Next Article in Journal
Nutritional, Antioxidant, Antimicrobial, and Anticholinesterase Properties of Phyllanthus emblica: A Study Supported by Spectroscopic and Computational Investigations
Previous Article in Journal
Sex Modifies the Impact of Type 2 Diabetes Mellitus on the Murine Whole Brain Metabolome
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Stable Isotope Dilution Analysis (SIDA) to Determine Metabolites of Furan and 2-Methylfuran in Human Urine Samples: A Pilot Study

Department of Chemistry, Division of Food Chemistry and Toxicology, RPTU Kaiserslautern-Landau, Erwin-Schrödinger-Str. 52, D-67663 Kaiserslautern, Germany
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Metabolites 2023, 13(9), 1011; https://doi.org/10.3390/metabo13091011
Submission received: 30 August 2023 / Revised: 10 September 2023 / Accepted: 11 September 2023 / Published: 14 September 2023
(This article belongs to the Section Nutrition and Metabolism)

Abstract

:
Furan and 2-methylfuran (2-MF) are food contaminants that are classified as potentially carcinogenic to humans. The main source of exposure for adults via food is coffee consumption. Furan and 2-MF are volatile, which complicates exposure assessment because their content measured in food prior to consumption does not afford a reliable dosimetry. Therefore, other ways of exposure assessment need to be developed, preferably by monitoring exposure biomarkers, e.g., selected metabolites excreted in urine. In this study, cis-2-buten-1,4-dial (BDA)-derived urinary furan metabolites Lys-BDA (l-2-amino-6-(2,5-dihydro-2-oxo-1H-pyrrol-1-yl)hexanoic acid), AcLys-BDA (l-2-(acetylamino)-6-(2,5-dihydro-2-oxo-1H-pyrrol-1-yl)hexanoic acid) and GSH-BDA (N-[4-carboxy-4-(3-mercapto-1H-pyrrol-1-yl)-1-oxobutyl]-l-cysteinyl-glycine cyclic sulfide), as well as acetyl acrolein (AcA, 2-oxo-pent-2-enal)-derived metabolites Lys-AcA (l-2-(acetylamino)-6-(2,5-dihydro-5-methyl-2-oxo-1H-pyrrol-1-yl)-hexanoic acid) and AcLys-AcA (l-2-amino-6-(2,5-dihydro-5-methyl-2-oxo-1H-pyrrol-1-yl)-hexanoic acid) and their stable isotopically labeled analogs, were synthesized and characterized through NMR and MS, and a stable isotope dilution analysis (SIDA) with UPLC-ESI-MS/MS was established. As a proof of concept, urinary samples of a four-day human intervention study were used. In the frame of this study, ten subjects ingested 500 mL of coffee containing 0.648 µmol furan and 1.059 µmol 2-MF. Among the furan metabolites, AcLys-BDA was the most abundant, followed by Lys-BDA and GSH-BDA. Exposure to 2-MF via the coffee brew led to the formation of Lys-AcA and AcLys-AcA. Within 24 h, 89.1% of the ingested amount of furan and 15.4% of the ingested amount of 2-MF were detected in the urine in the form of the investigated metabolites. Therefore, GSH-BDA, Lys-BDA, AcLys-BDA, Lys-AcA and AcLys-AcA may be suitable as short-term-exposure biomarkers of furan and 2-MF exposure.

1. Introduction

Furan and its alkylated analog 2-methylfuran (2-MF) are highly flammable aromatic-oxygen-containing heterocycles. In addition to their industrial use as basic chemicals, they are formed during combustion processes and thermal food processing. The formation of furan and 2-MF in food is complex because they may derive from compounds such as ascorbic acid, sugar, amino acids, polyunsaturated fatty acids and carotenoids [1,2,3] during food processing and cooking [4]. In particular, canned and baby food as well as coffee are high in furan and 2-MF [5]. In fact, for adults, coffee is the main source of dietary furan and 2-MF exposure, with 2-MF being the higher of the two [2]. The extent of formation of furan and 2-MF during coffee roasting depends on the roasting conditions [6]. Due to the volatility of furan, the use of a low temperature and long roasting time reduces the final furan and/or 2-MF content. In addition to process-dependent factors, storage conditions and household preparation methods of coffee beverages influence the content [7]. It has been estimated that only about 10% of the furan produced during roasting may be available for intake by coffee drinkers [6,8]. However, a medium-roast coffee brew may still contain, on average, 40 µg/kg of furan and 173 µg/kg of 2-MF [2,5]. Becalski et al. analyzed the mean concentrations of furan (38.7 µg/kg) and 2-MF (172 µg/kg) in 48 different non-brewed and brewed coffee samples [9]. They found a loss of furans of about 27–85% in brewed samples compared to non-brewed samples. Also, the concentration of 2-MF exceeded the concentration of furan in roasted coffee and the respective brews by approximately fourfold [9,10]. For risk assessment, it is important to establish valid biomarkers that reflect actual exposure as well as monitor the furan and 2-MF content in food.
Furan has been classified by the International Agency for Research on Cancer (IARC) as possibly carcinogenic to humans (group 2B) [11]. Furan and 2-MF have been shown to exert dose-dependent liver toxicity and carcinogenic effects in rodent studies [2,12]. However, it is not yet clear whether furan or 2-MF exerts a genotoxic effect, as the results of in vivo and in vitro genotoxicity tests are inconclusive [2,13]. After oral intake, furan and 2-MF are absorbed from the gastrointestinal tract. It has been reported that when radiolabeled [2,5-14C]-furan (8 mg/kg bodyweight (bw)) was administered to rats via gavage, approximately 80% of the radioactivity was eliminated within 24 h, of which 14% was exhaled as furan and 26% as CO2, and 20% was detected in urine and 22% in feces [14]. After absorption, furan is metabolized by cytochrome P450 2E1 to cis-2-buten-1,4-dial (BDA) [15].
BDA is highly reactive and reacts with cellular nucleophiles, such as proteins, amino acids, nucleosides and glutathione (GSH), as shown in Figure 1 [1]. Metabolites detected in urine have been shown to be related to BDA. A renal metabolite in the urine of furan-treated rats was identified as GSH-BDA (N-[4-carboxy-4-(3-mercapto-1H-pyrrol-1-yl)-1-oxobutyl]-l-cysteinylglycine cyclic sulfide) [16]. GSH-BDA has also been detected in rat urine through metabolite profiling together with other metabolites, such as AcLys-BDA (l-2-(acetylamino)-6-(2,5-dihydro-2-oxo-1H-pyrrol-1-yl)hexanoic acid) [17]. GSH-BDA can further react with amino acids such as Lys to form crosslinks. The degradation of these crosslinks can generate the metabolite AcCys-BDA-AcLys and its sulfoxide [18,19]. In rat urine, it was found that 0.6–1.1% of the applied furan dose was excreted renally as GSH-BDA and 1.4–2.1% as AcLys-BDA within 24 h [20].
However, no quantitative data have been reported on excretion in humans after dietary furan exposure. After furan inhalation through tobacco smoking, it was found that human acetylate Cys-BDA-Lys crosslinks mostly at the cysteine site. The corresponding sulfoxide (AcCys-BDA-Lys sulfoxide) showed a strong correlation with smoking and its level decreased after smoking cessation, revealing delayed kinetics and a potential contribution from the degradation of protein adducts [23]. These findings were confirmed through an analysis of urine samples from water pipe tobacco smokers [24].
In another study, when 14C-radiolabeled 2-MF (50–200 mg/kg bw) was administered to rats intraperitoneally (i.p.), radioactivity was detected in the liver, kidneys, lungs and blood [25]. After absorption, 2-MF is metabolized to 3-acetyl acrolein (AcA, 2-oxopent-2-enal) in rodents [2,25]. AcA is highly reactive and can be trapped with semicarbazide, but as a Michael acceptor, it can react with cellular nucleophiles, such as proteins, amino acids, nucleosides and glutathione (GSH).
A proposed metabolic route from 2-MF to Lys-AcA and AcLys-AcA is shown in Figure 2.
To date, only limited nonquantitative data have been obtained through high-resolution ToF-MS on excretion in humans after dietary 2-MF exposure [27]. A correlation between smoking and methylfuran excretion was found by analyzing the volatile compounds furan, 2-MF and 3-methylfuran in urine samples [28].
The main aims of our study were to develop a quantitative method for the analysis of furan and 2-MF biomarkers in human urine and to provide first insights into furan and 2-MF metabolism after coffee intake.

2. Materials and Methods

2.1. Materials

For use in syntheses, 2,5-dihydro-2,5-dimethoxyfuran (97%, mixture of cis and trans, DMDHF), 2,5-dihydro-2,5-dimethoxy-2-methylfuran (97%, mixture of cis and trans, MDMDHF), l-Lys (≥98%), sodium hydroxide, sodium bicarbonate, acetic acid anhydride, GSH (≥98%), [Gly-13C2,15N]-GSH trifluoroacetate (15N ≥ 98%, 13C ≥ 99%, CP ≥ 95%) and N-α-acetyllysine were purchased from Sigma-Aldrich (St. Louis, USA), whereas acetic acid (99%) was from Avantor Performance Materials B.V. (Radnor, USA) and [13C6,15N2]-Lys x 2HCl (13C6 ≥ 99%; 15N2 ≥ 99%, CP ≥ 98%) was from Carl Roth (Karlsruhe, Germany). For preparative purposes, HPLC-grade formic acid and acetonitrile were purchased from Sigma-Aldrich (St. Louis, MO, USA) and Fisher (Schwerte, Germany), respectively, whereas MS-grade solvents formic acid and acetonitrile were purchased from Carl Roth (Karlsruhe, Germany) and Merck (Darmstadt, Germany), respectively. Ammonium formate was obtained from Fisher (Schwerte, Germany); ammonium acetate, D2O, DMSO-d6, sodium acetate and ammonia were from Sigma-Aldrich (St. Louis, MO, USA); ammonium hydroxide was from Grüssing (Filsum, Germany); and citric acid was from Carl Roth (Karlsruhe, Germany).

2.2. Syntheses

2.2.1. Synthesis of l-2-Amino-6-(2,5-dihydro-2-oxo-1H-pyrrol-1-yl)hexanoic acid (Lys-BDA) and [13C6,15N2]-l-2-Amino-6-(2,5-dihydro-2-oxo-1H-pyrrol-1-yl)hexanoic acid ([13C6,15N2]-Lys-BDA)

An amount of 33.2 µL (273 µmol) of DMDHF was dissolved in 3 mL of 0.1% aqueous acetic acid and stirred for 24 h at room temperature (RT). This initial BDA formation method was based on Alves et al. [29], but acetic acid was used instead of HCl. An equimolar amount of l-Lys (40 mg, 273 µmol) or [13C6,15N2]-Lys (50 mg, 273 µmol) was dissolved in 1 mL of H2Odd, added to the BDA reaction solution and then stirred for another 24 h at RT. In the case of [13C6,15N2]-Lys, the hydrochloride was neutralized with NaOH before adding to the BDA reaction solution. Purification of the desired Lys-BDA adduct was performed in two runs through semi-preparative HPLC using a 1200 series Agilent HPLC system (Waldbronn, Germany) equipped with a Synergi 4 µm Polar-RP column (250 × 10 mm, Phenomenex, Torrance, CA, USA). In the first run, the synthesis mixture was diluted 1:5 with H2Odd and injected in aliquots of 2 mL, and a mobile phase of 0.1% aqueous formic acid (A) and acetonitrile (B) was used at a flow rate of 5 mL/min. The gradient started at 1% B and was kept isocratic for 1.5 min, and then B was increased to 30% over 13 min. Afterward, the HPLC column was rinsed with 50% B for 2.8 min to prepare for the next cycle, which was started after at least 2.8 min of equilibration under the initial solvent conditions. Fractions of 0.25 min were collected and analyzed through HPLC-ESI-MS/MS (HPLC 1100 series, Agilent, coupled to a Sciex MS API 2000 (AB SCIEX, Darmstadt, Germany)). The desired product 2-amino-6-(2-oxo-5H-pyrrol-1-yl)-hexanoic acid (Lys-BDA) eluted in the time span of 6.25–9.00 min. Pooled fractions were purified again through the HPLC method described above, but H2Odd was used as solvent A. The product eluted in a time interval of 6.00–8.25 min. After lyophilization, the yield was 21%. Characterization of Lys-BDA was performed through MS and NMR. [13C6,15N2]-Lys-BDA was only examined through MS. Purity was >95%.
l-2-Amino-6-(2,5-dihydro-2-oxo-1H-pyrrol-1-yl)hexanoic acid (M = 212.245 g/mol): 1H-NMR (400 MHz, D2O) δ [ppm]: 7.26 (dt, 1H, C3-H), 6.08 (dt, 1H, C4-H), 4.12 (bs, 2H, C5-H2), 3.64 (t, 1H, Lys α-CH), 3.41 (t, 2H, Lys ε-CH2), 1.90–1.72 (m, 2H, Lys β-CH2), 1.60 (quin, 2H, Lys δ-CH2), 1.39–1.20 (m, 2H, Lys γ-CH2); HPLC-ESIpos-MS2 of m/z 213 (CE = 30 eV): 82, 84, 132, 149, 167; tR at 9.7 min.
[13C6,15N2]-l-2-amino-6-(2,5-dihydro-2-oxo-1H-pyrrol-1-yl)hexanoic acid (M = 220.188 g/mol): HPLC-ESIpos-MS2 of m/z 221 (CE = 30 eV): 88, 90, 138, 156, 174.

2.2.2. Synthesis of l-2-(acetylamino)-6-(2,5-dihydro-2-oxo-1H-pyrrol-1-yl)hexanoic acid (Nα-AcLys-BDA) and [13C6,15N2]-l-2-(acetylamino)-6-(2,5-dihydro-2-oxo-1H-pyrrol-1-yl)hexanoic acid ([13C6,15N2]-Nα-AcLys-BDA)

Isolated Lys-BDA or [13C6,15N2]-Lys-BDA was dissolved in 0.5 mL of H2Odd and mixed with 0.5 mL of a saturated sodium bicarbonate solution. Next, a 100-fold excess of acetic anhydride was added and then the mixture was stirred for 5 min at RT. Isolation was performed using the two-step HPLC method described above. In the first run, the desired product 2-acetamido-6-(2-oxo-5H-pyrrol-1-yl)hexanoic acid (Nα-AcLys-BDA) eluted between 11.5 and 13.0 min and in the second run between 7.0 and 12.0 min. After drying, a white powder was obtained with a yield of 64%. Purity (1H-NMR) was 96%.
l-2-(acetylamino)-6-(2,5-dihydro-2-oxo-1H-pyrrol-1-yl)hexanoic acid (M = 254.282 g/mol): 1H-NMR (400 MHz, D2O) δ [ppm]: 7.23 (dt, 1H, C3-H), 6.05 (dt, 1H, C4-H), 4.17 (dd, 1H, Lys α-CH), 4.07 (s, 2H, C5-H2), 3.37 (t, 2H, Lys ε-CH2), 1.91 (s, 3H, COCH3), 1.8–1.71 (m, 1H, Lys β-CHa), 1.69–1.59 (m, 1H, Lys β-CHb), 1.57–1.51 (m, 2H, Lys δ-CH2), 1.29–1.19 (m, 2H, Lys γ-CH2); HPLC-ESIpos-MS2 of m/z 255 (CE = 30 eV): 82, 84, 167, 209, 237; tR at 13.2 min.
[13C6,15N2]-l-2-(acetylamino)-6-(2,5-dihydro-2-oxo-1H-pyrrol-1-yl)hexanoic acid: (M = 262.224 g/mol); HPLC-ESIpos-MS2 of m/z 263 (CE = 30 eV): 88, 90, 174, 216, 245.

2.2.3. Synthesis of N-[4-carboxy-4-(2-or3-mercapto-1H-pyrrol-1-yl)-1-oxobutyl]-l-cysteinylglycine cyclic sulfide (GSH-BDA) and [Gly-13C2,15N]-N-[4-carboxy-4-(2-or3-mercapto-1H-pyrrol-1-yl)-1-oxobutyl]-l-cysteinylglycine cyclic sulfide ([Gly-13C2,15N]-GSH-BDA)

As described by Alves et al. [29], to the BDA solution prepared as described above, 1 mL of H2Odd was added in which an equimolar amount of GSH or [Gly-13C2,15N]-GSH was dissolved. The reaction mixture was stirred for 24 h at RT. After 1:2 dilution with H2Odd, 2 mL of the solution was injected into the preparative HPLC system described above. Using a flow rate of 5 mL/min, the gradient started at 10% B, was kept isocratic for 1 min and was then raised to 40% B over 8 min. Afterward, the column was rinsed with 90% B for 5 min and equilibrated for an additional 5 min under the initial conditions. In the first cycle, 0.1% aqueous formic acid was used as solvent A, whereas in the second cycle, it was changed to H2Odd. The desired compounds N-[4-carboxy-4-(3-mercapto-1H-pyrrol-1-yl)-1-oxobutyl]-l-cysteinylglycine cyclic sulfide and N-[4-carboxy-4-(2-mercapto-1H-pyrrol-1-yl)-1-oxobutyl]-l-cysteinylglycine cyclic sulfide (BDA-GSH) eluted between 5.0 and 5.5 min. After freeze-drying, fractions containing a 1:1 mixture of both isomers were obtained with a yield of 2%. Purity was >95%.
BDA-GSH (M = 355.366 g/mol): 1H NMR (400 MHz, D2O) δ [ppm]: 6.89 (s, 1 H, pyrrole C5-Hα), 6.76 (s, 1 H, pyrrole C5-Hb), 6.71 (s, 1 H, pyrrole C2-H), 6.60 (s, 1 H, pyrrole C3-H), 6.27 (s, 1 H, pyrrole C4-Hb), 6.10 (s, 1 H, pyrrole C4-Hα), 3.89–3.75 (m, 4 H, Gly CH), 3.35–3.25 (m, 1 H, Cys β-CH2A), 2.9 (d, 1 H, Cys β-CH2B), 2.86–2.70 (m, 2 H, Cys β-CH1), 2.50–1.85 (m, 8 H, Glu β-CH2 + Glu γ-CH2); HPLC-ESIneg-MS2 of m/z 354 (CE = −31 eV): 98, 124, 141, 185. tR at 12.7 min.
[13C2,15N]-BDA-GSH (M = 358.344 g/mol): HPLC-ESIneg-MS2 of m/z 357 (CE = −31 eV): 98, 124, 143, 188.

2.2.4. Synthesis of l-2-Amino-6-(2,5-dihydro-5-methyl-2-oxo-1H-pyrrol-1-yl)hexanoic acid (Lys-AcA) and [13C6,15N2]-l-2-Amino-6-(2,5-dihydro-5-methyl-2-oxo-1H-pyrrol-1-yl)hexanoic acid ([13C6,15N2]-Lys-AcA)

An amount of 39.5 µL (273 µmol) of MDMDHF was dissolved in 3 mL of 0.1% aqueous acetic acid and stirred for 1 h at RT to obtain a 3-acetyl acrolein (AcA) reaction solution, similarly to the formation of the BDA reaction solution. An equimolar amount of l-Lys (40 mg, 273 µmol) or [13C6,15N2]-Lys (50 mg, 273 µmol) was dissolved in 1 mL of H2Odd, added to the AcA reaction solution and then stirred for another 24 h at RT. In the case of [13C6,15N2]-Lys, the hydrochloride was neutralized with NaOH before adding to the AcA reaction solution. Purification of the desired Lys-AcA adduct was performed in two runs through semi-preparative HPLC using an Agilent 1200 series HPLC system (Waldbronn, Germany) equipped with a Synergi 4 µm Polar-RP column (250 × 10 mm, Phenomenex, Torrance, USA). In the first run, the synthesis mixture was diluted 1:2 with H2Odd and injected in aliquots of 2 mL using a mobile phase of 0.1% aqueous formic acid (A) and acetonitrile (B) at a flow rate of 6 mL/min. The gradient started at 1% B and was kept isocratic for 2.5 min, and then the percentage of B was increased to 8% over 7 min. Afterward, the HPLC column was rinsed with 80% B for 4.0 min to prepare for the next cycle, which was started after at least 5.0 min of equilibration under the initial solvent conditions. Fractions of 0.5 min were collected and analyzed through HPLC-ESI-MS/MS (HPLC 1100 series, Agilent, coupled to a Sciex MS API 2000 (AB SCIEX, Darmstadt, Germany)). The desired product l-2-amino-6-(2,5-dihydro-5-methyl-2-oxo-1H-pyrrol-1-yl)-hexanoic acid (Lys-AcA) eluted in a time span of 4.5–5.5 min. Pooled fractions were purified again via the HPLC method described above, but H2Odd was used as solvent A. The product eluted in a time interval of 4.5–8.0 min. After lyophilization, the yield was 18%. Characterization of Lys-AcA was performed through MS and NMR. [13C6,15N2]-Lys-AcA was only examined through MS. Purity was > 95%.
l-2-Amino-6-(2,5-dihydro-5-methyl-2-oxo-1H-pyrrol-1-yl)hexanoic acid (M = 226.272 g/mol): 1H NMR (600 MHz, D2O) δ [ppm]: 3.62 (m, 1H, C5-H), 6.08 (dt, 1H, C4-H), 4.12 (bs, 2H, C5-H2), 3.15 (m, 1H, Lys α-CH), 2.45–2.33 (m, 2H, Lys ε-CH2), 2.17–2.00 (m, 2H, Lys β-CH2), 1.78–1.50 (m, 2H, Lys δ-CH2), 1.53–1.31 (m, 2H, Lys γ-CH2), 1.43 (d, 3H, CH3); HPLC-ESIpos-MS2 of m/z 227 (CE = 30 eV): 82 (69%), 84 (100%), 130 (19%), 163 (14%), 181 (25%). tR at 8.8 min.
[13C6,15N2]-l-2-amino-6-(2,5-dihydro-5-methyl-2-oxo-1H-pyrrol-1-yl)hexanoic acid (M = 233.215 g/mol): HPLC-ESIpos-MS2 of m/z 235 (CE = 25 eV): 88, 90, 130, 170, 188.

2.2.5. Synthesis of l-2-(acetylamino)-6-(2,5-dihydro-2-oxo-5-methyl-1H-pyrrol-1-yl)hexanoic acid (AcLys-AcA) and [13C6,15N2]-l-2-(acetylamino)-6-(2,5-dihydro-2-oxo-5-methyl-1H-pyrrol-1-yl)hexanoic acid ([13C6,15N2]-AcLys-AcA)

As described in Alves et al. [29] but with modifications, the isolated Lys-AcA or [13C6,15N2]-Lys-AcA was dissolved in 1 mL of H2Odd and mixed with 1 mL of a saturated sodium bicarbonate solution. Next, a 100-fold excess of acetic anhydride was added, and the mixture was stirred for 5 min at RT. Isolation was performed using the two-step HPLC method described above. Purification of the desired Lys-AcA adduct was performed in two runs through semi-preparative HPLC using an Agilent 1200 series HPLC system (Waldbronn, Germany) equipped with a Synergi 4 µm Polar-RP column (250 × 10 mm, Phenomenex, Torrance, CA, USA). For the first run, the synthesis mixture was diluted 1:2 with H2Odd and injected in aliquots of 2 mL using a mobile phase of 0.1% aqueous formic acid (A) and acetonitrile (B) at a flow rate of 6 mL/min. The gradient started at 1% B and was kept isocratic for 1.5 min, and the percentage of B was increased to 11% over 7 min and then to 20% B over 3 min. Afterward, the HPLC column was rinsed with 80% B for 2.5 min to prepare for the next cycle, which was started after at least 5.0 min of equilibration under the initial solvent conditions. Fractions of 0.5 min were collected and analyzed through HPLC-ESI-MS/MS (HPLC 1100 series, Agilent, coupled to a Sciex MS API 2000 (AB SCIEX, Darmstadt, Germany)). Pooled fractions were purified again via the HPLC method described above, but H2Odd was used as solvent A. The product eluted between 9.0 and 12.0 min. In the first run, the desired product l-2-acetylamino-6-(2,5-dihydro-2-oxo-5-methyl-5H-pyrrol-1-yl)hexanoic acid (AcLys-AcA) eluted between 9.5 and 11.5 min and, in the second run, between 8.0 and 12.0 min. After drying, a white powder was obtained with a yield of 68%. Purity (1H-NMR) of > 98%.
l-2-(acetylamino)-6-(2,5-dihydro-2-oxo-5-methyl-1H-pyrrol-1-yl)hexanoic acid (M = 268.308 g/mol): 1H-NMR (600 MHz, D2O): δ[ppm] 4.29 (dd, J = 8.7, 4.8, 1H) 3.26–3.17 (m, 2H), 3.15 (t, J = 6.7 Hz, 0H), 2.83 (t, J = 6.6 Hz, 0H), 2.56–2.48 (m, 1H), 2.45–2.36 (m, 1H), 2.26–2.20 (m, 1H), 2.19 (s, 0H), 2.12–2.04 (m, 1H), 2.01 (s, 3H), 1.91–1.81 (m, 1H), 1.78–1.69 (m, 1H), 1.63–1.54 (m, 2H), 1.50 (d, J = 10.0 Hz, 2H), 1.44–1.31 (m, 2H); HPLC-ESIpos-MS2 of m/z 269 (CE = 25 eV): HPLC-ESIpos-MS2 of m/z 269 (CE = 25 eV): 82 (26%), 84 (23%), 164 (100%), 181 (31%), 223 (11%), 227 (12%); tR at 12.0 min.
[13C6,15N2]-l-2-(acetylamino)-6-(2,5-dihydro-5-methyl-2-oxo-1H-pyrrol-1-yl)hexanoic acid: (M = 276.251 g/mol); HPLC-ESIpos-MS2 of m/z 277 (CE = 25 eV): 88, 90, 171, 188, 230, 234.

2.3. Stability Study

To explore the influence of pH on the storage stability of the metabolites detected in the urine samples, a fresh urine sample was collected from a volunteer 3 to 4 h after they had consumed 500 mL of freshly brewed coffee. The urine sample was divided into 4 mL aliquots, and each was mixed with either 4 mL of 100 mM ammonium formate buffer at pH 10 or 4 mL of 100 mM ammonium acetate buffer at pH 5.4. The adjusted aliquots were stored at −20 °C and analyzed on the same day as collection and after 4, 8, 12, 24, 36, 48 and 60 weeks.

2.4. Human Intervention Study

This study was originally conducted to investigate the excretion of niacin and alkylpyrazine metabolites in 2015 and has been described in detail elsewhere [30,31]. However, due to the appropriate study design, it was also used as a pilot study for the investigation of renal excretion of furan and 2-MF metabolites after coffee intake. The human intervention study was approved by the ethical commission of the Rhineland-Palatinate medical commission in Mainz (no. 837.427.15(10195)) and performed in accordance with the ethical standards of the Declaration of Helsinki (1964).
Ten healthy, nonsmoking volunteers of Caucasian origin (six females and four males; age 22–27 years, BMI 22.6 ± 2.6 kg m−2) who were accustomed to drinking coffee on a daily basis participated in this study and gave written consent before joining this study. Other inclusion criteria were as follows: age between 20 and 50 years, BMI between 19 and 25 kg/m2, not taking medication and/or dietary supplements, not pregnant, not performing competitive sports, not participating in other studies or a regular blood donor. This study lasted four days, during which the participants received standardized meals prepared without canned foods and heated exclusively in open pots to minimize heat induced contaminant intake. For each participant, a first overnight urine sample was collected during the last 12 h of a two-day run-in period, followed by a spot urine sample in the morning prior to coffee consumption. Subsequently, each participant consumed 500 mL of freshly brewed coffee prepared from 30 g of medium–dark-roasted Arabica coffee (provided by Tchibo GmbH, Hamburg, Germany) using a SENSEO pad coffee maker (Philips, Amsterdam, Netherlands). An aliquot of the coffee brew was stored at −20 °C and then analyzed using GC-MS to determine its furan (44.11 µg, 0.648 µmol per 500 mL coffee) and 2-MF (86.94 µg, 1.059 µmol per 500 mL coffee) content. After coffee consumption, total urine was collected over a 36 h period divided into ten collection intervals: 0–1 h, 1–2 h, 2–3 h, 3–4 h, 4–5 h, 5–6 h, 6–8 h, 8–12 h, 12–24 h and 24–36 h. Volumes of urine were calculated by weighing the samples and using an assumed density of 1 kg/L. Aliquots were stored at −20 °C until sample preparation.

2.5. Sample Preparation

An internal standard working solution was prepared in H2Odd containing 1.0 µM [13C6,15N2]-Lys-BDA, 2.5 µM [13C6,15N2]-AcLys-BDA, 1.0 µM [Gly-13C2,15N]-GSH-BDA, 2.5 µM [13C6,15N2]-Lys-AcA and 1.0 µM [13C6,15N2]-AcLys-AcA and was stored at −20 °C. Urine samples from the human intervention study were thawed at RT. An amount of 1 mL of each sample was spiked with 20 µL of the internal standard working solution. After acidification with 50 µL of formic acid, the samples were mixed well and centrifuged (15.000× g, 5 min). Afterward, the supernatants were transferred into brown glass vessels through a syringe filter (0.22 µm). Aliquots of 2 µL were analyzed through the HPLC-ESI-MS/MS method described below.

2.6. UPLC-ESI-MS/MS

An Agilent 1290 series HPLC system (Waldbronn, Germany) was used coupled to a QTRAP 5500 mass spectrometer (AB SCIEX, Darmstadt, Germany) equipped with an electrospray (ESI) source. The metabolites Lys-BDA, AcLys-BDA, GSH-BDA, Lys-AcA and AcLys-AcA and their stable-isotope-labeled analogs used as standards were separated using a Synergi 2.5 µm Polar-RP column (100 × 2 mm, Phenomenex, Torrance, CA, USA) equipped with an appropriate guard column at 50 °C. The mobile phase comprised 0.1% aqueous formic acid (A) and 0.1% formic acid in acetonitrile (B) at a flow rate of 400 μL/min. The content of B in the eluent was initially held for 0.2 min at 1% and then was raised to 7% over 0.8 min and held for 1.9 min. Afterward, the content of B was increased to 35% over 0.3 min. The final conditions were held for 0.7 min and then the column was rinsed with 95% B for 1.4 min and equilibrated for 2.5 min with 1% B. Due to different ionization conditions, two different MRM periods in the positive mode were used. Table 1 shows the MS parameters for all analyzed compounds, whereas Table 2 shows LOD and LOQ, which were defined as 1:3 and 1:10 of the signal-to-noise ratio, respectively [32].
On this basis, standard solutions for calibration were prepared in H2Odd with concentration ranges of 5–250 nM for Lys-BDA, 10–500 nM for AcLys-BDA, 1–50 nM for GSH-BDA, 5–250 nM for Lys-AcA and 2–100 nM for AcLys-ACA containing 20 nM [13C6,15N2]-Lys-BDA, 50 nM [13C6,15N2]-AcLys-BDA, 20 nM [Gly-13C2,15N]-GSH-BDA, 50 nM [13C6,15N2]-Lys-AcA and 20 nM [13C6,15N2]-AcLys-AcA. For calibration, the observed peak area ratios were plotted against the concentration ratios, resulting in a correlation coefficient of R2 > 0.99. The method’s precision was determined by repeatedly analyzing standards at LOQ, at the highest concentration used for calibration and at the midway point between these two values (Table 2). The intra-day coefficients of variation were ≤4.1% for the intermediate and high concentrations, and ≤5.7% for LOQ. The corresponding inter-day coefficients of variation were ≤5.1% based on analyses performed on five consecutive days. Using a spiked urine sample, the robustness was shown to be below 8.4% for all analytes under study. Recoveries were between 99 and 103%.

2.7. Data Analysis/Statistics

Each urine sample was analyzed in triplicate. The HPLC-ESI-MS/MS system was operated using the Analyst 1.7.2 software package (AB Sciex, Darmstadt, Germany), and chromatographic data were processed using Multiquant 3.0.3 (AB Sciex, Darmstadt, Germany). Quantitative data for the urine samples are reported as the mean ± standard deviation (SD) for all ten participants. Two-sample t-tests were performed using Origin, Version 2023 (OriginLab Corporation, Northampton, MA, USA); statistical significance was determined using p-value thresholds of p ≤ 0.05, p ≤ 0.01 and p ≤ 0.001.

3. Results

In a previous untargeted approach [27], urine samples were analyzed and for the first time shown to contain metabolites of 2-MF, i.e., AcLys-AcA, Lys-AcA and AcCys-AcA, and additionally acetylcysteine butendial lysine (AcCys-BDA-Lys), acetylcysteine-SO-butendial acetyllysine (AcCys-(SO)-BDA-AcLys), methanethiol butendial glutamic acid (methanethiol-BDA-Glu), glutathione butendial lysine (GSH-BDA-Lys) and glutathione butendial glutamine (GSH-BDA-Gln). These metabolites were only identified through high-resolution MS but could not be quantified. In the present study, urine samples were analyzed, with the most abundant peaks identified as Lys-BDA, AcLys-BDA and GSH-BDA as metabolites of furan and AcLys-AcA, Lys-AcA and AcCys-AcA as metabolites of 2-MF using UPLC-ESI-MS/MS. The proposed formation mechanisms are shown in Figure 1 and Figure 2. Afterward, these analytes and their isotopically labeled analogs were synthesized, and a robust UPLC-ESI-MS/MS method was established. Other metabolites identified by Stegmüller et al. [27] using high resolution could not be clearly confirmed by the respective synthesis of standards, e.g., AcCys-BDA-Lys, and due to occurring mass defects, a clear identification was not possible. Urine samples collected during the human intervention study were assayed using the method described above and analytes were quantified through SIDA in the MRM mode (see experimental section). To accurately determine the amounts of furan metabolites Lys-BDA, AcLys-BDA and GSH-BDA as well as 2-MF metabolites Lys-AcA and AcLys-AcA in human urine samples after coffee intake, a stable isotope dilution analysis (SIDA) with HPLC-ESI-MS/MS detection was developed. Therefore, reference material for analytes that were not commercially available, as well as the corresponding stable-isotope-labeled analogs, was synthesized as internal standards for the SIDA.

3.1. Synthesis

The most pronounced 2-MF metabolites AcLys-AcA and Lys-AcA as well as the isotopically labeled analogs were synthesized in accordance with the method already established for the synthesis of AcLys-BDA, Lys-BDA and GSH-BDA published by Alves et al. [29] with modifications.
According to protocols in the literature, BDA can be synthesized from furan using dimethyldioxirane [33] and 2,5-diacetoxy-2,5-dihydrofuran (DMDHF) [34]. However, for the syntheses in this work, BDA was generated from DMDHF under acidic conditions [29]. The reaction of BDA with amino groups can yield both 3-pyrrolin-2-one and 4-pyrrolin-2-one derivatives. In preliminary experiments using 2,5-diacetoxy-2,5-dihydrofuran and Lys, these two isomers were formed at each of the two amino groups of Lys (data not shown). The elution order of the four isomers on the HPLC column (Synergi 4 µm Polar-RP, Phenomenex, Torrance, CA, USA) was 4-pyrrolin-2-one-α-Lys, 4-pyrrolin-2-one-ε-Lys, 3-pyrrolin-2-one-α-Lys and 3-pyrrolin-2-one-ε-Lys. In addition, other products whose mass spectra suggested BDA-Lys-BDA conjugates were observed during the synthesis. In contrast, the reaction of DMDHF produced only the desired 3-pyrrolin-2-one derivatives in a ratio of about 8:2 (ε-Lys/α-Lys). In the presence of strong acids and bases, a tautomeric equilibrium of the pyrrolin-2-one-derivatives was previously observed [35]. The opening of the pyrrolin-2-one ring structure under basic conditions has already been discussed by Alves et al. as a possible mechanism [36]. This is probably the reason why lyophilization under acidic conditions led to rearrangement of the desired compound and necessitated a solvent change after purification (first preparative HPLC run). A tautomeric rearrangement also occurs during acetylation. Hence, the isomers were purified again. The measured 1H-NMR spectrum of AcLys-BDA is in accordance with the literature data [37]. During the synthesis and purification of GSH-BDA, it was not possible to separate the two isomers (7-[(carboxymethyl)carbamoyl]-5-oxo-9-thia-1,6-diazabicyclo [8.2.1]trideca-10(13),11-diene-2-carboxylic acid and 3-[(carboxymethyl)carbamoyl]-5-oxo-3,4,5,6,7,8-hexahydro-2H-pyrrolo [2,1-b][1,3,8]thiadiazecine-8-carboxylic acid). According to NMR spectra, the isomers were present in a 1:1 mixture and the data are consistent with the literature [21]. Since the isomers could not be distinguished by either chromatography or mass spectrometry, it was not possible to clarify the ratio in which the two isomers were metabolically formed. The syntheses of the corresponding isotopically labeled compounds were analogous and confirmed through comparison of mass spectrometric data.

3.2. Urine Sample Analysis

A representative chromatogram is shown in Figure 3. During sample preparation, only isotopically labeled standards were added, and the samples were vortexed, acidified and filtered prior to UPLC-ESI-MS/MS measurement to avoid spurious results.
All three metabolites of furan (AcLys-BDA, Lys-BDA and GSH-BDA) were present in urine samples collected during the run-in period as well as after coffee consumption. The most abundant analyte was AcLys-BDA, followed by Lys-BDA and GSH-BDA. All three metabolites showed similar kinetics, as shown in Figure 4. For Lys-BDA, a slight but non-significant increase was detected in the first hour after coffee consumption. During the urine collection periods 4–5 h, 5–6 h, 6–7 h and 8–12 h, the excretion rate was significantly higher than the background value prior coffee ingestion, with a maximum at 6–8 h (5.4 ± 1.7 nmol/h). Within the first 24 h, 10.6 ± 4.4% of the ingested amount of furan was excreted as Lys-BDA. In contrast, the excretion rate of AcLys-BDA reached its maximum already after 1 h (44.3 ± 14.7 nmol/h), which was also highly significant compared to the background rate. A second maximum was detected in the period of 6–8 h (33.4 ± 8.7 nmol/h), which was comparable to the other two analytes. In total, the participants excreted 78 ± 18% (0.505 ± 0.117 µmol) of the ingested furan as AcLys-BDA in urine in the 24 h following intake. In furan-treated rats, the excretion of 1.4–2.1% of the dose was detected as AcLys-BDA within 24 h [20]. This seems low, but rats have been shown to only excrete 20% of the furan dose via urine [14]. The trend in kinetics of the metabolite GSH-BDA was very similar to that of Lys-BDA. There was a non-significant increase after 1 h (0.22 ± 0.18 nmol/h) and a significant maximum during the urine collection period of 6-8 h (0.18 ± 0.09 nmol/h). Overall, the participants excreted 89.1 ± 21.1% (0.577 ± 0.137 µmol) of the ingested furan in the urine within 24 h.
For 2-MF, the two metabolites Lys-AcA and AcLys-AcA were monitored after coffee consumption, which both peaked already 1 h after ingestion. As shown in Figure 5, the metabolites showed similar kinetics. The maximum during 0–1 h for Lys-AcA was 8.0 ± 2.4 nmol and, for AcLys-AcA, 11.7 ± 5.4 nmol/h. Overall, 8.3 ± 2.4% (0.088 ± 0.025 µmol) of the ingested amount of 2-MF was excreted as Lys-AcA and 7.1 ± 2.8% (0.075 ± 0.030 µmol) as AcLys-AcA. In total, 15.4 ± 4.8% (0.163 ± 0.051 µmol) of the ingested 2-MF was excreted in the form of the metabolites under study. Interestingly, for the 2-MF metabolites, no second increase was observed in the kinetics, in contrast to the furan metabolites, which showed a slight second increase after 4 h. To the best of our knowledge, the kinetics of 2-MF metabolites in humans have not yet been reported. However, in 1985, Ravindranath et al. showed that AcA might be formed when scavenging agents such as N-acetyl cysteine (NAC), semicarbazide and cysteine are used [25]. Our study provides the first quantitative data on 2-MF metabolites, revealing similarities to the metabolism of furan.
At the time of analysis, the samples were more than two years old and had been frozen unbuffered. The pH of the collected samples was not measured, but the pH of human urine typically ranges from 4.5 to 8.0 [38]. To verify whether pH-sensitive metabolites were degraded during storage and to optimize sample storage for further studies, the stability at different pH values was investigated over a period of 60 weeks (Figure 6). Storage at pH 10 resulted in significant losses of Lys-BDA and AcLys-BDA in the urine samples after only 4 weeks, and their levels continued to decrease thereafter. A possible mechanism for rearrangement of the pyrrolin-2-one ring structure has been discussed previously [36]. In contrast, storage at a slightly acidic pH of 5.4 did not result in significant reductions. Since the latter is a typical pH of natural human urine, it can be assumed that no significant losses occurred during the storage period. These results demonstrate the importance of controlling the pH during storage.

4. Discussion

Our pilot study had several limitations as this study was designed to determine niacin in urine after coffee consumption. The focus was on niacin and afterward on pyrazin metabolism in humans. As a first step, this study was used to identify markers of furan and alkylfuran metabolism. When interpreting metabolization rates, these facts have to be considered. For example, the subjects excreted furan and 2-MF metabolites above the baseline excretion even on day 4 without coffee consumption in the days before. At the end of the run-in phase, excretion of the metabolites was clearly detectable. This can probably be explained by the study design; although the subjects had a controlled diet that was low in heated foods to minimize levels of furan and 2-MF, they were not given any behavioral guidelines or restrictions. Thus, it is possible that exposure to furan and 2-MF may have occurred, e.g., through passive smoking, living and/or walking on a busy street, open fires or other unknown sources. This may also be the reason for interindividual variations. Grill et al. previously detected metabolites Lys-BDA and AcLys-BDA in the urine of nonsmokers [23]. Our results suggest that the furan metabolite AcLys-BDA and 2-MF metabolite AcLys-AcA may be suitable as biomarkers for short-term exposure to diet-associated levels of furan and 2-MF. Their significant increase in excretion immediately after the consumption of 500 mL of the coffee beverage may be indicative of the rapid detoxification of BDA and AcA via conjugation. In a recent study examining urine samples of waterpipe tobacco smokers in the morning before and after smoking, no significant differences were detected in the levels of six BDA metabolites, but levels of Lys-BDA and AcLys-BDA were higher after smoking than before [24]. This suggests that the latter metabolites are generated during smoking (Cys-BDA-Lys crosslinks and their derivatives), particularly as they were also found to be higher in the urine of smokers than of nonsmokers. Therefore, they may be suitable as biomarkers of long-term exposure because they could be released from proteins. In contrast, in our study, the increased excretion of AcLys-BDA immediately after coffee consumption suggests that it could be used as a biomarker for short-term exposure and originates from the reaction with free amino acids. However, further research is needed to clarify these findings [20]. Interestingly, in addition to the metabolites GSH-BDA and AcLys-BDA reported in rat urine samples after ingestion of furan [20], we identified and quantitated Lys-BDA.

5. Conclusions

This study provides the first insights into human metabolism and the urinary excretion of diet-related furan and 2-MF consumed in coffee. Despite this study’s limitations, AcLys-BDA, AcLys-AcA and Lys-AcA might have the potential to be viable biomarkers. However, there is an abundance of possible metabolites, as recently demonstrated in an untargeted analysis [27]. Therefore, further research is needed to identify and quantify additional metabolites of furan and 2-MF, e.g., dimethylfuran formed during food processing, and to perform human intervention studies considering, e.g., the influence of the food matrix on furan/2-MF bioavailability.

Author Contributions

E.R. supervised the study and designed the human intervention study; E.R. and T.B. planned and carried out the human intervention study; D.K., C.K., J.R., H.T., L.L., V.K. and J.I.K. synthesized and characterized the analytical and isotopically labeled standard substances; H.B., D.B., D.K. and J.I.K. developed the sample preparation; J.I.K., D.B., S.S., D.B., V.K. and D.K. developed the analytical method and performed the measurements; D.K. performed the stability studies; E.R., D.K., S.S. and J.I.K. evaluated the data statistically and wrote the manuscript; all the authors critically reviewed the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was partly funded by the Deutsche Forschungsgemeinschaft (DFG) project numbers RI 1176/12-1 and RI 1176/13-1.

Institutional Review Board Statement

The study was conducted in accordance with the Declaration of Helsinki and approved by the ethical commission of the Rhineland-Palatinate medical commission in Mainz (no. 837.427.15(10195)).

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author. Data available on request due to restrictions e.g., privacy or ethical.

Acknowledgments

We gratefully acknowledge Gerhard Eisenbrand for helpful discussions and suggestions. We thank the Tchibo GmbH for providing the coffee pad samples and the SENSEO pad machine.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Moro, S.; Chipman, J.K.; Wegener, J.-W.; Hamberger, C.; Dekant, W.; Mally, A. Furan in heat-treated foods: Formation, exposure, toxicity, and aspects of risk assessment. Mol. Nutr. Food Res. 2012, 56, 1197–1211. [Google Scholar] [CrossRef]
  2. Knutsen, H.K.; Alexander, J.; Barregård, L.; Bignami, M.; Brüschweiler, B.; Ceccatelli, S.; Cottrill, B.; Dinovi, M.; Edler, L.; Grasl-Kraupp, B.; et al. Risks for public health related to the presence of furan and methylfurans in food. EFSA J. 2017, 15, e05005. [Google Scholar] [CrossRef] [PubMed]
  3. Perez Locas, C.; Yaylayan, V.A. Origin and mechanistic pathways of formation of the parent furan—A food toxicant. J. Agric. Food Chem. 2004, 52, 6830–6836. [Google Scholar] [CrossRef] [PubMed]
  4. Crews, C.; Castle, L. A review of the occurrence, formation and analysis of furan in heat-processed foods. Trends Food Sci. Technol. 2007, 18, 365–372. [Google Scholar] [CrossRef]
  5. EFSA. Update on furan levels in food from monitoring years 2004–2010 and exposure assessment. EFSA J. 2011, 9, 2347. [Google Scholar] [CrossRef]
  6. Guenther, H.; Hoenicke, K.; Biesterveld, S.; Gerhard-Rieben, E.; Lantz, I. Furan in coffee: Pilot studies on formation during roasting and losses during production steps and consumer handling. Food Addit. Contam. Part A Chem. Anal. Control Expo. Risk Assess. 2010, 27, 283–290. [Google Scholar] [CrossRef] [PubMed]
  7. Rahn, A.; Yeretzian, C. Impact of consumer behavior on furan and furan-derivative exposure during coffee consumption. A comparison between brewing methods and drinking preferences. Food Chem. 2019, 272, 514–522. [Google Scholar] [CrossRef]
  8. Altaki, M.S.; Santos, F.J.; Galceran, M.T. Occurrence of furan in coffee from Spanish market: Contribution of brewing and roasting. Food Chem. 2011, 126, 1527–1532. [Google Scholar] [CrossRef]
  9. Becalski, A.; Hayward, S.; Krakalovich, T.; Pelletier, L.; Roscoe, V.; Vavasour, E. Development of an analytical method and survey of foods for furan, 2-methylfuran and 3-methylfuran with estimated exposure. Food Addit. Contam. Part A Chem. Anal. Control Expo. Risk Assess. 2010, 27, 764–775. [Google Scholar] [CrossRef]
  10. Becalski, A.; Halldorson, T.; Hayward, S.; Roscoe, V. Furan, 2-methylfuran and 3-methylfuran in coffee on the Canadian market. J. Food Compos. Anal. 2016, 47, 113–119. [Google Scholar] [CrossRef]
  11. IARC. Dry cleaning, some chlorinated solvents and other industrial chemicals. In IARC Monographs on the Evaluation of Carcinogenic Risks to Humans; IARC: Lyon, France, 1995; pp. 393–407. [Google Scholar]
  12. National Toxicology Program. Toxicology and Carcinogenesis Studies of Furan (CAS No. 110-00-9) in F344 Rats and B6C3F1 Mice (Gavage Studies). Natl. Toxicol. Program Tech. Rep. Ser. 1993, 402, 1–286. [Google Scholar]
  13. Rietjens, I.M.C.M.; Dussort, P.; Günther, H.; Hanlon, P.; Honda, H.; Mally, A.; O’Hagan, S.; Scholz, G.; Seidel, A.; Swenberg, J.; et al. Exposure assessment of process-related contaminants in food by biomarker monitoring. Arch. Toxicol. 2018, 92, 15–40. [Google Scholar] [CrossRef] [PubMed]
  14. Burka, L.T.; Washburn, K.D.; Irwin, R.D. Disposition of 14Cfuran in the male F344 rat. J. Toxicol. Environ. Health 1991, 34, 245–257. [Google Scholar] [CrossRef]
  15. Chen, L.J.; Hecht, S.S.; Peterson, L.A. Identification of cis-2-butene-1,4-dial as a microsomal metabolite of furan. Chem. Res. Toxicol. 1995, 8, 903–906. [Google Scholar] [CrossRef] [PubMed]
  16. Peterson, L.A.; Cummings, M.E.; Chan, J.Y.; Vu, C.C.; Matter, B.A. Identification of a cis-2-butene-1,4-dial-derived glutathione conjugate in the urine of furan-treated rats. Chem. Res. Toxicol. 2006, 19, 1138–1141. [Google Scholar] [CrossRef] [PubMed]
  17. Kellert, M.; Wagner, S.; Lutz, U.; Lutz, W.K. Biomarkers of furan exposure by metabolic profiling of rat urine with liquid chromatography-tandem mass spectrometry and principal component analysis. Chem. Res. Toxicol. 2008, 21, 761–768. [Google Scholar] [CrossRef] [PubMed]
  18. Lu, D.; Sullivan, M.M.; Phillips, M.B.; Peterson, L.A. Degraded protein adducts of cis-2-butene-1,4-dial are urinary and hepatocyte metabolites of furan. Chem. Res. Toxicol. 2009, 22, 997–1007. [Google Scholar] [CrossRef] [PubMed]
  19. Lu, D.; Peterson, L.A. Identification of furan metabolites derived from cysteine-cis-2-butene-1,4-dial-lysine cross-links. Chem. Res. Toxicol. 2010, 23, 142–151. [Google Scholar] [CrossRef]
  20. Karlstetter, D.; Mally, A. Biomonitoring of heat-induced food contaminants: Quantitative analysis of furan dependent glutathione- and lysine-adducts in rat urine as putative biomarkers of exposure. Food Chem. Toxicol. 2020, 143, 111562. [Google Scholar] [CrossRef]
  21. Peterson, L.A.; Cummings, M.E.; Vu, C.C.; Matter, B.A. Glutathione trapping to measure microsomal oxidation of furan to cis-2-butene-1,4-dial. Drug Metab. Dispos. 2005, 33, 1453–1458. [Google Scholar] [CrossRef]
  22. Hamberger, C.; Kellert, M.; Schauer, U.M.; Dekant, W.; Mally, A. Hepatobiliary toxicity of furan: Identification of furan metabolites in bile of male f344/n rats. Drug Metab. Dispos. 2010, 38, 1698–1706. [Google Scholar] [CrossRef]
  23. Grill, A.E.; Schmitt, T.; Gates, L.A.; Lu, D.; Bandyopadhyay, D.; Yuan, J.-M.; Murphy, S.E.; Peterson, L.A. Abundant rodent furan-derived urinary metabolites are associated with tobacco smoke exposure in humans. Chem. Res. Toxicol. 2015, 28, 1508–1516. [Google Scholar] [CrossRef] [PubMed]
  24. Kassem, N.O.F.; Peterson, L.A.; Liles, S.; Kassem, N.O.; Zaki, F.K.; Lui, K.-J.; Vevang, K.R.; Dodder, N.G.; Hoh, E.; Hovell, M.F. Urinary metabolites of furan in waterpipe tobacco smokers compared to non-smokers in home settings in the US. Toxicol. Lett. 2020, 333, 202–210. [Google Scholar] [CrossRef] [PubMed]
  25. Ravindranath, V.; McMenamin, M.G.; Dees, J.H.; Boyd, M.R. 2-Methylfuran toxicity in rats--role of metabolic activation in vivo. Toxicol. Appl. Pharmacol. 1986, 85, 78–91. [Google Scholar] [CrossRef] [PubMed]
  26. Ravindranath, V.; Burka, L.T.; Boyd, M.R. Reactive metabolites from the bioactivation of toxic methylfurans. Science 1984, 224, 884–886. [Google Scholar] [CrossRef] [PubMed]
  27. Stegmüller, S.; Beißmann, N.; Kremer, J.I.; Mehl, D.; Baumann, C.; Richling, E. A New UPLC-qTOF approach for elucidating furan and 2-methylfuran metabolites in human urine samples after coffee consumption. Molecules 2020, 25, 5104. [Google Scholar] [CrossRef]
  28. Mochalski, P.; Unterkofler, K. Quantification of selected volatile organic compounds in human urine by gas chromatography selective reagent ionization time of flight mass spectrometry (GC-SRI-TOF-MS) coupled with head-space solid-phase microextraction (HS-SPME). Analyst 2016, 141, 4796–4803. [Google Scholar] [CrossRef]
  29. Alves, J.C.F. Preliminary studies towards the preparation of reactive 3-pyrrolin-2-ones in conjugate addition reactions for the syntheses of potentially bioactive 2-pyrrolidinones and pyrrolidines. J. Braz. Chem. Soc. 2007, 18, 855–859. [Google Scholar] [CrossRef]
  30. Kremer, J.I.; Pickard, S.; Stadlmair, L.F.; Glaß-Theis, A.; Buckel, L.; Bakuradze, T.; Eisenbrand, G.; Richling, E. Alkylpyrazines from coffee are extensively metabolized to pyrazine carboxylic acids in the human body. Mol. Nutr. Food Res. 2019, 63, e1801341. [Google Scholar] [CrossRef]
  31. Kremer, J.I.; Gömpel, K.; Bakuradze, T.; Eisenbrand, G.; Richling, E. Urinary excretion of niacin metabolites in humans after coffee consumption. Mol. Nutr. Food Res. 2018, 62, e1700735. [Google Scholar] [CrossRef]
  32. Shrivastava, A.; Gupta, V. Methods for the determination of limit of detection and limit of quantitation of the analytical methods. Chron Young Sci 2011, 2, 21. [Google Scholar] [CrossRef]
  33. Peterson, L.A.; Naruko, K.C.; Predecki, D.P. A reactive metabolite of furan, cis -2-butene-1,4-dial, is mutagenic in the Ames assay. Chem. Res. Toxicol. 2000, 13, 531–534. [Google Scholar] [CrossRef] [PubMed]
  34. Byrns, M.C.; Vu, C.C.; Peterson, L.A. The formation of substituted 1,N6-etheno-2’-deoxyadenosine and 1,N2-etheno-2’-deoxyguanosine adducts by cis-2-butene-1,4-dial, a reactive metabolite of furan. Chem. Res. Toxicol. 2004, 17, 1607–1613. [Google Scholar] [CrossRef] [PubMed]
  35. Mondelli, R.; Bocchi, V.; Gardini, G.P.; Chierici, L. NMR spectra of Δ3- and Δ4-pyrrolin-2-one. Org. Magn. Reson. 1971, 3, 7–22. [Google Scholar] [CrossRef]
  36. Alves, J.C.F. 2-Pyrrolidinones and 3-pyrrolin-2-ones: A study on the chemical reactivity of these structural moieties. Org. Chem. Int. 2011, 2011, 803120. [Google Scholar] [CrossRef]
  37. Chen, L.J.; Hecht, S.S.; Peterson, L.A. Characterization of amino acid and glutathione adducts of cis-2-butene-1,4-dial, a reactive metabolite of furan. Chem. Res. Toxicol. 1997, 10, 866–874. [Google Scholar] [CrossRef]
  38. Whelton, A.; Watson, A.J.; Rock, R.C. Tietz Textbook of Clinical Chemistry, 2nd ed.; Burtis, C.A., Ashwood, E.R., Tietz, N.W., Eds.; Saunders: Philadelphia, PA, USA, 1994; pp. 1513–1575. ISBN 0721644724. [Google Scholar]
Figure 1. Metabolic pathway and formation of the investigated renal furan metabolites [17,21,22].
Figure 1. Metabolic pathway and formation of the investigated renal furan metabolites [17,21,22].
Metabolites 13 01011 g001
Figure 2. Proposed metabolic pathway and formation of the investigated renal 2-methylfuran (2-MF) metabolites (AcA formation first described by Ravindranath et al. [26]).
Figure 2. Proposed metabolic pathway and formation of the investigated renal 2-methylfuran (2-MF) metabolites (AcA formation first described by Ravindranath et al. [26]).
Metabolites 13 01011 g002
Figure 3. Total Ion Chromatogram (TIC) of a representative urine sample after ingestion of 500 mL of coffee brew containing 44.11 µg of furan and 86.94 µg of 2-MF (P3, time point 2–3 h). Respective retention times of Lys-BDA, AcLys-BDA and GSH-BDA (metabolites of furan), with AcLys-AcA and Lys-AcA (metabolites of 2-MF) indicated.
Figure 3. Total Ion Chromatogram (TIC) of a representative urine sample after ingestion of 500 mL of coffee brew containing 44.11 µg of furan and 86.94 µg of 2-MF (P3, time point 2–3 h). Respective retention times of Lys-BDA, AcLys-BDA and GSH-BDA (metabolites of furan), with AcLys-AcA and Lys-AcA (metabolites of 2-MF) indicated.
Metabolites 13 01011 g003
Figure 4. Excretion kinetics (nmol/h) of (A) Lys-BDA, (B) AcLys-BDA and (C) GSH-BDA from 12 h before to 36 h after ingestion of 500 mL of coffee brew. (n = 10) Boxes, Perc 25/75, whisker, 95% confidence interval; squares, mean; dash, median; times symbol, minimum and maximum; dashed gray line, coffee intake.
Figure 4. Excretion kinetics (nmol/h) of (A) Lys-BDA, (B) AcLys-BDA and (C) GSH-BDA from 12 h before to 36 h after ingestion of 500 mL of coffee brew. (n = 10) Boxes, Perc 25/75, whisker, 95% confidence interval; squares, mean; dash, median; times symbol, minimum and maximum; dashed gray line, coffee intake.
Metabolites 13 01011 g004
Figure 5. Excretion kinetics (nmol/h) of (A) Lys-AcA and (B) AcLys-AcA from 12 h before to 36 h after ingestion of 500 mL of coffee brew. (n = 10) Boxes, Perc 25/75, whisker, 95% confidence interval; squares, mean; dash, median; times symbol, minimum and maximum; dashed gray line, coffee intake.
Figure 5. Excretion kinetics (nmol/h) of (A) Lys-AcA and (B) AcLys-AcA from 12 h before to 36 h after ingestion of 500 mL of coffee brew. (n = 10) Boxes, Perc 25/75, whisker, 95% confidence interval; squares, mean; dash, median; times symbol, minimum and maximum; dashed gray line, coffee intake.
Metabolites 13 01011 g005
Figure 6. Stability of (A) Lys-BDA, (B) AcLys-BDA and (C) GSH-BDA stored at pH 5.4 and 10 over a period of 60 weeks at −20 °C. Data for GSH-BDA after 24 weeks are missing.
Figure 6. Stability of (A) Lys-BDA, (B) AcLys-BDA and (C) GSH-BDA stored at pH 5.4 and 10 over a period of 60 weeks at −20 °C. Data for GSH-BDA after 24 weeks are missing.
Metabolites 13 01011 g006
Table 1. Mass spectrometric parameters for Lys-BDA, AcLys-BDA, GSH-BDA, Lys-AcA and AcLys-AcA, and their stable isotope (IS)-labeled analogs. Mass transitions labeled with * were used for quantification.
Table 1. Mass spectrometric parameters for Lys-BDA, AcLys-BDA, GSH-BDA, Lys-AcA and AcLys-AcA, and their stable isotope (IS)-labeled analogs. Mass transitions labeled with * were used for quantification.
Compound-Specific ParametersRt
(min)
Q1/Q3
(m/z)
Dwell
(msec)
DP
(V)
EP
(V)
CE (V)CXP (V)
Period 1Lys-AcA1.39226.921/146.1 *150106133114
226.921/181.275106131718
Lys-AcA-IS 234.887/90.0 *15086102710
234.887/188.110086101728
Lys-BDA1.51212.847/84.1 *125318298
212.847/132.01253172920
Lys-BDA-IS 221.037/90.0 *12511102914
221.037/174.112511101720
Period 2AcLys-AcA2.98268.859/164.0 *90131132516
268.859/82.07013173112
AcLys-AcA-IS 276.944/170.1 *90136113314
276.944/88.070136113718
GSH-BDA3.16355.953/210.0 *17536103116
355.953/136.01253687314
GSH-BDA-IS 358.829/209.8 *17536103716
358.829/136.01253610718
AcLys-BDA3.62254.962/84.0 *9056123910
254.962/209.17056101728
AcLys-BDA-IS 262.946/90.1 *9041103510
262.946/216.07041101718
MS-specific parametersDuration
(min)
CUR
(psi)
IS
(V)
TEM (°C)GS1
(psi)
GS2 (psi)CAD
Period 12.552015005504555−2
Period 23.452055005504555−3
IS, ionspray voltage; GS1, nebulizer gas; TEM, source temperature; GS2, heater gas; CUR, curtain gas; CAD, collisionally activated dissociation gas; DP, declustering potential; EP, entrance potential; CE, collision energy; CXP, cell exit potential.
Table 2. Parameters for validation of the method.
Table 2. Parameters for validation of the method.
LOD
(nM)
LOQ
(nM)
Standard
conc. (nM)
Recovery
(%)
Lys-BDA0.140.471.2 102
AcLys-BDA0.130.442.4 103
GSH-BDA1.24.010 99
Lys-AcA0.260.851.6 101
AcLys-AcA0.210.702.0 102
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kremer, J.I.; Karlstetter, D.; Kirsch, V.; Bohlen, D.; Klier, C.; Rotermund, J.; Thomas, H.; Lang, L.; Becker, H.; Bakuradze, T.; et al. Stable Isotope Dilution Analysis (SIDA) to Determine Metabolites of Furan and 2-Methylfuran in Human Urine Samples: A Pilot Study. Metabolites 2023, 13, 1011. https://doi.org/10.3390/metabo13091011

AMA Style

Kremer JI, Karlstetter D, Kirsch V, Bohlen D, Klier C, Rotermund J, Thomas H, Lang L, Becker H, Bakuradze T, et al. Stable Isotope Dilution Analysis (SIDA) to Determine Metabolites of Furan and 2-Methylfuran in Human Urine Samples: A Pilot Study. Metabolites. 2023; 13(9):1011. https://doi.org/10.3390/metabo13091011

Chicago/Turabian Style

Kremer, Jonathan Isaak, Dorothea Karlstetter, Verena Kirsch, Daniel Bohlen, Carina Klier, Jan Rotermund, Hannah Thomas, Lukas Lang, Hanna Becker, Tamara Bakuradze, and et al. 2023. "Stable Isotope Dilution Analysis (SIDA) to Determine Metabolites of Furan and 2-Methylfuran in Human Urine Samples: A Pilot Study" Metabolites 13, no. 9: 1011. https://doi.org/10.3390/metabo13091011

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop