You are on page 1of 10

Article

pubs.acs.org/JAFC

Polymethoxyflavones Isolated from the Peel of Miaray Mandarin


(Citrus miaray) Have Biofilm Inhibitory Activity in Vibrio harveyi
Ram M. Uckoo, G. K. Jayaprakasha,* Amit Vikram, and Bhimanagouda S. Patil*
Vegetable and Fruit Improvement Center, Department of Horticultural Sciences, Texas A&M University, 1500 Research Parkway
Suite A120, College Station, Texas 77845, United States
*
S Supporting Information

ABSTRACT: Citrus fruits are a good source of bioactive compounds with numerous beneficial biological activities. In the
present study, fruits of the unexplored Miaray mandarin were used for the isolation of 10 bioactive compounds. Dried peels were
sequentially extracted with hexane and chloroform in a Soxhlet-type apparatus for 8 h. The extracts were concentrated under
vacuum and separated by flash chromatography to obtain nine polymethoxyflavones and a limonoid. The purity of each
compound was analyzed by high-performance liquid chromatography (HPLC), and the compounds were identified by spectral
analysis using MALDI-TOF-MS and NMR. The isolated compounds were identified as 5-hydroxy-3,7,3′,4′-tetramethoxyflavone,
5,6,7,8,4′-pentamethoxyflavone (tangeretin), 3,5,6,7,8,3′,4′-heptamethoxyflavone, 5,6,7,8,3′,4′-hexamethoxyflavone (nobiletin),
3,5,7,8,3′,4′-hexamethoxyflavone, 3,5,7,3′,4′-pentamethoxyflavone (pentamethylquercetin), 5,7,4′-trimethoxyflavone, 5,7,8,4′-
tetramethoxyflavone, 5,7,8,3′,4′-pentamethoxyflavone, and limonin. These compounds were further tested for their ability to
inhibit cell−cell signaling and biofilm formation in Vibrio harveyi. Among the evaluated polymethoxyflavones, 3,5,6,7,8,3′,4′-
heptamethoxyflavone and 3,5,7,8,3′,4′-hexamethoxyflavone inhibited autoinducer-mediated cell−cell signaling and biofilm
formation. These results suggest that Miaray mandarin fruits are a good source of polymethoxyflavones. This is the first report on
the isolation of bioactive compounds from Miaray mandarin and evaluation of their biofilm inhibitory activity as well as isolation
of pentamethylquercetin from the Citrus genus.
KEYWORDS: citrus, flash chromatography, identification, pentamethylquercetin, polymethoxyflavones, biofilm

■ INTRODUCTION
Citrus fruits including oranges, mandarins, grapefruits, limes,
antitumor, antiviral, and anti-inflammatory activities.12−14
Nobiletin, a polymethoxyflavone present in the majority of
and lemons are some of the most-consumed fruits in the world. citrus fruits, is reported to have potential antidementia activity
These fruits are excellent sources of phytochemicals such as in both in vitro and in vivo studies.15 However, a major
flavonoids, limonoids, organic acids, and carotenoids.1−4 The bottleneck in investigating the potential biological activity of
mandarin (Citrus reticulata) citrus cultivars have a loose peel, polymethoxyflavones is their unavailability in large-scale
bright orange-red color, and juicy, succulent segment quantity and the relative impurity of most commercial
membranes. They are among the most rapidly increasing citrus preparations. Prior research from our laboratory reported the
varieties in terms of worldwide production and consumption.5 development of chromatographic methods for the analysis of
Native to Southeast Asia, mandarins are commercially amines in different mandarin species and the isolation of
cultivated in the temperate regions of the world,6 with annual polymethoxyflavones from mandarin, grapefruit, and or-
production estimated at 20.3 million metric tons.7 Due to the ange.16,17 Our results indicate that mandarins are a rich source
economic benefits of cultivation of mandarins, these fruits have of polymethoxyflavones and could be exploited for large-scale
been extensively studied to enhance their yield and quality. isolation. Polymethoxyflavones primarily occur in the fruit peels
Mandarins typically show high diversity due to hybridization and to a lesser extent in the juice. These compounds are also
and somatic mutations.8−10 These factors have led to significant reported to be present in high levels in the fruits of Shiikuwasha
phenotypic and genetic variation, resulting in numerous species. (Citrus depressa) and Calamondin (Citrus madurensis).18,19
Miaray mandarin (Citrus miaray Tan.) is used as a rootstock for In recent years, advances in chromatographic techniques
citrus propagation.11 Miaray mandarin fruits are round, bright have led to improved methods for the rapid separation and
yellow in color, and 5−7 cm in diameter, with sour, juicy isolation of bioactive compounds. Simultaneously separation
segment membranes. Only a few scientific studies have and spectroscopic detection methods, termed hyphenated
examined this species, and most of these studies have focused chromatography, can be used to elucidate the tentative identity
on evaluation of genetic heritability. These studies suggest that of the compounds of interest.20−22 Among separation
Miaray mandarins are unique, with distinct genotypic character- techniques, flash chromatography provides several advantages,
istics in comparison to other mandarin species.10
Polymethoxyflavones are a group of flavonoids that have Received: January 2, 2015
three or more methoxyl moieties attached to their basic flavone Revised: June 27, 2015
structure. Polymethoxyflavones have been extensively studied Accepted: July 3, 2015
for their biological properties such as anticancer, antilipogenic,

© XXXX American Chemical Society A DOI: 10.1021/acs.jafc.5b02445


J. Agric. Food Chem. XXXX, XXX, XXX−XXX
Journal of Agricultural and Food Chemistry Article

Figure 1. Flash chromatograms obtained by the separation of extracts of Miaray mandarin peel using silica gel columns. The individual peaks A−G
yielded compounds 1−7, and H−N yielded compounds 1, 3, 4, and 7−10.

including large sample volumes, a wide range of solvents, biofilm inhibitory activity in a dose-dependent manner. Citrus
automated detection, and inline ultraviolet−visible spectral has approximately 28 structurally distinct reported polymethox-
scanning (UV−vis). This improved technique enables medium- yflavones, with variation in the number and position of the
pressure liquid chromatographic separations as well as real-time methoxyls attached to the flavone backbone.30 These structural
monitoring of the compounds by their unique UV−vis spectral variants may potentially demonstrate a wide range of inhibitory
characteristics. Flash chromatography has been successfully activities on bacterial biofilms and help in the identification of
applied for the separation and isolation of several bioactive novel antimicrobial agents.
compounds such as curcuminoids, flavonoids, and coumarins In the present study, nine polymethoxyflavones and one
from a wide range of plant extracts and industrial by- limonoid were isolated for the first time from Miaray mandarins
products.17,23,24 using rapid flash chromatography along with 3,5,7,3′,4′-
The antimicrobial activity of citrus peel extracts has been pentamethoxyflavone (pentamethylquercetin). These com-
intensively investigated,25,26 but only a few studies have pounds were characterized by spectral analysis using high-
evaluated the effect of pure polymethoxyflavones, primarily performance liquid chromatography (HPLC), matrix-assisted
due to commercial unavailability. Furthermore, polymethoxy- laser desorption/ionization-time-of-flight-mass spectrometry
(MALDI-TOF), and 1D and 2D nuclear magnetic spectroscopy
flavones do not demonstrate potent bactericidal or bacterio-
(NMR). All identified compounds were evaluated for their
static effects. For example, a recent study showed that
biofilm inhibitory activity using Vibrio harveyi.


polymethoxyflavones nobiletin and tangeretin exhibit low
antimicrobial activities against six strains of microorganisms,
including Escherichia coli, Staphylococcus sp., and Salmonella MATERIALS AND METHODS
typhi.26 An alternative mechanism to attenuate bacterial Plant Material. Mature Citrus miaray mandarin fruits (Figure S1)
pathogenicity is interference with bacterial cell−cell communi- were harvested in December 2010 from the Texas A&M University−
cation pathways. We have previously demonstrated that several Kingsville Citrus Center orchard (Weslaco, TX, USA). A voucher
specimen (249960) has been submitted to the S. M. Tracy Herbarium,
citrus limonoids and flavonoids interfere with bacterial signaling Texas A&M University (College Station, TX, USA). The peels were
cascades, leading to attenuation of various virulence mecha- separated and dried to ≤5% moisture. The peels were blended to
nisms including biofilm formation.27−29 Our prior study obtain 40−60 mesh size powder in a Vita-prep blender (Vita-Mix
showed that 3′,4′,5,6,7-pentamethoxyflavone (sinensitin) exerts Corp., Cleveland, OH, USA) and used for extraction.

B DOI: 10.1021/acs.jafc.5b02445
J. Agric. Food Chem. XXXX, XXX, XXX−XXX
Journal of Agricultural and Food Chemistry Article

Reagents and Instrumentation. Solvents used for analysis were compounds. Evaporation of the solvent from the pooled fractions of
of HPLC grade and obtained from Fisher Scientific (Pittsburgh, PA, peaks H−N yielded crystallized compounds 1, 3 4, 5, 6, 7, and 10.
USA). Nanopure water (NANOpure, Barnstead/Thermolyne, Dubu- Liquid Chromatography. The HPLC system consisted of a
que, IA, USA) was used for HPLC analysis. The solvents used for flash Waters 1525 HPLC series (Milford, MA, USA) connected to a
chromatography were of analytical grade and purchased from Fisher photodiode array detector. A Gemini C18 column (3 μm, 250 × 4.6
Scientific. The separation of polymethoxyflavones was carried out on a mm) (Phenomenex, Torrance, CA, USA) was used for the separations.
flash chromatography system (Combiflash Rf, Teledyne ISCO, A gradient mobile phase of 3 mM phosphoric acid (A) and acetonitrile
Lincoln, NE, USA). Silica gel (particle size 35−60 μm) columns (B) was used for the separations at a flow rate of 1 mL/min. Initially,
(220 g) were purchased from ISCO Inc. (RediSep Rf ISCO Inc.). elution was started with a gradient of 5% B followed by a linear
Soxhlet Extraction of Peels. Eight hundred grams of dried peel increase to 50% in 20 min and then returned to 5% in 5 min. The
powder was loaded into a Soxhlet-type apparatus and extracted with injection volume was set at 10 μL, and the polymethoxyflavones were
hexane and chloroform sequentially for 16 h each. The extracts were detected at λ280 nm and λ340 nm. Chromatographic data were collected
concentrated to obtain 31 and 23 g of hexane and chloroform extracts, and processed using Empower2 software (Waters).
respectively. Matrix-Assisted Laser Desorption/Ionization-Time of Flight-
Sample Preparation and Chromatographic Separation. The Mass Spectrometry (MALDI-TOF-MS) Analysis. The samples for
hexane extract (31 g) was dissolved in 50 mL of hexane, impregnated MS analysis were prepared by dissolving the isolated compounds (1−
with 20 g of silica gel, and dried at room temperature in a hood. A 10) in acetonitrile and mixed with 2′,4′,6′-trihydroxyacetophenone
similar procedure was followed for the chloroform extract and used for (THAP) matrix. Half a microliter of the matrix mixture was spotted on
purification. The chromatography system Teledyne ISCO CombiFlash a MALDI sample plate and air-dried. MALDI-TOF mass spectra were
Rf 4x system equipped with a fiber optic spectrometer, a UV−vis acquired using a Voyager DE-Pro (Applied Biosystems, Carlsbad, CA,
(λ200 nm−λ700 nm) detector, and a 220 g silica gel flash column was used USA) mass spectrometer in positive reflector ion mode. After time-
for the separation. This detector can detect and monitor UV spectra of delayed extraction of 275 ns, the ions were accelerated to 20 kV for
the eluent after separation from the flash column using PeakTrak 2.0.0 TOF mass spectrometric analysis. A total of 100 laser shots were
software. acquired with the signal averaged per mass spectrum.
Hexane Extract. The silica gel impregnated hexane extract of NMR Analysis. 1H and 13C NMR attached proton test (APT)
Miaray mandarin was subjected to flash chromatography on a silica gel spectra were recorded at 400 and 100 MHz, respectively, by FT NMR
column. The column was equilibrated with hexane for 3 column (JEOL USA, Inc., Peabody, MA, USA). The isolated compounds were
volumes prior to separations. Polymethoxyflavones were separated dissolved in chloroform (CDCl3) except compound 8, which was
with a 65 min gradient program of solvent A (hexane) and solvent B dissolved in deuterated dimethyl sufoxide (DMSO-d6). Compounds 1,
(chloroform): 100% A held for 4 min, linearly increased to 10% B over 5, 6, 8, and 10) were identified by using 1D and 2D NMR data
1 min, held for 12 min, linearly increased to 40% B over 6 min, held including heteronuclear single-quantum correlation spectroscopy
for 14 min, linearly increased to 100% B over 6 min, held for 12 min, (HSQC); heteronuclear multiple-bond correlation spectroscopy
and then finally returned to the initial conditions and held for 10 min. (HMBC), double quantum filter correlation spectroscopy (DQF-
The flow rate was maintained at 150 mL/min, and individual fractions COSY), and nuclear Overhauser effect spectroscopy (NOSEY).
were collected by monitoring the eluting analytes at λ210 nm and λ340 nm. Bacterial Strains and Media. V. harveyi strains BB170
The eluent from the column was also monitored using an inline UV− (luxN::Tn5), BB886 (luxPQ::Tn5), BB120 (wild-type), JAF483
vis detector for monitoring the absorption spectra of the separating (luxO D47A), and BNL258 (hfq::Tn5lacZ) were kindly provided by
compounds. Individual peaks were tentatively identified on the basis of B. L. Bassler (Princeton University, Princeton, NJ, USA).31−34 E. coli 5,
the distinct absorption spectrum of each fraction. Seven major peaks an environmental isolate, was used as a positive control for
(A−G) were observed and collected in individual fractions (Figure autoinducer-2 (AI-2) activity.35 Autoinducer bioassay (AB) or Luria
1A). The retention times of the separated peaks were as follows: peak Marine (LM) media were used to culture the V. harveyi strains.36
A, 22.5−25 min; peak B, 25−29.5 min; peak C, 29.5−34 min; peak D, Growth Assay. Overnight cultures of V. harveyi BB120 were
34−37.5 min; peak E, 39−42 min; peak F, 42−45 min; peak G, 45−50 diluted 100-fold in AB media and treated with polymethoxyflavones
min. The individual peak fractions were further analyzed by HPLC and (50 μM) or an equivalent volume of DMSO. The cultures were grown
pooled on the basis of their similarity in retention time and matching for 16 h, and OD570 was measured every 15 min by using a Synergy
UV spectral data. Evaporation of the solvent from the pooled fractions HT multimode microplate reader (BioTek Instruments, Winooski,
of peaks A−G yielded crystallization of compounds 1−5, 8, and 9, VT, USA). The instrument was set to maintain a temperature of 30
respectively. °C, and plates were constantly shaken at medium speed between
Chloroform Extract. The silica gel impregnated chloroform extract readings. The data are presented as the mean of three biological
was subjected to flash chromatography on a silica gel column (220 g). replicates.
The column was equilibrated with hexane for 3 column volumes prior Bioluminescence Assay. The bioluminescence was measured
to separations. Polymethoxyflavones were separated with a 75 min using the method described previously from our laboratory.27 In brief,
gradient program of solvent A and solvent B: 100% A held for 6 min, E. coli 5 and V. harveyi BB120 were cultured overnight in Luria−
linearly increased to 10% B over 4 min, held for 15 min, linearly Bertani (LB) and LM media, respectively, to obtain high
increased to 40% B over 5 min, held for 15.5 min, linearly increased to concentrations of autoinducer activity. The overnight cultures were
60% B over 5 min, held for 4.5 min, linearly increased to 100% B over centrifuged at 10000 rpm for 10 min in a microcentrifuge and filtered
6 min, held for 12 min, and then finally returned to the initial using a 0.2 μm cellulose acetate membrane filter to obtain clear cell-
conditions and held for 2 min. The flow rate was maintained at 150 free supernatant (CFS). The CFSs were stored at −20 °C until use.
mL/min, and individual fractions were collected by monitoring the Inhibition of autoinducer [harveyi autoinducer (HAI) and AI-2]-
eluting analytes at λ210 nm and λ340 nm. The eluent from the column was mediated bioluminescence was measured in a 96-well plate assay.27
also monitored using an inline UV−vis detector for monitoring the The final concentrations of polymethoxyflavones tested were 12.5, 25,
absorption spectra of the separating compounds. Individual peaks were and 50 μM. Diluted (2500-fold) overnight cultures (900 μL) of
tentatively identified on the basis of the distinct absorption spectrum reporter strains BB886 (for HAI) and BB170 (for AI-2) were
of each fraction. Seven major peaks (H−N) were observed and incubated with 5 μL of CFS, 0.5 μL of polymethoxyflavones or
collected (Figure 1B). The retention times of the separated peaks were DMSO, and 4.5 μL of sterile Autoinducer Bioassay (AB) medium at
as follows: peak H, 30−32.5 min; peak I, 32.5−35 min; peak J, 35−39 30 °C with shaking at 100 rpm. Light production was measured by a
min; peak K, 39−44 min; peak L, 51−55 min; peak M, 55−59 min; Victor2 1420 multilabel counter (Beckman Coulter) in luminescence
peak N, 62−70 min. Twenty microliters of each fraction was diluted mode. The values were recorded as relative light units and used for
with 500 μL of acetone and analyzed by HPLC for detection of calculations. The relative activity was calculated as the ratio of
polymethoxyflavones and further pooling of fractions with similar luminescence of the test sample to the control (DMSO) sample.36

C DOI: 10.1021/acs.jafc.5b02445
J. Agric. Food Chem. XXXX, XXX, XXX−XXX
Journal of Agricultural and Food Chemistry Article

Biofilm Assay. The biofilm assay was conducted using the method The chromatograms of the HPLC analysis of the isolated
described previously from our laboratory.29 Briefly, an overnight compounds are presented in Figure 2. The absence of other
culture of V. harveyi BB120 was diluted 1:50 in LM medium, and 190
μL of this fresh culture was incubated with 7 μL of sterile medium and peaks demonstrates the purity of the isolated compounds. The
0.5 μL of DMSO or polymethoxyflavones (12.5, 25, 50 μM) dissolved purity of the isolated compounds ranged between ≥95 and
in DMSO. The biofilm mass was quantified by washing with ≤98%. The absorption spectra of the compounds were
phosphate buffer (0.1 M, pH 7.4), followed by staining with 0.3%
Crystal Violet (Fisher) for 20 min. The dye associated with biofilm obtained using a photodiode array detector. All of the isolated
was dissolved with 200 μL of 33% acetic acid, and A570 was measured.
The mean ± standard deviation of three biological replicates is
presented.

■ RESULTS AND DISCUSSION


Extraction. Successive extractions of mandarin peels using
nonpolar hexane and medium-polar chloroform in a Soxhlet
type apparatus yielded 3.9 and 2.9% of concentrated crude
extract, respectively. HPLC analysis of the extracts (Figure S2)
suggested that the extracts contain high levels of polymethoxy-
flavones. Additionally, the chromatograms suggested that the
successive extractions with hexane and chloroform enabled
fractionation of low-polar and medium-polar molecules,
respectively. The individual polymethoxyflavones were purified
by flash chromatography.
Purification of Polymethoxyflavones. Hexane Extract.
The step gradient elution using hexane and acetone enabled
clear separation of individual polymethoxyflavones (Figure 1a).
The eluent from the column was monitored using an inline
UV−vis detector to show the absorption spectra of the
compounds, which were subsequently collected in fractions.
The initial isocratic elution using 10% acetone separated the oil
from the crude extract. After separation of the oils, the linear
change in gradient elution to 60:40 hexane/acetone resulted in
good separation of the low-polarity polymethoxyflavones,
whereas medium-polar polymethoxyflavones were separated
by the gradual increase in the solvent polarity to 100% acetone.
On the basis of HPLC analysis, fractions showing similar peak
profiles were pooled and concentrated under vacuum. The
concentrated peak (A−G) fractions yielded crystallized
compounds, which were collected, their purity was analyzed
by HPLC, and they were identified by spectroscopic data.
Chloroform Extract. The impregnated chloroform extract
was subjected to flash chromatographic separation using a step
gradient of hexane and acetone solvents with a total run time of
75 min. To enable good separation of medium- and low-polar
polymethoxyflavones, an additional step gradient of 40:60
hexane/acetone was used, followed by a linear increase to 100%
acetone (Figure 1b). The step gradient elution resulted in
separation of peaks H−N. Similar to the chromatographic
separation of the hexane extract, the eluent from the column
was monitored using an inline UV−vis detector. Individual
peaks with distinct absorption spectra were collected in
individual fractions. The fractions collected for individual
peaks were analyzed by HPLC and concentrated to obtain
crystallized compounds. These were further subjected to
spectral analysis using MS and NMR to identify the
compounds.
Altogether, 10 purified compounds were isolated from
hexane and chloroform extracts. The yields of the isolated
compounds were as follows: 1, 18 mg; 2, 76 mg; 3, 8628 mg; 4,
2012 mg; 5, 210 mg; 6, 121 mg; 7, 1077 mg; 8, 79 mg; 9, 810
mg; and 10, 820 mg.
Identification. The identification and structural elucidation Figure 2. HPLC chromatograms of isolated compounds (1−10) along
of the isolated compounds from Miaray mandarin were with their UV spectra. The separation was conducted on a Gemini C18
conducted using spectral analysis by HPLC, MS, and NMR. column using gradient mobile phase.

D DOI: 10.1021/acs.jafc.5b02445
J. Agric. Food Chem. XXXX, XXX, XXX−XXX
Journal of Agricultural and Food Chemistry Article

Figure 3. 1 H NMR spectra of purified compounds (1, 5, 6, 8, and 10) recorded on a JEOL ECS NMR spectrometer at 400 MHz. All compounds
were dissolved in CDCl3 except 8, which was dissolved in DMSO-d6. The solvent peak is denoted by an asterisk (*).

compounds had distinct UV maxima in the range of ments (Figure S4). These data were helpful when assigning the
λ325 nm−λ353 nm, which is characteristic of polymethoxyflavones. chemical shifts of another three pentamethoxyflavones of
The results are consistent with earlier studies.16,37,38 The compounds 6 and 10.
MALDI-TOF mass spectral data for compounds 2, 3, 4, 7, and Compounds 4 and 5 are hexamethoxyflavones, whereas
9 were found to be 373.04, 433.16, 403.31, 470.21, and 343.26, compound 2, 6, and 10 are pentamethoxyflavones with [M +
respectively (Figure S3). The NMR chemical shifts (not 1]+ of 403 and 373, respectively. The structural confirmation is
shown) of these compounds were matched to reported not very accurate using only mass spectral data. Therefore, we
values.16,39 Using these data, the structures of compounds 2, have used 1D and 2D NMR spectral data to validate the
3, 4, 7, and 9 were confirmed and identified as 5,6,7,8,4′- findings (Figures 3 and 4 and Figures S4−S9). 1H NMR signals
pentamethoxyflavone (tangeretin), 3,5,6,7,8,3′,4′-heptamethox- (Figure 3) of compounds 1, 5, 6, 8, and 10 showed at δ 3.5−4.0
yflavone, 5,6,7,8,3′,4′-hexamethoxyflavone (nobiletin), limonin, and 6.3−8.0, indicating that all compounds had structures
and 5,7,8,4′-tetramethoxyflavone, respectively. The structure of typical of polymethoxyflavones. Their characteristics and
tangeretin was also confirmed by 2D NMR spectral assign- assignments were made using 13C APT NMR (Figure 4) and
E DOI: 10.1021/acs.jafc.5b02445
J. Agric. Food Chem. XXXX, XXX, XXX−XXX
Journal of Agricultural and Food Chemistry Article

Figure 4. 13C NMR attached proton test (APT) spectra of the isolated compounds (1, 5, 6, 8, and 10) recorded on a JEOL ECS spectrometer at 100
MHz. All compounds were dissolved in CDCl3 except 8, which was dissolved in DMSO-d6.

2D experiments including HMBC, HMQC, DQFCOSY, and substitution. A sharp singlet at δ 6.4 was assigned to the sixth
NOSEY. position, which was evidenced from the δ 92.32 signal. The B-
Compound 1: The presence of a downfield proton resonance at ring proton signals showed a pattern similar to ABX coupling
12.5 ppm on 1H NMR is a characteristic indicator of a chelated for three protons at H-2′, H-5′, and H-6′ (Figure S5).
hydroxyl group at the fifth position of the flavone skeleton. Due However, the H-2′ displayed a broad singlet due to NOE effect
to the presence of the hydroxyl group at C-5, the carbonyl caused by methoxyl substitution at the C-3 position. Using
signal at the fourth position in 13C NMR showed at δ 178.5. these spectral data and MALDI-TOF-MS (m/z at 403.3491),
Moreover, the lack of a resonance at δ 6.6 for the olefinic compound 5 was identified as 3,5,7,8,3′,4′-hexamethoxyflavone
proton in 1H NMR indicated that there is a substituent at the 3- and signals were matched to reported values.45
position. The 1H NMR spectrum showed the ABX-type Compound 6: The 1H NMR spectrum showed the presence of
aromatic proton signals indicating the existence of 1′,3′,4′- five methoxyl groups, and those were shown from the 13CNMR
trisubstituted flavones. 13C NMR showed the presence of three APT spectrum to be similar to compound 5, with substitutions
methyl groups at around δ 56 and one distinct signal at δ 60.2 at 3-, 5-, 7-, 3′-, and 4′-positions. These data suggested that
for the 7-methoxyl group. The MALDI-TOF-MS spectrum compound 6 is a pentamethoxylated flavonoid. In 1H NMR two
showed the [M + 1]+ at m/z 359.3011. Using these spectral aromatic broad singlets at δ 6.34 and 6.48 well correlated in
data identified compound 1 as 5-hydroxyl-3,7,3′,4′-tetrame- DQFCOSY to two different neighboring methoxy groups at δ
thoxyflavone. The chemical shifts show good agreement with 3.88 and 3.93, which indicates meta substitution in the A-ring
the earlier papers by Li et al.40 (Figure S6). On the other hand, in the proton NMR spectrum,
Compound 5: The 1H NMR spectrum showed the presence of the δ 6.34 proton had a meta coupling to δ 6.48 (J = 1.8 Hz),
six methoxyl groups at δ 3.88−3.98, and those were tentatively which was assigned to H-6 and H-8. Spectra in the B-ring had
assigned to the 3-, 5-, 7-, 8-, 3′-, and 4′-positions. These ABX type aromatic proton signals, which confirms the presence
assignments were further confirmed from 13C NMR APT of three protons, and it was confirmed by integrated area of
chemical shifts with two signals observed at low field (δ 59.9 three protons. The size of the coupling constant (J = 1.1 and
and 61.5) for two typical methoxyl groups at 3- and 8- 8.4 Hz) is characteristic of meta and ortho coupling as found in
F DOI: 10.1021/acs.jafc.5b02445
J. Agric. Food Chem. XXXX, XXX, XXX−XXX
Journal of Agricultural and Food Chemistry Article

3′,4′-methoxylated flavonoids, whereas such correction were


not found in tangeretin due to the absence of ABX coupling
(Figure S7). Moreover, proton signals for H-2′ displayed
downfield at δ 7.9 as compared to PMF (10) due to the
presence of methoxyl substitution at the C-3-position, which
will cause NOE effect. A similar trend was observed for
hexamethoxyflavone (5). Using this NMR and mass spectral
data ([M + 1]+ − 373.3003), we confirmed compound 6 as
3,5,7,3′,4′-pentamethoxyflavone (pentamethylquercetin).
These chemical shifts were identical to those from the scientific
literature.42 This compound was previously identified in the
bark extract of Melicope subunifoliolata.43
Compound 8: The TOF-MS spectrum showed a molecular ion
peak at m/z 313.2550. The 1H NMR (DMSO-d6) spectral
analysis showed the presence of three methoxyl signals at δ
3.78, 3.8, and 3.86, which were confirmed in the 13C NMR APT
spectrum at δ 55.5, 55.9, and 56.5. Two broad singlets for one
proton in the aromatic region showed the presence of two
meta-coupled doublets at δ 6.38 (J = 2 Hz) and δ 6.55 (J = 2
Hz) in the A-ring for H-6 and H-8 protons. Another two ortho-
coupled doublets at δ 7.05 (J = 9 Hz, 2H) and δ 7.92 (J = 9 Hz,
2H) were assigned to the B-ring (H-3′, H-5′, H-2′, H-6′)
(Figure 3). This compound had one proton singlet at δ 6.61,
which was assigned to the olefinic proton at H-3.44 Thus,
compound 8 was determined to be 5,7,4′-trimethoxyflavone,
and the spectral data were consistent with data from the
literature.30,41
Compound 10 had a TOF-MS m/z 373.2725 [M + 1]+,
which indicates the compound is a pentamethoxyflavone. The
1
H NMR spectrum showed the presence of five methoxyls at δ
3.91−3.99 and two proton singlets at δ 6.4 and 6.6 (Figure 3).
In the APT spectrum (Figure 4), the A-ring carbon signals
coincided well with those of compound 9 (tetramethoxy- Figure 5. Structures of isolated compounds (1−10) from Miaray
flavone).17 The 1H NMR spectrum revealed the presence of an mandarins. Compounds 1, 5, 6, 8, and 10 were identified by NMR and
ABX system for the three aromatic protons in the B-ring at H- mass spectral analysis as 5-hydroxy-3,7,3′,4′-tetramethoxyflavone,
2′, H-5′, and H-6′ as demonstrated by coupling constant signals 3,5,7,8,3′,4′-hexamethoxyflavone, 3,5,7,3′,4′-pentamethoxyflavone
at δ 7.39 (d, 1H, J = 1.8 Hz), δ 6.95 (d, 1H, J = 8.2 Hz), and δ (pentamethylquercetin), 5,7,4′-trimethoxyflavone, and 5,7,8,3′,4′-pen-
7.6 (d, 1H, J = 8.2 Hz), respectively (Figure 3 and Figure S8). tamethoxyflavone, respectively. Compounds 2, 3, 4, 7, and 9 were
identified by mass spectral analysis as tangeretin, heptamethoxyflavone,
These results indicate the presence of methoxyls at the 3′- and nobiletin, limonin, and tetramethoxyflavone, respectively.
4′-positions in the B-ring. The singlets at δ 6.58 and 6.4 were
assigned to the H-3 and H-6 protons, respectively, using review describing the various compounds present in citrus
DQFCOSY data in Figure S8. 1HNMR spectra of four species by Manthey et al.47
pentamethoxyflavones for understanding the chemical shifts Antimicrobial Activity of Polymethoxyflavones. We
and structural assignments of all signals are presented in Figure next examined whether these compounds affect bacterial
S9. Although sinensetin is not isolated from Miaray mandarins, growth by treating V. harveyi with polymethoxyflavones. The
the spectrum was given for comparison purposes from our kinetic growth curve was calculated by recording the OD570 for
earlier study. Using these spectral data, compound 10 was 16 h of growth at optimal temperature. Figure 6 illustrates the
determined to be 5,7,8,3′,4′-pentamethoxyflavone, and the growth kinetics of V. harveyi BB120 after treatment with 50 μM
chemical shifts were compared to reported data.33,45 of the eight polymethoxyflavones. The sigmoid bacterial growth
In summary, results from the spectral analysis of the isolated curve observed indicates the polymethoxyflavones do not
compounds confirm the identity of the isolated compounds as inhibit the bacterial growth at 50 μM concentration. Similar
5-hydroxy-3,7,3′,4′-tetramethoxyflavone (1), 5,6,7,8,4′-pentam- results were also observed in other studies evaluating the
ethoxyflavone (tangeretin) (2), 3,5,6,7,8,3′,4′-heptamethoxy- antimicrobial activity of polymethoxyflavones (nobiletin and
flavone (3), 5,6,7,8,3′,4′-hexamethoxyflavone (nobiletin) (4), tangeretin), which had minimum inhibitory concentration
3,5,7,8,3′,4′-hexamethoxyflavone (5), 3,5,7,3′,4′-pentamethoxy- (MIC) of ≥1600 μg/mL,26 suggesting that these compounds
flavone (pentamethylquercetin) (6), limonin (7), 5,7,4′- are not bactericidal.
trimethoxyflavone (8), 5,7,8,4′-tetramethoxyflavone (9), and Inhibition of Biofilm Formation and Bioluminescence.
5,7,8,3′,4′-pentamethoxyflavone (10), respectively (Figure 5). Biofilms are surface adherent colonies of bacteria that form an
The results are consistent with the reported values.38,45,46 To extracellular polymeric matrix to protect themselves. Control of
the best of our knowledge, this is the first report of the isolation biofilm formation remains a concern among medical personnel,
of 3,5,7,3′,4′-pentamethylquercetin from the genus Citrus. It is primarily related to medical devices and implants. The bacterial
worth noting that this compound was previously reported in a biofilms in medical devices and implants are difficult to control
G DOI: 10.1021/acs.jafc.5b02445
J. Agric. Food Chem. XXXX, XXX, XXX−XXX
Journal of Agricultural and Food Chemistry Article

Figure 6. Growth curve of V. harveyi BB120 in the presence of


polymethoxyflavones. The evaluated polymethoxyflavones were
tangeretin, 3,5,6,7,8,3′,4′-heptamethoxyflavone (Hepta-MF), nobiletin,
5,7,4′-trimethoxyflavone (Tri-MF), 3,5,7,8,3′,4′-hexamethoxyflavone
(Hexa-MF), 5,7,8,4′-tetramethoxyflavone (Tet-MF), 3,5,7,3′,4′-pen-
tamethoxyflavone (Penta-MF1), and 5,7,8,3′,4′-pentamethoxyflavone
(Penta-MF2).

due to increased antimicrobial resistance of biofilms.48 As


biofilm formation is regulated by several factors including
quorum sensing, interference with the quorum sensing pathway
is currently thought to be a viable strategy to control biofilms.
Quorum sensing is a population-dependent phenomenon
that is mediated through the concentration of autoinducers, Figure 7. HAI-1 and AI-2 induced bioluminescence inhibition in V.
signaling molecules synthesized by bacteria.49 Quorum sensing harveyi BB886 (A) and BB170 in the presence of polymethoxyflavones
regulated bioluminescence production in V. harveyi and its (B). The evaluated polymethoxyflavones were tangeretin (TANG),
signal transduction pathway has been well studied. The 3,5,6,7,8,3′,4′-heptamethoxyflavone (Hep-MF), nobiletin (NOB),
autoinducers are detected by three coincidence detectors, 5,7,4′-trimethoxyflavone (Tri-MF), 3,5,7,8,3′,4′-hexamethoxyflavone
LuxN, LuxPQ, and CqsS, that phosphorylate LuxU, a (Hex-MF), 5,7,8,4′-tetramethoxyflavone (Tet-MF), 3,5,7,3′,4′-pen-
phosphorelay protein. LuxU then phosphorylates the response tamethoxyflavone (Penta-MF-1), and 5,7,8,3′,4′-pentamethoxyflavone
(Penta-MF-2).
regulator LuxO that in turn regulates the expression of
downstream genes including bioluminescence genes.31−34
Furthermore, interference with V. harveyi bioluminescence
has been widely used as a model to study and identify quorum
sensing inhibitors.27−29 Panels A and B of Figure 7 illustrate the
results obtained from the analysis of bioluminescence in BB886
and BB170 strains of V. harveyi, respectively. Among the
evaluated polymethoxyflavones, only 3,5,6,7,8,3′,4′-heptame-
thoxyflavone and 3,5,7,8,3′,4′-hexamethoxyflavone seem to
interfere with AI-2-mediated bioluminescence in BB170. The
treatment of BB886 strain with polymethoxyflavones demon-
strated <40% inhibition over control (DMSO).
The inhibition of bioluminescence suggests that certain
polymethoxyflavones may interfere with biofilm formation. Figure 8. Inhibition of V. harveyi BB120 biofilm after treatment with
Consistent with the bioluminescence interference results, only polymethoxyflavones. The evaluated polymethoxyflavones were
3,5,6,7,8,3′,4′-heptamethoxyflavone and 3,5,7,8,3′,4′-hexame- tangeretin, 3,5,6,7,8,3′,4′-heptamethoxyflavone (Hepta-MF), nobiletin,
5,7,4′-trimethoxyflavone (Tri-MF), 3,5,7,8,3′,4′-hexamethoxyflavone
thoxyflavone inhibited biofilm formation in a dose-dependent
(Hexa-MF), 5,7,8,4′-tetramethoxyflavone (Tet-MF), 3,5,7,3′,4′-pen-
manner (Figure 8). Approximately 61% (with respect to tamethoxyflavone (Penta-MF1), and 5,7,8,3′,4′-pentamethoxyflavone
control−DMSO treatment) of biofilm formation was inhibited (Penta-MF2).
by 3,5,7,8,3′,4′-hexamethoxyflavone at a concentration of 50
μM. The IC50 value of 3,5,7,8,3′,4′-hexamethoxyflavone was
calculated to be 37.38 μM. Although 3,5,6,7,8,3′,4′-heptame- compounds might inhibit biofilm formation. Therefore,
thoxyflavone inhibited biofilm, the percent inhibition was constitutively luminescent V. harveyi mutants were investigated
approximately 35% at the maximum evaluated concentration of after treatment with 50 μM 3,5,7,8,3′,4′-hexamethoxyflavone
50 μM. Other polymethoxyflavones did not have >35% (Figure 9). The results demonstrated an increase in bio-
inhibition of biofilm with respect to control. luminescence in the luxO mutant strain (JAF483), but showed
Analyzing the bioluminescence in specific mutant strains no change in the hfq mutant (BNL258), suggesting an effect on
after treatment with polymethoxyflavones may enable us to LuxO. However, further studies are needed to validate the effect
understand the possible pathway through which these on LuxO. The results indicate that 3,5,7,8,3′,4′-hexamethoxy-
H DOI: 10.1021/acs.jafc.5b02445
J. Agric. Food Chem. XXXX, XXX, XXX−XXX
Journal of Agricultural and Food Chemistry Article

composition is modulated by household processing techniques. J.


Food Sci. 2012, 77 (9), C921−C926.
(2) Uckoo, R. M.; Jayaprakasha, G. K.; Somerville, J. A.;
Balasubramaniam, V. M.; Pinarte, M.; Patil, B. S. High pressure
processing controls microbial growth and minimally alters the levels of
health promoting compounds in grapefruit (Citrus paradisi Macfad)
juice. Innovative Food Sci. Emerging Technol. 2013, 18, 7−14.
(3) Kim, J.; Jayaprakasha, G. K.; Uckoo, R. M.; Patil, B. S. Evaluation
of chemopreventive and cytotoxic effect of lemon seed extracts on
Figure 9. Bioluminescence inhibition in V. harveyi mutants JAF483 human breast cancer (MCF-7) cells. Food Chem. Toxicol. 2012, 50 (2),
(luxO) and BNL258 (hfq) by 3,5,7,8,3′,4′-hexamethoxyflavone (50 423−430.
μm). (4) Uckoo, R. M.; Jayaprakasha, G. K.; Patil, B. S. Phytochemical
analysis of organic and conventionally cultivated Meyer lemons (Citrus
meyeri Tan.) during refrigerated storage. J. Food Compos. Anal. 2015,
42, 63−70.
flavone is a potent inhibitor of V. harveyi quorum sensing and (5) Androula, G. Evaluation of rootstocks for ‘Clementine’ mandarin
appears to have larger effect on the AI-2 system. The effect on in Cyprus. Sci. Hortic. 2002, 93 (1), 29−38.
luxO mutants seems to suggest that 3,5,7,8,3′,4′-hexamethoxy- (6) Gmitter, F.; Hu, X. The possible role of Yunnan, China, in the
flavone may be a nonspecific inhibitor. Further studies are origin of contemporary citrus species (rutaceae). Econ. Bot. 1990, 44
required to evaluate the ability of polymethoxyflavones to (2), 267−277.
interfere with quorum sensing-mediated biofilm formation in (7) USDA. Citrus: World Markets and Trade; U.S. Department of
Agriculture, Foreign Agricultural Service, Government Printing Office:
human pathogens. Washington, DC, USA, 2011.
In summary, 10 bioactive compounds were successfully (8) Wu, G. A.; Prochnik, S.; Jenkins, J.; Salse, J.; Hellsten, U.; Murat,
isolated from Miaray mandarin using flash chromatography, F.; Perrier, X.; Ruiz, M.; Scalabrin, S.; Terol, J. Sequencing of diverse
and their structural identification was confirmed by spectral mandarin, pummelo and orange genomes reveals complex history of
analysis using MALDI-TOF and NMR. Among the isolated admixture during citrus domestication. Nat. Biotechnol. 2014, 32 (7),
compounds, 3,5,7,3′,4′-pentamethoxyflavone is reported here 656−662.
for the first time from Citrus spp.. Among the evaluated (9) Luro, F.; Gatto, J.; Costantino, G.; Pailly, O. Analysis of genetic
polymethoxyflavones, 3,5,7,8,3′,4′-hexamethoxyflavone demon- diversity in Citrus. Plant Genet. Resour. 2011, 9 (2), 218−221.
strated inhibitory activity on autoinducer-mediated cell−cell (10) Coletta Filho, H. D.; Machado, M. A.; Targon, M. L. P. N.;
signaling and biofilm formation in V. harveyi. Further research is Moreira, M. C. P. Q. D. G.; Pompeu, J. Analysis of the genetic diversity
required to understand the role of polymethoxyflavones in the among mandarins (Citrus spp.) using RAPD markers. Euphytica 1998,
102 (1), 133−139.
inhibition of biofilm in pathogenic bacteria. Results from this (11) Hutchison, D. J.; Hearne, C. J.; Bistline, F. W. The performance
study could enable the development of strategies using of ‘Valencia’ orange trees on 21 rootstocks in the Florida flatwoods.
polymethoxyflavones for preventing bacterial pathogenicity. Proc. Fla. State Hortic. Soc. 1992, 105, 60−63.
Moreover, identification of potent antimicrobials from citrus (12) Kurowska, E.; Manthey, J.; Casaschi, A.; Theriault, A.
species will provide added economic benefits to both producers Modulation of hepG2 cell net apolipoprotein B secretion by the
and the citrus-processing industry.


citrus polymethoxyflavone, tangeretin. Lipids 2004, 39 (2), 143−151.
(13) Li, S.; Pan, M.-H.; Lai, C.-S.; Lo, C.-Y.; Dushenkov, S.; Ho, C.-
ASSOCIATED CONTENT T. Isolation and syntheses of polymethoxyflavones and hydroxylated
*
S Supporting Information polymethoxyflavones as inhibitors of HL-60 cell lines. Bioorg. Med.
Additional figures as cited in the text. The Supporting Chem. 2007, 15 (10), 3381−3389.
(14) Sergeev, I. N.; Li, S.; Ho, C.-T.; Rawson, N. E.; Dushenkov, S.
Information is available free of charge on the ACS Publications
Polymethoxyflavones cctivate Ca2+-dependent apoptotic targets in
website at DOI: 10.1021/acs.jafc.5b02445.


adipocytes. J. Agric. Food Chem. 2009, 57 (13), 5771−5776.
(15) Nakajima, A.; Ohizumi, Y.; Yamada, K. Anti-dementia activity of
AUTHOR INFORMATION nobiletin, a citrus flavonoid: a review of animal studies. Clin.
Corresponding Authors Psychopharmacol. Neurosci. 2014, 12 (2), 75−82.
*(B.S.P.) Phone: (979) 862-4521. Fax: (979) 862-4522. E- (16) Uckoo, R. M.; Jayaprakasha, G. K.; Patil, B. S. Rapid separation
mail: b-patil@tamu.edu. method of polymethoxyflavones from citrus using flash chromatog-
*(G.K.J.) E-mail: gkjp@tamu.edu. raphy. Sep. Purif. Technol. 2011, 81 (2), 151−158.
(17) Uckoo, R. M.; Jayaprakasha, G. K.; Patil, B. S. Hyphenated flash
Funding
chromatographic separation and isolation of coumarins and poly-
This project is based upon work supported by the USDA-NIFA methoxyflavones from byproduct of citrus juice processing industry.
2010-34402-20875, “Designing Foods for Health”, through the Sep. Sci. Technol. 2013, 48 (10), 1467−1472.
Vegetable & Fruit Improvement Center. (18) Yamamoto, K.; Yahada, A.; Sasaki, K.; Ogawa, K.; Koga, N.;
Notes Ohta, H. Chemical markers of shiikuwasha juice adulterated with
The authors declare no competing financial interest. calamondin juice. J. Agric. Food Chem. 2012, 60 (44), 11182−11187.


(19) Lou, S.-N.; Hsu, Y.-S.; Ho, C.-T. Flavonoid compositions and
ACKNOWLEDGMENTS antioxidant activity of calamondin extracts prepared using different
solvents. J. Food Drug Anal. 2014, 22 (3), 290−295.
We acknowledge Chandra Mohan, Shiva Rani, and Kalpana for (20) Wilson, I. D.; Brinkman, U. A. T. Hyphenation and hypernation:
assisting in harvesting and peeling the fruits.


the practice and prospects of multiple hyphenation. J. Chromatogr., A
2003, 1000 (1−2), 325−356.
REFERENCES (21) Griffiths, P. R.; Pentoney, S. L.; Giorgetti, A.; Shafer, K. H. The
(1) Uckoo, R. M.; Jayaprakasha, G. K.; Balasubramaniam, V. M.; hyphenation of chromatography and FT-IR. Anal. Chem. 1986, 58
Patil, B. S. Grapefruit (Citrus paradisi Macfad) phytochemicals (13), 1349A−1366A.

I DOI: 10.1021/acs.jafc.5b02445
J. Agric. Food Chem. XXXX, XXX, XXX−XXX
Journal of Agricultural and Food Chemistry Article

(22) Marston, A.; Hostettmann, K. Natural product analysis over the (40) Li, S.; Yu, H.; Ho, C.-T. Nobiletin: efficient and large quantity
last decades. Planta Med. 2009, 75, 672−682. isolation from orange peel extract. Biomed. Chromatogr. 2006, 20 (1),
(23) Uckoo, R. M.; Jayaprakasha, G. K.; Patil, B. S. Rapid separation 133−138.
method of polymethoxyflavones from citrus using flash chromatog- (41) Machida, K.; Osawa, K. On the flavonoid constituents from the
raphy. Sep. Purif. Technol. 2011, 81 (2), 151−158. peels of Citrus hassaku Hort. ex Tanaka. Chem. Pharm. Bull. 2012, 37
(24) Jayaprakasha, G. K.; Negi, P. S.; Sikder, S.; Mohanrao, L. J.; (4), 1092−1094.
Sakariah, K. K. Antibacterial activity of Citrus reticulata peel extracts. Z. (42) Silva, T.; Carvalho, M.; Braz-Filho, R. Estudo espectroscópico
Naturforsch., C: J. Biosci. 2000, 55, 1030. em elucidaçaõ estrutural de flavonóides de Solanum jabrense Agra &
(25) Nannapaneni, R.; Muthaiyan, A.; Crandall, P. G.; Johnson, M. Nee e S. paludosum Moric. Quim. Nova 2009, 32, 1119−1128.
G.; O’Bryan, C. A.; Chalova, V. I.; Callaway, T. R.; Carroll, J. A.; (43) Chung, L. Y.; Yap, K. F.; Goh, S. H.; Mustafa, M. R.; Imiyabir, Z.
Arthington, J. D.; Nisbet, D. J.; Ricke, S. C. Antimicrobial activity of Muscarinic receptor binding activity of polyoxygenated flavones from
commercial citrus-based natural extracts against Escherichia coli Melicope subunifoliolata. Phytochemistry 2008, 69 (7), 1548−1554.
O157:H7 isolates and mutant strains. Foodborne Pathog. Dis. 2008, 5 (44) Sugiyama, S.; Umehara, K.; Kuroyanagi, M.; Ueno, A.; Taki, T.
(5), 695−699. Studies on the differentiation inducers of myeloid leukemic cells from
(26) Yi, Z.; Yu, Y.; Liang, Y.; Zeng, B. In vitro antioxidant and Citrus species. Chem. Pharm. Bull. 1993, 41 (4), 714−719.
antimicrobial activities of the extract of Pericarpium Citri Reticulatae of (45) Koichi, M.; Keisuke, O. On the flavonoid constituents from the
a new Citrus cultivar and its main flavonoids. LWT−Food Sci. Technol. peels of Citrus hassaku Hort. ex Tanaka. Chem. Pharm. Bull. 1989, 37
2008, 41 (4), 597−603. (4), 1092−1094.
(27) Vikram, A.; Jesudhasan, P. R.; Jayaprakasha, G. K.; Pillai, S. D.; (46) Raman, G.; Cho, M.; Brodbelt, J. S.; Patil, B. S. Isolation and
Patil, B. S. Citrus limonoids interfere with Vibrio harveyi cell−cell purification of closely related Citrus limonoid glucosides by flash
signalling and biofilm formation by modulating the response regulator chromatography. Phytochem. Anal. 2005, 16 (3), 155−160.
LuxO. Microbiology 2011, 157 (1), 99−110. (47) Manthey, J. A.; Guthrie, N.; Grohmann, K. Biological properties
(28) Vikram, A.; Jesudhasan, P. R.; Jayaprakasha, G. K.; Pillai, S. D.; of citrus flavonoids pertaining to cancer and inflammation. Curr. Med.
Jayaraman, A.; Patil, B. S. Citrus flavonoid represses Salmonella Chem. 2001, 8 (2), 135−153.
pathogenicity island 1 and motility in S. Typhimurium LT2. Int. J. Food (48) Djeribi, R.; Bouchloukh, W.; Jouenne, T.; Menaa, B.
Microbiol. 2011, 145 (1), 28−36. Characterization of bacterial biofilms formed on urinary catheters.
(29) Vikram, A.; Jayaprakasha, G. K.; Jesudhasan, P. R.; Pillai, S. D.; Am. J. Infect. Control 2012, 40 (9), 854−859.
Patil, B. S. Suppression of bacterial cell−cell signalling, biofilm (49) Reading, N. C.; Sperandio, V. Quorum sensing: the many
formation and type III secretion system by citrus flavonoids. J. Appl. languages of bacteria. FEMS Microbiol. Lett. 2006, 254 (1), 1−11.
Microbiol. 2010, 109 (2), 515−527.
(30) Uckoo, R. M.; Jayaprakasha, G. K.; Patil, B. S. Chromatographic
techniques for the separation of polymethoxyflavones from citrus. In
Emerging Trends in Dietary Components for Preventing and Combating
Disease; Patil, B. S., Jayaprakasha, G. K., Murthy, K. N. C., Seeram, N.,
Eds.; ACS Symposium Series 1093; American Chemical Society:
Washington, DC, USA, 2012.
(31) Surette, M. G.; Bassler, B. L. Quorum sensing in Escherichia coli
and Salmonella typhimurium. Proc. Natl. Acad. Sci. U. S. A. 1998, 95
(12), 7046−7050.
(32) Freeman, J. A.; Bassler, B. L. A genetic analysis of the function of
LuxO, a two-component response regulator involved in quorum
sensing in Vibrio harveyi. Mol. Microbiol. 1999, 31 (2), 665−677.
(33) Freeman, J. A.; Bassler, B. L. Sequence and function of LuxU: a
two-component phosphorelay protein that regulates quorum sensing
in Vibrio harveyi. J. Bacteriol. 1999, 181, 899−906.
(34) Lenz, D. H.; Mok, K. C.; Lilley, B. N.; Kulkarni, R. V.;
Wingreen, N. S.; Bassler, B. L. The small RNA chaperone Hfq and
multiple small RNAs control quorum sensing in Vibrio harveyi and
Vibrio cholerae. Cell 2004, 118 (1), 69−82.
(35) Qin, Y.; Luo, Z.-Q.; Farrand, S. K. Domains formed within the
N-terminal region of the quorumsensing activator TraR are required
for transcriptional activation and direct interaction with RpoA from
Agrobacterium. J. Biol. Chem. 2004, 279 (39), 40844−40851.
(36) Lu, L.; Hume, M. E.; Pillai, S. D. Autoinducer-2-like activity
associated with foods and its interaction with food additives. J. Food
Prot. 2004, 67 (7), 1457−1462.
(37) Yao, X.; Xu, X.; Fan, G.; Qiao, Y.; Cao, S.; Pan, S.
Determination of synergistic effects of polymethoxylated flavone
extracts of Jinchen orange peels (Citrus sinensis Osberk) with amino
acids and organic acids using chemiluminescence. Eur. Food Res.
Technol. 2009, 229 (5), 743−750.
(38) Wang, D.; Wang, J.; Huang, X.; Tu, Y.; Ni, K. Identification of
polymethoxylated flavones from green tangerine peel (Pericarpium
Citri Reticulatae Viride) by chromatographic and spectroscopic
techniques. J. Pharm. Biomed. Anal. 2007, 44 (1), 63−69.
(39) Perez, J. L.; Jayaprakasha, G. K.; Valdivia, V.; Munoz, D.;
Dandekar, D. V.; Ahmad, H.; Patil, B. S. Limonin methoxylation
influences the induction of glutathione S-transferase and quinone
reductase. J. Agric. Food Chem. 2009, 57 (12), 5279−5286.

J DOI: 10.1021/acs.jafc.5b02445
J. Agric. Food Chem. XXXX, XXX, XXX−XXX

You might also like