ArticlePDF Available

Quinolizidine alkaloids and phomopsins in lupin seeds and lupin containing food

Authors:

Abstract and Figures

In recent years there has been growing interest in replacing (genetically modified) soya by lupin. Lupin seeds, flours and lupin containing food have been analyzed in order to assess the relevance of a potential health hazard given by mycotoxins and/or naturally occurring alkaloids. Since not all important alkaloids used for quantitation were commercially available, isolation of lupanine, 13alpha-hydroxylupanine and angustifoline from lupin flours of high alkaloid contents was performed. Alkaloids were analyzed by GC-MS/GC-FID in parallel, while the phomopsin mycotoxins were analyzed by ELISA, since chromatographic methods were not sensitive enough and required time-consuming sample cleanup. The analyzed lupin containing foods were free of phomopsins. In foods where lupin was only a minor constituent the alkaloid content was of no concern. However, roasted lupin beans intended as coffee surrogate had alkaloid contents close to the Australian intervention limit of 200 microg/g.
Content may be subject to copyright.
Journal of Chromatography A, 1112 (2006) 353–360
Quinolizidine alkaloids and phomopsins in lupin seeds
and lupin containing food
Hans Reinhard , Heinz Rupp, Fritz Sager, Michael Streule, Otmar Zoller
Swiss Federal Office of Public Health, Division of Food Science, CH-3003 Bern, Switzerland
Available online 15 December 2005
Abstract
In recent years there has been growing interest in replacing (genetically modified) soya by lupin. Lupin seeds, flours and lupin containing
food have been analyzed in order to assess the relevance of a potential health hazard given by mycotoxins and/or naturally occurring alkaloids.
Since not all important alkaloids used for quantitation were commercially available, isolation of lupanine, 13-hydroxylupanine and angustifoline
from lupin flours of high alkaloid contents was performed. Alkaloids were analyzed by GC–MS/GC–FID in parallel, while the phomopsin
mycotoxins were analyzed by ELISA, since chromatographic methods were not sensitive enough and required time-consuming sample cleanup.
The analyzed lupin containing foods were free of phomopsins. In foods where lupin was only a minor constituent the alkaloid content was
of no concern. However, roasted lupin beans intended as coffee surrogate had alkaloid contents close to the Australian intervention limit of
200 g/g.
© 2005 Elsevier B.V. All rights reserved.
Keywords: Preparative chromatography; Lupins; Quinolizidine alkaloids; Phomopsins; Dual column GC–MS/GC–FID; LC–DAD–ECD; LC–MS/MS; ELISA
1. Introduction
Lupins are leguminous plants like soybeans. Since lupin and
soya have comparable nutritive characteristics [1], lupins have
been discussed as possible substitute of genetically modified
soya in human foodstuffs. From the genus Lupinus more than
400 species are known, from which only four are of agronomic
interest: Lupinus albus (white lupin), Lupinus angustifolius (nar-
row leaf or blue lupin), Lupinus luteus (yellow lupin) and Lupi-
nus mutabilis (Andean lupin). Main producing countries are
Australia (L. angustifolius), Russia (L. luteus) and Poland (L.
luteus). In Europe especially L. albus and L. luteus are used
as green forage, as manure or are intended for human nutri-
tion. Apart from their benefits [1], lupins contain antinutritive
components such as saponins, tannins and flavonoids in varying
concentrations [2], but also naturally occurring bitter principles.
More than 150 alkaloids of the quinolizidine, piperidine and
indole group (Fig. 1) are known at concentrations up to 6%, con-
ferring the plant resistance to pathogens and herbivores. Though,
plant breeders have developed so called sweet lupins with alka-
Corresponding author. Tel.: +41 31 324 9386; fax: +41 31 322 9574.
E-mail address: hans.reinhard@bag.admin.ch (H. Reinhard).
loid contents below 0.05%, adequate for animal and human
consumption. The alkaloids vary in toxicity, i.e. teratogenicity
of anagyrine [3] and ammodendrine [4] as well as neurotoxicity
of other alkaloids has been documented [5]. Moreover, seeds
of L. angustifolius may additionally be infected by the fungus
Phomopsis leptostromiformis, which could result in contamina-
tion with phomopsin mycotoxins [6] and give rise to lupinosis, a
severe hepatic disease [7–9]. From the five phomopsins reported
in literature [10], only the structures of phomopsin A and B have
yet been elucidated [11] (Fig. 2). Regulations in Australia and
Great Britain demand to comply with a maximum contamination
of 200 g/g alkaloids and of 5ng/g phomopsins in lupin-based
food [12]. Apart from toxicity, the possible allergenic potential
of lupins is under discussion; cross-allergenicity of peanut and
lupin has been shown [13]. Recently, several cases of allergenic
reactions in humans have been reported in Switzerland [14] but
lupin has not been included in the new EU and Swiss legislation
on the mandatory labeling of allergens [15].
The aim of the present work was to unambiguously detect
and quantify lupin alkaloids and mycotoxins in order to assess
a possible health hazard for food products intended for human
consumption on the Swiss market. To reach this objective, it
was necessary to isolate non-commercial alkaloids and develop
analytical methods for a variety of analytes.
0021-9673/$ – see front matter © 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.chroma.2005.11.079
354 H. Reinhard et al. / J. Chromatogr. A 1112 (2006) 353–360
Fig. 1. Exemplary structures of important representatives from quinolizidine,
piperidine and indole alkaloids found in lupins.
Fig. 2. Structure of phomopsin A (PhA), one of the five documented pho-
mopsins. In phomopsin B (PhB), the chlorine atom is replaced by hydrogen.
2. Experimental
2.1. Chemicals and immunoreagents
Gramine (Ord.# G1,080-6) was from Aldrich, lupinine
(Ord.# 01659L) and -isolupanine perchlorate (Ord.# 14343L)
were form Apin (Abingdon, GB), citric acid monohydrate
(Ord.# 27490), N,O-bis(trimethylsilyl)trifluoroacetamide
(BSTFA)/1% trimethylchlorosilane (TMCS) (Ord.# 15238),
N,O-bis(trimethylsilyl)acetamide (BSA, Ord.# 15242),
orthophosphoric acid (85%, ord.# 79620) and pyridine (Ord.#
82702) were from Fluka (Buchs, Switzerland), bovine serum
albumin (BSA, Ord.# A9085), cytisine (Ord.# C2899),
sparteine (Ord.# S2126), phomopsin A (PhA, Ord.# P4829),
tetramethylbenzidine (TMB, Ord.# T2885), Tween 20 (Ord.#
P1379) and phosphate buffered saline tablets (PBS; Ord.#
P4417) were from Sigma, hydrogen peroxide (30%, ord.#.
8597), sodium azide (Ord.# 822335), sodium acetate trihydrate
(Ord.# 6267), sodium hydrogencarbonate (Ord.# 6329), sodium
carbonate (Ord.# 6392), sodium chloride (Ord.# 6404), sodium
hydroxide (Ord.# 6498) and sulfuric acid (Ord.# 731) were
from VWR Int. All these chemicals and the solvents used were
of analytical grade and applied without prior purification. High
purity water was produced by an EasyPure UV system (Mod.
D7402, Barnstead, Dubuque, US). Anti-phomopsin IgG and
phomopsin-horseradish peroxidase enzyme conjugate were
donations from CSIRO, Australian Animal Health Laboratory,
Geelong, Australia.
2.2. Instrumentation
2.2.1. GC
Gas chromatographic separations were performed on a
HP6890 (Agilent), coupled to a mass spectrometer (HP5973,
Agilent) and a flame ionization detector (FID, Agilent). Two
identical capillary columns (HP-1MS, length 30 m, i.d. 0.25 mm,
film thickness 0.25 m, Agilent) were fitted by a two-hole fer-
rule to the injector and to the mass spectrometer and the FID,
respectively. The combined data acquisition of both detectors
resulted in different retention times of the corresponding ana-
lytes. This was overcome through a tool in the acquisition
software (MS ChemStation G1701BA rev. B.02.00, operated
in enhanced mode, Agilent) which requires a two point linear fit
to align the signals of two detectors. As alignment peaks hex-
adecane (C16) and hexacosane (C26) were chosen, also used
as reference peaks for retention index calculations (see Section
2.3.4). Derivatized samples were analyzed on a HP5890 gas
chromatograph coupled to a HP5972 mass spectrometer (both
from Agilent) and separated on a HP-1MS capillary column as
cited above.
2.2.2. Preparative LC
The system consisted of a Merck-Hitachi L6200 gradient
pump, an L3000 DAD (both from VWR Int.), a fraction collec-
tor (FC 203, Gilson) and a degasser (Gastorr GT-103, Omnilab,
Mettmenstetten, Switzerland).
2.2.3. Analytical LC
The chromatographic system consisted of two LC-10AD
high pressure gradient LC pumps (Shimadzu, Kyoto, Japan),
a static mixer (Lee Micro Mixer 10 l, Westbrok, US) a SIL-
10A auto injector (Shimadzu) and a SPD-M10AVp UV diode
array detector (DAD, Shimadzu) with an electrochemical detec-
tor (ECD, Coulochem II, Esa, Chelmsford, US) in series or of a
triple quadrupole mass spectrometer system (MS/MS, API 3000,
ABI-MSD Sciex, Concord, Canada) as detector, respectively. A
Gastorr GT-104 degasser (Omnilab) was used and the temper-
ature of the column was kept at 20 C by a Pelcooler column
oven (Portmann, Biel-Benken, Switzerland). Reversed phase
materials used were custom fills from Macherey-Nagel (Oensin-
H. Reinhard et al. / J. Chromatogr. A 1112 (2006) 353–360 355
gen, Switzerland) and included the stationary phases Inertsil
ODS3 and Spherisorb ODS1, both with 5 m particle size,
which were all in columns of 250 mm ×4 mm and additionally
equipped with a pre-column Spherisorb ODS2 (8 mm ×4 mm,
5m particles) for LC–DAD–ECD. For LC–MS/MS Inertsil
ODS2 columns, 125 mm ×2 mm, 3 m particles, were used.
2.3. Alkaloids
2.3.1. Isolation of reference alkaloids
For extraction 50 g of freshly ground lupins (particle size
<0.5 mm, milled on a Mod. ZM1, Retsch, Haan, Germany) of
a bitter variety of L. angustifolius were homogenized at high
speed in 500 ml 0.5 M hydrochloric acid for 5 min with a poly-
tron mixer (Kinematica, Littau, Switzerland). The slurry was
centrifuged at 4600 rpm for 25 min at 15 C (Labofuge 400R,
Heraeus, Z¨
urich, Switzerland). A slightly turbid extract resulted,
which was adjusted to pH 9.1 with concentrated ammonia solu-
tion. The extract, ca. 250 ml, was then partitioned over 5 ×50 ml
ChemElut cartridges (Ord.# 1219-8009, Varian) and after 20 min
eluted with 2 ×100 ml of dichloromethane. The combined elu-
ates were evaporated to dryness. Depending on the lupin species
chosen the extraction resulted 300–600 mg crude isolate.
For pre-cleaning, crude isolate was dissolved in portions of
ca. 100 mg in diethyl ether saturated with 5% ammonia (v/v) and
charged on a 5 g silica cartridge (Ord.# 36950, Waters) which
had been conditioned with diethyl ether. The coated isolate was
then washed with 10 ml of solvent. The main component of
this elution fraction was lupanine. Subsequently, methanol was
added to the solvent in four steps (10, 25, 50 and 75%, v/v,
respectively) and elution continued with increasing methanol
content. The 10% methanol fraction contained lupanine, 13-
hydroxylupanine and angustifoline, the 25% methanol frac-
tion contained mainly 13-hydroxylupanine with portions of
lupanine and angustifoline, while the fractions with higher
methanol content were discarded. This pre-cleaning procedure
was applied several times until about 300–400 mg substance was
gained per fraction.
In a next step, isocratic column chromatography was run
on a 250 mm ×10 mm Nucleosil C18 column or on the better
performing 250 mm ×4.6 mm C18ec column (both with parti-
cle size 5 m, Macherey-Nagel) with 5mM KH2PO4-methanol
(94:6) as mobile phase. Injection volumes were 20–100 mg iso-
lates. Fractions of about 20 ml were collected and analyzed.
From several separations, fractions of identical content were
combined and concentrated. This procedure led to an alkaloid
purity of about 90%. The methanol content of the mobile phase
was then changed to 12% (v/v, pH 2.7), resulting in shorter
retention times. In the subsequently resulting fractions only the
main zones were collected, while the front, end and mixed zones
were fed as collected fraction to repeated chromatographic sep-
aration, from which again the main zones were gained. The
collected main fractions of all three alkaloids had colors from
yellow to light brown. A further cleanup on silica cartridges
as described above largely removed this inconvenience. At this
point alkaloid purity had reached about 95%. Next, gradient
chromatographic separation on a 300 mm ×7.8 mm silica col-
umn (filled with C-Gel C640, particle size 15–35 m, Zeochem,
Uetikon, Switzerland) with diethyl ether saturated with ammo-
nia (5%, v/v)–methanol as mobile phase was applied, varying the
eluent from (90:10) to (70:30). The collected main zones were
concentrated under nitrogen flow at <40 C and remaining water
and ammonia removed through azeotropic evaporation with
ethanol. The residues were dissolved in methanol and passed
through a syringe filter (0.45 m, PTFE, Titan, Infochroma,
Zug, Switzerland) in order to remove insoluble particles. Stor-
age of methanolic alkaloid solutions at 5 C overnight resulted in
white, voluminous precipitates for all three alkaloids, identified
as fatty acids by GC–MS. To remove the precipitates, a cleanup
over magnesium silicate cartridges (0.5 g, Florisil Sep-Pak, ord.#
43405, Waters) was carried out with diethyl ether saturated with
ammonia (5%, v/v)–methanol. Methanol content was varied
from 0 to 5% (v/v). This procedure removed also the residual
slight coloration. The resulting isolates were of sticky, syrup-like
consistency. After removing insoluble components (<1%), crys-
tallization from hexane resulted in pseudo-crystalline angusti-
foline and 13-hydroxylupanine, while lupanine remained oily.
2.3.2. Sample preparation
Seeds were frozen in liquid nitrogen and milled to a particle
size of <0.5 mm (Mod. ZM1, Retsch). One gram of dry lupin
flour or up to 2.5 g of fresh lupin food product was weighed
in a centrifuge tube. Fifteen milliliters 0.5 M hydrochloric acid
were added and samples were kept for 10 min at room temper-
ature. The suspension was then homogenized for 3 min with
a polytron mixer, chilled to 5C and subsequently centrifuged
at 6000 rpm for 15 min at 5 C (Labofuge 400R). The super-
natant was decanted in a tared flask and the procedure described
above was repeated with addition of hydrochloric acid. Super-
natants were combined, adjusted to pH 10 by adding ammonia
(aqueous solution 25%, v/v) and weighed. Twenty milliliters of
the alkalized solution were applied to a 20 ml ChemElut car-
tridge (Varian) and after 15 min the alkaloids were eluted by
two-fold 40 ml dichloromethane application. The solvents were
evaporated to dryness under reduced pressure. The residue was
dissolved in 5 ml methanol, transferred into a vial, evaporated to
dryness under nitrogen flow and stored at 20 C. For GC anal-
ysis, 1 ml methanol containing 50 g/ml caffeine for response
factor calculation, 5 g/ml C16 and 5 g/ml C26 for retention
index calculation was added. Subsequently, 1 l of this solution,
bitter lupin extracts only after further dilution, was injected.
2.3.3. Derivatization
In addition to the sample preparation described above (Sec-
tion 2.3.2.), derivatization with BSA/TMCS/pyridine 8:1:1
(v/v/v), 10 min at 60 C was used in cases were the analysis
was hampered by matrix peaks or lack of analytes detectability
without derivatization (see Section 3.1.2).
2.3.4. GC conditions
The GC equipped with two identical capillary columns in
parallel (HP-1MS, length 30 m, i.d. 0.25 mm, film thickness
0.25 m) was operated with a temperature program starting at
80 C held for 1 min, increasing to 300 Cat15C/min and held
356 H. Reinhard et al. / J. Chromatogr. A 1112 (2006) 353–360
for 7 min at the final temperature. As carrier gas helium at a con-
stant flow of 1 ml/min was used. The injector was operated in
split mode (20:1) at a temperature of 250 C. Injected volumes
were 1 l. The same conditions applied for the GC HP5890, on
which derivatized samples were separated. The FID was oper-
ated with hydrogen at 40 ml/min, air at 450 ml/min, nitrogen as
makeup gas at 45 ml/min at a temperature of 250 C.
Standard solutions were prepared by dissolving 2–5 mg alka-
loid as free base in 10 ml methanol. While gramine solution had
to be prepared daily, quinolizidine standards proved to be stable
for several days. Quantification was carried out by evaluating
the signals from the FID, unless derivatization was necessary.
In matrix spoilt samples 13-hydroxylupanine and angustifo-
line had to be quantified through peak evaluation from mass
spectra at m/z= 246 and 265, respectively (see Section 3.1.2).
For FID quantification, prior determination of response fac-
tors (RF) was required: peak areas of standard alkaloids were
related to caffeine (RF1), resulting in RF-values for lupa-
nine (0.45), 13-hydroxylupanine (0.58), angustifoline (0.59),
gramine (0.69) and sparteine (0.37). For -isolupanine the RF
of lupanine was used.
Retention indices (RI) were calculated according to [16], with
hexadecane (C16, RI1600) and hexacosane (C26, RI2600)
as reference peaks. Unexpected trace alkaloids were identified
by RI and MS [24].
2.4. Phomopsins
2.4.1. ELISA conditions
The competitive ELISA was carried out as described by Than
et al. [17]. Briefly, microtiter plates (Ord.# 439454, 96 F bottom
wells from Nunc, Denmark) were coated with anti-phomopsin
antiserum in aqueous carbonate/NaN3coating buffer. The plates
were then incubated overnight at 4 C. After 17–20 h of adsorp-
tion, the plates were washed with NaCl/Tween 20 aqueous
solution. Standards (range 5–1280 pg PhA/100 l) or samples
(50 l/well) and diluted phomopsin-enzyme conjugate (dilution
1:500, 0.5% BSA in PBS) were added and incubated after short
agitation for 3 h at room temperature. Following washing, TMB
substrate solution was added and after 15 min of incubation the
reaction was stopped by adding 0.5 M sulfuric acid. The opti-
cal density was measured at 450 nm (EIA Reader Mod. 2550,
BioRad) and a logistic dose response curve fitted through the
phomopsin standards data from which the PhA + B concentra-
tion of unknown samples was calculated. All determinations
were performed in four-fold.
According to Than [18], samples were prepared as fol-
lows: a representative sample of seed, e.g. 50g, was soaked
in 250 ml methanol–water (4:1, v/v) overnight at room temper-
ature and homogenized the next day for 3 min at high speed.
The resulting paste was stirred for 2 h and then centrifuged
at 2000 rpm for 10 min (Labofuge 400R). The supernatants
were stored at 18 C until analysis. Samples were diluted
1:10 with assay buffer before analysis. Flour samples were
treated correspondingly to the above, only omitting the unnec-
essary homogenization, but were stirred during at least 2 h of
soaking.
2.4.2. LC–DAD–ECD
The mobile phase consisted of methanol-133 mM orthophos-
phoric acid pumped at a flow rate of 1 ml/min. Gradient elution
was achieved as follows: eluent composition changed from
(80:20) to (40:60) in 30 min, raised to (80:20) within 1 min
and the system conditioned for 5 min at (80:20). The DAD
was set to detection wavelengths 210 and 290nm, respectively.
PhA exhibits UV absorption of λMeOH
max nm (ε): 210(52297) [this
work], 210(50023) [19] and 290(16154) [this work], 290(16030)
[19], respectively. The ECD was equipped with an amperomet-
ric analytical cell operated in DC mode (Mod. 5040 from Esa,
cell potential 600 mV, gain range 100 nA).
2.4.3. LC–MS/MS
The mobile phase consisted of methanol-formic acid (0.1%,
v/v) at a flow rate of 0.2 ml/min. During linear gradient elution,
the eluent composition changed from (70:30) to (10:90) within
15 min, was then raised to (70:30) and kept constant for 5 min.
Under these conditions PhA eluted after 6 min. The mass spec-
trometer was equipped with an electrospray ion source operated
in positive ion mode at an ionization voltage of 5kV, heated at
350 C. The nebulizer gas flow was set to 8l N2/min. The mass
transitions used for quantification were (precursor product):
789 323 m/zand 789 226 m/z, acquired during 150 ms
(dwell time) each. Collision energies to generate the product
ion 323 and 226 m/zwere 39 and 57 eV, respectively. Nitrogen
was used as collision gas at a pressure of 1 Pa within the colli-
sion quadrupole (readout correct according to [20]). Both mass
selective quadrupoles were operated at unit resolution condi-
tions.
3. Results and discussion
3.1. Alkaloids
3.1.1. Physicochemical properties of isolates
To confirm identity and purity of isolates, several
physicochemical investigations were applied (Table 1).
Chromatographic purity was sufficient for lupanine and
13-hydroxylupanine, while angustifoline contained minor
impurities. Elemental analysis revealed a somewhat increased
oxygen content for 13-hydroxylupanine and angustifoline.
Retention index, response factor, melting point and mass
spectral fragments compare well with literature, angustifoline’s
melting point being somewhat low, though. The RF of 13-
hydroxylupanine was sensitive to the degree of GC injector
liner contamination and could subsequently rise to an RF value
of 0.75. The susceptibility of 13-hydroxylupanine had been
reported before [25], where a vast relative standard deviation
(RSD) for RF determination was described.
3.1.2. GC–FID and GC–MS analysis
The developed method for alkaloid determination proved to
be reliable and stable. Unfortunately, it was not possible to quan-
tify alkaloids by FID and confirm identity by MS. Lupin flours
from Australian provenience, i.e. from L. angustifolius, required
sample derivatization, since matrix peaks in the chromatograms
H. Reinhard et al. / J. Chromatogr. A 1112 (2006) 353–360 357
Table 1
Physicochemical properties of isolated alkaloids
Lupanine 13-Hydroxylupanine Angustifoline
GC–MS purity, cold on column (%) >99 >99 97
GC–MS purity, BSTFA derivatization (%) >99 >99 94
LC-DAD purity at 200 and 220 nm (%) >99 >99 97
Elemental analysis (calc. value, conformity) C 72.45 (72.54, 99.87%), H
9.66 (9.74, 99.18%), N 11.19
(11.28, 99.20%), O 6.34 (6.44,
98.44%)
C 67.77 (68.15, 99.44%), H 9.17
(9.15, 100.21%), N 10.17 (10.06,
95.94%), O 12.75 (12.10, 105.37%)
C 71.23 (71.76, 99.26%), H 9.31
(9.46, 98.41%), N 11.49 (11.95,
96.15%), O 6.17 (6.83, 119.7%)
Melting point (ref. value) (C) 38 (41 [21]) 171–172 (169.5 [21]) 74 (79–80 [22])
IR νKBr
max (cm1) 2930, 1640, 1440 ([23]) 3420, 2925, 2360, 1630, 1440 3074, 2910, 1640, 1440
Retention index (ref. value)a2204 (2165 [24]) 2447 (2410 [24]) 2111 (2083 [24])
Response factor (ref. value)b0.45 (0.51 [25]) 0.58 (0.63 [25]) 0.55 (0.50 [25])
MS [m/z] from EI (relative intensity, %) M+= 248 (59); 136 (100), 149
(60), 248 (59), 150 (44), 247
(38). [24,26-28]
M+= 264 (41); 152 (100), 246 (55),
165 (44), 264 (41), 134 (38).
[24,26–28]
M+= 234 (0.2); 193 (100), 112
(54), 150 (17), 194 (13), 55 (10).
[24,27, 28]
aAccording to Kovats [16], with C16 and C26 as reference peaks.
bCompared to caffeine [25].
masked the alkaloid content (Fig. 3). Derivatization led to peak
shifts for 13-hydroxylupanine and angustifoline, allowing their
unambiguous detection. Lupinine could only be detected after
derivatization, being one of the minor constituents (0.6–2% of
Fig. 3. GC–MS chromatograms for underivatized seeds of L. luteus,L. albus and
L. angustifolius. Note that the ordinate is in log scale in order to visualize small
peaks: (1) C16; (2) gramine; (3) caffeine (IS); (4) sparteine; (5) ammodendrine;
(6) angustifoline; (7) -isolupanine; (8) lupanine; (9) 13-hydroxylupanine and
(10) C26.
total alkaloid content), though. Derivatized alkaloids were quan-
tified by MS although FID would have been more favourable due
to better reproducibility. The limit of detection for trimethylsi-
lyl derivatives was a factor of 6–12 lower than for underivatized
alkaloids. We first used derivatization with BSTFA/TMCS, but
since for angustifoline two products formed and conversion was
incomplete, this procedure was dropped. Instead, sample deriva-
tization with BSA/TMCS/pyridine was used (see Section 2.3.3),
giving still rise to two angustifoline products but at complete
conversion, though. The first product mass spectrum showed
the expected loss of the allyl side chain as base peak and high-
est mass ion (m/z= 265, M-41+), while the second product
spectrum contained the ions m/z= 335 (14%), 309 (60%) and
265 (100%). Silylation artifacts are well known from literature
and in our case it seems that a CO2-incorporation, documented
especially for amino acids [29], i.e. carboxylate formation had
happened. The fragments of the second product are tentatively
interpreted as M-15+(loss of methyl), M-41+(loss of allyl
side chain) and M-85+(loss of CO2and allyl side chain after
rearrangement), respectively.
From the teratogenic alkaloids reported, only ammodendrine
could be detected in seeds of L. albus in trace amounts, but was
not detectable in ready-to-consume food.
Recoveries from a flour sample spiked at a 10 ng level
were 115% for lupanine, 43% for 13-hydroxylupanine, 49%
for angustifoline, 93% for cytisine and 139% for sparteine.
For trimethylsilylated 13-hydroxylupanine a recovery of 90%
resulted. From a standard solution recoveries (±RSD, n=3)
were: 99% (±6%) for lupanine, 94% (±11%) for 13-
hydroxylupanine and 50% (±28%) for angustifoline. The lim-
its of detection (LOD, S/N= 3) related to flour were 1 g/g
for lupanine, 2 g/g for angustifoline and 6 g/g for 13-
hydroxylupanine. For trimethylsilylated angustifoline and 13-
hydroxylupanine the LOD was 0.3 and 0.5 g/g, respectively.
To check reproducibility, a lupin flour was processed five
times and the RSD found was 3% for lupanine and 7% for
13-hydroxylupanine, while total alkaloid content was 97 and
77 g/g, respectively. For a trimethylsilylated lupin flour pro-
358 H. Reinhard et al. / J. Chromatogr. A 1112 (2006) 353–360
Table 2
Content of alkaloids and phomopsin A and B in investigated seeds and flours
Seeds and flours AlkaloidsaPhomopsin A + B
“Bitter” (g/g) n“Sweet” (g/g) n(ng/g) n
L. albus 202 1 143–226 5 <0.1 4
L. luteus 0 500–895 4 0
L. angustifolius 10800–19800 3 44–2120 median: 138 18 <0.1–20b8
aSum of standard alkaloids (see Sections 2.3.4 and 3.1.2).
bOne unique positive sample, see text.
cessed five times, 13-hydroxylupanine was found to have an
RSD of 14% at a total alkaloid content of 70 g/g.
The found alkaloid contents of seeds and flours are given in
Table 2, while Fig. 3 depicts typical chromatograms of the three
species studied. The samples have been grouped according to
their origin in L. albus,L. luteus and L. angustifolius and into
bitter or sweet varieties, according to supplier’s indications. For
quantitation, the found standard alkaloids (see Section 2.3.4)
were summed. For samples from the L. albus group, included
alkaloids were lupanine, 13-hydroxylupanine and angustifo-
line. The first two were directly quantifiable by FID, the lat-
ter was quantified by MS after trimethylsilylation. Traces of
-isolupanine, dihydroxylupanine, ammodendrine, albine and
13-angeloyloxylupanine or 13-tigloyloxylupanine were also
found, identified by RI and MS [24].IntheL. luteus group
gramine and sparteine were the predominant alkaloids directly
quantified from FID, while lupinine was present in minor con-
centrations of 3–10 g/g, quantified by MS after trimethylsily-
lation. The samples of L. angustifolius showed an origin related
behavior: Samples from Australia had high matrix load and
were quantified similar to samples from L. albus. The remain-
ing samples contained noticeable amounts of -isolupanine
(range 40–90 g/g for the sweet, 237 g/g for the bitter vari-
ety), which could be directly quantified together with lupanine,
13-hydroxylupanine and angustifoline by FID.
For L. albus, the alkaloid content of bitter and sweet varieties
was comparable, while for L. angustifolius there was a 10–20-
fold increase in alkaloid content for the bitter varieties (Table 2).
In samples of the L. albus group lupanine varied from 69 to
111 g/g, 13-hydroxylupanine from 48 to 117 g/g and angus-
tifoline from 9 to 12 g/g. In L. luteus gramine varied in the range
450–893 g/g, sparteine 2–50 g/g and lupinine 3–12 g/g.
The bitter varieties of L. angustifolius contained lupanine in
the range 4.0–12.9 mg/g, 13-hydroxylupanine 5.0–6.6 mg/g
and -isolupanine 213–237 g/g, while sweet lupin seeds
and flours had 18–862 g/g of lupanine, 18–943 g/g of
13-hydroxylupanine, 7–226 g/g of angustifoline and those
not from Australian provenience additionally 37–90 g/g -
isolupanine.
Table 3 shows the alkaloid contents found in foods on
the Swiss market. According to the alkaloid pattern, mainly
lupins from L. albus and exceptionally from L. angustifolius
but not from L. luteus have been used, i.e. lupanine, 13-
hydroxylupanine and angustifoline were the principal alkaloids
found. -Isolupanine, 13-angeloyloxylupanine or 13-
tigloyloxylupanine were present in trace amounts in two of the
coffee products, in the curryburger, the lupin tofu products and in
one of the lupin flour containing breads. For quantification, in all
food samples except coffee, lupanine and 13-hydroxylupanine
could be directly quantified from FID. The matrix in coffee
samples implied trimethylsilylation of 13-hydroxylupanine
and angustifoline. Unfortunately, even after trimethylsilylation
of angustifoline the matrix in all food samples except the lupin
snack product, where all three alkaloids were directly quan-
tifiable from FID, prevented data interpretation. Since alkaloid
ratios in our investigated seeds and flour measurements were
relatively constant within one species and have also been used
as taxonomical markers [30], it can be assumed, that the lost
angustifoline content is in the order of 5–10% of the summed
up value for lupanine and 13-hydroxylupanine. The values in
Table 3 have not been correspondingly corrected, though.
As for the found alkaloid contents, the coffee samples show
outstanding results and almost reach the Australian maximum
level for human exposure set at 200g/g of derived lupin prod-
ucts. While the other commercial products contain lupin flour or
isolates in limited content, the coffee powder is made of roasted
lupin grains alone. Alkaloid transfer from roasted lupin beans to
the ready-to-consume drink was about 50–70% at a water extrac-
tion temperature of 90 C. Therefore, the possible advantage of
the absence of caffeine has to be weighed against an important
intake of toxic alkaloids.
3.2. Phomopsins
3.2.1. ELISA
The applied ELISA-kit proved to be sensitive and reliable,
but is not commercially available, yet. The kit detects both PhA
Table 3
Lupin containing foods investigated. Values of found alkaloids and phomopsin
A and B are given
Food products AlkaloidsaPhomopsin A + B
(g/g) n(ng/g) n
Lupin coffee 136–182 4 <0.1 3
Curryburger 4 1 <0.1 1
Lupin tofu untreated 12–15 2 <0.1 2
Lupin tofu with tomato 3 1 <0.1 1
Lupin bread (lupin flour
content 7-10%)
8–12 2 <0.1 1
Falafel with lupin tofu 3 1 0
Lupin snack 6 1 0
aSum of standard alkaloids (see Sections 2.3.4 and 3.1.2).
H. Reinhard et al. / J. Chromatogr. A 1112 (2006) 353–360 359
and PhB [17]. PhA is two to five times more toxic than PhB
[31], therefore it would be desirable to have a chromatographic
confirmation method being able to distinguish between the two
mycotoxins in the low ng/g range.
Typical RSD for samples was ±20%. The best sensitivity for
this assay was in the range 20–320 pg PhA/well and the limit
of detection was 0.1 ng/g related to the samples. The recovery
from lupin flour spiked with PhA was in the range 89–111% at
a concentration of 7 ng/g PhA.
The found PhA + B contents of seeds and flours are given
in Table 2. All samples except one were free of phomopsins,
expected especially in seeds and flours from L. angustifolius. The
only positive probe was optically striking, since attack of pho-
mopsis results in darkened grain hulls [6]. Thus, it was already
foreseen to dispose the positive sample and its corresponding
fodder batch.
Table 3 shows the PhA + B contents found in foods on
the Swiss market. Phomopsins were absent in all of the
analyzed food samples, i.e. no value was in the order of
magnitude of the Australian limit value of 5 ng/g for fin-
ished food products. For both L. albus and L. angus-
tifolius resistant varieties against phomopsis are known,
though. In the future, genetically modifications could enhance
resistance to lupin diseases, such as phomopsis infection
[32].
3.2.2. LC
The attempted chromatographic methods for phomopsins
detection failed, since the attainable limit of detection was too
high compared to those of ELISA. The detection limits for
PhA for DAD (at 290 nm) and ECD were 450 and 50 ng/g,
respectively, for a signal to noise ratio (S/N) of three from a
standard solution injected on a Spherisorb ODS1 column. For
LC–MS/MS, the achievable limit of detection was 1ng/g PhA
(S/N= 3).
Additionally, sample cleanup turned out to be crucial, while
for the ELISA procedure no cleanup was required. Although
several attempts were made to establish an efficient cleanup
method for PhA, i.e. cleanup of lupin flour samples over
C18, SAX, SiOH and Oasis according to the manufacturer’s
instructions, all efforts were unsuccessful. An LC-UV method
to detect PhA at a 500 ng/g level from lupin stubble had
been published before, using an Amberlite XAD-2 and a
cation exchange cleanup, requiring an analysis time of 2 days
[33].
In a lupin flour sample spiked with PhA chromatographic
interference between PhA and genisteine occurred, i.e. genis-
teine coeluted with PhA. The estimated genisteine content was
3.5 g/g related to the lupin flour sample from L. angustifolius,
completely masking any potential PhA presence.
Precision of LC–MS/MS was evaluated by injecting 10l
aliquots of standard solution containing 3.2 ng PhA in a gradient
run. The ion transition 789 226 m/zwas monitored after 4 h of
instrument’s warm-up, all other parameters were set as described
in Section 2.4.3. Reproducibility of retention time and peak area
over a 10 h run resulted in an RSD of 0.5% (n= 10) and 8.9%
(n= 10), respectively.
4. Conclusions
There’s no urgent call for action concerning the alkaloid
and phomopsin A + B contents in foods on the Swiss market.
Though, products made of seeds alone, such as roasted lupin
beans promoted as caffeine free coffee surrogate, have consid-
erable alkaloid contents and almost reach the Australian limit
for finished food products. Toxicological interpretation might
be indicated for these products. The Australian New Zealand
Food Authority has set a tolerable level of exposure to lupin
alkaloids for humans at 35 g/kg/day [34].
From the alkaloids found in the Swiss market products it can
be concluded that mainly L. albus, rarely L. angustifolius and
never L. luteus had been used in production.
The allergenic potential of lupins seems to be of major con-
cern [14]. An appropriate declaration at inclusions of lupin
in wheat flour was suggested due to the crossed peanut–lupin
allergy potential [13].
In our hands, trimethylsilylation of alkaloids was not ade-
quate, since silylation artifacts were noted. Here, a CO2-
incorporation occurred, which according to [29] could be
avoided by adding diluted hydrochloric acid during the final
stage of drying the samples prior to derivatization. Further,
acylation instead of silylation could be more adequate for the
analysis of amino compounds [35].
Additional efforts should be taken to develop a confirma-
tion method for phomopsins in the low ng/g range, especially to
identify which one of the five phomopsins is present since the
differences in their toxicity is of relevance [31].
Acknowledgments
We gratefully acknowledge the help and support from John
A. Edgar and Khin A. Than, both from CSIRO, Australian Ani-
mal Health Laboratory, Australia, who obligingly offered us the
phomopsin ELISA-kit they had developed. We are also indebted
to Bernhard Zimmerli for his farsighted planning and accompa-
nying of the project and for sharing his versatile knowledge and
to Michael Beer for valuable contributions to this article’s style.
References
[1] R.J. van Barneveld, Nutr. Res. Rev. 12 (1999) 203.
[2] C.R. Sirtori, M.R. Lovati, C.M. Manzoni, S. Castiglioni, M. Duranti, C.
Magni, S. Morandi, A. D’Agostina, A. Arnoldi, J. Nutr. 134 (2004) 18.
[3] R.F. Keeler, E.H. Cronin, J.L. Shupe, J. Toxicol. Environm. Health 1
(1976) 899.
[4] R.F. Keeler, K.E. Panter, Teratology 40 (1989) 423.
[5] J. Pothier, S.L. Cheav, N. Galand, D. Dormeau, C. Viel, J. Pharm.
Pharmacol. 50 (1998) 949.
[6] J.A. Edgar, in: M., Dracup, J., Palta, (Eds.), Proceeding of the First Aus-
tralian Lupin Technical Symposium, Department of Agriculture, South
Perth, AU, 1994, p. 85.
[7] M.R. Gardiner, D.S. Petterson, J. Comp. Pathol. 82 (1972) 5.
[8] M.V. Jago, J.E. Peterson, A.L. Payne, D.G. Campbell, Aust. J. Exp.
Biol. Med. Sci. 60 (1982) 239.
[9] E.M. T¨
onsing, P.S. Steyn, M. Osborn, K. Weber, Eur. J. Cell Biol. 35
(1984) 156.
[10] P.H. Mortimer, J.W. Ronaldson, in: R.F. Keeler, A.T. Tu (Eds.), Hand-
book of Natural Toxins, vol. 1, Dekker, New York, US, 1983, p. 390.
360 H. Reinhard et al. / J. Chromatogr. A 1112 (2006) 353–360
[11] M.F. Mackay, A. van Donkelaar, C.C.J. Culvenor, J. Chem. Soc. Chem.
Commun. (1986) 1219.
[12] Anonymous, ACNFP Report on Seeds from the Narrow Leafed Lupin,
Appendix IX, MAFF Publications, London, GB, 1996, p. 107.
[13] D.A. Moneret-Vautrin, L. Guerin, G. Kanny, J. Flabbee, S. Fremont, M.
Morisset, J. Allergy Clin. Immunol. 104 (1999) 883.
[14] B. W ¨
uthrich, D. Mittag, B.K. Ballmer-Weber, Allergologie 27 (2004)
495, article in German.
[15] Anonymous, Directive 2003/89/EC and 2005/26/EC of the European
Parliament and Council, Brussels, BE, 2003 and 2005, respectively.
[16] E. Kovats, Helv. Chim. Acta 206 (1958) 1915.
[17] K.A. Than, A.L. Payne, J.A. Edgar, in: L.F. James, E.M. Bailey, P.R.
Cheeke, M.P. Hegarty (Eds.), Poisonous Plants, Iowa State University
Press, Iowa, US, 1992, p. 259.
[18] K.A. Than, CSIRO, Australian Animal Health Laboratory, Australia,
private communication.
[19] C.C. Culvenor, A.B. Beck, M. Clarke, P.A. Cockrum, J.A. Edgar, J.L.
Frahn, M.V. Jago, G.W. Lanigan, A.L. Payne, J.E. Peterson, D.S. Pet-
terson, L.W. Smith, R.R. White, Aust. J. Biol. Sci. 30 (1977) 269.
[20] B.A. Thomson, D.J. Douglass, J.J. Corr, J.W. Hager, C.L. Jolliffe, Anal.
Chem. 67 (1995) 1696.
[21] D. Lide. (Ed.), Properties of Organic Compounds on CD-ROM V6,
Chapman&Hall/CRC Press, Boca Raton, US, 2000.
[22] M. Wiewi´
orowski, J. Skolik, Roczniki chemii 33 (1959) 461, article in
German.
[23] F. Tosun, M. Tanker, T. ¨
Ozden, A. Tosun, Planta Med. 52 (1986) 242.
[24] M. Wink, in: P. Waterman (Ed.), Methods in Plant Biochemistry, vol. 8,
Academic Press, London, GB, 1993, p. 197.
[25] C.R. Priddis, J. Chromatogr 261 (1983) 95.
[26] M. Muzquiz, C. Cuadrado, G. Ayet, C. de la Cuadra, C. Burbano, A.
Osagie, J. Agric. Food Chem. 42 (1994) 1447.
[27] T. Aniszewski, J. Sci. Food Agric. 61 (1993) 409.
[28] F.W. McLafferty (Ed.), Wiley Registry of Mass Spectral Data, seventh
ed., John Wiley and Sons Inc., Hoboken, US, 2000.
[29] J.L. Little, J. Chromatogr. A 844 (1999) 1.
[30] T. Aniszewski, Sci. Legumes 1 (1994) 111.
[31] J.G. Allen, G.R. Hancock, J. Appl. Toxicol. 9 (1989) 83.
[32] M. Shankar, M.W. Sweetingham, W.A. Cowling, Euphytica 125 (2002)
35.
[33] G.R. Hancock, P. Vogel, D.S. Petterson, Aust. J. Exp. Agric. 27 (1987)
73.
[34] Anonymous, Australian New Zealand Food Authority, Technical
Report Series No. 3, 2001, p. 1. (http://www.foodstandards.gov.au/
srcfiles/TR3.pdf).
[35] R.P. Evershed, in: K. Blau, J.M. Halket (Eds.), Handbook of Derivatives
for Chromatography, second ed., Wiley, Chichester, GB, 1993, p. 51.
... According to C. P. Smith's book, at least 500 species Lupinorum, as he called them, had been described up to that time. The same number is also referred to by Wink et al. [5], while Gresta et al. [6] listed only 170 species and Reinhard et al. [7] over 400. Similar differences are found in many more studies. ...
... Most articles about the analysis and quantitation of QAs use gas chromatogra equipped with a flame ionization detection (GC-FID) [28,29] or mass spectrometry MS) [7,[30][31][32][33][34], while HPLC application for that purpose is extremely limited [35][36][37] instance, a GC-MS based method was developed from Resta et al. using lupanine a ternal standard in samples of lupin-based products [31] and from Romeo et al. using s teine [33]. The quantitation of Lupinus alkaloids is complicated by the fact that comme standards are not easily available. ...
... products or seeds. Historically, the quantitation of QAs has relied on chromatographic methods, which, although effective, entail prolonged analysis times and extensive extraction or pretreatment procedures [7,[28][29][30][31][32][33][34][35][36][37]39]. Analytical methods based on capillary high resolution gas chromatography have been in widespread use since the early 1980s, with minimal evolution over time. ...
Article
Full-text available
Quinolizidine alkaloids (QAs) are toxic secondary metabolites of the Lupinus species, the presence of which limits the expansion of lupin beans consumption, despite their high protein content. Evaluation of the level of alkaloids in edible Lupinus species is crucial from a food safety point of view. However, quantitation of QAs is complicated by the fact that not all important alkaloids used for quantitation are commercially available. In this context, we developed a method for the simultaneous quantitation of eight major lupin alkaloids using quantitative NMR spectroscopy (qNMR). Quantitation and analysis were performed in 15 different seed extracts of 11 Lupinus spp. some of which belonged to the same species, with different geographical origins and time of harvest, as well as in all aerial parts of L. pilosus. The mature seeds of L. pilosus were found to be a uniquely rich source of multiflorine. Additionally, we developed a protocol using adsorption or ionic resins for easy, fast, and efficient debittering of the lupine seeds. The protocol was applied to L. albus, leading to a decrease of the time required for alkaloids removal as well as water consumption and to a method for QA isolation from the debittering wastewater.
... GC-MS is a common method used to detect quinolizidine alkaloids in different variety of lupin-based products. Lupanine, 13α-hydroxylupanine, angustifoline and other pyrrolizidine alkaloids have been found in lupin seeds, lupin flours and different lupin species Reinhard et al., 2006;Resta et al., 2008;Wink et al., 1995). GC-MS and GC-FID can also be applied in parallel for this analysis (Reinhard et at., 2006). ...
... Lupanine, 13α-hydroxylupanine, angustifoline and other pyrrolizidine alkaloids have been found in lupin seeds, lupin flours and different lupin species Reinhard et al., 2006;Resta et al., 2008;Wink et al., 1995). GC-MS and GC-FID can also be applied in parallel for this analysis (Reinhard et at., 2006). ...
... GC-MS/GC-FID was used in parallel for the analysis of the quinolizidine alkaloids lupanine, 13α-hydroxylupanine, α-isolupanine and angustifoline in lupin flours with high alkaloid contents. The total alkaloid concentrations in the bitter varieties ranged from 10,800 to 19,800 mg/kg and those in the sweet varieties ranged from 44 to 2120 mg/kg (Reinhard et al., 2006). ...
Article
Alkaloids have been utilized by humans for years. They have diverse applications in pharmaceuticals. They have been proven to be effective in treating a number of diseases. They also form an important part of regular human diets, as they are present in food items, food supplements, diet ingredients and food contaminants. Despite their obvious importance, these alkaloids are toxic to humans. Their toxicity is dependent on a range of factors, such as specific dosage, exposure time and individual properties. Mild toxic effects include nausea, itching and vomiting while chronic effects include paralysis, teratogenicity and death. This review summarizes the published studies on the toxicity, analytical methods, occurrence and risk assessments of six major alkaloid groups that are present in food, namely, ergot, glycoalkaloids, purine, pyrrolizidine, quinolizidine and tropane alkaloids.
... In the literature, one of the earliest used analytical methods for the determination of PHO-A is Enzyme-Linked-Immuno-Sorbent Assay (ELISA), as described by Reinhard et al., providing a sensitive and reliable method; the assay was not able to discriminate PHO-A from PHO-B in the low ng/g range [9]. Bioassays for the detection of PHO-A, such as Nursling rat bioassay (NRB) [10] have been also developed, however, since they are time-consuming, use experimental animals and require staff expertise, it is hard to consider them as a standard screening approach for PHOs detection in foods [4]. ...
... In the last two decades, confirmatory analytical methods were developed for the detection of PHOs in food samples using high performance liquid chromatography (HPLC) coupled to different detectors such as Ultraviolet (UV) [11], diode array with an electrochemical detector (DAD-ECD) [9] and tandem mass spectrometry (MS/MS). The latter has recently emerged as a powerful tool for this goal, since it exhibits a significant sensitivity and specificity, needed to meet accurately the maximum regulatory levels (MRL); for this reason, it is widely used for the quantification of mycotoxins in food and feed [12]. ...
... It is possible to distinguish two main types of lupins, sweet and bitter varieties, which differ in the alkaloid content [4]. In the first type, developed by plant breeders, the alkaloid amount ranges between 100 and 800 mg Kg −1 of dry weight; consumption by humans and animals is allowed [8]. Bitter lupins are characterized by a higher concentration of alkaloids, usually between 5000 and 40,000 mg Kg −1 of dry weight [9]. ...
... Although different approaches are presented in the literature, according to the German Federal Institute for Risk Assessment (BfR) Opinion No 003/2017, only of a few methods for the quantification of the alkaloids in lupins and derived foods are described and validated in the literature because of the limited availability of standard substances [7]. Most of these methods are mainly based on GC-MS [5,8,13,16,26] and GC with flame ionization detection (GC-FID) [27,28]. A study that aimed to determine the LAs in different Lupinus species by non-aqueous capillary electrophoresis (NACE) by using ultra-violet (UV) and MS detection was also reported [29]. ...
Article
Full-text available
Lupin alkaloids (LAs) represent a class of toxic secondary metabolites in plants, in particular in Lupinus spp.; they are produced as a defense mechanism due to their strong bitter taste and are very dangerous for human and animals. In this work, a sensitive and reliable high performance liquid chromatography—tandem mass spectrometry (HPLC-MS/MS) analytical method for the identification and quantification of thirteen lupin alkaloids was developed and validated according to FDA guidelines. Efficient extraction and clean-up steps, carried out by solid-phase extraction, were finely tuned on the basis of the characteristics of the analytes and lupin samples, providing good selectivity with minimized matrix interference. The effectiveness of the method was proven by the satisfactory recovery values obtained for most of the analytes and a matrix effect ≤23% for all tested levels. In addition, a sensitive and reliable determination of the target compounds was obtained; LOQs were between 1 and 25 µg Kg−1, i.e., below the requested maximum levels (<200 mg Kg−1). The method was applied to evaluate the LAs profile in different batches of raw L. albus L. samples, varying in size and across farming treatments.
... The method was slightly modified described by [19]. From dry matter (DM) with particles smaller than 0.5 mm; To 2 g of bitter sample or 5 g of debittered sample is homogenized with 25 mL of 0.5 M HCl in an ultrasound for 30 minutes at room temperature. ...
Article
Full-text available
Lupinus mutabilis Sweet grains, commonly known as Tarwi in the Bolivian highlands, are garnering increased attention as an emerging food source due to their elevated protein and fat content. Originally, these grains possess a bitter taste attributed to the presence of quinolizidine-type alkaloids, necessitating a debittering process. In this investigation, we assessed the most abundant alkaloids (sparteine and lupanine) in six Tarwi cultivars both before and after debittering, employing water and a sodium bicarbonate solution. The findings reveal that the debittering process effectively reduces alkaloid concentrations to permissible levels and may even confer health benefits. Additionally, a proximal analysis was carried and swift, straightforward qualitative method was devised to estimate quinolizidine alkaloids through the formation of an orange complex resulting from iodide-alkaloid ion interaction, that could be applied by the indigenous communities that produce this food.
... Lupins are annual plants belonging to the Fabaceae family, genus Lupinus. More than 400 species are known [1], originating from the Mediterranean region, North Africa and North and South America. Four Lupinus species, i.e., white lupin (Lupinus albus L.), yellow lupin (Lupinus luteus L.), blue or narrow leaf lupin (Lupinus angustifolius L.) and Andean lupin (Lupinus mutabilis Sweet.) are widely grown as green manure, forage and food. ...
Article
Full-text available
Four species of lupin (white lupin, yellow lupin, blue lupin and Andean lupin) are widely cropped thanks to the excellent nutritional composition of their seeds: high protein content (28–48 g/100 g); good lipid content (4.6–13.5 g/100 g, but up to 20.0 g/100 g in Andean lupin), especially unsaturated triacylglycerols; and richness in antioxidant compounds like carotenoids, tocols and phenolics. Particularly relevant is the amount of free phenolics, highly bioaccessible in the small intestine. However, the typical bitter and toxic alkaloids must be eliminated before lupin consumption, hindering its diffusion and affecting its nutritional value. This review summarises the results of recent research in lupin composition for the above-mentioned three classes of antioxidant compounds, both in non-debittered and debittered seeds. Additionally, the influence of technological processes to further increase their nutritional value as well as the effects of food manufacturing on antioxidant content were scrutinised. Lupin has been demonstrated to be an outstanding raw material source, superior to most crops and suitable for manufacturing foods with good antioxidant and nutritional properties. The bioaccessibility of lupin antioxidants after digestion of ready-to-eat products still emerges as a dearth in current research.
... 14 Similar to other varieties of L. angustifolius, 13-α-hydroxylupanine and lupanine were the most abundant alkaloids, followed by a lower presence of angustifoline. 14,24,36,37 3.5. Protein Extraction Recovery and Purity. ...
Article
Full-text available
Lupin is a promising protein source with a high protein concentration. Breeding efforts have resulted in the development of varieties low in quinolizidine alkaloids. The objective of this work was to evaluate 22 different blue lupin genotypes for a high protein concentration and low content of antinutritional alkaloids. These genotypes were grown under uniform controlled environmental and soil conditions, and the harvested seeds were evaluated for their composition. The low phosphorus content confirmed that the phytic acid presence was low in lupin, especially compared to other legumes. Furthermore, some of the varieties had less than 200 ppm alkaloids. Lupin proteins were rich in leucine and lysine, with the lowest amino acid concentration being methionine. There were significant differences in the protein concentration and recovery. This work demonstrated that an approach for selection of genotypes should be based on not only agronomic yields but also nutritional phenotypes, driving better decision making on future varietal selection.
... The content of alkaloids is very responsive to the impact of environmental factors, such as droughts, air temperature, geographic location, insolation level, agricultural practices, and the presence of pathogens (Christiansen et al., 1997;Cowling, Tarr, 2004;Ageeva et al., 2020). Moreover, the concentration of alkaloids in seeds of the same genotype under different growing conditions can show at least twofold variation, reaching even a tenfold increase, thus exceeding the required alkaloid content threshold and turning lupine genotypes traditionally classified as sweet into bitter ones (Cowling, Tarr, 2004;Reinhard et al., 2006;Romanchuk, Anokhina, 2018). ...
Article
Full-text available
Alkaloid content was assessed in the seeds of 59 narrow-leafed lupine (Lupinus angustifolius L.) accessions from the VIR collection in the environments of Leningrad Province. The selected set included accessions of different statuses (wild forms, landraces, and advanced cultivars) and different years of introduction to the collection. Alkaloids were analyzed using gas-liquid chromatography coupled with mass spectrometry. Concentrations of main alkaloids: lupanine, 13-hydroxylupanine, sparteine, angustifoline and isolupanine, and their total content were measured. The total alkaloid content variability identified in the seeds of the studied set of accessions was 0.0015 to 2.017 %. In most cases, the value of the character corresponded to the accession’s status: modern improved cultivars, with the exception of green manure ones, entered the group with the range of 0.0015–0.052 %, while landraces and wild forms showed values from 0.057 to 2.17 %. It is meaningful that the second group mainly included accessions that came to the collection before the 1950s, i. e., before the times when low-alkaloid cultivars were intensively developed. Strong variability of the character across the years was observed in the accessions grown under the same soil and climate conditions in both years. In 2019, the average content of alkaloids in the sampled set was 1.9 times higher than in 2020. An analysis of weather conditions suggested that the decrease in alkaloid content occurred due to a significant increase in total rainfall in 2020. Searching for links between the content of alkaloids and the type of pod (spontaneously non-dehiscent, or cultivated, spontaneously dehiscent, or wild, and intermediate) showed a tendency towards higher (approximately twofold in both years of research) total alkaloid content in the accessions with the wild pod type and the nearest intermediate one compared to those with the pod non-dehiscent without threshing. The correlation between the average total alkaloid content and seed color, reduced to three categories (dark, or wild, light, or cultivated, and intermediate), was significantly stronger in the group with dark seeds (5.2 times in 2019, and 3.7 times in 2020). There were no significant differences in the percentage of individual alkaloids within the total amount either between the years of research or among the groups with different pod types or the groups with different seed coat colors.
... Considering compounds 32, 36, 37 and 57, all gave a molecular ion peak in the positive mode ion spectrum followed by a highly abundant fragment ion formed by the loss of an ally group (M-41). This is characteristic for the tricyclic quinolizidine alkaloid known as angustifoline and N-alkyl and N-alkoxy derivatives (Reinhard, Rupp, Sager, Streule, & Zoller, 2006). Compound 57 appeared at m/z 235 (M+ H + ) corresponding to the molecular formula C 14 H 22 N 2 O. ...
Article
The current study attempts to illustrate how the chemical and biological profile of white lupine seeds varies throughout the course of various germination days using UHPLC-QqQ-MS combined to chemometrics. Abscisic acid showed maximum level in the un-germinated seeds and started to decline with seed germination accompanied by an increase in the levels of gibberellins which were undetectable in un-germinated seeds. Coumaronochromones were the most prevalent constituents detected in un-germinated seeds while day 2 sprouts showed significant accumulation of flavones. The levels of alkaloids showed significant increase upon germination of the seeds reaching its maximum in day 14 sprouts. The OPLS model coefficients plot indicated that lupinalbin D and F, apigenin hexoside, kaempferol hexoside, albine, and hydoxylupanine showed strong positive correlation to the alpha amylase inhibitory activity of the tested samples while lupinalbin A, lupinisoflavone, lupinic acid and multiflorine were positively correlated to the inhibition of alpha glycosidase activity. The results obtained indicated that seed germination has a profound effect on the chemical profile as well as the in-vitro antidiabetic activity of lupine seeds.
Article
Foods derived from lupin are attracting consumers and producers for their nutritional value and low cost of production. Lupin products may pose a health risk to consumers due to the presence of toxic alkaloid compounds. To monitor alkaloids in lupin seeds, we developed and validated a rapid LC-MS/MS method for the determination of five quinolizidine alkaloids and one indole alkaloid in lupin seeds. The final method involves extraction by water/acetonitrile, ultrasonication, separation using NaCl and MgSO4, alkalinisation by NaOH and analysis using HPLC-MS/MS. The separation was achieved in 6 min using HPLC directly coupled to MS/MS on a 6500 + QTRAP in ESI-positive mode. The method showed acceptable recovery for all tested compounds in soybean grain at a wide range from 0.55 (LOQ) to 55 mg/kg, with a mean recovery of 93% for the total alkaloid content (SD <2.5%). The method was tested on a commercial lupin flour sample yielding an alkaloid concentration of 130 mg/kg (SD 3.6%). Alkaloid profiles were assessed across different cultivars of five Lupinus species (Lupinus angustifolius, L. cosentinii, L. albus, L. luteus, and L. mutabilis). Using the method presented, toxic alkaloid levels can be monitored in lupin-based products to ensure their safety for human consumption. The method was deployed to measure the alkaloid content across different seeds of narrow-leaf lupin from the same farm, revealing that the alkaloid content varied widely from seed to seed from the same farm.
Article
Lupine (Lupinus albus) belongs to the family of the legumes (Leguminosae) and, therefore, demonstrates strong cross-reactivity with peanuts, chickpeas and white beans. Due to the high protein and lipid content, lupine flour is extensively used in food industry. It is added to cereal flours and used as emulsifier. Concomitant with the increasing intake of lupine, a rising prevalence of allergic reactions after ingestion as well as inhalation of lupine flour has been observed. We here present a patient with known severe peanut allergy who reacted after eating a pizza with severe respiratory symptoms, which required an intubation in the emergency room. Our investigations revealed that the pizza dough used in this "take-away" pizza was enriched with lupine flour. The sensitization to lupine could be proven by means of the skin prick test (SPT) with native lupine flour and also serologically as 44.30 kU/l (class 4) of specific IgE antibodies were measured by means of CAP-FEIA to lupine seed (f335, Pharmacia Diagnostics). A peanut contamination of the pizza was excluded by the manufacturer as well as by analysis in an immunoassay. IgE binding to major proteins at 20, 34, 60 and 62 kDa was detected by means of immunoblotting. In inhibition experiments, IgE-binding to lupine proteins was partially inhibited by peanut extract. Shortly after that first observation of a lupine allergy, a patient was admitted to our clinic who experienced allergic reactions including edema in the face and abdominal pain after ingestion of gingerbread that was declared to contain lupine flour. Apart from a slight pollinosis in childhood, the patient had no history of allergic reactions and no history of food allergy at all. She revealed a positive reaction to lupine flour in skin prick test and was positive for specific IgE against lupine flour in CAP-FEIA (class 3; 7.59 kU/l). Contrary to the first case, the IgE binding to lupine flour was not inhibited by peanut extract. On the basis of our observations in the two patients and further cases described in the literature, we conclude that allergic reactions to lupine can be mediated by cross-reacting IgE antibodies to peanut, chickpea and white beans or may occur due to an isolated sensitization as well. Because of the increasing use of lupine flour in the food industry and its high allergenic properties, lupine flour should be regarded as a highly allergenic food. Similar to peanut and soy, lupine should be added to the list of allergenic foods of the EU and Swiss Food Regulation that have to be mandatory declared.
Article
A crystallographic study of phompsin A, the hexapeptide mycotoxin of Phomopsis leptostromiformis responsible for lupinosis disease in animals, has shown that it is a linear peptide, modified by an ether bridge in place of the hydroxy groups of the N-methyl-3-(3-chloro-4,5-dihydroxyphenyl)-3-hydroxyalanine and 3-hydroxyisoleucine units, and thus containing a 13-membered ring.
Article
An assay using high performance liquid chromatography to measure phomopsin A, the principal mycotoxin responsible for lupinosis is described. Samples of lupin stubble are extracted with methanol: water and purified by partitioning between n-butanol and water, chromatography on Amberlite XAD-2 and by cation exchange chromatography. The analysis is performed on a reverse phase C18 column using a methanol:water gradient and UV detection. The limit of detection for this procedure is 0.5 mg phomopsin A per kg stubble. Improvements to the extraction and purification procedures were made and total analysis time was reduced to 2 days. The mean (± s.d.) recovery for the purification procedure was 64.3 ± 4.5% over a wide range of concentrations.
Article
Forty-nine genotypes of Lupinus albus seeds from different countries have been analyzed for their alkaloid content by thin-layer chromatography and gas chromatography-mass spectrometry. Twenty samples were sweet, while 29 were bitter, and the taste was positively correlated with the alkaloid content. The composition of the-alkaloids showed that lupanine was the main alkaloid present, and variable amounts of albine, alpha-isolupanine, multiflorine, and 13-hydroxylupanine were also detected in the seeds.
Article
Collsionally activated dissociation efficiency and fragment ion mass resolution have been characterized on a triple quadrupole mass spectrometer equipped with an enclosed collision cell which is operated at relatively high pressure, The higher pressure results in improved efficiency (and therefore improved ion transmission) and better mass resolution than has been obtained previously with collision cells operated at lower pressure, Efficiencies of 30-50%, measured as the fractions of original precursor ion in Q1 which are detected as resolved fragment ions, are typical for singly charged ions, compared with 5-10% achieved previously, Mass resolution of fragment ions is improved to the point that the isotopes from quadruply charged fragment ions can be resolved.
Article
An alkaloid-poor line of Washington lupin (Lupinus polyphyllus Lindl var SF/TA) was developed in an experiment started in 1982. The nutritive quality (alkaloid content, protein and amino acids, fat and fatty acids. macro- and micronutrients, fibre, sugars) yields, and seed quality of this line were studied. The results show that the total alkaloid content was low and varied in different seeds from 226 μg g−1 to 366 μg g−1 of dry matter. The main alkaloid was lupanine, but 16 other alkaloids (including sparteine and gramine) were also present. The var SF/TA cannot yet be used for human nutrition without processing although it would be a valuable protein crop. The results confirm that seeds which look different also vary in chemical composition.
Article
For the characterization of organic substances in gas chromatography a number termed «retention index» is proposed. A simple relation exists between the retention index of a compound on a non-polar stationary phase and its boiling point.