Você está na página 1de 7

Food Chemistry 135 (2012) 2253–2259

Contents lists available at SciVerse ScienceDirect

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

Enzymatic bioconversion of citrus hesperidin by Aspergillus sojae naringinase:


Enhanced solubility of hesperetin-7-O-glucoside with in vitro inhibition of
human intestinal maltase, HMG-CoA reductase, and growth of Helicobacter pylori
Young-Su Lee a, Ji-Young Huh a, So-Hyun Nam a, Sung-Kwon Moon b, Soo-Bok Lee a,⇑
a
Department of Food and Nutrition, Brain Korea 21 Project, Yonsei University, Seoul 120-749, Republic of Korea
b
Department of Food Biotechnology, Chungju National University, Chungju 380-702, Republic of Korea

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

Article history: Hesperetin-7-O-glucoside (Hes-7-G) was produced by the enzymatic conversion of hesperidin by Asper-
Received 20 March 2012 gillus sojae naringinase due to the removal of the terminal rhamnose. Extracts from orange juice and peel
Received in revised form 31 May 2012 containing the hesperidin were so treated by this enzyme that the hesperidin could also be converted to
Accepted 2 July 2012
Hes-7-G. The solubility of Hes-7-G in 10% ethanol was enhanced 55- and 88-fold over those of hesperidin
Available online 14 July 2012
and hesperetin, respectively, which may make Hes-7-G more bioavailable. Hes-7-G was 1.7- and 2.4-fold
better than hesperidin and hesperetin, respectively, in the inhibition of human intestinal maltase. Hes-7-
Keywords:
G was more potent by 2- and 4-fold than hesperidin in the inhibition of human HMG-CoA reductase.
Hesperidin
Hesperetin-7-O-glucoside
Additionally, Hes-7-G exhibited more effective inhibition of the growth of Helicobacter pylori than hes-
Enzymatic bioconversion peretin, while its effectiveness was similar to that of hesperidin. Therefore, the results suggest that bio-
Naringinase converted Hes-7-G is more effective and bioavailable than hesperidin, as it has enhanced inhibitory and
Solubility solubility properties.
Ó 2012 Elsevier Ltd. All rights reserved.

1. Introduction hesperidin helps reduce abnormal capillary leakiness or excess


swelling in the legs due to fluid accumulation (Garg et al., 2001).
Flavonoids are the most abundant polyphenolic compounds, Both hesperidin and hesperetin provide health benefits such as
characterised by a C6–C3–C6 carbon skeleton, natural occurrence antioxidant, anti-inflammatory, and antimicrobial effects and pre-
in fruits, vegetables, and other plant-derived products, and their vent bone loss (Chiba et al., 2003; Garg et al., 2001).
health-promoting activities (Garg, Garg, Zaneveld, & Singla, 2001; Hesperidin can be deglycosylated by a-rhamnosidase to yield
González-Barrio et al., 2004). The health beneficial properties of rhamnose and hesperetin 7-O-glucoside (Hes-7-G), which is
flavonoids include biological activities such as anticarcinogenic, known to remove the bitterness from citrus juices and to improve
anti-diabetic, anti-inflammatory, antibacterial, immune-stimulat- the aroma in grape juices and wines (Kamiya, Esaki, & Tanaka,
ing, antiviral, bone-loss preventive and estrogenic effects, based 1985; Manzanares et al., 2003). The enzymatic bioconversion of
on the free radical scavenging and antioxidant activities of these hesperidin to Hes-7-G can change the absorption site and enhance
compounds (Chiba et al., 2003; Havsteen, 2002; Liu, 2004; Tapiero, its bioavailability in juices and tea (Arts, Sesink, Faassen-Peters, &
Tew, Nguyen Ba, & Mathé, 2002; Tripoli, La Guardia, Giammanco, Hollman, 2004; Bredsdorff et al., 2010; González-Barrio et al.,
Di Majo, & Giammanco, 2007). Among flavonoids, the flavanone 2004; Nielsen et al., 2006). The sugar moiety is proposed to be a
hesperidin is the major flavonoid present in citrus fruits such as major determinant of the absorption site and bioavailability of
sweet oranges and lemons (Brand et al., 2008; Bredsdorff et al., the flavonoid. Hes-7-G and hesperetin aglycone, produced by the
2010; González-Barrio et al., 2004). In immature oranges, the hydrolysis of enzymatic treatments, are better absorbed from the
amount of hesperidin may make up to 14% of the fresh weight of small intestine compared to hesperidin glycoside with rhamnose
the fruit (Barthe, Jourdan, McIntosh, & Mansell, 1988). Hesperidin (Nielsen et al., 2006). The bioavailability of the flavonoid has been
is structurally a flavanone b-glycosides at C-7 of hesperetin agly- demonstrated to increase when hesperidin is enzymatically trans-
cone, which is 30 ,5,7-trihydroxy-40 -methoxy flavanone. Its sugar formed into Hes-7-G (Bredsdorff et al., 2010; Day et al., 2000; Hab-
moiety is a disaccharide rutinose composed of rhamnose and auzit et al., 2009; Manach, Morand, Gil-Izquierdo, Bouteloup-
glucose, that is, O-a-L-rhamnosyl-(1 ? 6)-glucose. Supplemental Demange, & Rémésy, 2003; Németh et al., 2003). However, it is
not clearly explained why Hes-7-G is more bioavailable than native
⇑ Corresponding author. Tel.: +82 2 2123 3124; fax: +82 2 312 5229. hesperidin. After it is absorbed, the intestinal intracellular hespere-
E-mail address: soobok@yonsei.ac.kr (S.-B. Lee). tin aglycone is subsequently conjugated into glucuronidated and

0308-8146/$ - see front matter Ó 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.foodchem.2012.07.007
2254 Y.-S. Lee et al. / Food Chemistry 135 (2012) 2253–2259

sulfated metabolites, which are detected in human blood and urine 2.2. Preparation and assay of naringinase from A. sojae
(Brand et al., 2008; Gardana, Guarnieri, Riso, Simonetti, & Porrini,
2007). The naringinase enzyme from the liquid culture of A. sojae was
Naringinase (EC 3.2.1.40) is commercially useful in the pharma- purified according to a previous method (Chang et al., 2011). The
ceutical and food industries, because of the hydrolysis bioconver- crude enzyme in the culture solution was precipitated with 80%
sion of sugar moieties in antibiotics, flavonoids, or glycolipids (w/v) (NH4)2SO4 (7000g for 30 min at 4 °C), dissolved in a 50 mM
(Ribeiro & Ribeiro, 2008). Naringinase is also used in debittering sodium citrate buffer (pH 6.0), and dialysed. The protein sample
citrus juices and wine (Puri & Banerjee, 2000; Vila-Real et al., was purified by the rechromatography of a Q-Sepharose column
2007, 2010). In fact, naringinase is known to catalyse the deglyco- in a 50 mM Tris–HCl buffer (pH 7.0) at a flow rate of 1 ml/min. Ac-
sylation of many natural glycosides, including flavonoids (Ribeiro tive fractions that exhibited a-rhamnosidase activity were pooled,
& Ribeiro, 2008). The enzyme has both a-L-rhamnosidase and b- dialysed, and concentrated with an ultrafiltration membrane (Mil-
D-glucosidase activities. Recently, it has been shown that the ruti- lipore Co., Bedford, MA, USA) for further investigation.
noside, O-a-L-rhamnosyl-(1 ? 2)-glucose of naringin is hydrolysed a-L-Ramnosidase activity of the purified naringinase was spec-
by the naringinase produced by an Aspergillus sojae strain isolated trophotometrically assayed under standard conditions at 25 °C in
from a traditional Korean fermented soybean, to preferably yield a a 50 mM sodium citrate buffer (pH 6.0) using pNPaR as the sub-
rhamnose and naringenin-7-O-glucoside, prunin (Chang et al., strate for hydrolysis (Chang et al., 2011). b-D-Glucosidase activity
2011). Prunin is considered to have a potential use as a sweetening of the enzyme was also spectrophotometrically determined at
agent in diabetes therapy without concurrent undesirable bitter- 400 nm using pNPbG in the same buffer at 25 °C. The concentration
ness, and it possesses many biological activities, such as antiviral, of the liberated p-nitrophenol was determined at 400 nm with a
anti-inflammatory, hypoglycemic, and hypocholesterolemic effects diode array spectrophotometer (DU 730; Beckman Inc., Fullerton,
(Choi, Yokozawa, & Oura, 1991; Jeon, Park, & Choi, 2004; Kaul, Mid- CA, USA). One unit (1 IU) of the enzyme activity was defined as
dleton, & Ogra, 1985). The naringinase from A. sojae has generally the amount of enzyme that releases 1 lmol of p-nitrophenol per
recognised as safe (GRAS) status (Chang et al., 2011) and thus min under standard conditions. The protein concentration was
can be used in food applications. This particular form has high a- determined according to the Bradford method with bovine serum
L-rhamnosidase activity and relatively very low b-D-glucosidase albumin (BSA) as a standard (Bradford, 1976).
activity, thus acting like a-L-rhamnosidase for the hydrolysis of
rhamnose of naringin in grapefruit. 2.3. Enzymatic bioconversion of hesperidin by naringinase
In this study, the hydrolysis of O-a-L-rhamnosyl-(1 ? 6)-glucose
of hesperidin by the naringinase from A. sojae was examined. The A substrate solution (0.1–1.0 mM) of hesperidin in a 50 mM so-
bioconversion of hesperidin to the intermediate Hes-7-G was mon- dium citrate buffer (pH 6.0) containing 10% dimethylsulfoxide
itored and the intermediate was produced by batch-type reactions. (DMSO) or 10–20% ethanol was incubated with naringinase at
The biological properties of the flavonoid, including solubility and 37 °C for 12–24 h. After the reaction was terminated with 0.1 M
in vitro inhibition of enzymes and Helicobacter pylori growth, were HCl solution, the reaction mixture was centrifuged at 12,000g for
investigated and compared based on the sugar moiety. 15 min and the resulting supernatant was filtered through a
0.22-lm pore-size membrane (Millipore Co., Bedford, MA, USA).
The bioconversion of hesperidin by A. sojae naringinase was then
2. Materials and methods investigated by both thin-layer chromatography (TLC) and high
performance liquid chromatography (HPLC). TLC analysis of the
2.1. Materials, enzymes, and microorganisms hesperidin hydrolysis was performed on Whatman K5F silica gel
plates (Whatman, Kent, UK) with isopropyl alcohol–ethyl ace-
Hesperidin, hesperetin, naringin, naringenin, L-rhamnose, tate–water (3:1:0.3, v/v/v) as the solvent system (Chang et al.,
p-nitrophenyl a-L-rhamnopyranoside (pNPaR), p-nitrophenyl b-D- 2011). The TLC plate was developed, dried, and visualised by dip-
glucopyranoside (pNPbG), b-nicotinamide adenine dinucleotide ping it in a solution containing 0.3% (w/v) N-(1-naphthyl)-ethy-
20 -phosphate tetrasodium salt (NADPH), DL-3-hydroxy-3-methyl- lenediamine and 5% (v/v) H2SO4 in methanol and heating it for
glutalyl coenzyme A sodium salt (HMG-CoA), and dithiothreitol 10 min at 110 °C. HPLC analysis was carried out qualitatively and
(DTT) were purchased from Sigma–Aldrich (St. Louis, MO, USA). quantitatively to determine the hesperidin and the hydrolysed
The juice and peel of oranges, which were purchased at a local products, such as Hes-7-G and hesperetin, on a Waters symmetry
supermarket, were prepared by slight modifications of the previ- C18 column (4.6  250 mm; Waters Co., Milford, MA, USA) con-
ous method (Gil-Izquierdo, Gil, & Ferreres, 2002). The juice of or- nected to a Waters model 510 system with a 996 photodiode array
anges was obtained by squeezing and centrifuging fruit to collect detector at 254 nm (Ribeiro & Ribeiro, 2008). The linear HPLC gra-
the supernatant. The peel of the oranges was blended, extracted dient was composed of two solvents, solvent A and solvent B, 10%
with 80% ethanol in an ultrasonic bath for 2 h at room temperature and 90% (v/v) acetonitrile, respectively, in water. After a 20 ll
and concentrated to remove ethanol by evaporisation. The juice injection of each sample, solvent B was increased from 10% to
supernatant and peel extract of oranges were lyophilised for fur- 30% in 20 min, increased to 100% in 5 min, held at 100% for
ther use. Naringinase from Penicillium decumbens was also ob- 10 min, and then reduced to 10% in 10 min. The time-dependent
tained from Sigma–Aldrich and dialysed against 50 mM Tris–HCl change in the composition of the reaction mixture that was period-
buffer (pH 7.0), followed by concentration for use. Human intesti- ically sampled was also analysed by HPLC. The hesperidin
nal maltase, which is an N-terminal catalytic domain of intestinal (1.3 mM) was incubated with naringinase (0.07 mg/ml) in a
maltase-glucoamylase, was kindly provided by Professor Doman 50 mM sodium citrate buffer (pH 6.0) containing 10% DMSO at
Kim at Chonnam National University after expression in Pichia pas- 37 °C for 12 h.
toris and subsequent purification (Ryu, Seo, Kang, Kim, & Kim, The lyophilised extracts of the juice and peel of oranges were
2011; Sim, Quezada-Calvillo, Sterchi, Nichols, & Rose, 2008). Hu- treated with 10% DMSO and 10% ethanol, respectively, to solve
man HMG-CoA reductase (the catalytic domain) was also obtained the hesperidin flavanone in orange. Hesperidin (0.17 mM dissolved
from Sigma–Aldrich. H. pylori ATCC 43526 was purchased from the in DMSO for the juice extract and 0.77 mM in ethanol for the peel
Korean Culture Center of Microorganisms (KCCM). All other chem- extract) was incubated at 37 °C for 24 h with crude enzyme solu-
icals used were of analytical grade. tion (1.7–2 mg/ml), which was prepared by the ammonium sulfate
Y.-S. Lee et al. / Food Chemistry 135 (2012) 2253–2259 2255

precipitation, dialysis, and concentration of the culture broth for A. compared (Chang et al., 2011). The enzyme was stored at 80 °C
sojae. After the reaction was terminated by 0.1 M HCl, the reaction until analysis. HMG-CoA-dependent oxidation of NADPH was mon-
solution was filtered through a 0.22 lm membrane filter (Millipore itored for 30 min at 340 nm with a DU 730 spectrophotometer
Co.) and analysed by HPLC. (Beckman Inc., Fullerton, CA, USA) equipped with a cell holder that
was maintained at 25 °C. The standard assay mixture (100 ll) con-
2.4. Repetitive batch-type production of Hes-7-G tained 200 lM HMG-CoA, 200 lM NADPH, 100 mM NaCl, 1.0 mM
EDTA, 10 mM DTT, and 100 mM sodium phosphate buffer (pH
To produce Hes-7-G from hesperidin and increase the enzyme 6.8). All reaction components, including the enzyme (10 lg/ml)
productivity, the enzymatic reaction was performed by a repetitive without HMG-CoA, were first monitored to detect potential
batch reaction that recycled the enzyme according to the previous HMG-CoA-independent oxidation of NADPH. The reaction was
method of other reports (Chang et al., 2011; Kragl et al., 1993). For then initiated by the addition of HMG-CoA (0.05–2 mM) to the
one cycle of the batch reaction in a total volume of 6 ml, naringin- reaction mixture. One unit of HMG-CoA reductase is defined as
ase (0.1 mg/ml) was incubated with 0.12 mM hesperidin in a the amount of enzyme that catalyses the oxidation of 1 lmol of
50 mM sodium citrate buffer containing 10% ethanol at 37 °C for NADPH per min. The HMG-CoA-dependent oxidation reactions
12 h. The enzyme was repeatedly used in an ultrafiltration mem- were performed with various concentrations (0–80 lM) of hesper-
brane (Amicon Ultra-15 device, molecular weight cut-off of idin, Hes-7-G, hesperetin, and naringin as the inhibitors.
10 kDa; Millipore Co.). The solution bearing the reaction product
was separated by a centrifugation (5000g) at the end of each batch.
2.7. In vitro growth inhibition of H. pylori
The batch reaction was also repeated by the addition of new sub-
strate solution. The filtrate for each cycle was monitored by HPLC
The effects of the hesperidin and naringin flavanones (0.05–
and freeze-dried. This repetitive batch process was continuously
0.5 mM), such as hesperidin, Hes-7-G, hesperetin, naringin, prunin,
performed for up to 17 cycles.
and naringenin, on the growth of H. pylori were examined. H. pylori
The major reaction products, Hes-7-G and rhamnose, were puri-
was cultured in a test tube containing 5-ml of sterilised brain–
fied from the reaction mixture using a Prevail Carbohydrate ES col-
heart-infusion broth (BHI; Difco Laboratories, Detroit, MI, USA)
umn (300  20 mm; Alltech Associates, Inc., Deerfield, IL, USA) that
supplemented with 5% horse serum (Sigma–Aldrich) with shaking
was connected to a Waters Delta Prep 4000 preparative chroma-
(150 rpm) at 37 °C for 3 days. The test tube was contained in a Gas-
tography system with a Waters™ 486 absorbance detector at
Pak EZ Pouch System (5–15% CO2; BD Diagnostics, MD, USA) under
254 nm (Waters Co.). A linear gradient in the HPLC was employed
microaerophilic conditions. At 3 days, 50 ll of culture solution was
using two solvents of 20% (v/v) acetonitrile in water (A) and 80% (v/
cultivated further on a sterilised 96 well-microplate (Nunclon™D
v) acetonitrile in water (B) for 60 min at a flow rate of 4 ml/min at
Surface, Thermo Fisher Scientific, Denmark) containing 180 ll
room temperature. The solvent B was increased from 10% to 30% in
BHI broth supplemented with 5% horse serum and a 20 ll flava-
25 min, also increased to 100% in 10 min, held at 100% for 10 min,
none solution (or distilled water in the control) per well. The 96
and then reduced to 10% in 10 min. The fraction containing Hes-7-
well-microplate with lid was incubated under 10% CO2 at 37 °C.
G was collected, evaporated to remove the acetonitrile, and freeze-
The growth of bacteria in the microplate was determined by mea-
dried. The homogeneity of the purified products was identified by
suring an optical density (OD) at 620 nm several times from 0 to
the TLC and HPLC analyses described above.
80 h.

2.5. Solubility
3. Results and discussion
Excess hesperidin or purified reaction product, such as Hes-7-G
and hesperetin, was mixed with 200 ll of distilled deionised water 3.1. Enzymatic bioconversion of hesperidin
(DDW) or 10% (v/v) ethanol in water by vortexing at 25 °C, fol-
lowed by incubation for 1 h in an ultrasonic bath (5510-DTH; Bran- Hesperidin is an abundant and inexpensive byproduct of citrus
son, Danbury, CT, USA) to maximise the solubility of the compound cultivation, which can be obtained from waste orange peel from
(Chang et al., 2011; Li, Park, Park, Park, & Park, 2004). After sonica- the citrus industry (Garg et al., 2001). The bioavailability of hesper-
tion, each sample was filtered through a 0.22-lm pore-size mem- idin was shown to increase in human subjects when hesperidin
brane (Millipore Co.) and the solution concentration was measured was converted to Hes-7-G (Bredsdorff et al., 2010). In this study,
by HPLC analysis. the A. sojae naringinase (GRAS) was the enzyme used to treat hes-
peridin. The purified naringinase derived from A. sojae converted
2.6. In vitro inhibition of human intestinal maltase and HMG-CoA the hesperidin mainly to the intermediate Hes-7-G (Fig. 1). The su-
reductase gar moiety of hesperidin is an O-a-L-rhamnosyl-(1 ? 6)-glucose,
which is attached to the 7-OH group of hesperetin aglycone. The
The in vitro inhibition of human intestinal maltase was investi- purified naringinase from A. sojae can hydrolyse naringin that con-
gated and compared by hesperidin and naringin flavanones, such tains an O-a-L-rhamnosyl-(1 ? 2)-glucose as the sugar moiety,
as hesperidin, Hes-7-G, hesperetin, naringin, punin, and naringe- mainly removing the rhamnose and producing the intermediate
nin, depending on the flavanone structure and the sugar moiety. prunin (Chang et al., 2011). Irrespective of a-L-rhamnnose linkage,
The reaction of human intestinal maltase was assayed using malt- A. sojae naringinase could hydrolyse the rutinoside disaccharide of
ose at 0.33–3.5 mM as a substrate, and the various flavanones at 0– hesperidin to convert it to a glucoside, which were determined by
5 mM as inhibitors. Reactions were carried out in 50 mM sodium HPLC (Fig. 2). Thus, hesperidin was efficiently converted to the
phosphate buffer (pH 7.0) at 25 °C. The released glucose in the intermediate Hes-7-G by the rhamnosidase activity of A. sojae
reaction was determined by using the modified glucose oxidase/ naringinase with scarcely any hesperetin aglycone produced. A
peroxidase method (Kim et al., 1999). The type of inhibition and time-dependent bioconversion of hesperidin by A. sojae naringin-
the inhibition constant (Ki) were determined using Dixon plots ase was monitored in 50 mM sodium citrate buffer (pH 6.0) con-
by a non-linear regression program, DNRPEASY (Duggleby, 1984). taining 10% DMSO at 37 °C for 12 h, as shown in Fig. 3A. The
The in vitro inhibition of human HMG-CoA reductase by hesper- results show that the enzymatic bioconversion of hesperidin by
idin, Hes-7-G, hesperetin, and naringin was also investigated and the purified naringinase produces gradually increasing levels of
2256 Y.-S. Lee et al. / Food Chemistry 135 (2012) 2253–2259

Hes-7-G
A 0.4
0.35 1
rhamnose
0.3
0.25
hesperetin 0.2
hesperidin 0.15
glucose
0.1
G1 0.05
G2 0
G3 -0.05 0 10 20 30 40 50
G4 Time (min)
G5
G6
G7 B 0.2
2
0.15 1

S 1 2 3 4 5 6 0.1

Fig. 1. TLC analysis of the reaction of A. sojae purified naringinase with hesperidin 0.05
at 37 °C. Lane S, standard maltooligosaccharides (G1–G7); lane 1, a-L-rhamnose;
lane 2, hesperidin; lane 3, reaction of A. sojae naringinase with hesperidin for 12 h;
0
lane 4, reaction of P. decumbens naringinase with hesperidin for 12 h; lane 5, Hes-7-
G; lane 6, hesperetin. The reaction was carried out with naringinase at 0.07 mg/ml 0 10 20 30 40 50
and hesperidin at 1 mM in 50 mM sodium citrate buffer (pH 6.0) containing 10% -0.05
DMSO.
Time (min)

Hes-7-G, reaching a plateau (1.2 mM) within 12 h under the de-


scribed condition. A very small amount of hesperidin and hespere- C 0.3
2
tin were detectable (each 0.07 mM, not more than 5%) at the end of 0.25
the reaction. The molar production yield of Hes-7-G based on the
0.2
molarity of the hesperidin used was found to be approximately
90%. To increase the enzyme productivity, the production of Hes- 0.15
7-G was effectively performed by repetitive batch reactions with 0.1
enzyme recycling. As shown in Fig. 3B, the hesperidin (0.13 mM) 3
was dissolved and reacted in each batch (6 ml) of 50 mM sodium 0.05
citrate buffer containing 10% ethanol. This organic solvent is gener- 0
ally used in the food industry and the hesperidin was easily con-
-0.05 0 10 20 30 40 50
verted mainly to Hes-7-G. The enzyme activity was almost
constantly maintained up to 12th batch cycle, and was then largely Time (min)
decreased by 40% of the original activity at the 13th cycle, a level
that was maintained up to the 17th cycle. The total molar produc- Fig. 2. HPLC analysis on the reaction of purified A. sojae naringinase with hesperidin
tion yield of Hes-7-G based on hesperidin (mM) was, on average, at 37 °C. (A) Hesperidin (B) reaction of A. sojae naringinase with hesperidin for 6 h
86% for 12 batch cycles. The stability of the enzyme might be de- (C) reaction of A. sojae naringinase with hesperidin for 12 h. 1, hesperidin; 2, Hes-7-
G; 3, hesperetin.
creased by the organic solution in the repetitive reaction. The
Hes-7-G produced from the reaction solution (5 ml) was easily
purified from the rhamnose product and unreacted hesperidin in et al., 2004). Our results suggest that the GRAS A. sojae naringinase
a single step by Prep-LC, respectively. The purity and homogeneity can be applied to the flavanones from citrus fruit containing hes-
of the separated Hes-7-G was estimated by HPLC to be above peridin as well as naringin to convert the flavanones (flavonoid
99.9%. The rhamnose of the rutinoside was concomitantly pro- rutinosides) to the corresponding glucosides.
duced in an equimolar amount and separately obtained.
Extracts from the peel and juice of oranges containing the
hesperidin were subjected to the treatment of the crude enzyme 3.2. Properties of bioconverted products
solution from A. sojae. As shown in Fig. 4, the hesperidin detected
in the juice and peel of oranges was efficiently converted to Hes-7- 3.2.1. Solubility
G within 12 h, with negligible hesperetin also produced. The molar The solubility of hesperidin was compared with that of both
production yields of Hes-7-G, based on the molarity of hesperidin Hes-7-G and hesperetin in distilled deionised water (DDW) and
existed in the extracts, was evaluated to be approximately 71% for 10% ethanol solution at 25 °C (Table 1). The solubility of Hes-7-G
orange juice and 78% for orange peel, respectively. In fact, the en- was remarkably higher than that of hesperidin approximately as
zyme solution of A. sojae contains some independent b-glucosidase 36-fold greater in DDW and 55-fold greater in 10% ethanol. The
which can convert Hes-7-G to hesperetin. If the purified naringin- solubility of Hes-7-G was also approximately 77-fold greater than
ase was used, the production yield of Hes-7-G would have hesperetin in DDW and 88-fold greater in 10% ethanol. In fact, a
increased. The enzymatic conversion of the flavanones in orange glycosylation of aglycone is well known to increase the water
juice was reported to be performed with 50% conversion of rutino- solubility of flavonoids and other aglycones. The hydrolysis of the
side to the corresponding flavonoid glucoside (González-Barrio a-L-rhamnosyl moiety in hesperidin likely results in the
Y.-S. Lee et al. / Food Chemistry 135 (2012) 2253–2259 2257

A A 0.2
1.5
Concentration (mM)

0.15

Concentration (mM)
1

0.1
0.5

0.05

0 2 4 6 8 10 12 14 0
0 4 8 12 16 20 24 28
Time (h)
Time (h)
B 0.15
B 1
Concentration (mM)

0.1 0.8

Concentration (mM)
0.6
0.05
0.4

0 0.2
0 2 4 6 8 10 12 14 16 18
0
Batch cycle 0 4 8 12 16 20 24 28
Fig. 3. Time progress of A. sojae purified naringinase reaction with hesperidin (A) Time (h)
and repetitive batch-type production of Hes-7-G (B). The reaction was performed in
50 mM sodium citrate buffer (pH 6.0) containing 10% DMSO (A) and 10% ethanol (B)
Fig. 4. Time course analyses of A. sojae crude naringinase reactions with hesperidin
with naringinase at 0.07 mg/ml at 37 °C for 12 h. The reaction mixture was analysed
in the extracts of orange juice (A) and orange peel (B). Hesperidin, j; Hes-7-G, d.
by HPLC. Hesperidin, j; Hes-7-G, d; hesperetin, D.
The reaction was performed in 50 mM sodium citrate buffer (pH 6.0) containing
10% DMSO (juice) or 10% ethanol (peel) with crude naringinase at 37 °C for 24 h.
The reaction mixture was analysed by HPLC.
enhancement of solubility. It is generally accepted that the solubil-
ity of naringin is largely increased by its conversion to prunin due
to the removal of a-L-rhamnose during the hydrolysis (Chang et al., Table 1
2011). The enhanced solubility of Hes-7-G is a desirable character- Comparison of solubilities of hesperidin, Hes-7-G, and hesperetin at 25 °C.
istic for practical applications in food or pharmacological indus- Compound Solubility in DDWa (lM) Solubility in 10% ethanol (lM)
tries, which might be capable of improving its bioavailability
Hesperidin 19.2 ± 0.1 37.5 ± 0.1
through greater absorption as compared to hesperidin and agly- Hes-7-G 688.4 ± 30.8 2063.3 ± 28.4
cone hesperetin. In addition, the absorption of Hes-7-G was sug- Hesperetin 8.9 ± 0.2 23.5 ± 0.1
gested to be actively performed in the small intestine rather than a
DDW indicates distilled deionised water.
the colon, which was understood on the absorption model through
deglycosylation of flavonoid glucoside at the epithelial cell mem-
brane of the small intestine (González-Barrio et al., 2004; Németh
with inhibition constants (Ki) of 1.0–4.4 mM (Table 2). The inhibi-
et al., 2003). The bioavailability of the hesperidin in orange juice
tion constants of the flavanone glucosides, Hes-7-G and prunin,
was increased by the enzymatic removal of the a-L-rhamnosyl
were slightly lower than those of the flavanone rutinosides and
moiety from the rutinosides to the glucosides (Bredsdorff et al.,
aglycones. Therefore, the intermediate compounds, Hes-7-G and
2010; González-Barrio et al., 2004). Therefore, the enzymatic bio-
prunin, were more potent than the other flavanones tested. These
conversion of hesperidin to Hes-7-G using A. sojae naringinase
results indicate that the flavanones were slightly different at bind-
may be useful for enhancing its potential bioavailability in citrus
ing the enzyme active site, depending on the type of sugar moiety
food products, as well as the bioconversion of naringin to prunin
of the flavanone competing with the maltose substrate. It has also
(Chang et al., 2011).
been indicated that the glucose moiety of the flavanones may be
preferred at the binding of the enzyme and thus effective at inhib-
3.2.2. In vitro inhibitions of human intestinal maltase and HMG-CoA iting the human brush border intestinal maltase and therefore
reductase potentially interfere with intestinal digestive functions.
Flavonoids including the flavanones were known to inhibit rat In addition, the in vitro inhibition of human HMG-CoA reductase
small intestinal and yeast a-glucosidases (Tadera, Minami, by Hes-7-G was compared to the inhibition by hesperidin, hespere-
Takamatsu, & Matsuoka, 2006). The flavanones, hesperidin and tin, and naringin. Ki values for these compounds were in the range
naringin, as well as their intermediate compounds and aglycones, of 10–44 lM (Table 2, Supplementary material). HMG-CoA reduc-
mildly inhibited human intestinal maltase. The flavanones tested tase is a target enzyme for cholesterol-lowering compounds,
all exhibited competitive inhibition of human intestinal maltase because it is known to be involved in a rate-limiting step in
2258 Y.-S. Lee et al. / Food Chemistry 135 (2012) 2253–2259

Table 2 would be a more effective mild anti-diabetic and cholesterol-low-


Comparison of inhibition constants (Ki) of flavanones against human intestinal ering flavanone with higher solubility than hesperidin. The greater
maltase and human HMG-CoA reductase.
amount of hesperidin in citrus food may be used by the enzymatic
Compound Human intestinal maltase Human HMG-CoA reductase conversion to Hes-7-G.
Inhibition patterna Ki (mM)b Inhibition pattern Ki (lM)
Hesperidin Competitive 3.1 ± 0.7 Non-competitive 19.0 ± 0.1 3.2.3. In vitro inhibitory effects of flavanones on the growth of H. pylori
Hes-7-G Competitive 1.8 ± 0.1 Non-competitive 9.8 ± 1.2 As shown in Fig. 5, the flavanones, hesperidin and naringin, as
Hesperetin Competitive 4.4 ± 0.6 Non-competitive 9.6 ± 0.4
well as their intermediate compounds and aglycones, showed
Naringin Competitive 2.8 ± 0.1 Non-competitive 43.7 ± 2.3
Prunin Competitive 1.0 ± 0.1 –c –
some inhibitory effects on the growth of H. pylori. The antibacterial
Naringenin Competitive 3.8 ± 0.9 – – activity of flavonoids is being increasingly investigated with
a,b
medicinal significance (Cushnie & Lamb, 2005). The flavonoids
Inhibition pattern and constants (Ki values) were determined by Dixon plots for
the activity of human intestinal maltase (HIM) or human HMG-CoA reductase (HCR)
have inhibitory effects on the growth of H. pylori, which is a path-
in the presence of flavanones (0–5 mM for HIM and 0–80 lM for HCR) as an ogenic bacterium that causes chronic gastric inflammation and
inhibitor. promotes carcinogenesis (Lee, Shin, & Hahm, 2008). The growth-
c
Not determined. inhibiting effect of Hes-7-G was significant and similar to that of
hesperidin, but better than hesperetin at 0.5 mM. Due to the lim-
ited solubility, hesperetin existed in an undissolved state, probably
resulting in relatively lower effects than Hes-7-G and hesperidin at
A 120 this concentration. In addition, the growth-inhibiting effect of
prunin was better than those of naringin and naringenin at
100 0.5 mM. Even though naringenin was restricted in solubility and
Relative growth (%)

undissolved, the growth-inhibiting effect of naringenin was


80 slightly better than that of naringin at 0.5 mM. The enhancement
of the intermediate compounds, Hes-7-G and prunin, at inhibiting
60 the growth of H. pylori might be mainly attributed to the increased
solubilities of the intermediate compounds. The more soluble form
40 of the flavanone might yield greater effects on the inhibition of
bacterial growth, even at the same concentration. The inhibition
20 mechanism of the flavanones against the growth of H. pylori is un-
clear. However, it has been suggested that the antibacterial mech-
0 anism of flavonoids, including some flavanones, is through their
0 12 24 36 48 60 72 84
inhibition of nucleic acid synthesis or cytoplasmic membrane
function (Cushnie & Lamb, 2005). For instance, quercetin exhibits
Time (h)
an inhibition of bacterial DNA gyrase (Ohemeng, Schwender, Fu,
& Barrett, 1993) and an increase in the permeability of the inner
B 120
bacterial membrane (Mirzoeva, Grishanin, & Calder, 1997), and
naringenin exhibits bacterial membrane interference and an
inhibition of bacterial motility (Mirzoeva et al., 1997; Tsuchiya &
100
Iinuma, 2000).
Relative growth (%)

In conclusion, Hes-7-G could be easily produced from hesperi-


80
din by the enzymatic reaction of A. sojae naringinase. Hesperidin
in the extracts of orange juice and peel could also be enzymatically
60
converted to the intermediate Hes-7-G compound. The solubility of
Hes-7-G was much higher than that of either hesperidin or hes-
40
peretin, which may make it more bioavailable. Hes-7-G was also
20
approximately 2-fold better than hesperidin in the inhibition of
human intestinal maltase and HMG-CoA reductase. In addition,
0 Hes-7-G showed superior inhibition of the growth of H. pylori com-
0 20 40 60 80 100 pared to hesperetin. Thus, the intermediate compound, Hes-7-G, is
suggested to be a more bioavailable and functional flavanone with
Time (h) enhanced solubility.

Fig. 5. Inhibitory effects of flavanones on the growth of H. pylori. (A) Relative


growth in the presence (0.5 mM) of hesperidin (j), Hes-7-G (d), and hesperetin (N) Acknowledgments
(B) relative growth in the presence (0.5 mM) of naringin (j), prunin (d), and
naringenin (N). Relative growth of the control () indicates 100% at 72-h (A) and 80- This research was supported by Basic Science Research Program
h (B) in the absence of flavanones.
through the National Research Foundation of Korea (NRF) Grant
funded by the Korean Government (MEST) (2009-0077358), and
partly by the Marine and Extreme Genome Research Center Pro-
cholesterol biosynthesis (Hilleman, Wurdeman, & Lenz, 2001). The
gram, Ministry of Maritime Affairs and Fisheries, Republic of Korea.
decrease in the inhibition constant of Hes-7-G in comparison with
that of hesperidin might imply that the glucose sugar moiety of the
intermediate flavanone more suitably bound to the enzyme than Appendix A. Supplementary data
the rutinose disaccharide moiety of hesperidin. The inhibitory ef-
fect of hesperidin was better than naringin, indicating that the Supplementary data associated with this article can be found, in
aglycone part might be partly preferable for the inhibition of the online version, at http://dx.doi.org/10.1016/j.foodchem.2012.
HMG-CoA reductase. Finally, our results suggest that Hes-7-G 07.007.
Y.-S. Lee et al. / Food Chemistry 135 (2012) 2253–2259 2259

References Kamiya, S., Esaki, S., & Tanaka, R. (1985). Synthesis of certain disaccharides
containing a a- or b-L-rhamnopyranosidic group and the substrate specificity of
a-L-rhamonosidase from Aspergillus niger. Agricultural and biological chemistry,
Arts, I. C., Sesink, A. L., Faassen-Peters, M., & Hollman, P. C. (2004). The type of sugar
49, 2351–2358.
moiety is a major determinant of the small intestinal uptake and subsequent
Kaul, T. N., Middleton, E., & Ogra, P. L. (1985). Antiviral effect of flavonoids on
biliary excretion of dietary quercetin glycosides. British Journal of Nutrition, 91,
human viruses. Journal of Medical Virology, 15, 71–79.
841–847.
Kragl, U., Gödde, A., Wandrey, C., Kinzy, W., Cappon, J. J., & Lugtenburg, J. (1993).
Barthe, G. A., Jourdan, P. S., McIntosh, C. A., & Mansell, R. L. (1988).
Repetitive batch as an efficient method for preparative scale enzymic synthesis
Radioimmunoassay for the quantitative determination of hesperidin and
of 5-azido-neuraminic acid and 15N-L-glutamic acid. Tetrahedron-Asymmetry, 4,
analysis of its distribution in Citrus sinensis. Phytochemistry, 27, 249–254.
1193–1202.
Brand, W., van der Wel, P. A., Rein, M. J., Barron, D., Williamson, G., van Bladeren, P.
Lee, S.-Y., Shin, Y. W., & Hahm, K.-B. (2008). Phytoceutical: Mighty but ignored
J., et al. (2008). Metabolism and transport of the citrus flavonoid hesperetin in
weapons against Helicobacter pylori infection. Journal of digestive diseases, 9,
caco-2 cell monolayers. Drug Metabolism and Disposition, 36, 1794–1802.
129–139.
Bredsdorff, L., Nielsen, I. L. F., Rasmussen, S. E., Cornett, C., Barron, D., Bouisset, F.,
Li, D., Park, J.-H., Park, J.-T., Park, C. S., & Park, K.-H. (2004). Biotechnological
et al. (2010). Absorption, conjugation and excretion of the flavanones,
production of highly soluble daidzein glycosides using Thermotoga maritima
naringenin and hesperetin from a-rhamnosidase-treated orange juice in
maltosyltransferase. Journal of Agricultural and Food Chemistry, 52, 2561–2567.
human subjects. British Journal of Nutrition, 103, 1602–1609.
Liu, R. H. (2004). Potential synergy of phytochemicals in cancer prevention:
Chang, H.-Y., Lee, Y.-B., Bae, H.-A., Huh, J.-Y., Nam, S.-H., Sohn, H.-S., et al. (2011).
Mechanism of action. Journal of Nutrition, 134, 3479–3485.
Purification and characterisation of Aspergillus sojae naringinase: The
Manach, C., Morand, C., Gil-Izquierdo, A., Bouteloup-Demange, C., & Rémésy, C.
production of prunin exhibiting markedly enhanced solubility with in vitro
(2003). Bioavailability in humans of the flavanones hesperidin and narirutin
inhibition of HMG-CoA reductase. Food Chemistry, 124, 234–241.
after the ingestion of two doses of orange juice. European Journal of Clinical
Chiba, H., Uehara, M., Wu, J., Wang, X., Masuyama, R., Suzuki, K., et al. (2003).
Nutrition, 57, 235–242.
Hesperidin, a citrus flavonoid, inhibits bone loss and decreases serum and
Manzanares, P., Orejas, M., Gil, J. V., de Graaff, L. H., Visser, J., & Ramon, D. (2003).
hepatic lipids in ovariectomized mice. Journal of Nutrition, 133, 1892–1897.
Construction of a genetically modified wine yeast strain expressing the
Choi, J. S., Yokozawa, T., & Oura, H. (1991). Improvement of hyperglycemia and
Aspergillus aculeatus rhaA gene, encoding an a-L-rhamnosidase of enological
hyperlipemia in streptozotocin-diabetic rats by a methanolic extract of Prunus
interest. Applied and Environmental Microbiology, 69, 7558–7562.
davidiana stems and its main component, prunin. Planta Medica, 57, 208–211.
Mirzoeva, O. K., Grishanin, R. N., & Calder, P. C. (1997). Antimicrobial action of
Cushnie, T. P. T., & Lamb, A. J. (2005). Antimicrobial activity of flavonoids.
propolis and some of its components: The effects on growth, membrane
International Journal of Antimicrobial Agents, 26, 343–356.
potential and motility of bacteria. Microbiological Research, 152, 239–246.
Day, A. J., Canada, F. J., Diaz, J. C., Kroona, P. A., Mclauchlana, R., Fauldsa, C. B., et al.
Németh, K., Plumb, G. W., Berrin, J. G., Juge, N., Jacob, R., Naim, H. Y., et al. (2003).
(2000). Dietary flavonoid and isoflavone glycosides are hydrolysed by the
Deglycosylation by small intestinal epithelial cell beta-glucosidases is a critical
lactase site of lactase phlorizin hydrolase. FEBS Letters, 468, 166–170.
step in the absorption and metabolism of dietary flavonoid glycosides in
Duggleby, R. G. (1984). Regression analysis of nonlinear Arrhenius plots: An
humans. European Journal of Nutrition, 42, 29–42.
empirical model and a computer program. Computers in Biology and Medicine,
Ohemeng, K. A., Schwender, C. F., Fu, K. P., & Barrett, J. F. (1993). DNA gyrase
14, 447–455.
inhibitory and antibacterial activity of some flavones (1). Bioorganic & Medicinal
Gardana, C., Guarnieri, S., Riso, P., Simonetti, P., & Porrini, M. (2007). Flavanone
Chemistry Letters, 3, 225–230.
plasma pharmacokinetics from blood orange juice in human subjects. British
Ribeiro, I. A., & Ribeiro, M. H. L. (2008). Naringin and naringenin determination and
Journal of Nutrition, 98, 165–172.
control in grapefruit juice by a validated HPLC method. Food Control, 19,
Garg, A., Garg, S., Zaneveld, L. J. D., & Singla, A. K. (2001). Chemistry and
432–438.
pharmacology of the citrus bioflavonoid hesperidin. Phytotherapy Research, 15,
Ryu, H.-J., Seo, E.-S., Kang, H.-K., Kim, Y.-M., & Kim, D. (2011). Expression,
655–669.
purification, and characterization of human intestinal maltase secreted from
Gil-Izquierdo, A., Gil, M. I., & Ferreres, F. (2002). Effect of processing techniques at
Pichia pastoris. Food Science and Biotechnology, 20, 561–565.
industrial scale on orange juice antioxidant and beneficial health compounds.
Sim, L., Quezada-Calvillo, R., Sterchi, E. E., Nichols, B. L., & Rose, D. R. (2008). Human
Journal of Agricultural and Food Chemistry, 50, 5107–5114.
intestinal maltase-glucoamylase: Crystal structure of the N-terminal catalytic
González-Barrio, R., Trindade, L. M., Manzanares, P., de Graaff, L. H., Tomás-
subunit and basis of inhibition and substrate specificity. Journal of Molecular
Barberán, F. A., & Espín, J. C. (2004). Production of bioavailable flavonoid
Biology, 375, 782–792.
glucosides in fruit juices and green tea by use of fungal alpha-L-rhamnosidases.
Tadera, K., Minami, Y., Takamatsu, K., & Matsuoka, T. (2006). Inhibition of a-
Journal of Agricultural and Food Chemistry, 52, 6136–6142.
glucosidase and a-amylase by flavonoids. Journal of Nutritional Science and
Habauzit, V., Nielsen, I. L., Gil-Izquierdo, A., Trzeciakiewicz, A., Morand, C., Chee, W.,
Vitaminology, 52, 149–153.
et al. (2009). Increased bioavailability of hesperetin-7-glucoside compared with
Tapiero, H., Tew, K. D., Nguyen Ba, G., & Mathé, G. (2002). Polyphenols: Do they play
hesperidin results in more efficient prevention of bone loss in adult
a role in the prevention of human pathologies ? Biomedicine & Pharmacotherapy,
ovariectomised rats. British Journal of Nutrition, 102, 976–984.
56, 200–207.
Havsteen, B. H. (2002). The biochemistry and medical significance of the flavonoids.
Tripoli, E., La Guardia, M., Giammanco, S., Di Majo, D., & Giammanco, M. (2007).
Pharmacology & Therapeutics, 96, 67–202.
Citrus flavonoids: Molecular structure, biological activity and nutritional
Hilleman, D. E., Wurdeman, R. L., & Lenz, T. L. (2001). Therapeutic change of HMG-
properties: A review. Food Chemistry, 104, 466–479.
CoA reductase inhibitors in patients with coronary artery disease.
Tsuchiya, H., & Iinuma, M. (2000). Reduction of membrane fluidity by antibacterial
Pharmacotherapy, 21, 410–415.
sophoraflavanone G isolated from Sophora exigua. Phytomedicine, 7, 161–165.
Jeon, S.-M., Park, Y. B., & Choi, M.-S. (2004). Antihypercholesterolemic property of
naringin alters plasma and tissue lipids, cholesterol-regulating enzymes, fecal
sterol and tissue morphology in rabbits. Clinical Nutrition, 23, 1025–1034.

Você também pode gostar