Studies On The Formation of Methylglyoxal From Dihydroxyacetone in Manuka (Leptospermum Scoparium) Honey - SPECTROPHOTOMETER

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 5

Carbohydrate Research 361 (2012) 7–11

Contents lists available at SciVerse ScienceDirect

Carbohydrate Research
journal homepage: www.elsevier.com/locate/carres

Studies on the formation of methylglyoxal from dihydroxyacetone in Manuka


(Leptospermum scoparium) honey
Julia Atrott, Steffi Haberlau, Thomas Henle ⇑
Institute of Food Chemistry, Technische Universität Dresden, D-01062 Dresden, Germany

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

Article history: Dihydroxyacetone (DHA) and methylglyoxal (MGO) are unique carbohydrate metabolites of manuka
Received 22 May 2012 honey. A method for the reliable quantification of DHA in honey samples was established, based on deriv-
Received in revised form 17 July 2012 atization with o-phenylenediamine (OPD) and subsequent RP-HPLC with UV detection. The previously
Accepted 31 July 2012
unknown reaction product of DHA and OPD was identified as 2-hydroxymethylquinoxaline by spectro-
Available online 8 August 2012
scopic means. DHA was exclusively determined in 6 fresh manuka honeys originating directly from
the beehive as well as 18 commercial manuka honey samples, ranging from 600 to 2700 mg/kg and
Keywords:
130 to 1600 mg/kg, respectively. The corresponding MGO contents varied from 50 to 250 mg/kg in fresh
Manuka honey
Carbohydrate metabolism
and 70 to 700 mg/kg in commercial manuka honey samples. A good linear correlation between DHA and
Dihydroxyacetone MGO values in commercial manuka honeys was observed, resulting in a mean ratio of DHA to MGO of 2:1.
Methylglyoxal In contrast to this, the DHA-to-MGO relation was much higher in fresh manuka honeys but approximated
Quinoxaline to a ratio of 2:1 while honey ripening. Heating experiments revealed that MGO formation based on ther-
Glycation mal treatment as a consequence, for example, of caramelization in honey does not occur. DHA and MGO
can serve as suitable unique quality parameter for manuka honey.
Ó 2012 Elsevier Ltd. All rights reserved.

1. Introduction course of caramelization. In a recent report, Adams et al.11 identi-


fied dihydroxyacetone (DHA) as direct precursor for MGO forma-
The presence of 1,2-dicarbonyl compounds such as 3-deoxyglu- tion in manuka honeys. In this study, the authors had
cosone (3-DG), glyoxal (GO), and methylglyoxal (MGO) in honey determined DHA in nectar and fresh honey derived from manuka
was first described in 2004 by Weigel et al.1 as a consequence of plants and proved that DHA is non-enzymatically transferred to
sugar degradation or caramelization. Over a wide range of different MGO during honey storage. However, the origin of DHA stills re-
honey types, 3-DG, which is the precursor for 5-hydroxymethyl- mains enigmatic.
furfural (HMF), could be determined in high amounts, varying be- DHA is a well-known degradation product of carbohydrates and
tween 80 and 1270 mg/kg and correlating with heating or storage was detected in caramelized mixtures next to other compounds as
conditions.1–3 Contents of GO and MGO generally were very low a result of the Maillard reaction.12 DHA was also found in naturally
(up to 5 mg/kg). However, surprisingly high amounts of MGO (up aged wine samples.13 Moreover, its presence was described in rela-
to 760 mg/kg) were found exclusively in manuka honey.2 This hon- tion to fermentation processes of selected microorganisms.14 Fur-
ey is derived from the flowers of the manuka tree (Leptospermum thermore, DHA is discussed as food additive, for example to
scoparium) in New Zealand. It was clearly demonstrated that the enhance browning in baked foods,15 or for food preservation.16 In
pronounced antibacterial activity of New Zealand manuka honey certain cosmetic products, namely self-tanning creams, DHA is
directly originates from MGO.2 The amount of MGO correlates with the functional ingredient.17
the antibacterial activity in manuka honey,4,5 and, therefore, can be Knowledge about the amount of DHA and MGO in honey is of
used as a tool for labelling the bioactivity of manuka honey. prime importance for manufacturers as well as for labelling pur-
Manuka honey was shown to inhibit a wide range of microor- pose. In the present study, a new method for the reliable quantifi-
ganisms, including multiresistant strains.6–8 Furthermore, wound cation of DHA via RP-HPLC using pre-column derivatization with o-
dressings containing manuka honey seem to be useful supports phenylenediamine (OPD) is reported. The previously unknown
in clinical applications for wound healing.9,10 derivative resulting from DHA and OPD is identified by spectro-
The origin and in particular the high concentration of MGO in scopic means. DHA and MGO was measured for a number of com-
manuka honey cannot be explained by sugar degradation in the mercially available and freshly produced manuka honeys in order
to obtain information about the ratio between DHA and MGO
⇑ Corresponding author. Fax: +49 351 463 34138. and the transformation of DHA to MGO during storage. Further-
E-mail address: Thomas.Henle@chemie.tu-dresden.de (T. Henle). more, the influence of a thermal treatment on MGO content in

0008-6215/$ - see front matter Ó 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.carres.2012.07.025
8 J. Atrott et al. / Carbohydrate Research 361 (2012) 7–11

manuka honey was studied in order to clarify whether the bioac- ketoamine (V) according to the conversion of Amadori products.
tive compound can be produced artificially in the final product. A second dehydration results in formation of an intermediate
(VI), which is converted into the quinoxaline (VII), putatively by
2. Results and discussion oxidation. Compared to dicarbonyl compounds like MGO or 3-
DG, complete derivatization of DHA with OPD takes more time
2.1. RP-HPLC analysis of DHA and characterization of DHA for development of the quinoxaline. Therefore, it can be supposed
quinoxaline that the formation of the keto group in (V) via enaminol (IV) is the
rate determining step of the reaction. The oxidation of intermedi-
The analysis of 1,2-dicarbonyl compounds such as 3-desoxy- ate (VI) to the quinoxaline cannot be proven yet, but it may due
glucosone (3-DG) or methylglyoxal (MGO) in honey is generally to the oxidative properties of the reaction mixture, for example
achieved by RP-HPLC of the corresponding quinoxalines resulting the influence of oxygen.
from derivatization with o-phenylenediamine (OPD). Under condi- Kinetic studies with varying incubation time, temperature, pH
tions applied for optimal derivatization of these compounds (phos- value, and molar ratio of DHA to OPD were performed in order to
phate buffer pH 6.5, incubation at room temperature), optimize derivatization of DHA to compound (VII) (data not
dihydroxyacetone (DHA) also formed an OPD-derivative eluting shown). Finally, it could be shown that best results are obtained
in the chromatogram at about 13 min (Fig. 1A). This reaction prod- after incubation for 16 h at 37 °C using an acetate buffer with a
uct was isolated using semi preparative RP-HPLC. The molar mass pH value of 4.0 for sample preparation. Due to these differences
of the isolate was determined with 160 g/mol by LC-ESI-TOF-MS. in methodical conditions, reliable simultaneous quantification of
The UV maximum at 319 nm was observed, varying only slightly DHA together with MGO was not possible. At pH value of 4.0,
when compared to the absorption maximum of 312 nm generally acid-induced formation of the MGO–quinoxaline was observed in
reported for quinoxalines. Based on the NMR data, the structure carbohydrate solutions, and hence, MGO values would be overesti-
was identified as 2-hydroxymethylquinoxaline (2-quinoxaline- mated. Otherwise, reaction of DHA and OPD is too slow at neutral
methanol, compound VII in Fig. 2). pH, and therefore, DHA does not react completely with OPD under
2-Hydroxymethylquinoxaline has already been described as these conditions. Consequently, to reach optimal quinoxaline for-
reaction product formed during incubation of sugars like glucose mation and best sensitivity for both analytes, two independent
or ribose, respectively, with OPD.18,19 The identified quinoxalines sample preparations and two chromatographic runs per sample
derived from this reaction mixture were explained by the buffer- are necessary.
induced degradation of hexoses to 1,2-dicarbonyl fragments.19 Performing DHA determination under these conditions, it be-
However, the authors did not propose DHA being involved in quin- came clear that notable amounts of DHA are also formed from
oxaline formation. Up to now, it was generally accepted that DHA the honey matrix itself, namely from glucose and fructose break-
does not react with OPD to a stable derivative, and, therefore, can- down, during derivatization (Fig. 1B and C). This reproducible basis
not be quantified by means of RP-HPLC following derivatization level of DHA has to be taken into account by using an artificial hon-
with OPD.20 ey matrix as blank value.21 Calibration of (VII) using this sugar ma-
A possible reaction mechanism for the formation of 2-hydrox- trix as 10% solution in acetate buffer showed a good linearity
ymethylquinoxaline from DHA and OPD is shown in Figure 2. First, (regression equation y = 0.15x + 5.66, R2 = 0.9944) in the range
a nucleophilic addition of one amino group of OPD (I) to the car- from 0.18 to 2.20 mmol/L DHA in sample solution, equivalent to
bonyl carbon of DHA (II) occurs. The iminol (III) is formed by dehy- DHA concentrations between 165 and 1980 mg/kg in honey. Quan-
dration and is able to convert via the enaminol (IV) into the tification of DHA in manuka honey (Fig. 1D) was finally performed

Figure 1. RP-HPLC after derivatization with OPD of (A) DHA standard solution in water, (B) solution of fructose and glucose simulating honey matrix, (C) rape honey, (D)
manuka honey.
J. Atrott et al. / Carbohydrate Research 361 (2012) 7–11 9

Figure 2. Proposed reaction mechanism for the reaction of DHA (II) with OPD (I) to 2-hydroxymethylquinoxaline (VII).

by matrix calibration using the above mentioned artificial honey


matrix. LOD (limit of detection) and LOQ (limit of quantification)
of DHA in test solutions were estimated with 0.01 and
0.12 mmol/L, respectively, representing a LOD of 10 mg/kg and
LOQ of 110 mg/kg in honey while using 10% sample solution.
Recovery of DHA added to the artificial honey matrix was deter-
mined with a mean of 95.3%. The relative repeatability for DHA
quantification was assessed by repetitive analysis of one manuka
honey sample and was at 7.8%.

2.2. Quantification of DHA and MGO in manuka honey

Within this work, 18 commercial as well as 6 fresh manuka


honey samples obtained directly from the beehive without further
processing were analysed for their contents of DHA and MGO. The
results are shown in Figure 3. In the fresh manuka honeys, DHA
varied from 611 to 2724 mg/kg (median 1622 mg/kg). These fresh
honey samples contained MGO in concentrations from 50 to
260 mg/kg (median 129 mg/kg), whereas in commercial manuka
honeys, MGO contents were ranging from 66 to 684 mg/kg with
a median value of 356 mg/kg, confirming the data in the litera-
ture.2,5 DHA in commercial samples varied from 127 to 1563 mg/
kg (median 735 mg/kg).
A good linear correlation (regression equation y = 0.47x + 11.57,
Figure 4. Relation of DHA to MGO in 18 commercial manuka honeys (dark squares),
R2 = 0.8977) between the corresponding concentration of DHA and compared with 6 fresh manuka honeys before (opened squares) and after storage at
MGO in commercial manuka honeys was observed (Fig. 4, black 37 °C for 6 weeks (grey triangles).
squares), resulting in a mean ratio of DHA to MGO of 2:1. Com-
pared to this, concentrations of DHA exceeded the values of MGO
During storage of these fresh samples at 37 °C for 6 weeks, the
in fresh manuka honeys seven to sixteenfold (Fig. 4, open squares).
amount of DHA decreased considerably due to formation of MGO
(Fig. 4, grey triangles). For stored fresh manuka honeys, a ratio of
DHA to MGO ranging from 5 to 9 was calculated. A longer honey
storage would be followed by a further increase of MGO content till
there is reached a plateau which can be also observed for data re-
ported by Adams et al.11 This conversion can be additionally char-
acterized by the ratio of DHA to MGO which is approximating
factor 2 in the course of honey ripening. Consequently, the discrim-
ination between fresh and ‘ripened’ manuka honeys on the basis of
defined DHA-to-MGO ratios seems to be a suitable tool for classifi-
cation of manuka honeys and for monitoring reactions occurring
during ripening. The simultaneous presence of DHA and MGO in
certain relation to each other is a unique quality parameter for
manuka honey.
The findings of this study are in agreement with the results of
Adams et al.11 These authors identified DHA as precursor for
MGO formation by illustrating curves of DHA decrease and MGO
increase while honey storage. Authors observed a similar relation
of DHA to MGO after honey incubation. Adams et al.,11 however,
had determined higher values of DHA and MGO in fresh manuka
honeys, which can be due to the quantification method used by
these authors.
Figure 3. Concentrations of DHA and MGO in 6 fresh and 18 commercial manuka Other floral or honeydew honeys which were analysed in this
honeys (box plot, Whiskers indicate min/max values,  = median value). study did not contain any DHA above LOD (Fig. 1C). The peak
10 J. Atrott et al. / Carbohydrate Research 361 (2012) 7–11

detectable for DHA in Figure 1C is due to the small amount of DHA not possible. In general, treatment of honey with high tempera-
formed from the honey matrix during derivatization (see above). tures is followed by massive changes in sensory properties like col-
Confirming the data of Weigel et al.,1 only small amounts of our, smell, and taste, respectively. Moreover, a remarkable increase
MGO ranging from 2 to 28 mg/kg were detected in these honeys. of 3-DG, resulting from degradation of glucose and fructose, and, in
In contrast, other honeys from Leptospermum species sources like a consequence, of 5-hydroxymethylfurfural (HMF) will occur due
jellybush honey from Australia (Leptospermum polygalifolium) can to the heating processes, making such honeys inacceptable.
also contain MGO and DHA in high amounts.20 Two samples of this
honey type analysed in this study revealed an analogous DHA-to- 2.4. Conclusion
MGO ratio of 2:1. By evaluation of the data of Windsor et al.20 in
the same way, it can be suggested that a similar conversion oc- Reliable analysis of DHA and MGO in honey was achieved by
curred in these honeys which can be illustrated by calculated pre-column derivatization with OPD, followed by RP-HPLC with
DHA-to-MGO relation ranging from 18.5 till 0.7. However, these UV-detection. DHA is transformed to MGO during ripening of man-
authors analysed partially very high values of MGO up to uka honey, resulting in a constant DHA-to-MGO ratio of 2:1, which
1723 mg/kg by doing a simultaneous determination of DHA and can be used as a quality index to monitor changes during storage
MGO, raising doubts concerning a possible overestimation of and to classify commercial products. DHA and MGO are unique
MGO as a result of ‘neoformation’ during sample preparation, and naturally occurring constituents of manuka honey.
which cannot be dispelled, as several methodical details are miss-
ing in this paper.20 3. Materials and methods
The phenomenon that DHA is not completely transferred to
MGO during honey storage and that a constant ratio between 3.1. Chemicals
DHA and MGO is obtained, might be due to the water content of
honey being the limiting factor. MGO formation occurs by dehy- Methylglyoxal (MGO, 40% in water) and dihydroxyacetone
dration of DHA. A back reaction presumably does not occur since (DHA) were obtained from Sigma–Aldrich (Steinheim, Germany)
this transformation of DHA to MGO was described to be irrevers- and the derivatizing agent o-phenylenediamine (OPD) from Fluka
ible.22 The water content of 10 commercial manuka honeys was (Munich, Germany). Sodium acetate, sodium dihydrogenphosphate
determined varied between 16.0% and 19.3% with a median value and disodium hydrogenphosphate were purchased from Grüssing
of 18.1%. This is in agreement with legislative regulations, which (Filsum, Germany). Methanol (HPLC grade) and acetic acid were or-
give very narrow limits for an acceptable water content in honey dered from VWR Prolabo (Leuven, Belgium). Water used for buffer
(maximum 20%).23 preparations and HPLC solvents was obtained using a Purelab plus
purification system (USFilter, Ransbach-Baumbach, Germany).
2.3. Influence of heat treatment on MGO levels in manuka
honey 3.2. Honey samples

Honey is not allowed to be heated.23 However, speculations Manuka honeys were kindly provided by Manuka Health Ltd, Te
concerning a fraudulent heat-treatment to increase the MGO con- Awamutu, New Zealand. In total, 18 commercial samples with
tent in order to obtain manuka honey with ‘optimized’ bioactivity, varying methylglyoxal content and 6 samples of fresh manuka
are conceivable. MGO may also increase while heating due to a po- honey from different regions originating directly from the beehives
tential release of ‘sugar-bound’ MGO.24 without any further treatment were obtained. Other honey sam-
To study the effect of thermal treatment on MGO content, two ples (rape, acacia, sunflower, bush flower, heather, chestnut, leath-
commercial samples of manuka honey with different MGO con- erwood, thyme, pine, and forest, in total 17 samples) were
tents and one sample of a rape honey were incubated at 70 °C for obtained from local supermarkets, or from Neuseelandhaus, Bergk-
10 min (resembling conditions of pasteurization) and up to 24 h amen, Germany (2 samples of jellybush honey), respectively.
for long-term heating. The results of these analyses, given in Ta- Water content of honey samples was measured refractrometrically
ble 1, show unambiguously that heat treatment of commercial according to Weigel et al.1 Additionally, an artificial honey matrix
rape honey did not lead to MGO formation, whereas the concentra- consisting of 46.5% fructose, 34.0% glucose, 1.5% sucrose, and 18.0%
tion of MGO in manuka honey remained constant while short heat- water was used.21
ing and decreased markedly during long-term incubation at 70 °C.
Moreover, the long-term storage of honey for 12 weeks at 37 °C re- 3.3. Honey storage and heat treatment
sulted in a decrease of MGO in manuka honey, which was more
pronounced in the sample which had been pre-heated for 10 min To simulate a long-term storage, the 6 fresh manuka honeys
at 70 °C. These results clearly show that a thermal generation of were incubated at 37 °C for 6 weeks. To investigate the effect of
MGO as a consequence of caramelization of sugars in honey is thermal treatment, one rape honey and two manuka honeys were
heated at 70 °C for 10 min as well as for 8 h and 24 h. Furthermore,
samples heated for 10 min at 70 °C were incubated afterwards at
37 °C for 12 weeks to assess the combination of heating and stor-
Table 1 age of honey compared to a non-heated sample.
Concentration of MGO after thermal treatment of manuka and rape honeys at 70 °C,
partially followed by storage at 37 °C for 12 weeks
3.4. Sample preparation
Incubation conditions MGO concentration (mg/kg)
70 °C 37 °C Manuka 1 Manuka 2 Rape The determination of MGO was performed according to Mavric
et al.2 with slight modifications. For this, 1 mL aliquots of 10% (w/v)
0 min — 100.7 307.2 2.3
0 min +12 weeks 90.6 268.1 5.3 honey solutions in 0.5 M phosphate buffer (pH 6.5) were mixed
10 min — 107.9 329.1 3.1 with 300 lL phosphate buffer (pH 6.5) and 300 lL OPD solution
10 min +12 weeks 67.4 241.8 5.5 (1% in phosphate buffer pH 6.5). Samples were incubated at room
8h — 68.7 237.8 3.8
temperature overnight and membrane filtered (0.45 lm). For anal-
24 h — 35.9 132.8 3.9
ysis of DHA, 1 mL aliquots of 10% honey solutions in 0.5 M acetate
J. Atrott et al. / Carbohydrate Research 361 (2012) 7–11 11

buffer (pH 4.0) were mixed with 300 lL acetate buffer (pH 4.0) and ESI-TOF mass spectrometer (PerSeptive Biosystem, Framingham,
300 lL OPD solution (1% in acetate buffer pH 4.0), followed by USA). The chromatographic conditions were the same as described
incubation for 16 h at 37 °C and membrane filtration (0.45 lm). for analytical HPLC, 50 lL sample solution was injected. Electro-
spray ionization was used in the positive ionization mode. Full scan
3.5. Analytical RP-HPLC mass spectra were measured in mass range 100–1000 m/z in the
tic-mode. For data acquisition the software Data Explorer Version
HPLC analyses were performed using an Äkta basic system from 4.0.0.1 (Applied Biosystems, Foster City, USA) was used.
1
Amersham Pharmacia Biotech (Uppsala, Sweden) with a pump P- H NMR spectrum was recorded on a Bruker DRX 500 instru-
900 and an online degasser K-5004 (Knauer, Berlin, Germany) as ment (Rheinstetten, Germany) at 500 MHz. For this, 5 mg of the
well as an UV detector UV-900 and an auto sampler A-900. Peak purified and lyophilized isolate was dissolved in 750 lL deuterium
evaluation was managed using the software UNICORN 4.11. The oxide. All chemical shifts are given in parts per million (ppm) rel-
separation of quinoxalines was realized on a stainless steel column ative to the internal HOD signal (4.70 ppm).
filled with Eurospher 100 RP18 material (250 mm  4.6 mm, 5 lm Data of the isolated quinoxaline of DHA were as follows. 1H
particle size, integrated pre column; Knauer, Berlin, Germany). The NMR (500 MHz, D2O), d [ppm]: 8.74 (s, H-3), 7.81–7.87 (m, H-5,
mobile phase were 0.075% acetic acid in water (solvent A) and a H-8), 7.68–7.73 (m, H-6, H-7), 4.82 (s, H-10 A, H-10 B). Mass spectros-
mixture of 80% methanol and 20% solvent A (solvent B). The gradi- copy gave a m/z of 161 for [MH]+. UV spectroscopy gave a kmax of
ent started with 40% solvent B for 1 min and then was elevated lin- 319 nm.
early to 100% B over a period of 20 min, was changed back to 40% B
in 4 min and was held there for 7 min. The flow rate was 0.8 mL/ Acknowledgements
min, the separation was done at 30 °C, 20 lL sample solution
was injected and peaks were detected by measurement of UV The authors thank Dr. Uwe Schwarzenbolz, Institute of Food
absorbance at 312 nm. Quantification was achieved by external Chemistry, for his help during the LC-ESI-TOF-MS measurements,
calibration with standard solution for MGO or by matrix calibra- and furthermore, the members of the Institute of Organic Chemis-
tion using an artificial honey matrix according to Wahdan21 for try, namely Dr. Margit Gruner and Anett Rudolph, for recording the
DHA, respectively. NMR data.
The limits of detection (LOD) and quantification (LOQ) of DHA
were calculated from blank values using artificial honey matrix References
(measurement of 10 independent blank values on one day).25 For
determination of repeatability, 10 samples of a manuka honey 1. Weigel, K. U.; Opitz, T.; Henle, T. Eur. Food Res. Technol. 2004, 218, 147–151.
2. Mavric, E.; Wittmann, S.; Barth, G.; Henle, T. Mol. Nutr. Food Res. 2008, 52, 483–
were derivatized on different days. The recovery of DHA was esti- 489.
mated by spiking 10% solutions of artificial honey with varying 3. Marceau, E.; Yaylayan, V. A. J. Agric. Food Chem. 2009, 57, 10837–10844.
concentrations of DHA and analysing as described above. 4. Adams, C. J.; Boult, C. H.; Deadman, B. J.; Farr, J. M.; Grainger, M. N.; Manley-
Harris, M.; Snow, M. J. Carbohydr. Res. 2008, 343, 651–659.
5. Atrott, J.; Henle, T. Czech. J. Food Sci. 2009, 27, 163–165.
3.6. Isolation and identification of quinoxaline resulting from 6. Cooper, R. A.; Molan, P. C.; Harding, K. G. J. Appl. Microbial. 2002, 93(5), 857–
DHA and OPD 863.
7. Molan, P. Honey: Antimicrobial Actions and Role in Disease Management. In
New Strategies combating bacterial infection, Ahmad, I., Aqil, F., Eds.; Wiley-VCH:
Semipreparative HPLC was performed with a system consisting Weinheim, Germany, 2008; pp 229–253.
of two pumps K-1001 with mixing chamber, automatically driven 8. George, N. M.; Cutting, K. F. Wounds 2010, 19, 231–236.
9. Molan, P. C. Int. J. Low. Extrem. Wounds 2006, 5, 40–54.
valves for injection and fractionation, online degasser and UV
10. Simon, A.; Traynor, K.; Santos, K.; Blaser, G.; Bode, U.; Molan, P. eCAM 2009, 6,
detector K-2501 (all from Knauer, Berlin, Germany). Data handling 165–173.
was managed using the software Eurochrom 2000. The separation 11. Adams, C. J.; Manley-Harris, M.; Molan, P. C. Carbohydr. Res. 2009, 344, 1050–
1053.
was realized on a stainless steel column filled with Eurospher 100-
12. Severin, T.; Hiebl, J.; Popp-Ginsbach, H. Z. Lebensm. Unters. Forsch. 1984, 178,
8 C18 (33 mm  8 mm with integrated pre column; Knauer, Berlin, 284–287.
Germany). A DHA standard solution in water (2500 mg/L) after 13. Laurie, V. F.; Waterhouse, A. L. J. Agric. Food Chem. 2006, 54, 4668–4673.
derivatization with OPD (1% in 0.5 M acetate buffer pH 4) for 14. Green, S. R. U.S. 2948658, 1960 Aug 09.
15. Kirk, D. L. U.S. 3220850, 1965 Nov 30.
16 h at 37 °C was used for fractionation. The derivative of DHA 16. Oborsh, E. V.; Barkate, J. A.; Wesu, C.; Owen, T. M. CA 1054434 A1, 1979 May
was separated with same solvents as mentioned above for analyt- 15.
ical separations. The gradient started with 40% B for 1.2 min, ele- 17. Monfrecola, G.; Prizio, E. Compr. Ser. Photosci. 2001, 3, 489–493.
18. Takagi, M.; Mizutani, M.; Tsuchiva, K. Bull. Univ. Osaka Prefecture, Ser. B: Agric.
vated on 86% B in 14.2 min, on 95% in further 3 min, and on 100% Life Sci. 1972, 24, 43–48.
in 1.8 min, then gradient changed back on 40% B in 3 min and 19. Rizzi, G. P. J. Agric. Food Chem. 2004, 52, 953–957.
was held there for further 7.8 min. The flow rate was set to 20. Windsor, S.; Pappalardo, M.; Brooks, P.; Williams, S.; Manley-Harris, M. J.
Pharmacog. Phytother. 2012, 4, 6–11.
1.6 mL/min. Injection volume was 2 mL. The eluate collected from 21. Wahdan, H. A. L. Infection 1998, 26, 30–35.
16.4 to 19.4 min was evaporated to dryness, taken up in 2 mL 22. Fedoronko, M.; Koenigstein, J. Collect. Czech. Chem. Commun. 1969, 34, 3881–
water and lyophilized. 3894.
23. Codex Alimentarius Commission: Revised Codex Standard for honey, Codex
Absorption spectrum of the isolate in water was recorded by a
STAN 12-1981, Rev.1 (1987), Rev.2 (2001).
Specord S100 diode array spectrophotometer (Carl Zeiss, Jena, Ger- 24. Stephens, J. C.; Schlothauer, R. C. WO 2010082845 A1, 2010 Jul 22.
many). LC–MS measurement was realized by using a LC system 25. Mocak, J.; Bond, A. M.; Mitchell, S.; Scollary, G. Pure Appl. Chem. 1997, 69, 297–
328.
1100 Series (Agilent Technologies, PaloAlto, USA) with Mariner

You might also like