Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

The Development of a Specific and Sensitive LC-MS-Based Method for the Detection and Quantification of Hydroperoxy- and Hydroxydocosahexaenoic Acids as a Tool for Lipidomic Analysis

  • Priscilla B. M. C. Derogis,

    Affiliation Departamento de Bioquímica, Instituto de Química, Universidade de São Paulo, São Paulo, SP, Brazil

  • Florêncio P. Freitas,

    Affiliation Departamento de Bioquímica, Instituto de Química, Universidade de São Paulo, São Paulo, SP, Brazil

  • Anna S. F. Marques,

    Affiliation Luiz Barssotti Application Laboratory, Waters Technologies from Brazil, São Paulo, SP, Brazil

  • Daniela Cunha,

    Affiliation Departamento de Bioquímica, Instituto de Química, Universidade de São Paulo, São Paulo, SP, Brazil

  • Patricia P. Appolinário,

    Affiliation Departamento de Bioquímica, Instituto de Química, Universidade de São Paulo, São Paulo, SP, Brazil

  • Fernando de Paula,

    Affiliation Luiz Barssotti Application Laboratory, Waters Technologies from Brazil, São Paulo, SP, Brazil

  • Tiago C. Lourenço,

    Affiliation Luiz Barssotti Application Laboratory, Waters Technologies from Brazil, São Paulo, SP, Brazil

  • Michael Murgu,

    Affiliation Luiz Barssotti Application Laboratory, Waters Technologies from Brazil, São Paulo, SP, Brazil

  • Paolo Di Mascio,

    Affiliation Departamento de Bioquímica, Instituto de Química, Universidade de São Paulo, São Paulo, SP, Brazil

  • Marisa H. G. Medeiros,

    Affiliation Departamento de Bioquímica, Instituto de Química, Universidade de São Paulo, São Paulo, SP, Brazil

  • Sayuri Miyamoto

    miyamoto@iq.usp.br

    Affiliation Departamento de Bioquímica, Instituto de Química, Universidade de São Paulo, São Paulo, SP, Brazil

Abstract

Docosahexaenoic acid (DHA) is an n-3 polyunsaturated fatty acid that is highly enriched in the brain, and the oxidation products of DHA are present or increased during neurodegenerative disease progression. The characterization of the oxidation products of DHA is critical to understanding the roles that these products play in the development of such diseases. In this study, we developed a sensitive and specific analytical tool for the detection and quantification of twelve major DHA hydroperoxide (HpDoHE) and hydroxide (HDoHE) isomers (isomers at positions 4, 5, 7, 8, 10, 11, 13, 14, 16, 17, 19 and 20) in biological systems. In this study, HpDoHE were synthesized by photooxidation, and the corresponding hydroxides were obtained by reduction with NaBH4. The isolated isomers were characterized by LC-MS/MS, and unique and specific fragment ions were chosen to construct a selected reaction monitoring (SRM) method for the targeted quantitative analysis of each HpDoHE and HDoHE isomer. The detection limits for the LC-MS/MS-SRM assay were 1−670 pg for HpDoHE and 0.5−8.5 pg for HDoHE injected onto a column. Using this method, it was possible to detect the basal levels of HDoHE isomers in both rat plasma and brain samples. Therefore, the developed LC-MS/MS-SRM can be used as an important tool to identify and quantify the hydro(pero)xy derivatives of DHA in biological system and may be helpful for the oxidative lipidomic studies.

Introduction

Docosahexaenoic acid [DHA, 22:6 n-3] is an n-3 fatty acid that is highly enriched in the brain. DHA is particularly enriched in synaptosomal membranes and synaptic vesicles, and this enrichment suggests a role for this fatty acid in the central nervous system [1]. DHA is critical for normal brain function, and changes in the quantity and/or oxidation of DHA are associated with neurodegenerative diseases [2-5].

DHA has six double bonds and is therefore highly susceptible to oxidation. In biological systems, DHA can be oxidized in two ways: (i) enzymatically via cyclooxygenases (COX) [6,7], lipoxygenases (LOX)[8] or cytochrome P450 (CYP450) [9]; and (ii) non-enzymatically, through reactions with reactive oxygen species (ROS)[10] and transition metal ions [11]. Both oxidation mechanisms produce a large variety of oxidative metabolites, which include a series of hydroperoxide and hydroxide positional isomers (see Figure 1).

thumbnail
Figure 1. Generation of hydroperoxy, hydroxy and epoxy derivatives of DHA by enzymatic and non-enzymatic mechanisms.

*19- and 5-HpDoHE and HDoHE are specifically formed by singlet oxygen oxidation.

https://doi.org/10.1371/journal.pone.0077561.g001

Enzymatically, COX-2 converts DHA to 13-hydroxy-DHA (13-HDoHE) [7]. In the presence of aspirin (acetylsalicylic acid, ASA), acetylated COX-2 produces a 17(R)-hydroperoxy-DHA (17(R)-HpDoHE) intermediate, which is reduced to 17(R)-HDoHE and then converted to one of the aspirin-triggered D series resolvins (At-RvD) [7,12]. DHA is also a substrate for LOX enzymes. The 15-LOX converts DHA to 17(S)-HpDoHE and 17S-HDoHE, which can be subsequently converted to D series resolvins (RvD) and protectin/neuroprotectin D1 (PD1) [13]. The 12- and 12/15-LOX convert DHA to 14(S)-H(p)DoHE, which undergoes 13(14)-epoxidation to form maresin 1 (MaR1) [14]. Alternatively, 12-LOX can also generate 11(S)-HDoHE [15] and 5-LOX converts DHA into 2 isomers: 4S-HDoHE, and 7(S)-HDoHE; the former is the major isomer [6,16]. The cytochrome P450 (CYP450) enzymes also catalyze the production of two HDoHE (21- and 22-HDoHE) and several epoxides [17].

DHA is also oxidized non-enzymatically by ROS. The oxidation mechanism can involve radical or non-radical species, such as singlet molecular oxygen. This excited species can be produced under inflammatory conditions by reactions that involve biological peroxides [18-22]. Both mechanisms lead to the production of HpDoHE as the primary products. Free radical-mediated oxidation of DHA generates ten positional isomers of HpDoHE (20-, 17-, 16-, 14-, 13-, 11-, 10-, 8-, 7- and 4-HpDoHE). In contrast, singlet oxygen-mediated oxidation generates twelve positional isomers, including the same ten isomers produced by radical-induced oxidation and the 5- and 19-HpDoHE isomers. Additionally, the non-enzymatic oxidation of DHA is also reported to generate a series of secondary products, including neuroprostanes, neurofurans and aldehydes (neuroketals and short-chain aldehydes) [23-27].

Several lines of evidence indicate that lipid hydro(pero)xides are increase under pathological conditions [28-30]. For this reason, particular attention has been focused on the study of the formation and pathophysiological role of lipid hydro(pero)xides. For example, recent studies on hydroperoxides and other oxy-derivatives of DHA (e.g., resolvins and neuroprotectins) have enhanced our knowledge of the resolution phase of inflammation and of the roles of ASA and n-3 in this process [13]. In this context, the identification and structural characterization of HpDoHE and HDoHE formed by enzymatic and/or non-enzymatic mechanisms are critical to reveal new enzymes, products and biological activities [31]. Thus, the aim of this study was to develop a specific and quantitative analytical method suitable for both HpDoHE and HDoHE positional isomers.

Liquid chromatography coupled to tandem mass spectrometry (LC-MS/MS) with the selected reaction monitoring (SRM) method has been widely used for the identification and quantitation of lipid oxidation products. In this study, we present a detailed description of the standardization of a method for qualitative and quantitative LC-MS/MS-SRM analysis of the twelve different positional isomers of both HpDoHE and HDoHE as an additional tool for use in lipidomic studies aiming to dectect DHA oxidation products in biological samples.

Experimental Section

Chemicals

4Z,7Z,10Z,13Z,16Z,9Z-docosahexaenoic acid (DHA), 5S-hydroxy-6E,8Z,11Z,14Z-eicosatetraenoic-5,6,8,9,11,12,14,15-d8 acid (5(S)-HETE-d8) and 12S-hydroxy-5Z,8Z,10E,14Z-eicosatetraenoic-5,6,8,9,11,12,14,15-d8 acid (12(S)-HETE-d8) were purchased from Cayman Chemicals Corp. (Ann Arbor, MI). HPLC-grade methanol, acetonitrile, isopropanol, hexane, chloroform, as well as standard grade ammonium hydroxide, phosphoric acid and formic acid were obtained from JT Baker (Avantor Performance Materials, Mexico). Potassium hydroxide (KOH), hydrochloric acid (HCl), 2,6-di-tert-butyl-p-cresol (BHT), deferoxamine mesylate salt, Chelex 100 sodium form and diethylenetriaminepentaacetic acid were purchased from Sigma-Aldrich Inc. (St. Louis, MO). Ketamine hydrochloride and xylazine was obtained from Vet Brands (Sespo Ind. e Com., Brazil). All aqueous solutions were prepared with ultrapure water purified by a Direct-Q3 system (Merck Millipore, Germany) and treated with Chelex 100 before use. Millex Filter Units (0.22 μm) were purchased from Merck Milipore, Germany.

Preparation of HpDoHE and HDOHE standards

HpDoHE was synthesized by photooxidation of DHA under an atmosphere saturated with O2, and methylene blue was used as the photosensitizer, as previously described (see supporting information method S1 and scheme S1) [21]. The conversion of HpDoHE to HDoHE was performed as described by Terao et al.[32] (see supporting information method S2). The HpDoHE were analyzed and then purified with a Prominence HPLC system (Shimadzu, Tokyo, Japan) equipped with two types of columns: a reverse phase, C18 semi-preparative column (Luna C18-2 100Å, 250 x 10 mm, 5 μm, Phenomenex Inc., Torrance, CA), eluted with a mobile phase of acetonitrile:water:formic acid (70:30:0.005, v/v/v) and water at 4.7 mL/min; and a normal phase, silica semi-preparative column (Luna Silica 100 Å, 250 x 10 mm, 5 μm, Phenomenex Inc., Torrance, CA), with an isocratic mobile phase of hexane:isopropanol:water (99:1:0.1, v/v/v) at 10 mL/min. The PDA detector was set to scan from 200 to 500 nm, and the hydroperoxides were monitored at 205 (all isomers) and 235 nm (isomers with dienes conjugates) (supporting information scheme S1). The fractions containing the isolated isomers were dried with a rotary evaporator, and the residue was solubilized in methanol and stored at -80 °C. An aliquot of each HpDoHE and HDoHE isomer was checked by HPLC- PDA and UV absorbance at 235 nm. HpDoHE concentration was also confirmed by iodometry [33].

Chromatographic standardization

The chromatographic method was developed by a screening study using an automated and integrated system consisting of Fusion Method Development software (S-Matrix Corp., Eureka, CA), Empower 3 chromatography data software (Waters Corp., Milford, MA) and an UHPLC system (Acquity UPLC HClass, Waters Corp., Milford, MA). The screening and optimization study of the chromatographic method was done considering the following parameters: number of peaks, resolution greater than 0.80 and tailing smaller than 2.0. The screening study was performed with four different reversed phase UPLC columns (50 x 2.1 mm, 1.7 µm) the HSS T3, HSS PFP, BEH C8 and CSH Phenyl hexil. Three pH values were evaluated: phosphoric acid 0.1 % (pH 2.5), phosphoric acid 0.05 % (pH 3.5) and ammonium hidroxide 0.1 % (pH 10). Acetonitrile and methanol were used as organic modifier. Gradients between 80 - 100 % of the organic solvent was performed from 5 to 10 minutes and three temperatures were also evaluated 25, 35 and 40 °C. The optimized chromatographic condition consisted of a BEH C8 column (100 x 2.1 mm, 1,7 µm) eluted with a gradient solvent system of A, 0.1 % ammonium hydroxide in water, pH 10; and B, 18 % methanol in acetonitrile at 0.5 mL/min. Elution was started with 30 % B, was held for 1.55 min, was followed by a gradient step to 69% B over 15 min; then, the percentage of B was maintained at 95 % for 2 min and was restored to 30 % for 4 min to allow equilibration. Column temperature was set at 40 °C.

MS/MS analysis of HpDoHE and HDoHE

The MS/MS fragmentation pattern for each HpDoHE and HDoHE isomer was initially analyzed with a Quattro II triple quadrupole mass spectrometer (Micromass, Manchester, UK) and API 4000 QTrap (Applied Biosystems Inc., Foster City, CA). In this preliminary step, each isolated isomer was identified through a comparison of the obtained fragment ions with the theoretical fragments (Tables S1 and S2 in the supporting information). The final fragmentation study and the establishment of the quantitative method were conducted with an UHPLC system (Acquity UPLC) coupled to a triple-quadrupole mass spectrometer (XEVO TQ-S, Waters Corp., Milford, MA). Based on the MS/MS spectra, the most intense and/or specific fragment ions were selected for the SRM method. The MS and MS/MS analyses were conducted in ESI negative mode. The source temperature was set to 150 °C, the desolvation temperature was 550 °C, and the capillary voltage was set to 3 kV. The dwell time was set automatically as 9 msec. The collision energy and cone voltage were optimized for each compound (see Tables 1 and 2) by the Intellistart tool from MassLynx software (Waters Corp., Milford, MA). Minor adjustments on collision energies were also performed manually for some of the analytes having poor fragment ion intensities. In both cases, the optimization was performed by direct infusion of the isolated standards. The cone energy and collision energy were chosen as the energies which generate the strongest precursor signal and the greatest intensity for the chosen fragment, respectively. Two SRM methods, one for HpDoHE and the other for HDoHE, were created separately to ensure maximum sensitivity in the detection and quantification of the isomers. Peak identification and quantification were performed with TargetLynx software (Waters Corp., Milford, MA). The qualitative SRM was set as target trace and the quantitative SRM was set as quantification trace. The analyte was only quantified in the presence of both transitions.

CompoundSRM (m/z)CV (V)CE (V)RT (min)LOD (pg)*r2Calibrated range (ng/µL)
20-HpDoHE (2)359.10 → 71.10Quantitative33146.794710.9980.25-10
341.10 → 71.10Qualitative3314
19-HpDoHE (1)341.30 → 83.10Quantitative41196.66230.9920.01-1
359.30 → 83.10Qualitative4119
17-HpDoHE (4)341.30 → 111.20Quantitative30127.0310.9910.01-1
359.30 → 111.20Qualitative3012
16-HpDoHE (3)341.20 → 233.20Quantitative31117.11490.990.01-1
359.20 → 233.20Qualitative3111
14-HpDoHE (6)341.10 → 151.10Quantitative29117.28290.9910.01-1
359.10 → 151.10Qualitative2911
13-HpDoHE (5)341.10 → 121.10Quantitative32167.2210.9960.01-1
359.10 → 121.10Qualitative3216
11-HpDoHE (7)341.10 → 243.20Quantitative29117.39760.9980.01-1
359.10 → 243.20Qualitative2911
10-HpDoHE (7)341.10 → 188.20Quantitative31127.396710.9830.01-1
359.10 → 188.20Qualitative3112
8-HpDoHE (8)359.30 → 108.30Quantitative30167.831160.9950.01-1
341.30 → 113.10Qualitative319
7-HpDoHE (8)341.10 → 201.20Quantitative25137.80960.9970.01-1
359.10 → 201.20Qualitative2512
5-HpDoHE (9)359.20 → 281.30Quantitative32108.731620.9810.01-1
359.20 → 147.20Qualitative328
4-HpDoHE (10)341.30 → 115.10Quantitative22158.845020.9990.01-1
359.20 → 115.10Qualitative2215
5(S)-HETE-d8327.20 → 116.1022147.25
12(S)-HETE-d8327.20 → 184.2030146.13

Table 1. Optimized mass conditions for LC-MS/MS method for HpDoHE.

Numbers in bold correspond to the peak visualized in the LC-MS analysis (Figure 2). * on column
Optimized mass transitions (m/z), cone voltage (CV), collision energy (CE), retention time (RT), limit of detection (LOD) and linearity (r2).
CSV
Download CSV
CompoundSRM (m/z)CV (V)CE (V)RT (min)LOD (pg)*r2Calibrated range (pg/µL)
20-HDoHE (12)343.10 → 241.10Quantitative36155.952.710.9980.62-100
343.10 → 285.10Qualitative3613
19-HDoHE (11)343.10 → 229.10Quantitative32156.105.020.9990.62-100
343.10 → 273.30Qualitative3214
17-HDoHE (14)343.10 → 201.10Quantitative27146.463.890.9971.25-100
343.10 → 245.30Qualitative2714
16-HDoHE (13)343.10 → 233.20Quantitative26136.361.270.9960.62-100
343.10 → 261.10Qualitative2613
14-HDoHE (16)343.10 → 161.20Quantitative30126.712.600.9960.25-100
343.10 → 205.10Qualitative3015
13-HDoHE (15)343.10 → 193.20Quantitative30126.581.160.9970.25-100
343.10 → 221.10Qualitative3013
11-HDoHE (18)343.10 → 149.20Quantitative28136.955.300.9930.25-100
343.10 → 165.10Qualitative2813
10-HDoHE (17)343.10 → 153.10Quantitative28136.842.070.9980.62-100
343.10 → 181.10Qualitative2815
8-HDoHE (19)343.10 → 189.20Quantitative32127.351.630.9970.25-100
343.10 → 113.10Qualitative3213
7-HDoHE (19)343.10 → 141.10Quantitative31157.360.460.9940.25-100
343.10 → 109.10Qualitative3113
5-HDoHE (20)343.20 → 85.10Quantitative26108.078.480.9941.25-100
343.20 → 93.10Qualitative2610
4-HDoHE (21)343.10 → 101.10Quantitative29138.400.500.9940.25-100
343.10 → 115.10Qualitative2916
5(S)-HETE-d8327.20 → 116.1022147.25
12(S)-HETE-d8327.20 → 184.2030146.13

Table 2. Optimized mass conditions for LC-MS/MS method for HDoHE.

Numbers in bold correspond to the peak visualized in the LC-MS analysis (Figure 2). * on column
Optimized mass transitions (m/z), cone voltage (CV), collision energy (CE), retention time (RT), limit of detection (LOD) and linearity (r2).
CSV
Download CSV

Calibration and validation experiments

The method validation was performed as in-house 3-day protocol to determine linearity, LOD and LOQ, inter- and intra-day variation, and recovery for all compounds. Stock standard solutions of all the HpDoHE and HDoHE isomers were prepared in acetonitrile:methanol:H2O (52:18:30, v/v/v) and stored in amber vials at -80 °C. These stock solutions were diluted to prepare the solutions for the calibration curves. The following concentrations were used for the calibration curve: 10, 4, 2, 1, 0.5, 0.25, 0.1, 0.05, 0.025 and 0.01 ng/µL for HpDoHE: and, 4, 2, 1, 0.1, 0.05, 0.025, 0.01, 0.0050, 0.0025, 0.0012, 0.0006 and 0.0003 ng/µL for HDoHE. Isotopically labeled internal standards were prepared in 100 µL of acetonitrile:methanol:H2O (52:18:30, v/v/v). For quantitative analysis, 10 µL of 5(S)-HETE-d8 (12.5 ng/µL) and 12(S)-HETE-d8 (12.5 ng/µL) were added to the samples. The injection volume was 10 µL. The calibration curves were constructed by plotting the ratio of the peak areas of the analyte and the internal standard as a function of analyte concentration with linear regression. The linearity of the method was assessed by performing 5 replicate analyses with 8 different concentrations. For recovery calculations the phosphate buffer saline solution (PBS) was spiked either before or after the extraction with 1 ng/µL of each analyte and the ratios of the peak areas were calculated. The same procedure was conducted with plasma and brain homogenate. The accuracy and precision of the assay were assessed by analyzing blank samples (methanol) spiked with 3 different concentrations in 3 replicates on the same day and on 3 consecutive days for intra- and inter-day precision and accuracy. Precision was calculated as the relative standard deviation (%) and accuracy was determined from the percentage ratio of the measured concentration to the expected concentration.

Ethics statement

The experimental procedures were conducted in accordance with the ethical principles for animal experimentation adopted by the Brazilian College of Animal Experimentation and were approved by the ethics committee on Animal Care and Use (Comissão de Ética em Cuidados e Uso Animal do Instituto de Química da Universidade de São Paulo – CEUA – IQ-USP) (Permit Number: 15/2011). All surgery was performed under anesthesia, and all efforts were made to minimize animal suffering.

Biological sample preparation

Plasma and brain samples were obtained from four-month-old Sprague-Dawley rats (n=3). The rats were maintained under a controlled temperature and light-dark cycle with food and water offered ad libitum. The animals were anesthetized with an intraperitoneal dose of ketamine hydrochloride (0.9 mL/kg body weight) and xylazine (0.5 mL/kg body weight). Blood was collected from the right atrium of the heart by cardiac puncture into a heparinized tube and centrifuged for 30 min at 4 °C and 1500 x g to separate the plasma. For the analysis, plasma from 3 rats was combined and stored at -20°C until use. Following the blood collection, the rats were immediately decapitated, their whole brains were rapidly excised, and the cortexes were separated and frozen at -20°C. Similarly, the cortex samples (0.2 - 0.3 g) from 3 rats were combined and were homogenized in ice by using an PowerGen 1000 homogeneizer (Fisher Scientific) for 20 s in 5 vol (1.0 - 1.5 mL) of PBS solution (10 mM, pH 7.4) for 2 min on ice.

Lipid extraction

Lipids were extracted from plasma and brain samples after saponification step by the Bligh and Dyer method, with modifications [34]. Ice-cold methanol containing BHT (100 µM) and 5(S)-HETE-d8 (0.25 ng/µL, 500 µL) and KOH (1 M) in methanolic solution (500 µL) were added to the samples (500 µL plasma or brain homogenate). The mixtures were incubated for 30 min in the dark at 37°C. The mixture was cooled on ice, acidified with 60 µL HCl (10 M) and then extracted with chloroform/methanol/water. The sample was mixed with a vortex mixer for 1 min and centrifuged at 1500 x g for 5 min at 4°C. The chloroform/methanol layer was dried with nitrogen, and the residue was resuspended in 100 µL of acetonitrile:methanol:H2O (52:18:30, v/v/v). Finally, 10 µL of 12(S)-HETE-d8 (12.5 ng/µL) in the same mixture of solvents was added, and the resulting solution was filtered through 0.22 μm Millex Filter Units before injection into the LC-MS/MS system. The injection volume was 10 µL.

Statistical analysis

The differences between the concentration levels obtained for each isomer (mean ± S.D.) were determined by one way analysis of variance (ANOVA) followed by the Tukey–Kramer multiple comparison test. A P value of 0.05 or less was used as the criterion for statistical significance.

Results and Discussion

The specific and quantitative analysis of the HpDoHE and HDoHE positional isomers is non-trivial due to the structural similarities and common fragmentation patterns of these molecules. For this reason, there have been few reports describing the analysis of HpDoHE and HDoHE by LC-MS/MS [6,35-38]. To the best of our knowledge, this study is the first analytical report describing the detailed MS/MS characterization and the development of an SRM method for reliably discriminating all twelve positional isomers of both HpDoHE and HDoHE.

Chromatographic separation of HpDoHE and HDoHE isomers

The UHPLC chromatographic condition optimized by the automated screening study consisted of a C8 column eluted with a gradient solvent system 0.1% ammonium hydroxide in water (pH 10) and methanol:acetonitrile (18:82, v/v).To confirm the resolution of the analytes using the optimized chromatographic condition, a mixture containing all twelve positional isomers of both HpDoHE and HDoHE was analyzed by UHPLC-MS. Figure 2 shows the chromatographic separation of the standardized condition that separates 10 peaks corresponding to the HpDoHE isomers (numbered sequentially from 1-10) and 11 peaks corresponding to HDoHE (numbered sequentially from 11-21).

thumbnail
Figure 2. Analysis of a mixture of HpDoHE and HDOHE isomers by UPLC-MS.

The HpDoHE and HDOHE isomers were detected through the selection of m/z at 359 (A) and 343 (B), respectively. The identities of the peaks numbered in red are shown in Tables 1 and 2 and Figure 5.

https://doi.org/10.1371/journal.pone.0077561.g002

Characterization of HpDoHE and HDoHE isomers by MS/MS analysis

For the MS/MS method development we initially focused on the characterization and optimization of MS/MS parameters to obtain the most selective and sensitive ionization and fragmentation conditions for all 24 positional isomers. The HpDoHE and HDoHE generated abundant deprotonated molecules [M-H] at m/z 359 and 343, respectively, which were selected as precursor ions for collision-induced dissociation (CID) in MS/MS analysis. Additionally, the HpDoHE were easily dehydrated [M-H-H20] during ionization to yield an intense ion at m/z 341. Thus, the MS/MS studies of HpDoHE were performed by selecting both the deprotonated (m/z 359) and dehydrated molecules (m/z 341). The fragmentation spectra of the HpDoHE and HDoHE isomers are shown in Figures 1 and 2, respectively. The MS/MS spectra obtained by selecting the precursor ions at m/z 359 and 341 were similar (data not shown).

The fragmentation of fatty acid hydro(pero)xide ions by CID-MS/MS yields fragment ions that are common to all or more than one isomer (non-specific fragments) as well as fragments that are specific and indicative of the position of the hydroperoxy or hydroxyl group (specific fragments) [39-41]. Non-specific fragments are usually formed through peripheral cleavages, such as the loss of water, CO2 or both, whereas isomer-specific fragments are formed through internal cleavages of the carbon-carbon bond. For the HpDoHE, non-specific fragment ions were observed at m/z 341, 315 and 297; for the HDoHE, non-specific fragments were observed at m/z 325, 299 and 281. In contrast, specific fragment ions were derived from the α- or β-cleavage of the carbon-carbon bond adjacent to the hydroperoxide/hydroxide group. The α- and β-cleavages can occur at the carboxy or methyl side, giving rise to four possible sites of fragmentation (Scheme S2, Supporting information). Moreover, each of these fragmentation paths yields two fragments; one containing the carboxy segment, and the other the methyl segment. For ease of reading, we named these fragment ions according to the three-letter code proposed by the Serhan group, with some modifications (Scheme S2)[37,41]. The expected theoretical fragments formed through these fragmentation paths were listed for each isomer (Tables S1 and S2, Supporting Information) and were compared with the obtained data.

Consistent with previous studies [36,37,39,42], the mono-hydroperoxy (Figure 3) and hydroxy (Figure 4) derivatives of DHA exhibited characteristic fragment ions derived from the α- and β-cleavages followed by one or two hydrogen shifts. MS/MS spectra of the HDoHE isomers showed intense specific fragment ions derived mostly from α-cleavage; in agreement with the data reported by Hong et al.[37] For instance, the 20-, 19-, 17-, 16-, 14-, 13-, 11-, 10- and 5-HDoHE isomers showed intense fragment ions corresponding to the [αcc+H] and [αcc+H-CO2] ions, whereas the 8-, 7- and 4-HDoHE isomers showed intense fragment ions corresponding to [αmc-H] and/or [αmm] (Figure 4). In contrast, fragmentation pattern for HpDoHE isomers was more variable. The 20-, 19-, 17-, 16-, 14- and 13-HpDoHE yielded intense ions that corresponded to [αcc+H], [αcc+H-CO2] and [βcm+H-H2O] and the other isomers (11-, 10-, 8-, 7-, 5- and 4-HpDoHE) exhibited the typical α- and β-cleavages, in addition to other types of chain fragmentations that were not explored further in this study.

thumbnail
Figure 3. Fragmentation pattern analysis of the HpDoHE isomers – MS/MS spectra obtained from m/z 359.

The selected fragments for SRM analysis are depicted in each isomer structure. The numbers in red correspond to the peaks observed in the LC-MS analysis (Figure 2).

https://doi.org/10.1371/journal.pone.0077561.g003

thumbnail
Figure 4. Fragmentation pattern analysis of the HDoHE isomers – MS/MS spectra obtained from m/z 343.

The selected fragments for SRM analysis are depicted in each isomer structure. The numbers in red correspond to the peaks observed in the LC-MS analysis (Figure 2).

https://doi.org/10.1371/journal.pone.0077561.g004

In summary, a comparison of the fragment data obtained in this study and the theoretically expected data allowed us to unambiguously identify and characterize each HpDoHE and HDoHE isomer.

Development of the selected reaction monitoring method

Based on the MS/MS spectra of each HpDoHE and HDoHE isomer we developed an SRM method. The selection was based on the intensity and specificity of the fragment ions observed in the MS/MS spectra (Figures 3 and 4).Tables S1 and S2 in the Supporting Information provide the selected fragments used for the SRM method (in red). Figure 5 shows the mass chromatograms obtained using the newly developed SRM method.

thumbnail
Figure 5. Representative chromatograms of individual SRM transitions selected for each HpDoHE and HDoHE isomer.

The numbers in red correspond to the peaks observed in the LC-MS analysis (Figure 2).

https://doi.org/10.1371/journal.pone.0077561.g005

Specific fragment ions selected for HDoHE isomers were all derived from α-cleavage, whereas for HpDoHE isomers, the selected fragment ion was derived mostly from β-cleavage. For some isomers, such as 11-, 7- and 5-HpDoHE, it was necessary to select less specific fragment ions to gain sensitivity in the analysis. As mentioned before, the ESI ionization of hydroperoxides favors the appearance of its corresponding dehydrated ion [39]. Therefore, to gain sensitivity in the analysis, the ion at m/z 341 was also selected as an alternative precursor ion for the analysis of the hydroperoxides.

Quantitative method

The fragment ions selected for the LC-MS/MS-SRM detection of the HpDoHE and HDoHE isomers were chosen to yield the best resolution, selectivity and the highest signal to noise ratio for each isomer. Differently to Yang et al. [43] we did not split the MS method containing all SRM transitions into different acquisition periods. Instead, we chose to split the analysis into two separate methods (one to analyze HpDoHE and the other to analyze the HDoHE isomers) to maintain an adequate number of data points across each chromatographic peak ( 15 points per peak) and thus increase sensitivity.

The selected mass transitions and the optimized conditions for the mass spectrometer are presented in Tables 1 and 2. Two mass transitions were selected for each isomer to enhance the detection and quantification selectivity. To determine the limit of detection (LOD) and the linear dynamic ranges for the different isomers, five batches of calibration curves containing 8 different concentrations of each analyte were used. The LOD values were calculated with a signal to noise ratio of 3. The LOD values were within 1−670 pg for HpDoHE and 0.5−8.5 pg for HDoHE injected onto the column, showing that the method is about 10-100 times more sensitive for the detection of the hydroxides compared to hydroperoxides. The batches used in the present work gave linear dynamic ranges of 12.5−1000 pg/µL for HpDoHE and 0.3−100 pg/µL for HDoHE. The r2 values determined by regression analysis were 0.99 or greater for each isomer.

The recoveries for all tested compounds in PBS were good and ranged from 80 to 120 %, with the exception of the 20-HpDoHE which showed a recovery of about 60 % (Table S3 from supporting information). In biological matrices, the average recovery for the HpDoHE isomers was poor (16 ± 22 % in brain and 22 ± 14 % in plasma) compared to the HDoHE isomers (84 ± 20 % in brain and 90 ± 13 % in plasma) which showed good recoveries for all isomers (Table S3 from supporting information). Considering that hydroperoxide loss due to sample processing is around 20 %, we can assume that most of the HpDoHE isomers was either reduced by the antioxidant machinery or degraded by some components of the biological sample. In contrast, recoveries for the hydroxides were good probably reflecting their greater stability in biological samples compared to the hydroperoxides.

The precision and accuracy of the method showed to be also good. All tested isomers with the exception of 5-HpDoHE and 5-HDoHE present relative standard deviation lower than 8 % and an accuracy higher than 90% (Table S4 and S5 from supporting information).

Application of the method to biological samples

To demonstrate the applicability of our method as a tool for HpDoHE and HDoHE lipidomic analysis, we used it to detect the basal levels of these isomers in rat plasma and brain samples. All cautions to avoid ex-vivo oxidations were taken, such as keeping samples at low temperatures and using antioxidants and chelating agents during sample preparation and lipid extraction.

As it would be expected from the recovery studies, the HpDoHE isomers were not detected in the tested biological samples. Among the HDoHE isomers, eleven were detected in the rat plasma. The most abundant isomer observed in plasma was the 14-HDoHE isomer (51.55 ± 9.45 ng/mL), a 12-LOX product, which was present at a 6-10-fold higher concentration than the other isomers (Figure 6A, P<0,001). A similar trend was previously observed by Gomolka et al., who also detected higher levels of 14-HDoHE (65.40 ± 15.84 ng/mL) in whole blood samples from mice [38].

thumbnail
Figure 6. HDoHE profile of rat plasma and brain homogenate samples taken under basal conditions.

The results are the mean ± SD of three independent analyses. Values not sharing common superscript are significantly different (P<0.05).

https://doi.org/10.1371/journal.pone.0077561.g006

In brain samples, all twelve HDoHE isomers were detected (Figure 6B). To date, the only studies describing the detection of HDoHE in the brain have been performed in models in which the brain homogenates were incubated with DHA or were challenged to produce the oxidized products [6,7,44,45]. To our knowledge, this is the first study describing the detection of the twelve isomers in brain homogenates at basal conditions.

Interestingly, we found relatively higher levels of HDoHE in the brain than in the plasma samples. This is most likely due to the presence of high concentrations of DHA in nervous tissue. Among the twelve isomers, 20-, 14-, 11- and 4-HDoHE were predominant, and these isomers were all present at similar levels in brain sample (Figure 6B). The 12-LOX is the major LOX in the brain, and this might explain the high levels of 14-HDoHE and 11-HDoHE found in this tissue [46,47]. A predominance of the 20- and 4-HDoHE isomers in brain was also previously reported by Kim et al [45]. These isomers seem to be preferentially accumulated through the non-enzymatic oxidation of DHA [36].

Among the less abundant HDoHE isomers, it should be pointed out that we could also detect the 19- and 5-HDoHE isomers, which are known to be specifically formed by singlet oxygen mediated oxidation. Despite the need of further investigations, their detection could serve as a fingerprint for singlet oxygen-mediated oxidation [48,49].

Conclusions

HpDoHE and HDoHE can act as important lipid mediators in many physiological and pathophysiological events. However, there is little literature describing the quantification of HpDoHE and HDoHE isomers in biological samples. Additionally, the studies that have sought to examine HpDoHE and HDoHE were based on the analysis of lower number of isomers. In this study, we have standardized an LC-MS/MS-SRM method for the analysis of 12 isomers of each HpDoHE and HDoHE. In this way, we are providing a broad and specific method for the analysis of HpDoHE and HDoHE isomers that can be applied to the studies that seek to understand the role of DHA and its oxidation products in biological systems.

Supporting Information

Method S1.

HpDoHE synthesis by photooxidation.

https://doi.org/10.1371/journal.pone.0077561.s001

(DOCX)

Scheme S1.

Representative UV spectra of the HpDoHE and HDoHE isomers with (A) and without (B) conjugated dienes.

https://doi.org/10.1371/journal.pone.0077561.s003

(TIFF)

Scheme S2.

The nomenclature for the HpDoHE and HDoHE chain-cut segments obtained from theoretical MS/MS fragmentation. This scheme was adapted from Serhan’s group proposed nomenclature for MS/MS ions generated from docosanoids. Each fragment is identified by a 3-letter abbreviation in which: (A) the first Greek letter (α or β) indicates the position of the carbon-carbon bond being cleaved relative to the hydroperoxide/hydroxide group; (B) the second letter indicates whether the cleavage occurred at the carboxy (c) or methyl (m) side; and (C) the third letter indicates the segment that corresponds to the fragment ion, the carboxy (c) or methyl (m) side segment. Thus, with this nomenclature, 4 types of fragment ions can be formed from α-carbon bond cleavage (αcc, αcm, αmc, αmm) or from β-carbon bond cleavage (βcc, βcm, βmc, βmm). For a more detailed explanation on the fragmentation mechanism, see Murphy et al.[42] and Hong et al.[37].

https://doi.org/10.1371/journal.pone.0077561.s004

(TIFF)

Table S1.

Mass of theoretical fragments and m/z values of the ions used for the identification of HpDoHE.

https://doi.org/10.1371/journal.pone.0077561.s005

(DOCX)

Table S2.

Mass of theoretical fragments and m/z values of the ions used for the identification of HDoHE.

https://doi.org/10.1371/journal.pone.0077561.s006

(DOCX)

Table S3.

Recovery of HpDoHE an HDoHE in different matrix (n=3).

https://doi.org/10.1371/journal.pone.0077561.s007

(DOCX)

Table S4.

Intra-day and inter-day precision (coefficients of variation) and accuracies for the twelve HpDoHE isomers.

https://doi.org/10.1371/journal.pone.0077561.s008

(DOCX)

Table S5.

Intra-day and inter-day precision (coefficients of variation) and accuracies for the twelve HDoHE isomers.

https://doi.org/10.1371/journal.pone.0077561.s009

(DOCX)

Acknowledgments

We thank Prof. Norberto Peporine Lopes (Faculty of Pharmaceutical Sciences of Ribeirão Preto, University of São Paulo) for technical revision of the main text and to Prof. Maurício da Silva Baptista (Institute of Chemistry, University of São Paulo) for the determination of the luminous intensity of the lamps used for HpDoHE synthesis. We also thank to Danilo Pereira (Waters Technologies from Brazil) and Izaura N. Toma and Fernanda M. Prado (Institute of Chemistry, University of São Paulo), for mass spectrometry technical assistance.

Author Contributions

Conceived and designed the experiments: PBMCD SM. Performed the experiments: PBMCD ASFM FPF DC PPA. Analyzed the data: PBMCD SM. Contributed reagents/materials/analysis tools: FP TCL PDM MHGM MM. Wrote the manuscript: PBMCD FPF ASFM DC PPA FP TCL MM PDM MHGM SM.

References

  1. 1. Salem N, Litman B, Kim HY, Gawrisch K (2001) Mechanisms of action of docosahexaenoic acid in the nervous system. Lipids 36: 945-959. doi:https://doi.org/10.1007/s11745-001-0805-6. PubMed: 11724467.
  2. 2. Horrocks LA, Yeo YK (1999) Health benefits of docosahexaenoic acid (DHA). Pharmacol Res 40: 211-225. doi:https://doi.org/10.1006/phrs.1999.0495. PubMed: 10479465.
  3. 3. Lukiw WJ, Bazan NG (2008) Docosahexaenoic Acid and the Aging Brain. J Nutr 138: 2510-2514. doi:https://doi.org/10.3945/jn.108.096016. PubMed: 19022980.
  4. 4. Bazan NG, Molina MF, Gordon WC (2011) Docosahexaenoic acid signalolipidomics in nutrition: significance in aging, neuroinflammation, macular degeneration, Alzheimer's, and other neurodegenerative diseases. Annu Rev Nutr 31: 321-351. doi:https://doi.org/10.1146/annurev.nutr.012809.104635. PubMed: 21756134.
  5. 5. Appolinário PP, Derogis PBMC, Yamaguti TH, Miyamoto S (2011) Metabolismo, oxidação e implicações biológicas do ácido docosahexaenoico em doenças neurodegenerativas. Química Nova 34: 1409-1416. doi:https://doi.org/10.1590/S0100-40422011000800021.
  6. 6. Hong S, Gronert K, Devchand PR, Moussignac RL, Serhan CN (2003) Novel docosatrienes and 17S-resolvins generated from docosahexaenoic acid in murine brain, human blood, and glial cells - Autacoids in anti-inflammation. J Biol Chem 278: 14677-14687. doi:https://doi.org/10.1074/jbc.M300218200. PubMed: 12590139.
  7. 7. Serhan CN, Hong S, Gronert K, Colgan SP, Devchand PR et al. (2002) Resolvins. J Exp Med 196: 1025-1037. doi:https://doi.org/10.1084/jem.20020760. PubMed: 12391014.
  8. 8. Bazan NG, Birkle DL, Reddy TS (1984) Docosahexaenoic acid (22:6, n-3) is metabolized to lipoxygenase reaction products in the retina. Biochem Biophys Res Commun 125: 741-747. doi:https://doi.org/10.1016/0006-291X(84)90601-6. PubMed: 6240268.
  9. 9. Fer M, Dréano Y, Lucas D, Corcos L, Salaün JP et al. (2008) Metabolism of eicosapentaenoic and docosahexaenoic acids by recombinant human cytochromes P450. Arch Biochem Biophys 471: 116-125. doi:https://doi.org/10.1016/j.abb.2008.01.002. PubMed: 18206980.
  10. 10. Yin H, Xu L, Porter NA (2011) Free radical lipid peroxidation: mechanisms and analysis. Chem Rev 111: 5944-5972. doi:https://doi.org/10.1021/cr200084z. PubMed: 21861450.
  11. 11. Cheng Z, Li Y (2007) What is responsible for the initiating chemistry of iron-mediated lipid peroxidation: an update. Chem Rev 107: 748-766. doi:https://doi.org/10.1021/cr040077w. PubMed: 17326688.
  12. 12. Sun YP, Oh SF, Uddin J, Yang R, Gotlinger KH et al. (2007) Resolvin D1 and its aspirin-triggered 17R epimer. J Biol Chem 282: 9323-9334. doi:https://doi.org/10.1074/jbc.M609212200. PubMed: 17244615.
  13. 13. Serhan CN, Petasis NA (2011) Resolvins and protectins in inflammation resolution. Chem Rev 111: 5922-5943. doi:https://doi.org/10.1021/cr100396c. PubMed: 21766791.
  14. 14. Serhan CN, Yang R, Martinod K, Kasuga K, Pillai PS et al. (2008) Maresins: novel macrophage mediators with potent antiinflammatory and proresolving actions. J Exp Med 206: 15-23. PubMed: 19103881.
  15. 15. Aveldano MI, Sprecher H (1983) Synthesis of hydroxy fatty acids from Aprel, 7, 10, 13, 16, 19-[1-14C] docosahexaenoic acid by human platelets. Journal of Biological Chemistry 258: 9339-9343.
  16. 16. Sapieha P, Stahl A, Chen J, Seaward MR, Willett KL et al. (2011) 5-Lipoxygenase metabolite 4-HDHA is a mediator of the antiangiogenic effect of omega-3 polyunsaturated fatty acids. Sci_Transl Med 3: 69ra12.
  17. 17. Westphal C, Konkel A, Schunck WH (2011) CYP-eicosanoids--a new link between omega-3 fatty acids and cardiac disease? Prostaglandins Other Lipid Mediat 96: 99-108. doi:https://doi.org/10.1016/j.prostaglandins.2011.09.001. PubMed: 21945326.
  18. 18. Miyamoto S, Martinez GR, Rettori D, Augusto O, Medeiros MHG et al. (2006) Linoleic acid hydroperoxide reacts with hypochlorous acid, generating peroxyl radical intermediates and singlet molecular oxygen. Proc Natl Acad Sci USA 103: 293-298. doi:https://doi.org/10.1073/pnas.0508170103. PubMed: 16387855.
  19. 19. Miyamoto S, Ronsein GE, Prado FM, Uemi M, Corrêa TC et al. (2007) Biological hydroperoxides and singlet molecular oxygen generation. IUBMB Life 59: 322-331. doi:https://doi.org/10.1080/15216540701242508. PubMed: 17505972.
  20. 20. Uemi M, Ronsein GE, Prado FM, Motta FD, Miyamoto S et al. (2011) Cholesterol hydroperoxides generate singlet molecular oxygen [O(2) ((1)Delta(g))]: near-IR emission, (18)O-labeled hydroperoxides, and mass spectrometry. Chemical Research in Toxicology 24: 887-895.
  21. 21. Miyamoto S, Martinez GR, Martins APB, Medeiros MHG, Di Mascio P (2003) Direct evidence of singlet molecular oxygen production in the reaction of linoleic acid hydroperoxide with peroxynitrite. J Am Chem Soc 125: 4510-4517. doi:https://doi.org/10.1021/ja029262m. PubMed: 12683821.
  22. 22. Miyamoto S, Martinez GR, Medeiros MH, Di Mascio P (2003) Singlet molecular oxygen generated from lipid hydroperoxides by the russell mechanism: studies using 18(O)-labeled linoleic acid hydroperoxide and monomol light emission measurements. J Am Chem Soc 125: 6172-6179. doi:https://doi.org/10.1021/ja029115o. PubMed: 12785849.
  23. 23. Roberts LJ 2nd, Montine TJ, Markesbery WR, Tapper AR, Hardy P et al. (1998) Formation of isoprostane-like compounds (neuroprostanes) in vivo from docosahexaenoic acid. J Biol Chem 273: 13605-13612. doi:https://doi.org/10.1074/jbc.273.22.13605. PubMed: 9593698.
  24. 24. Bernoud-Hubac N, Davies SS, Boutaud O, Montine TJ, Roberts LJ 2nd (2001) Formation of highly reactive gamma-ketoaldehydes (neuroketals) as products of the neuroprostane pathway. J Biol Chem 276: 30964-30970. doi:https://doi.org/10.1074/jbc.M103768200. PubMed: 11413140.
  25. 25. Esterbauer H, Schaur RJ, Zollner H (1991) Chemistry and biochemistry of 4-hydroxynonenal, malonaldehyde and related aldehydes. Free Radic Biol Med 11: 81-128. doi:https://doi.org/10.1016/0891-5849(91)90192-6. PubMed: 1937131.
  26. 26. Kawai Y, Takeda S, Terao J (2006) Lipidomic analysis for lipid peroxidation-derived aldehydes using gas chromatography-mass spectrometry. Chem Res Toxicol 20: 99-107. PubMed: 17226932.
  27. 27. Song WL, Lawson JA, Reilly D, Rokach J, Chang CT et al. (2008) Neurofurans, novel indices of oxidant stress derived from docosahexaenoic acid. J Biol Chem 283: 6-16. PubMed: 17921521.
  28. 28. Jira W, Spiteller G, Schramm A (1996) Increase in hydroxy fatty acids in human low density lipoproteins with age. Chem Phys Lipids 84: 165-173. doi:https://doi.org/10.1016/S0009-3084(96)02635-7. PubMed: 9022221.
  29. 29. Niki E (2009) Lipid peroxidation: physiological levels and dual biological effects. Free Radic Biol Med 47: 469-484. doi:https://doi.org/10.1016/j.freeradbiomed.2009.05.032. PubMed: 19500666.
  30. 30. Praticò D, Zhukareva V, Yao Y, Uryu K, Funk CD et al. (2004) 12/15-lipoxygenase is increased in Alzheimer's disease: possible involvement in brain oxidative stress. Am J Pathol 164: 1655-1662. doi:https://doi.org/10.1016/S0002-9440(10)63724-8. PubMed: 15111312.
  31. 31. Schneider C (2009) An update on products and mechanisms of lipid peroxidation. Mol Nutr Food Res 53: 315-321. doi:https://doi.org/10.1002/mnfr.200800131. PubMed: 19006094.
  32. 32. Terao J, Shibata SS, Matsushita S (1988) Selective quantification of arachidonic acid hydroperoxides and their hydroxy derivatives in reverse-phase high performance liquid chromatography. Anal Biochem 169: 415-423. doi:https://doi.org/10.1016/0003-2697(88)90306-5. PubMed: 3382013.
  33. 33. Buege JA, Aust SD (1978) Microsomal lipid peroxidation. Methods Enzymol 52: 302-310. doi:https://doi.org/10.1016/S0076-6879(78)52032-6. PubMed: 672633.
  34. 34. Bligh EG, Dyer WJ (1959) A rapid method of total lipid extraction and purification. Can J Biochem Physiol 37: 911-917. doi:https://doi.org/10.1139/o59-099. PubMed: 13671378.
  35. 35. Seal JR, Porter NA (2004) Liquid chromatography coordination ion-spray mass spectrometry (LC-CIS-MS) of docosahexaenoate ester hydroperoxides. Anal Bioanal Chem 380: 1007-1013. PubMed: 1533809314634702.
  36. 36. Lyberg AM, Adlercreutz P (2006) Monitoring monohydroperoxides in docosahexaenoic acid using high-performance liquid chromatography. Lipids 41: 67-76. doi:https://doi.org/10.1007/s11745-006-5072-z. PubMed: 16555474.
  37. 37. Hong S, Lu Y, Yang R, Gotlinger KH, Petasis NA et al. (2007) Resolvin D1, protectin D1, and related docosahexaenoic acid-derived products: Analysis via electrospray/low energy tandem mass spectrometry based on spectra and fragmentation mechanisms. J Am Soc Mass Spectrom 18: 128-144. doi:https://doi.org/10.1016/j.jasms.2006.09.002. PubMed: 17055291.
  38. 38. Gomolka B, Siegert E, Blossey K, Schunck WH, Rothe M et al. (2011) Analysis of omega-3 and omega-6 fatty acid-derived lipid metabolite formation in human and mouse blood samples. Prostaglandins Other Lipid Mediat 94: 81-87. doi:https://doi.org/10.1016/j.prostaglandins.2010.12.006. PubMed: 21236358.
  39. 39. MacMillan DK, Murphy RC (1995) Analysis of lipid hydroperoxides and long-chain conjugated keto acids by negative ion electrospray mass spectrometry. J Am Soc Mass Spectrom 6: 1190-1201. doi:https://doi.org/10.1016/1044-0305(95)00505-6.
  40. 40. Kerwin JL, Torvik JJ (1996) Identification of monohydroxy fatty acids by electrospray mass spectrometry and tandem mass spectrometry. Anal Biochem 237: 56-64. doi:https://doi.org/10.1006/abio.1996.0200. PubMed: 8660537.
  41. 41. Lu Y, Hong S, Tjonahen E, Serhan CN (2005) Mediator-lipidomics: databases and search algorithms for PUFA-derived mediators. J Lipid Res 46: 790-802. doi:https://doi.org/10.1194/jlr.D400020-JLR200. PubMed: 15722568.
  42. 42. Murphy RC, Fiedler J, Hevko J (2001) Analysis of nonvolatile lipids by mass spectrometry. Chem Rev 101: 479-526. doi:https://doi.org/10.1021/cr9900883. PubMed: 11712255.
  43. 43. Yang J, Schmelzer K, Georgi K, Hammock BD (2009) Quantitative Profiling Method for Oxylipin Metabolome by Liquid Chromatography Electrospray Ionization Tandem Mass Spectrometry. Anal Chem 81: 8085-8093. doi:https://doi.org/10.1021/ac901282n. PubMed: 19715299.
  44. 44. Hong S, Tjonahen E, Morgan EL, Lu Y, Serhan CN et al. (2005) Rainbow trout (Oncorhynchus mykiss) brain cells biosynthesize novel docosahexaenoic acid-derived resolvins and protectins-Mediator lipidomic analysis. Prostaglandins Other Lipid Mediat 78: 107-116. doi:https://doi.org/10.1016/j.prostaglandins.2005.04.004. PubMed: 16303609.
  45. 45. Kim HY, Karanian JW, Shingu T, Salem N Jr. (1990) Stereochemical analysis of hydroxylated docosahexaenoates produced by human platelets and rat brain homogenate. Prostaglandins 40: 473-490. doi:https://doi.org/10.1016/0090-6980(90)90110-H. PubMed: 2147773.
  46. 46. Bendani MK, Palluy O, Cook-Moreau J, Beneytout JL, Rigaud M et al. (1995) Localization of 12-lipoxygenase mRNA in cultured oligodendrocytes and astrocytes by in situ reverse transcriptase and polymerase chain reaction. Neurosci Lett 189: 159-162. doi:https://doi.org/10.1016/0304-3940(95)11482-C. PubMed: 7542757.
  47. 47. Hambrecht GS, Adesuyi SA, Holt S, Ellis EF (1987) Brain 12-HETE formation in different species, brain regions, and in brain microvessels. Neurochem Res 12: 1029-1033. doi:https://doi.org/10.1007/BF00970932. PubMed: 3120027.
  48. 48. Iuliano L (2011) Pathways of cholesterol oxidation via non-enzymatic mechanisms. Chem Phys Lipids 164: 457-468. doi:https://doi.org/10.1016/j.chemphyslip.2011.06.006. PubMed: 21703250.
  49. 49. Niki E, Yoshida Y, Saito Y, Noguchi N (2005) Lipid peroxidation: mechanisms, inhibition, and biological effects. Biochem Biophys Res Commun 338: 668-676. doi:https://doi.org/10.1016/j.bbrc.2005.08.072. PubMed: 16126168.