeas, 30 min in
case of lentil, 20 min in case of white beans, and 30 min in case
of pink-mottled cream beans. The soaked-cooked seeds were dehydrated in a forced-air tunnel at 75 3 C for 6 h. Samples were taken at each step and they were named as follows: S (soaked ours),
S + C (soaked and cooked ours) and S + C + D (soaked, cooked and
dehydrated ours).
2.3. Phytic acid
The individual inositol phosphates (IP3-IP6) were extracted
according Burbano, Muzquiz, Osagie, Ayet, and Cuadrado (1995)
with modications and measured by HPLC (Lehrfeld, 1994). Analysis was with a Beckman System Gold HPLC equipped with a refractive index. The column was a macrosporous polymer PRP-1
(150  4.1 mm, 5 lm) heated at 45 C and was equilibrated with
the mobile phase for 1 h. The mobile phase was 515 ml of methanol with 485 ml of water. Eight millilitre of tetrabutylammonium
hydroxide (40% in water), 1 ml 5 M sulphuric acid, 0.5 ml 91% formic acid and 100 ll of a phytic acid hydrolysate (6 mg ml 1) were
added sequentially. The pH was adjusted to 4.3 with 9 M sulphuric
acid. The mobile phase was ltered through a Millipore lter (0.45
lm) and degassed under a vacuum. The ow rate was 1.2 ml min 1
and the injection volume was 20 ll. The standard used was sodium
phytate (Sigma Chemicals, USA).
2.4. Enzyme inhibitors
Seed ours were extracted (1:10 w/v) by stirring with 0.02 M
sodium phosphate buffer pH 7.0 containing NaCl (8 g l 1) overnight at +1 C and centrifuged (50,000g for 25 min). The resultant
clear supernants were stored at 20 C.
Estimations of protease-inhibitor content were carried out
essentially as described previously (Grant et al., 1986). Seed extracts were added to excess trypsin or chymotrypsin and the mixtures left on ice to react. The level of remaining trypsin activity in
the reaction mixture was assessed using N-a-benzoyl-DL-p-nitroanilide as substrate (Kakade, Simons, & Liener, 1969) while the level of
uninhibited chymotrypsin was evaluated using glutaryl-L-phenylalanine-p-nitroanilide as substrate (Erlanger, Copper, & Bendich,

Haemagglutinating activity in the pH 7.0 buffer extracts was


estimated by a serial dilution procedure using rat blood cells
(Grant, 1991). The amount of material (g) which caused agglutination of 50% of the erythrocytes was dened as that containing 1
haemagglutinating unit (HU). The assays were reproducible to 1
dilution and values in the text were the mean of four separate measurements and Processor (kidney bean) was included in each assay
as a control. In addition to the haemagglutination assay, a competitive indirect ELISA assay for quantication of PHA (Phaseolus vulgaris lectin) in bean samples was performed according to Hajs
et al. (1995) with some modications. Plates were coated overnight at 4 C with 0.1 lg ml 1 PHA in 0.01 M phosphate-buffered
saline (PBS). Standard PHA diluted in PBST (0.01 M PBS containing
0.1% Tween 20, pH 7.4) or bean samples with unknown content of
lectin were added, followed by rabbit anti-PHA IgG antibody (Sigma Chemical Co, St Louis, Mo, USA, diluted 1: 32,000 with PBST).
After incubation for 1 h, goat anti-rabbit IgG biotin conjugate (Sigma Chemical Co, St Louis, Mo, USA) diluted with PBST (1:2000 v/v)
was added. After washing, Extravidin-peroxidase (Sigma Chemical
Co, St Louis, Mo, USA) diluted 1:500 was added. After incubation
for 1 h a solution of OPD (o-phenylene-diamine)H2O2 was added
and the reaction was stopped by adding 50 ll of 3 M H2SO4 and the
optical density measured at 492 nm.
2.6. Polyphenolic compounds
The similar structural groups of phenolic compound content are
evaluated in both raw and processed legume ours in order to get a
global value which will allow us to analyse the changes of these
components undergo as a consequence of processing.
For extracting polyphenolic compounds, ground seeds were
macerated with 2  30 ml of a solution of methanolClH (1)/
water (80:20 v/v) in orbital shaker, at room temperature (Dueas,
Hernndez, & Estrella, 2004). The two macerates were combined
and in the resulting liquid the quantication of the phenolic compound families was done.
In the methanol solution, the total polyphenols were quanticated by the FolinCiocalteu reactant (Singleton & Rossi, 1965),
catechins with vanillin/HCl (Swain & Hills, 1959) and proanthocyanidins through hydrolysis with butanol/HCl (Ribreau-Gayn &
Stonestreet, 1966).
2.7. In vitro protein digestibility
In vitro protein digestibility (IVPD) was evaluated according to
McDonough et al. (1990). The protein solutions, 4 mg ml 1 in distilled water, were added to 2.5 ml 0.2 N NaOH while stirring at
37 C in a water bath. After 30 min, 5 ml 0.075 N HCl was added
to the solution and pH was adjusted to pH 8.0. The reaction was
started by adding 1 ml of the multienzyme system consisting of
trypsin (23,100 U ml 1), chymotrypsin (186 U ml 1) and peptidase
(0.052 U ml 1). The amount of 0.1 N NaOH required to maintain
the pH 7.98 for exactly 5 min was recorded. The percent of

1065

M.A. Martn-Cabrejas et al. / Food Chemistry 114 (2009) 10631068

in vitro protein digestibility was calculated using the following


equation: IVPD = 79.28 + 40.74 B, where B = ml 0.1 N NaOH used
during 5 min.
2.8. Statistical analysis
Results were analysed using Duncans multiple range test
(DMRT) (Bender, Douglass, & Kramer, 1989). Differences were considered signicant at p 6 0.05.
3. Results and discussion
Results of individual and total inositol phosphates of legume
ours are showed in Table 1. The total inositol phosphate contents
determined by ion-pair HPLC exhibited differences among legumes
ranging from 6.75 mg g 1 DM (lentil) to 13.30 mg g 1 DM (pinkmottled cream bean). Similar contents were found in Castellano
and Sinaloa chickpea (7.15 and 7.90 mg g 1 DM, respectively).
The phytic acid composition of the studied legumes agreed with
previously published data on legumes (Daz-Batalla, Widholm, Fahey, Castao-Tostado, & Paredes-Lpez, 2006; Rubio et al., 2006;
Trugo et al., 1999), although chickpea and lentil showed lower levels. The relative percentage values obtained of IP3IP6 indicate
that the legumes contained more than 70% of its inositol phosphates in the form IP6, in case of lentil reached 91%. The relative
proportion of IP5 ranges from 16% in pink-mottled cream bean to
21% in Sinaloa chickpea, except for lentil (6%). The relative percentages of IP3 and IP4 are low and never higher than 9%. Only the
highly phosphorylated inositol phosphates IP6 and IP5 have a negative effect on the bioavailability of minerals, the others hydrolytic
products formed had a poor capacity to bind mineral (Lnnerdal,
2002).
Soaking process only reduced phytic acid content in lentil (38%)
and did not show any relevant changes in white, pink-mottled
cream beans and Sinaloa chickpea, whereas an important increase
was observed in Castellano chickpea (33%). IP6 was the major inositol phosphate found in all legumes and the relative proportion of

Table 1
Inuence of processing on inositol phosphate content (mg g

IP3-IP5 was low indicating a low degree of phytic acid hydrolysis in


the seeds, except for lentil. In most studies, soaking induced an
activation of the endogenous phytase and the diffusion of the products, but the extent of reduction is depending on the legume seed
(Alonso, Aguirre, & Marzo, 2000). A direct correlation between phytate reduction and phytase activity was found during germination
of lentils, however in chickpeas no correlation was established
(Greiner et al., 1998).
Regarding to cooking process, different behaviour is shown in
the studied legumes. Lentil, white and pink-mottled cream beans
exhibited a signicant reduction of phytic acid level, being more
relevant in lentil (40%) than in bean varieties (13% in white beans
and 10% in pink-mottled cream beans). The percentages of inositol
phosphorylated revealed a decrease of IP6 and the corresponding
increase of IP5-IP3 due to the hydrolysis of the IP6 during the cooking process. Other authors reported both lower and similar reductions in cooked legumes (Khatoon & Prakash, 2004; Rehman &
Shah, 2005). However, both chickpea varieties showed higher levels of phytic acid in the cooked ours, being more accentuated in
Castellano chickpea. These seed phytates somehow presented
more resistance to the process.
Regarding to dehydration process after cooking, it is observed a
general decline of phytic acid, being the most accentuated in case
of lentil (44%). Chickpea varieties reduced their phytic acid levels
respect to cooked samples but they were still higher than raw
levels. In case of beans the decrease of phytic acid in dehydrated
samples was more signicant in white beans (14%) than in pinkmottled cream beans (6%). Therefore, the dehydrated lentil our
seemed to present a better mineral availability when compared
to other legume ours. Hence, phytic acid was only partly affected
by the processing (soaking, cooking and dehydration) due to their
heat-stable nature. The results of this study were consistent to
those reported for cooking, autoclaving and extrusion as processing (Alonso et al., 2000; El-Hady & Habiba, 2003; Rehman & Shah,
2005).
a-Amylase and protease inhibitors are widely distributed in legumes. The studied legumes exhibited different contents of these

dry matter) of legume ours.

Sample

IP3

IP4

IP5

IP6

Total

Castellano chickpea
Raw
S
S+C
S+C+D

0.15 0.01a
0.19 0.01a
0.17 0.01a
0.17 0.00a

0.28 0.02c
0.38 0.01b
0.40 0.03b
0.61 0.01a

1.18 0.06d
1.65 0.01c
2.17 0.01b
2.53 0.03a

5.54 0.09d
7.33 0.02b
7.90 0.01a
6.25 0.03c

7.15 0.06c
9.55 0.01b
10.65 0.02a
9.55 0.00b

Sinaloa chickpea
Raw
S
S+C
S+C+D

0.17 0.00a
0.17 0.00a
0.18 0.00a
0.20 0.00a

0.45 0.01c
0.42 0.01c
0.50 0.01b
0.61 0.02a

1.68 0.00b
1.62 0.00b
1.97 0.05a
2.14 0.04a

5.60 0.01b
5.72 0.01a
5.71 0.02a
4.89 0.05c

7.90 0.01b
7.93 0.01b
8.36 0.06a
7.85 0.03b

Lentil
Raw
S
S+C
S+C+D

0.25 0.01c
0.45 0.01a
0.30 0.00b
0.24 0.01c

0.37 0.00c
0.72 0.01a
0.71 0.00a
0.60 0.00b

1.06 0.00a
0.64 0.01d
0.94 0.00b
0.87 0.00c

5.07 0.00a
2.41 0.07b
2.10 0.00c
2.04 0.02d

6.75 0.00a
4.22 0.10b
4.04 0.00c
3.75 0.03d

White bean
Raw
S
S+C
S+C+D

0.18 0.00d
0.21 0.00c
0.31 0.01b
0.35 0.00a

0.41 0.00d
0.43 0.00c
1.07 0.00b
1.06 0.01a

2.18 0.00d
2.22 0.01c
3.03 0.00b
3.07 0.01a

9.27 0.02a
9.18 0.02b
6.04 0.01c
5.89 0.01d

12.05 0.02a
12.05 0.01a
10.45 0.05b
10.38 0.03b

Pink-mottled cream bean


Raw
S
S+C
S+C+D

0.16 0.00a
0.16 0.00a
0.16 0.00a
0.17 0.00a

0.23 0.00d
0.26 0.00c
0.33 0.01b
0.44 0.00a

0.80 0.01d
1.06 0.00c
1.82 0.01b
2.33 0.07a

12.12 0.10a
11.73 0.07b
9.66 0.04c
9.55 0.01d

13.30 0.09a
13.20 0.07a
11.96 0.03c
12.50 0.07b

Mean values of each column followed by different superscript letter signicantly differ when subjected to Duncans multiple range test (p < 0.05).

1066

M.A. Martn-Cabrejas et al. / Food Chemistry 114 (2009) 10631068

Table 2
Inuence of processing on protease inhibitors content (mg g
ours.
Sample

a-Amylase inhibitor

dry matter) of legume

Trypsin inhibitor

Chymotrypsin inhibitor

Castellano chickpea
Raw
0.013 0.002b
S
0.022 0.001a
S+C
n.d.
S+C+D
n.d.

0.895 0.003b
1.007 0.002a
0.058 0.009c
0.055 0.007c

0.461 0.007b
0.752 0.010a
n.d.
n.d.

Sinaloa chickpea
Raw
n.d.
S
n.d.
S+C
n.d.
S+C+D
n.d.

0.912 0.011b
0.997 0.013a
0.074 0.008c
0.074 0.009c

n.d.
n.d.
n.d.
n.d.

Lentil
Raw
S
S+C
S+C+D

n.d.
n.d.
n.d.
n.d.

n.d.
n.d.
n.d.
n.d.

n.d.
n.d.
n.d.
n.d.

White bean
Raw
S
S+C
S+C+D

1.430 0.013a
1.200 0.015b
0.068 0.010c
0.061 0.003c

1.270 0.023b
1.550 0.018a
0.012 0.003c
0.010 0.002c

0.980 0.010b
1.700 0.018a
n.d.
n.d.

Pink-mottled
Raw
S
S+C
S+C+D

cream bean
0.990 0.012a
0.660 0.011b
0.082 0.013c
0.110 0.020c

1.090 0.022b
1.290 0.014a
0.009 0.003c
0.009 0.003c

0.713 0.012b
1.001 0.009a
n.d.
n.d.

Mean values of each column followed by different superscript letter signicantly


differ when subjected to Duncans multiple range test (p < 0.05).
n.d.: not detected.

antinutrients (Table 2). Beans contained nutritionally signicant


amounts of enzyme inhibitors, white bean showed higher levels.
Both chickpea varieties exhibited similar contents of trypsin inhibitor and only low content of chymotrypsin inhibitor in case of Castellano chickpea. Lentil did not show any enzyme inhibitor levels.
These antinutrient factors need to be inactivated by a suitable
pre-treatment before they can be safely used as a food source
(Hajs & Osagie, 2004).
Table 2 shows that various processing methods such as cooking
and dehydration signicantly (p < 0.05) reduced the levels of protease inhibitors, further to negligible concentrations. Compared
to the raw seeds, almost total reductions in the levels of antinutrients were observed in the thermal processing following the pattern
usually found in literature (Kataria, Chauhan, & Punia, 1989). The
increase of enzymatic inhibitors observed in soaking legumes
was more accentuated in case of beans than the rest of legumes.
It might be partially attributed to the leaching-out effect during
hydration.
An initial evaluation of lectin content in all legume samples was
carried out by using the haemagglutination assay. The results indicated that the highest haemagglutination level corresponded to
beans, 3.92 mg 100 mg 1 of Phaseolus lectin (PHA), in accordance
with the literature (Grant, 1991). The haemagglutination activity
was very low in case of lentil, 0.27 mg 100 mg 1 of Lens culinaris
lectin (LCL) and in case of chickpea was not determined due to
the irrelevance for the present study, thus the inuence of processing was studied in the rest of legumes. Soaking processing do not
modied the lectin content whereas heat-processing reduced drastically the level, being the dehydration process the methodology
which decreases lectins at minimum levels in all legume samples
(94% in beans and 92% in lentil). In addition to the haemagglutination test, indirect competitive ELISA was also applied to the analysis of beans. The results showed that pink-mottled cream bean had
higher content of lectin (1.30 mg 100 mg 1 ELISA PHA) than white
bean (1.19 mg 100 mg 1 ELISA PHA). This was in agreement with

the values detected by means of the haemagglunitation assay.


Soaked beans exhibited higher levels of lectin (18%) than the raw
one, similar trend was observed in the rest of antinutrient factors
(phytic acid and enzyme inhibitors). It might conrm a concentration effect. Moreover, the impact of processing is clear, no lectin
content was found in processed beans.
As shown in Table 3, total phenols signicantly varied among
the studied legumes. Raw lentil and pink-mottled cream bean
showed the highest levels (5.58 and 5.38 mg g 1 DM, respectively),
followed by Castellano and Sinaloa chickpeas, while white bean
exhibited the least phenol contents (2.08 mg g 1 DM). As was observed for other ANFs compounds, the inuence of processing is
relevant on polyphenol contents. Soaking did not cause relevant
changes in the total phenol content of most of legumes, except
for lentil which showed a remarkable decrease (76%). It could be
due to the presence of phenols on the periphery of the grain that
may pass out into the water through the seed coat. However, cooking of soaked seeds brought about a further reduction in the total
phenol contents of all pulses, which varied from 21% to 50%. The
results are in agreement with those reported by earlier authors
(Alonso et al., 2000; Manach, Scalbert, Morand, Rmsy, & Jimnez,
2004). The results of the present study reveal that dehydration did
not cause further effects than the ordinary cooking in reduction of
the concentration of polyphenolic compounds of seeds, except for
lentil that exhibited higher reduction (60%). The levels of total
phenolics found by Asami, Hong, Barret, and Mitchell (2003) in
air-dried strawberries, marionberries and corn were 1643% lower
than those found in frozen or freeze-dried samples. Decreases in
phenol content of the legume grains during cooking and dehydration may be ascribed to the binding of polyphenols with other
compounds (proteins) or the alterations in the chemical structure
of polyphenols which can not be extracted and determined by
available methods (Davis, 1981).
Table 3
Inuence of processing on total phenols. Total catechins and total procyanidins of
legume ours.
Total
proanthocyanidins
(mg g 1)

Total catechins
(mg g 1)

PC/
CAT

chickpea
3.04 0.23a
3.05 0.14a
1.74 0.06b
1.69 0.01b

1.96 0.03a
2.12 0.14a
0.75 0.05c
1.80 0.04b

0.58 0.01c
1.08 0.03a
0.95 0.05b
1.13 0.04a

3.37
1.96
0.79
1.59

Sinaloa chickpea
Raw
2.79 0.11a
S
2.56 0.15a
S+C
1.83 0.18b
S+C+D
1.84 0.03b

2.19 0.07a
2.07 0.11a
1.61 0.02c
1.28 0.08b

0.32 0.01c
0.61 0.03a
0.59 0.02a
0.42 0.05b

6.84
3.39
2.73
3.05

Lentil
Raw
S
S+C
S+C+D

5.91 0.28a
0.76 0.19b
1.17 0.26b
1.18 0.01b

3.24 0.05a
2.54 0.01b
1.60 0.05c
1.42 0.01d

1.82
0.31
0.73
0.83

White bean
Raw
2.08 0.01a
S
2.11 0.01a
S+C
1.64 0.04b
S + C + D 2.13 0.03a

n.d.
n.d.
n.d.
n.d.

0.54 0.07d
0.67 0.04c
1.32 0.02b
1.46 0.03a

Pink-mottled cream bean


Raw
5.38 0.06a
S
4.71 0.12b
S+C
2.84 0.18c
S + C + D 2.95 0.03c

6.81 0.51a
5.53 0.12b
n.d.
n.d.

4.38 0.08a
2.76 0.09b
2.28 0.11c
2.24 0.03c

Sample

Castellano
Raw
S
S+C
S+C+D

Total phenols
(mg g 1)

5.58 0.06a
1.35 0.13d
2.79 0.05b
2.19 0.02c

1.55
0.81

Mean values of each column followed by different superscript letter signicantly


differ when subjected to Duncans multiple range test (p < 0.05).
PC/CAT: total proanthocyanidins/total catechins.
n.d.: not detected.

1067

M.A. Martn-Cabrejas et al. / Food Chemistry 114 (2009) 10631068


Table 4
Inuence of processing on the in vitro protein digestibility (%) in legume ours.a
Sample

Raw
S
S+C
S+C+D
a

Castellano Chickpea

Sinaloa Chickpea

Lentil

White bean

Pink-mottled cream bean

80.0
81.0
85.7
90.8

81.1
81.2
89.7
90.9

85.0
85.5
91.7
95.5

79.1
79.9
88.6
90.0

79.6
80.0
89.2
91.3

(1)
(7)
(14)

(11)
(12)

(2)
(8)
(12)

(1)
(11)
(14)

(1)
(12)
(15)

Figures in parentheses indicate the percent increase over the values of the corresponding raw seed.

The changes that may occur in the content of the different phenolic families as a consequence of processing depend on the type of
legume and cultivar. The compounds that may be affected by the
thermal processing are proanthocyanidins and also catechins, since
there is a structural relationship between both compounds. Similar
to phenol contents, the content of proanthocyanidins ranged from
6.81 mg g 1 DM in pink-mottled cream bean and 5.91 mg g 1 DM
in lentil to 2.19 and 1.96 mg g 1 DM in Sinaloa and Castellano
chickpeas, respectively. No proanthocyanidin level was detected
in white bean. Our results showed a relevant relationship between
the seed coat colour and the contents of proanthocyanidins. However, this relationship is controversial (Beninger & Hoseld, 1999;
Daz-Batalla et al., 2006).
Soaking process decreased the levels of proanthocyanidins in
case of lentil (87%) and pink-mottled cream bean (67%), although
chickpea varieties did not exhibit reductions. Regarding to cooking,
this thermal processing caused important reductions of proanthocyanidin contents in all pulses and moreover these compounds
were not detected in pink-mottled cream bean. Dehydration process brought about similar results in proanthocyanidin content
compared to cooked seeds. The data agree with those found by
Alonso et al. (2000) and El-Hady and Habiba (2003) in other legumes using extrusion as processing. In general, air-drying at temperatures >60 C is regarded as unfavourable to extraction
efciency of phenolics due to the possibility of inducing oxidative
condensation or decomposition of thermolabile compounds such
as (+)-catechin (Asami et al., 2003).
The ratio proanthocyanidin/catechin, a relative approximation
of the polymerisation degree of proanthocyanidins, showed a similar tendency in all samples. It is important from a nutritional point
of view that the degree of polymerisation of proanthocyanidin be
low, because the extent of proanthocyanidinprotein interactions
increases with the degree of polymerisation (Davis, 1981; Ricardo-da-Silva, Cheynier, Souquet, & Moutounet, 1991). In these samples the degree of polymerisation decreased signicantly as a
consequence of soaking and cooking, while an increase occurred
during dehydration process. The increase in this last step of the
process could be related to several factors such as high temperatures, changes in their real solubility and chemical reactivity that
may modify the accessibility of the analysis (Tabera, Fras, Estrella,
Villa, & Vidal-Valverde, 1995). An increase in the high polymerised
proanthocyanidins was also observed in the germination and fermentation of lentils (Bartolom, Estrella, & Hernndez, 1997).
In vitro protein digestibility as inuenced by processing is shown
in Table 4. Data revealed that values are the lowest in white bean
and the highest in lentil and chickpea. A considerable variation
has been reported for legume protein digestibility in the literature
(Duranti & Guis, 1997). The effects of soaking were mostly inconsistent and did not seem to enhance the IVPD of either of the pulses reported in the present study. The observed values are in consonance
with recent published work (In-Hwa-Han, Swanson, & Byung,
2007). Cooking after soaking promoted higher protein digestibility
(from 7% to 12% improvement), similar percentage of IVPD improvement under cooking has been noticed previously (Clemente,
Snchez-Vioque, Bautista, & Milln, 1998; Rehman & Shah, 2005).

Compared with raw legumes, dehydration after cooking exhibited a signicant improvement of IVPD in all legumes from 12% to
15%. Dehydrated legume ours exhibit higher IVPD than legumes
processed by other methods such as ordinary cooking, microwave-cooking, ultrasound or high hydrostatic pressure (Clemente
et al., 1998; In-Hwa-Han et al. 2007; Khatoon & Prakash, 2004).
The effects are similar to those found applying autoclaving, pressure-cooking, or extrusion (Alonso et al., 2000; El-Hady & Habiba,
2003; Rehman & Shah, 2005). Improvement of protein digestibility
after processing could be attributed to the reduction or elimination
of different antinutrients. Phytic acid, as well as condensed tannins
and polyphenols are known to interact with protein to form complexes. These interactions could increase the degree of cross-linking, decreasing the solubility of proteins making protein
complexes which impair protease access to labile peptide bonds
(Genovese & Lajolo, 1996). In addition, thermal processing promoted structural changes of protein such as globulin, thereby
increasing chain exibility and accessibility to proteases (Swaisgood & Catignani, 1991).
Considering the processing methods attempted in the present
study, dehydration appears to be effective in reducing levels of
all the investigated antinutritional compounds, and improves the
protein digestibility of all studied seeds. Hence, a clear improvement in the nutritional characteristics of the studied legume seeds
is promoted by dehydration and it can be adopted as viable, costeffective processing method for the manufacture of texture-modied foods which aim to meet the needs of people with impaired
chewing and/or swallowing.

Acknowledgment
The authors also wish to thank to Vegenat SA for their nancial
support.

References
Aguilera, J. M., Chiralt, A., & Fito, P. (2003). Food dehydration and product structure.
Trends in Food Science and Technology, 14, 432437.
Alonso, R., Aguirre, A., & Marzo, F. (2000). Effects of extrusion and traditional
processing methods on antinutrients and in vitro digestibility of protein and
starch in faba and kidney beans. Food Chemistry, 68, 159165.
Asami, D. K., Hong, Y. J., Barret, D. M., & Mitchell, A. E. (2003). Comparison of the
total phenolic and ascorbic acid content of freeze-dried and air-dried
marionberry, strawberry and corn grown using conventional, organic and
sustainable agricultural practices. Journal of Agricultural and Food Chemistry, 51,
12371241.
Bartolom, B., Estrella, I., & Hernndez, T. (1997). Changes in phenolic compounds in
lentils (Lens culinaris) during germination and fermentation. Zeitschrift fur
Lebensmittel-Untersuchung und-Forschung, 205, 290294.
Bender, F. E., Douglass, L. W., & Kramer, A. (1989). Statistical methods for food and
agriculture. New York: Food Products Press.
Beninger, C. W., & Hoseld, G. L. (1999). Flavonoids composition of three genotypes
of dry bean (Phaseolus vulgaris) differing in seed coat color. Journal of American
Society of Horticultural Science, 124, 514518.
Burbano, C., Muzquiz, M., Osagie, A., Ayet, G., & Cuadrado, C. (1995). Determination
of phytate and lower inositol phosphates in Spanish legumes by HPLC
methodology. Food Chemistry, 52, 321325.
Clemente, A., Snchez-Vioque, R., Bautista, J., & Milln, F. (1998). Effect of cooking
on protein quality of chickpea (Cicer arietinum) seeds. Food Chemistry, 62, 16.

1068

M.A. Martn-Cabrejas et al. / Food Chemistry 114 (2009) 10631068

Davis, K. R. (1981). Effect of processing on composition and tetrahymena relative


nutritive-value of green and yellow peas, lentils and white pea beans. Cereal
Chemistry, 58, 454460.
Daz-Batalla, L., Widholm, J. M., Fahey, G. C., Castao-Tostado, E., & Paredes-Lpez,
O. (2006). Chemical components with health implications in wild and cultivated
Mexican common bean seeds (Phaseolus vulgaris L.). Journal of Agricultural and
Food Chemistry, 54, 20452052.
Dueas, M., Hernndez, T., & Estrella, I. (2004). Occurrence of phenolic compounds
in the seed coat and the cotyledon of peas (Pisum sativum L.). European Food
Research and Technology, 219, 116123.
Duranti, M., & Guis, C. (1997). Legume seeds: Protein content and nutritional value.
Field Crops Research, 53, 3145.
El-Hady, E. A. A., & Habiba, R. A. (2003). Effect of soaking and extrusion conditions
on antinutrients and protein digestibility of legumes seeds. LebensmittelWissenschaft und-Technologie, 13, 155175.
Erlanger, F. B., Copper, A. J., & Bendich, A. J. (1964). On the heterogeneity of threetimes crystalised a-chymotrypsin. Biochemistry, 3, 18801883.
Genovese, M. I., & Lajolo, F. M. (1996). In vitro digestibility of albumin proteins from
Phaseolus vulgaris L.). Effect of chemical modication. Journal of Agricultural and
Food Chemistry, 44, 30223028.
Grant, G. (1991). Lectins. In J. P. F. DMello, C. M. Duffus, & J. H. Duffus (Eds.), Toxic
substances in crop plants (pp. 4967). Cambridge: The Royal Society of
Chemistry.
Grant, G., McKenzie, N. H., Watt, W., Stewart, J. C., Dorward, P. M., & Pusztai, A.
(1986). Nutritional evaluation of soybeans (Glycine max): Nitrogen balance and
fractionation studies. Journal of the Science of Food and Agriculture, 37,
10011010.
Greiner, R., Pedrosa, M. M., Muzquiz, M., Ayet, G., Cuadrado, C., & Burbano, C. (1998).
Effect of germination on phytate content and phytase activity in legumes. In
Proceedings of the third European conference on Grain Legumes: Opportunities for
high quality, healthy and added-value crops to meet European demands
(pp. 8283). Wageningen: EAAP.
Hajs, G., Gelencsr, E., Pusztai, A., Grant, G., Sakhri, M., & Bardocz, S. (1995).
Biological effects and survival of trypsin inhibitors and the agglutinin from
soybean in the small intestine of the rat. Journal of Agricultural and Food
Chemistry, 43, 165170.
Hajs, G., & Osagie, A. U. (2004). Technical and biotechnological modications of
antinutritional factors in legumes and oilseeds. In Proceedings of the fourth
international workshop on antinutritional factors in legume seeds and oilseeds
(pp. 293305). Wageningen: EAAP.
In-Hwa-Han, A., Swanson, B. C., & Byung, K. B. (2007). Protein digestibility of
selected legumes treated with ultrasound and high hydrostatic pressure during
soaking. Cereal Chemistry, 84, 518521.
Kakade, M. L., Simons, N., & Liener, I. E. (1969). An evaluation of natural vs synthetic
substrates for measuring the antitrypsic activity of soybean sample. Cereal
Chemistry, 46, 518526.
Kataria, A., Chauhan, B. M., & Punia, D. (1989). Antinutrients and protein
digestibility (in vitro) of mung bean as affected by domestic processing and
cooking. Food Chemistry, 32, 917.
Khatoon, N., & Prakash, J. (2004). Nutritional quality of microwave-cooked and
pressure-cooked legumes. International Journal of Food Science and Nutrition, 55,
441448.
Lehrfeld, J. (1994). HPLC separation and quantication of phytic acid and some
inositol phosphates in foods: Problems and solutions. Journal of Agricultural and
Food Chemistry, 42, 27262730.
Lnnerdal, B. (2002). Phytic acid-trace element (Zn, Cu, Mn) interactions.
International Journal of Food Science and Technology, 37, 749758.

Manach, C., Scalbert, A., Morand, C., Rmsy, C., & Jimnez, L. (2004). Polyphenols:
Food sources and bioavailability. American Journal of Clinical Nutrition, 79,
727747.
Martn-Cabrejas, M. A., Aguilera, Y., Bentez, V., Moll, E., Lpez-Andru, F. J., &
Esteban, R. M. (2006). Effect of industrial dehydration on the soluble
carbohydrates and dietary bre fractions in legumes. Journal of Agricultural
and Food Chemistry, 54, 76527657.
McDonough, F. E., Sarwar, G., Steinke, F. H., Slump, P., Garca, S., & Boisen, S. (1990).
In vitro assay for protein digestibility: Interlaboratory study. Journal Association
Ofcial of Analytical Chemistry, 73, 622625.
Messina, M. J. (1999). Legumes and soybeans: Overview of their nutritional proles
and health effects. American Journal of Clinical Nutrition, 70, 439450.
Mota, M., & Empis, J. (2000). Novel foods and food ingredients: What is the mission
of scientists and technologists? Trends in Food Science and Technology, 11,
161168.
Nickels, N. C. (2006). Process for manufacturing dehydrated precooked legumes.
Patent US 2006/0263509 A1.
Piergiovanni, A. R. (1992). Effect of some experimental parameters on activity of
cowpea a-amylase inhibitors. Lebensmittel-Wissenschaft und-Technologie, 25,
321324.
Pusztai, A., Bardocz, S., & Martn-Cabrejas, M. A. (2004). The mode of action of ANFs
on the gastrointestinal tract and its microora. In Proceedings of the fourth
international workshop on antinutritional factors in legume seeds and oilseeds
(pp. 87100). Wageningen: EAAP.
Rehman, Z., & Shah, W. H. (2005). Thermal heat processing effects on antinutrients,
protein and starch digestibility of food legumes. Food Chemistry, 91, 327331.
Ribreau-Gayn, P., & Stonestreet, E. (1966). Dosage des tannins du vin rouges et
dtermination du leur structure. Chemie Analytique, 48, 188196.
Ricardo-da-Silva, J. M., Cheynier, V., Souquet, J. M., & Moutounet, M. (1991).
Interation of grape seed procyanidins with various proteins in relation to wine
ning. Journal of Science and Food Agriculture, 57, 111125.
Rothenborg, E., Ekman, S., Bulow, M., Moller, K., Svantesson, J., & Wendin, K. (2007).
Texture-modied meat and carrot for elderly people with dysphagia:
Preference in relation to health and oral status. Food and Nutrition Research,
51, 141147.
Rubio, L. R., Pedrosa, M. M., Cuadrado, C., Gelencser, E., Clemente, A., Burbano, C.,
et al. (2006). Recovery at the terminal ileum of some legume non-nutritional
factors in cannulated pigs. Journal of the Science of Food and Agriculture, 86,
979987.
Singleton, V. L., & Rossi, J. A. (1965). Colorimetry of total phenolics with
phosphomolibdic phosphotungstic acid reagent. American Journal of Enology
and Viticulture, 16, 144158.
Swain, T., & Hills, W. E. (1959). The phenolic constituents of Prunus domestica. I. The
quantitative analysis of phenolic constituents. Journal of the Science of Food
Agriculture, 10, 6368.
Swaisgood, E., & Catignani, L. G. (1991). Protein digestibility: In vitro methods of
assessment. In J. E. Kinsella (Ed.). Advances in Food and Nutrition Research (vol.
35, pp. 185230). San Diego: Academic Press.
Tabera, J., Fras, J., Estrella, I., Villa, R., & Vidal-Valverde, C. (1995). Natural
fermentation of lentils. Inuence of time, concentration and temperature on
protein content, trypsin inhibitor activity and phenolic compounds content.
Zeitschrift fur Lebensmittel-Untersuchung und-Forschung, 201, 587591.
Trugo, L. C., Muzquiz, M., Ayet, G., Burbano, C., Cuadrado, C., & Cavieres, C. (1999).
Inuence of malting on selected components of soya bean, black bean, chickpea
and barley. Food Chemistry, 65, 8590.
Vega-Mercado, H., Gngora-Nieto, M. M., & Barbosa-Cnovas, G. V. (2001). Advances
in dehydration of foods. Journal of Food Engineering, 49, 271289.

You might also like