Download Disposition of Glutathione Conjugates in Rats by a Novel Glutamic

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Basal metabolic rate wikipedia , lookup

Glycolysis wikipedia , lookup

Fatty acid metabolism wikipedia , lookup

Genetic code wikipedia , lookup

Biochemical cascade wikipedia , lookup

Mass spectrometry wikipedia , lookup

Nucleic acid analogue wikipedia , lookup

Citric acid cycle wikipedia , lookup

Matrix-assisted laser desorption/ionization wikipedia , lookup

Metabolic network modelling wikipedia , lookup

Fatty acid synthesis wikipedia , lookup

Biochemistry wikipedia , lookup

Peptide synthesis wikipedia , lookup

Metabolism wikipedia , lookup

Butyric acid wikipedia , lookup

Isotopic labeling wikipedia , lookup

15-Hydroxyeicosatetraenoic acid wikipedia , lookup

Biosynthesis wikipedia , lookup

Metalloprotein wikipedia , lookup

Amino acid synthesis wikipedia , lookup

Specialized pro-resolving mediators wikipedia , lookup

Glutathione wikipedia , lookup

Pharmacometabolomics wikipedia , lookup

Hepoxilin wikipedia , lookup

Metabolomics wikipedia , lookup

Transcript
0022-3565/00/2942-0735$03.00/0
THE JOURNAL OF PHARMACOLOGY AND EXPERIMENTAL THERAPEUTICS
Copyright © 2000 by The American Society for Pharmacology and Experimental Therapeutics
JPET 294:735–745, 2000 /2299/841110
Vol. 294, No. 2
Printed in U.S.A.
Disposition of Glutathione Conjugates in Rats by a Novel
Glutamic Acid Pathway: Characterization of Unique Peptide
Conjugates by Liquid Chromatography/Mass Spectrometry and
Liquid Chromatography/NMR
ABDUL E. MUTLIB, JOHN SHOCKCOR, ROBERT ESPINA, NILSA GRACIANI, ALICIA DU, and LIANG-SHANG GAN
Drug Metabolism and Pharmacokinetics Section, DuPont Pharmaceuticals Company, Stine-Haskell Research Center, Newark, Delaware
(A.E.M., J.S., R.E., A.D., L.-S.G.); and Chemical and Physical Sciences Division, DuPont Pharmaceuticals Company, Experimental Station,
Wilmington, Delaware (N.G.)
This paper is available online at http://www.jpet.org
ABSTRACT
With the advent of liquid chromatography/mass spectrometry
and liquid chromatography/NMR, it has become easier to characterize metabolites that were once difficult to isolate and
identify. These techniques have enabled us to uncover the
existence of an alternate pathway for the disposition of glutathione adducts of several structurally diverse compounds.
Studies were carried out using acetaminophen as a model
compound to investigate the role of the glutamic acid pathway
in disposition of the glutathione adducts. Although the mercapturic acid pathway was the major route of degradation of the
glutathione adducts, it was found that the conjugation of the
glutathione, cysteinylglycine, and cysteine adducts of acetaminophen with the ␥-carboxylic acid of the glutamic acid was
both interesting and novel. The coupling of the glutathione
adduct and the products from the mercapturic acid pathway
Glutathione (GSH) is one of the most important molecules
in the cellular defense against chemically reactive toxic compounds or oxidative stress. This protective function is due in
part to its involvement in conjugation reactions. GSH is a
tripeptide composed of L-glutamic acid, L-cysteine, and Lglycine (␥-Glu-Cys-Gly). The presence of cysteine in the tripeptide provides a sulfhydryl group that is nucleophilic, and
hence GSH reacts with electrophiles as the thiolate ion, GS⫺.
These electrophiles may be chemically reactive and are produced as metabolic products of a phase 1 reaction, or they
may be more stable foreign compounds. Hence, GSH protects
cells by removing reactive metabolites. The GSH adducts
may then either be excreted, usually into the bile, or the
conjugate may undergo further metabolism via the mercapturic acid pathway. This involves sequential removal of the
Received for publication November 16, 1999.
with the glutamic acid led to unusual peptide conjugates. The
natures of these adducts were confirmed unequivocally by
comparisons with synthetic standards. This pathway (addition
of glutamic acids) led to larger peptides, in contrast to the
mercapturic acid pathway, in which the glutathione adducts are
broken down to smaller molecules. The enzyme responsible for
the addition of glutamic acid to the different elements of the
mercapturic acid pathway is currently unknown. It is postulated
that the ␥-carboxylic acid is activated (perhaps by ATP) before
enzymatic addition to the ␣-amino group of cysteine or glutamate takes place. The discovery of these peptide conjugates of
acetaminophen represents a novel disposition of glutathione
adducts of compounds. The formation of such conjugates may
represent yet another pathway by which drugs could produce
covalent binding via their reactive intermediates.
glutamyl and glycine groups and acetylation of the cysteine
amino group to produce the N-acetylcysteine conjugate or
mercapturic acid. Alternate metabolic pathways or catabolism of these GSH conjugates has not been described before.
In this report, we describe a unique and novel metabolic
pathway for GSH adducts of acetaminophen. Acetaminophen
is a widely used analgesic known to cause hepatotoxicity in
animals and in humans exposed to high doses of this compound. The formation of the GSH conjugate from acetaminophen has been well documented (Hinson et al., 1982). The
degradation of the GSH adduct via the mercapturic acid
pathway has been the subject of several publications and
reviews (Commandeur et al., 1995, and the references cited
therein).
A previously unreported metabolic pathway for these GSH
adducts, leading to the formation of peptide conjugates, is
described in this report. These polar metabolites were iso-
ABBREVIATIONS: GSH, glutathione; LC, liquid chromatography; FMOC, N-␣-fluorenylmethoxycarbonyl; MS, mass spectrometry; SPE, solid
phase extraction; COSY, correlated spectroscopy; TOCSY, total correlated spectroscopy; HMBC, heteronuclear multiple bond coherence; HSQC,
heteronuclear single quantum coherence.
735
Downloaded from jpet.aspetjournals.org at ASPET Journals on May 7, 2017
Accepted for publication April 26, 2000
736
Mutlib et al.
Materials and Methods
Chemicals and Supplies. Trifluoroacetic acid (TFA), phenol,
ethanedithiol, thioanisole, acetaminophen, p-acetamidophenol-␤-Dglucuronide, N-acetylbenzoquinone imine (NAPQI), cysteinylglycine,
and ␥-glutamylcysteine were purchased from Sigma-Aldrich Chemical Co. (St. Louis, MO). Resin Fmoc-Gly and Fmoc-Cys (Trt) were
obtained from Advanced ChemTech (Louisville, KY). Fmoc-L-glutamic acid-␣-t-butyl ester was obtained from NovaBiochem (San
Diego, CA). All other amino acids and synthesizer reagents were
obtained from PE Biosystems (Foster City, CA). Bond-Elut C18
cartridges (10 g/60 ml) were obtained from Varian Sample Preparation Products (Harbor City, CA). All general solvents and reagents
were of the highest grade commercially available.
Synthesis of ␥-Glu-␥-Glu-Cys-Gly (Tetrapeptide) and ␥-Glu␥-Glu-Cys (Tripeptide). The tetrapeptide ␥-Glu-␥-Glu-Cys-Gly and
the tripeptide ␥-Glu-␥-Glu-Cys were made using L-isomers of each
amino acid. The ␣-carboxylic acid was protected on the glutamic acid,
allowing the ␥-carboxylic acid to be coupled with the amino functional group of the coupling amino acid. The peptides were synthesized on solid phase using Fmoc protection and 2-(1H-benzotriazol1-yl)-1, 1, 3, 3, -tetramethyluronium hexafluorophosphate (HBTU)
activation on a peptide synthesizer (model ABI 433A; PE Biosystems). The TFA cleavage of the peptide from the resin using reagent
K (84% TFA, 4% H2O, 4% thioanisole, 2% EDTA, 6% phenol; King et
al., 1990) was carried out, and the products were precipitated from
ethyl ether. After lyophilization of the crude products, HPLC purification was performed on a Rainin Dynamax HPLC system using a
semipreparative C18 column (21.4 ⫻ 250 mm). The mobile phase
consisted of two components: A (0.1% TFA in water) and B (10:90 v/v
acetonitrile/0.1% TFA). The proportion of component B was increased from 0 to 15% over 15 min at a flow rate of 18 ml/min. A total
of 7.6 mg of ␥-Glu-␥-Glu-Cys-Gly (tetrapeptide) and 1.7 mg of ␥-Glu␥-Glu-Cys (tripeptide) were obtained after lyophilizing the collected
fractions.
Synthesis of Tetrapeptide Conjugate (M6) of Acetaminophen. The synthesis of 3-(␥-glutamyl-glutathion-S-yl)acetaminophen (M6, see Scheme 1) was done by reacting freshly prepared
solution of NAPQI with ␥-Glu-␥-Glu-Cys-Gly. NAPQI (1.3 mg) was
dissolved in 167 ␮l of acetonitrile in a test tube, and 135 ␮l was
transferred to a vessel containing 2.2 ml of ice-cold water and 270 ␮l
of 0.2 M ammonium acetate solution (pH 6.0). ␥-Glutamyl-␥-glutamylcysteinylglycine (3.0 mg) was dissolved in 135 ␮l of water and
added to the NAPQI solution. The reaction container was covered
with aluminum foil to protect NAPQI from light, and the solution
was stirred for 1 h at room temperature. After removal of the solvent
under nitrogen at room temperature, the residue was reconstituted
in water (500 ␮l) and chromatographed on a semipreparative HPLC
column (Waters Symmetry C18, 7.8 ⫻ 300 mm, 7 ␮m). The solvent
system consisted of a mixture of acetonitrile and 0.05% TFA. The
solvent composition was maintained at 5% acetonitrile for the first 5
min, followed by a linear ramp to 15% acetonitrile in the next 8 min.
The percentage organic was increased to 60% after 13 min and
maintained at this composition for the next 3 min. The column was
re-equilibrated with 5% acetonitrile in 0.1% TFA for 10 min before
the next injection. The solvent was delivered at a rate of 3.5 ml/min.
Approximately 4.5 mg of the tetrapeptide conjugate was recovered
after HPLC purification. Mass spectral and NMR analyses were
performed to confirm the identity of the isolated product.
Synthesis of Cysteinylglycine Conjugate (M4) of Acetaminophen. The synthesis of 3-(glycinylcystein-S-yl)acetaminophen (M4,
see Scheme 1) was done by reacting freshly prepared solution of
NAPQI with cysteinylglycine in the same manner as described for
metabolite M6. NAPQI (15 mg) was dissolved in 2 ml of acetonitrile
in a scintillation vial, and the entire sample was transferred to a
vessel containing 5 ml of ice-cold water and 1 ml of 0.2 M ammonium
acetate solution (pH 6.0). Cysteinylglycine (18 mg) was dissolved in
2 ml of water and added to the NAPQI solution. The reaction was
carried out as described, after which the solvent was removed under
nitrogen at room temperature. The residue was reconstituted in
water and chromatographed on a semipreparative HPLC column.
The HPLC system used to separate the dipeptide conjugate was
same as that described for isolation of the tetrapeptide conjugate, M6
(see earlier). A second HPLC purification was done using an isocratic
mobile phase (10:90 v/v acetonitrile/0.05% TFA) delivered at a rate of
3.5 ml/min. Approximately 18 mg of the cysteinylglycine conjugate
was recovered after purification.
Synthesis of Dipeptide Conjugate (M8) of Acetaminophen.
The synthesis of 3-(␥-glutamylcystein-S-yl)acetaminophen (M8, see
Scheme 1) was done by reacting freshly prepared solution of NAPQI
with ␥-glutamylcysteine. NAPQI (2.5 mg) was dissolved in 330 ␮l of
acetonitrile in a test tube, and 300 ␮l was transferred to a vessel
containing 4.8 ml of ice-cold water and 600 ␮l of 0.2 M ammonium
acetate solution (pH 6.0). ␥-Glutamylcysteine (9.1 mg) was dissolved
in 0.5 ml of water, and 300 ␮l was added to the NAPQI solution. The
reaction container was covered with aluminum foil to protect NAPQI
from light, and the solution was stirred for 1 h at room temperature.
After removal of the solvent under nitrogen at room temperature, the
residue was reconstituted in water and chromatographed on a semipreparative HPLC column. The HPLC system used to separate the
dipeptide conjugate was same as that described for isolation of tetrapeptide conjugate M6 (see earlier). Approximately 3 mg of the
dipeptide conjugate was recovered after HPLC purification. Mass
spectral and NMR analyses were performed to confirm the identity of
the isolated product.
Synthesis of Tripeptide Conjugate (M9) of Acetaminophen.
The synthesis of 3-(␥-glutamyl-␥-glutamylcystein-S-yl)acetaminophen (M9, see Scheme 1) was done by reacting freshly prepared
solution of NAPQI with ␥-glutamyl-␥-glutamylcysteine in the same
manner as described for metabolites M6 and M8. Due to the low yield
of the product, the tripeptide conjugate was not isolated for NMR
analyses. Instead, the product was analyzed directly by LC/MS, and
comparisons were made with the in vivo metabolite.
LCMS. LC/MS was carried out by coupling a Hewlett-Packard
HPLC system (HP 1100) to a Finnigan LCQ ion trap mass spectrometer. The HPLC eluent was introduced to the mass spectrometer
using a pneumatically assisted electrospray source. The mass spectrometer was operated in both the negative and positive ion modes.
MSn (MS/MS experiments with n ⫽ 0 – 4) were done with relative
collision energy set to 20 to 25%. Metabolites that could not be
isolated in pure form for NMR analyses were subjected to several
LC/MSn experiments for the purpose of structure elucidation.
Downloaded from jpet.aspetjournals.org at ASPET Journals on May 7, 2017
lated from bile by solid phase extraction and by reversed
phase HPLC. Often, the structures of such polar metabolites
are not elucidated due to the difficulty in isolating and separating them from endogenous components in the biological
matrices. With the advent of liquid chromatography/mass
spectrometry (LC/MS) and LC/NMR, it has become possible
to elucidate the structures of minor metabolites. Characterization of these polar conjugates can provide very useful
information regarding the metabolic pathways and evidence
for any potentially reactive metabolites of a compound. This
report describes the unequivocal identification of novel peptide conjugates resulting from conjugation of GSH adducts
and its breakdown products with glutamic acid using LC/MS/
MS, LC/NMR, and NMR approaches. The ␥-carboxylic acid
coupling of glutamic acid to other amino acids of GSH-derived adducts was demonstrated by synthesizing authentic
standards and comparing spectral data with isolated metabolites. These metabolites are biologically novel, and the formation of these types of metabolites may represent a previously unappreciated combination of drug metabolism
reactions.
Vol. 294
2000
Disposition of Glutathione Adducts by Glutamic Acid Pathway
737
For analyses of the acetaminophen metabolites, HPLC was carried
out using a Hewlett-Packard HPLC system (HP 1100) coupled in
sequence to a Waters symmetry C18 column (150 ⫻ 3.9 mm, 5 ␮m).
The metabolites were separated by a gradient solvent system consisting of a mixture of acetonitrile and 0.05% TFA with the flow rate
set at 0.7 ml/min. The initial conditions consisted of a mixture of
acetonitrile and 0.05% TFA (5:95 v/v) and was maintained for 5 min
after the sample was injected. The percentage of acetonitrile was
increased from 5 to 30% in the next 15 min. After 20 min, the column
was washed with 90% acetonitrile for 5 min before re-equilibrating
with the initial mobile phase. Aliquots of bile and urine samples were
injected directly onto the HPLC, and the eluent was introduced into
the source of the mass spectrometer.
To detect the metabolites in the fractions from C18 cartridges and
from semipreparative HPLC column, aliquots (20 –50 ␮l) were introduced into the mass spectrometer using the flow injection analysis
method. The mobile phase consisted of a mixture of acetonitrile and
10 mM ammonium formate (1:1 v/v) delivered at a rate of 0.25
ml/min.
LC/NMR. The tetrapeptide conjugate of acetaminophen was characterized by LC/NMR. All of the spectra were obtained using a
Bruker Avance 600 MHz NMR spectrometer equipped with an 1H/
13
C LC/NMR flow probe with a cell volume of 120 ␮l. Suppression of
the residual water and acetonitrile signals was carried out using the
WET solvent suppression method in all of the LC/NMR experiments.
Chemical shifts were referenced to DMSO at ␦ of 2.49 ppm and to
acetonitrile at ␦ of 2.0 ppm. An HP 1100 LC system was used with a
Bruker Diode-Array detector set at 254 nm. A 3.9 ⫻ 150 mm HPLC
Waters Symmetry C18 5-mm column was used. The initial HPLC
conditions consisted of isocratic elution using 95% D2O and 5%
acetonitrile-d3 for first 5 min, followed by a gradient from 95% D2O
and 5% acetonitrile-d3 to 70% D2O and 30% acetonitrile-d3 over 15
min at a flow rate of 0.7 ml/min. Both solvents contained 0.05% TFA.
The spectra of isolated metabolites of acetaminophen were obtained using a 2.5-mm 1H/13C inverse probe. The samples were
dissolved in DMSO-d6 and filtered to remove particulate matter. The
structures of these metabolites (approximately 1–2 mg of each) were
determined from proton and carbon one-dimensional NMR as well as
two-dimensional proton-proton correlated spectroscopy (COSY), proton-carbon heteronuclear single quantum coherence (HSQC), and
long-range proton-carbon heteronuclear multiple bond coherence
(HMBC) NMR experiments.
Animal Studies. Six male Sprague-Dawley rats with cannulated
bile ducts (weighing between 250 – 400 g) were dosed orally with
acetaminophen (100 mg/ml, made in propylene glycol/water 8:2 v/v)
once daily at 500 mg/kg for 2 days. Serial bile and urine samples
were collected over ice and pooled from all of the animals on a daily
basis and stored frozen at ⫺20°C until analyzed. The dosing volume
was 5 ml/kg. In another study, a single male Sprague-Dawley rat
was dosed i.v. (60 mg/kg) with the cysteinylglycine conjugate of
acetaminophen (M4), and bile and urine were collected over 0- to 6and 6- to 24-h time intervals.
Isolation of GSH-Derived Conjugates from Rat Bile. Bile
obtained from the rats was pooled, diluted 1:1 with distilled water,
and loaded onto a C18 cartridge (10 g/60 ml). The sample was
allowed to elute under gravity at a rate of less than 1 ml/min. After
Downloaded from jpet.aspetjournals.org at ASPET Journals on May 7, 2017
Scheme 1. Proposed disposition of GSH adducts of acetaminophen via the mercapturic acid and the glutamic acid pathways.
738
Mutlib et al.
Vol. 294
Results
LC/MS and NMR Characterization of Acetaminophen Metabolites in Rat Bile. LC/MS analyses of bile
samples from rats dosed with acetaminophen showed that
the polar metabolites, including the glucuronide (M1), sulfate
(M2), cysteine (M3), cysteinylglycine (M4), and GSH (M5)
conjugates, were well resolved from each other and from the
endogenous materials using a gradient HPLC system. The
metabolites were monitored by using the LCQ mass spectrometer and a variable-wavelength UV detector. A representative total ion current (TIC) and an HPLC/UV trace obtained simultaneously during the analysis of a bile sample
Fig. 1. Structure of acetaminophen.
are shown in Fig. 2. A number of unusual metabolites (M6 –
M9) were also found during the analysis. These metabolites
either coeluted with other known metabolites of acetaminophen and/or were present in very low quantities. The extracted ion chromatograms showing the presence of these
metabolites in bile are shown in Fig. 2.
Metabolites M1 to M6 found in the bile of rats dosed with
acetaminophen produced pseudomolecular ions ([M ⫹ H]⫹)
at m/z 328, 232, 271, 328, 457, and 586, respectively, were
isolated in sufficient quantities for unambiguous confirmation of the structures by NMR experiments. Metabolites M7
to M9 were isolated chromatographically during LC/MS analyses for several MS/MS experiments.
Metabolite M1 was found to be the glucuronide conjugate
of acetaminophen (MH⫹ at m/z 328). The identity of this
metabolite was confirmed by comparing the retention time
and mass spectral fragmentation pattern with commercially
available synthetic standard.
Metabolite M2 was determined to be the sulfate conjugate
of acetaminophen with MH⫹ at m/z 232. The pseudomolecular ion at m/z 230 in the negative ion mode showed a characteristic loss of sulfate moiety (Weidolf et al., 1988). The 1H
NMR data showed proton signals of an intact aromatic ring [␦
7.05 (d) and 7.43 (d)]. The aromatic signals were characteristic of a para-substituted ring system. The signal for the
amide proton appeared at ␦ 9.82, whereas the acetyl group
appeared as a singlet at ␦ 2.0 ppm.
Metabolite M3 was found to be the cysteine conjugate with
MH⫹ at m/z 271. The 1H NMR data and the mass spectral
fragmentation pattern confirmed the structure of this metabolite. The 1H NMR spectrum showed signals at ␦ 7.60 (1H,
Ar-H, d), 7.24 (1H, Ar-H, dd), 6.80 (1H, Ar-H, d), 4.00 (1H, cys
␣, m), 3.30 (1H, cys ␤, m), 3.20 (1H, cys ␤⬘, m), 2.00 (3H,
COCH3, s). The aromatic proton signals had changed as a
result of substitution on the ring.
Metabolite M4 was found to be the cysteinylglycine adduct
of acetaminophen. The 1H NMR showed signals for the cysteine and glycine protons as reported for cysteinylglycine
conjugates (Mutlib et al., 1999). The 1H NMR spectrum
showed signals at ␦ 7.63 (1H, Ar-H, d), 7.24 (1H, Ar-H, dd),
6.82 (1H, Ar-H, d), 4.00 (1H, cys ␣, m), 3.78 (2H, gly ␣, m),
3.30 (1H, cys ␤, m), 3.18 (1H, cys ␤⬘, m), 2.00 (3H, COCH3, s).
The aromatic proton signals were found to be similar to those
obtained for the cysteine conjugate. The MH⫹ for the cystei-
Downloaded from jpet.aspetjournals.org at ASPET Journals on May 7, 2017
the sample had been loaded, the column bed was washed with 30 ml
of deionized water followed by elution with 5, 10, and 15% methanol
in water (20 ml volume of each). Aliquots from all the fractions were
analyzed by LC/MS. The fractions containing the GSH related adducts were subsequently loaded onto C18 cartridges (10 g/60 ml) for
further repurification. The fractions containing the metabolites were
pooled and dried under vacuum. The dried residues were reconstituted in water and rechromatographed on a semipreparative HPLC
column as described later.
Semipreparative HPLC separation of the polar metabolites of
acetaminophen isolated from bile was carried out on a Waters Symmetry C18 (300 ⫻ 7.5 mm, 7 ␮m) using acetonitrile and 0.05% TFA
as the eluent. The initial HPLC conditions consisted of isocratic
elution using 95% H2O and 5% acetonitrile for first 5 min, followed
by a gradient from 95% H2O and 5% acetonitrile to 85% H2O and
15% acetonitrile over 8 min. The percentage of acetonitrile was
increased to 60% over the next minute and maintained at this for an
additional 5 min before re-equilibrating the column with the initial
mobile phase. The flow rate was set at 3.5 ml/min. The eluent was
monitored at 254 nm using a variable wavelength detector. Five
peaks were collected corresponding to metabolites M1 (m/z 328), M2
(m/z 232), M3 (m/z 271), M4 (m/z 328), and M5 (m/z 457). Samples
corresponding to each metabolite were pooled and dried under vacuum. Final purification was done by rechromatographing each isolated metabolite on C18 cartridges. Repurification of M5 on an analytical column (3.9 ⫻ 150 mm, Waters Symmetry C18, 5 ␮m) resulted
in separation of the tetrapeptide conjugate, M6 with MH⫹ at m/z 586
(see later). After pooling the fractions containing the metabolites, the
samples were dried under vacuum and submitted for NMR analyses.
Approximately 1 to 2 mg of each metabolite M1 to M5 were isolated
for NMR analyses.
Isolation of Tetrapeptide Conjugate of Acetaminophen
from Rat Bile. The tetrapeptide conjugate, M6 (MH⫹ at m/z 586),
coeluted with the GSH adduct of acetaminophen during the initial
purification steps. This metabolite was isolated from the GSH adduct
on an analytical column (Waters Symmetry C19, 3.9 ⫻ 150 mm)
using an isocratic mobile phase consisting of a mixture of acetonitrile
and 0.05% TFA (5:95 v/v) delivered at 0.7 ml/min. The peak corresponding to this metabolite (11.2 min) was collected, pooled from
several injections, dried, and submitted for LC/NMR analyses.
Incubations of 3-(␥-Glutathion-S-yl)acetaminophen with
Rat Liver/Kidney Subcellular Fractions and Bile. The purified
GSH conjugate of acetaminophen, M5, was incubated in the presence
of rat bile and rat liver/kidney subcellular fractions. The absence of
any other metabolites in the purified metabolite, M5, was confirmed
by LC/MS experiments before incubations. Approximately 250 ␮g of
M5 was incubated with 0.5 ml of freshly collected rat bile or with
freshly prepared liver/kidney S9 fractions (5 mg/ml). The samples
were incubated at 37°C for 1 h, after which the samples were analyzed by LC/MS. Bile samples (25 ␮l) were analyzed by LC/MS
without any sample treatment. The liver/kidney incubations were
precipitated with 3 ml of acetonitrile, and the supernatant was dried,
reconstituted in the HPLC mobile phase, and injected onto LC/MS.
2000
Disposition of Glutathione Adducts by Glutamic Acid Pathway
739
nylglycine adduct was found at m/z 328. The LC/MS retention times and mass spectral fragmentation patterns for the
metabolite and the synthetic standard were found to be identical.
Metabolite M5 was the GSH adduct of acetaminophen with
the mass spectral data showing the pseudomolecular ion
(MH⫹) at m/z 457. The 1H NMR of the isolated metabolite
clearly showed the presence of the GSH and the loss of one of
the aromatic protons. The assignment of the proton signals
were done by TOCSY experiments; ␦ 7.45 (1H, Ar-H, d), 7.22
(1H, Ar-H, dd), 6.75 (1H, Ar-H, d), 4.48 (1H, cys ␣, m), 3.98
(1H, glu ␣, m), 3.76 (2H, gly ␣, d), 3.23 (1H, cys ␤, m), 2.98
(1H, cys ␤⬘, m), 2.41 (1H, glu ␥, m), 2.35 (1H, glu ␥⬘, m) 2.05
(2H, glut ␤, ␤⬘, m), 2.00 (3H, COCH3, s). The 1H NMR results
obtained for M5 was found to be consistent with data reported previously (Hinson et al., 1982). The MS/MS fragmentation ions are shown in Fig. 3. A number of MSn experiments with the fragment ions derived from MH⫹ (m/z 457)
and [M ⫺ H]⫺ (m/z 455) were done, and the results are shown
in Tables 1 and 2. Hence, M5 was confirmed as 3-(␥-glutathion-S-yl)acetaminophen.
Metabolite M6 was the tetrapeptide conjugate formed by
condensation of ␥-carboxyl group of the glutamic acid with
the ␣-amino group of the glutamate present in the GSH
adduct of acetaminophen. The proposed structure of this
metabolite is shown in Scheme 1. The mass spectral data
showed MH⫹ at m/z 586, which was consistent with an addition of a glutamic acid onto the GSH adduct of acetaminophen. LC/MS analysis of bile samples obtained from rats
dosed with acetaminophen showed that this metabolite
eluted after the GSH conjugate on the HPLC column. The
structure of this adduct was supported by MSn experiments
that showed that the glutamic acid was linked to the GSH
Fig. 3. MS/MS fragmentation patterns of the GSH adduct (M5) and the
tetrapeptide conjugate (M6). The fragment ions a (m/z 382) and b (m/z
182) are formed by an elimination of glutamic acid and by cleavage at the
COS bond of cysteine moiety, respectively. The mass spectral analysis
was done in the positive ion mode.
Downloaded from jpet.aspetjournals.org at ASPET Journals on May 7, 2017
Fig. 2. LC/MS and HPLC/UV profiles (A and B, respectively) of acetaminophen metabolites present in rat bile showing the presence of glucuronide
conjugate (M1), sulfate conjugate (M2), cysteine (M3), cysteinylglycine (M4), GSH adduct (M5), and tetrapeptide and pentapeptide conjugates (M6 and
M7, respectively). The extracted ion chromatograms of metabolites M6 (m/z 586), M7 (m/z 715), M8 (m/z 400), and M9 (m/z 529) showed that these
metabolites coeluted with each other or with other metabolites of acetaminophen.
740
Mutlib et al.
Vol. 294
TABLE 1
Mass spectral data (MS/MS in positive ion mode) for the glutathione (M5) and the peptide conjugate M6 of acetaminophen
The corresponding fragments are shown in Fig. 3.
Metabolite M5
(MH⫹457)
4573
382,328a,311,269,225,182
45733283
311a,269,225,208,182,140
457332833113
269,225,208,182a,140
457332831823
140
Metabolite M6
(MH⫹586)
5863
568,511,457a,382,328,311,269
58635113
493,448,382a,365,337,328,310,241
58634573
493,382,328a,311,269,225,182
586345733283
311a,269,225,208,182,140
3, the fragment ions originating from the particular ion.
a
Base peak in the mass spectrum.
TABLE 2
Mass spectral data (MS/MS in the negative ion mode) for the glutathione (M5) and peptide conjugates (M6 and M7) of acetaminophen
4573
272a,254,210,182,179
45532723
254a,210,179,143,128
455327232543
210,179a,167,126,87
Metabolite M6
[M ⫺ H]⫺584
5843
566,548,455,401a,383,365,339,272,254,210
58434013
383a,365,339,272,254
58434553
272a,254,182,179
Metabolite M7
[M ⫺ H]⫺ 713
7133
695,619,584a,566,530,512,455,401,272
71335843
516,455,401a,383,272,254
713358434553
272a,254,210,182
Glutathione
[M ⫺ H]⫺306
3063
288,272a,254,210,179,143,128
30632723
254a,210,179,143
306327232543
210,179a,143
3, the fragment ions originating from the particular ion.
a
Base peak in the mass spectrum.
adduct via the ␥-carboxylic acid. The analyses were carried
out in both positive and negative ion modes to support the
proposed structure. The fragment ions obtained from these
experiments are listed in Tables 1 and 2. The fragmentation
pathways confirming the structure of this conjugate is shown
in Fig. 3. Based on the product ion spectra of a series of model
GSH adducts, Haroldsen et al. (1988) identified routes of
fragmentation of these conjugates. Subsequent studies with
other GSH adducts have confirmed the generality of these
pathways, regardless of ionization technique adopted to generate the pseudomolecular ion. The MS/MS favors the cleavage of specific peptide bonds in the GSH residue, notably the
␥-glutamyl linkage to yield ion a (Fig. 3). In the positive ion
mode, MS/MS of the pseudomolecular ion (m/z 457) of the
GSH conjugate produced the major fragment ion at m/z 328
formed by the loss of the glutamate moiety. Loss of glycine
residue (⫺75 amu) from the parent ion led to a fragment ion
at m/z 382. This characteristic loss of glycine residue from the
GSH adduct has been described previously (Haroldsen et al.,
1988) and was found to be very diagnostic in confirming the
structures of M6 and M7 (see later). The ion at m/z 182 is
derived from the cleavage of carbon-sulfur bond leading to
retention of sulfur on the aromatic ring of acetaminophen.
Such fragmentation pathway has been described for GSH
adducts of aromatic compounds (Haroldsen et al., 1988). The
aromatic thioether conjugates usually display a fragment
resulting from the cysteinyl COS bond with proton transfer
to the aromatic moiety to yield a stabilized aryl-SH2⫹ species
(ion b, Fig. 3). The MS/MS fragmentation of the pseudomolecular ion (m/z 586) of the tetrapeptide produced ions at m/z
457 and 511 formed by losses of glutamate and glycine residues, respectively. Loss of glutamate from ion m/z 511
yielded a significant ion at m/z 382. The fragmentation pattern produced from different ions after several MSn experiments was consistent with the structure proposed in Fig. 3
and Scheme 1. The negative ion MS showed a fragmentation
pattern dominated by fragments resulting from the cleavage
of amino acid residues from the Cys-Gly-Glu entity. For comparison, the MS/MS data for the GSH adduct is shown in
Table 2. The MS/MS of the pseudomolecular ion ([M ⫺ H]⫺ at
m/z 584) of the tetrapeptide conjugate M6 produced a major
fragment ion at m/z 401 (Table 2). This ion represents the
Gly-Cys-Glu-Glu fragment with the sulfur of cysteine being
retained by the aromatic ring of the acetaminophen (fragment ion at m/z 182). The MS/MS data obtained from MSn
experiments with different fragment ions are shown in Table
2. The MS/MS data for intact reduced GSH is included in
Table 2 for comparison purposes. The fragment ions arising
from MS/MS of ions at m/z 401 and 272 confirmed that the
GSH moiety linked to acetaminophen was intact (compare
data with those of M5 and GSH). Results from further
MS/MS experiments confirming the structure of the tetrapeptide is shown in Table 2.
The 1H NMR and TOCSY data for the tetrapeptide conjugate M6 isolated from rat bile are shown in Figs. 4 and 5,
respectively. The 1H NMR spectrum clearly showed the presence of additional resonances for the extra glutamic acid at ␦
4.28 (1H, glu ␣, m), 2.3 (2H, glu ␥, m), and 1.90 –2.00 (2H, glu
␤, m), in addition to the usual GSH protons. The correlation
between the ␣, ␤, and ␥ protons of the two glutamic acids in
the molecule is shown by TOCSY data (see Fig. 5). The
LC/MS and NMR data obtained for the synthetic standard
were found to be identical with those obtained from M6.
Based on the spectroscopic data and comparison with synthetic standard, M6 was confirmed as 3-(␥-glutamyl-glutathion-S-yl)acetaminophen.
Metabolite M7 was found to be the pentapeptide conjugate
of acetaminophen with [M ⫺ H]⫺ at m/z 713. The structure of
this conjugate was determined to be 3-(␥-glutamyl-␥-glutamyl-glutathion-S-yl)acetaminophen as shown in Scheme 1.
MS/MS of ion at m/z 713 produced sequential losses of two
glutamates resulting in ions at m/z 584 and 455 (see Table 2).
Further MS/MS studies with ions at m/z 584 and 455 supported the proposed structure shown in Scheme 1. M7 was
not produced in sufficient quantities to be isolated for characterization by NMR.
Metabolite M8 was found to be the dipeptide conjugate of
acetaminophen with MH⫹ at m/z 400. The structure of this
conjugate was confirmed as 3-(␥-glutamylcystein-S-yl)acetaminophen (see Fig. 6). MS/MS of the pseudomolecular ion at
Downloaded from jpet.aspetjournals.org at ASPET Journals on May 7, 2017
Metabolite M5
[M ⫺ H]⫺ 455
2000
Disposition of Glutathione Adducts by Glutamic Acid Pathway
741
m/z 400 produced an ion at m/z 271 formed by the characteristic loss of glutamate residue. The other fragments included
ions at m/z 140, 182, 208, 225, and 254. An ion at m/z 337 was
postulated to be formed by subsequent losses of NH3 and
formic acid from the ␣-position of the attached glutamic acid.
Hence, the data suggested that the glutamic acid was linked
to the amino group of the cysteine via the ␥-carboxylic terminal group. The structure of this conjugate was confirmed
by synthesizing the dipeptide conjugate of acetaminophen.
The 1H NMR (see Fig. 6) and HMBC experiments confirmed
the structure of the synthetic standard. The LC/MS retention
time and mass spectral data of the metabolite matched those
of the standard as shown in Fig. 7.
Metabolite M9 was found to be the tripeptide conjugate of
acetaminophen with MH⫹ at m/z 529. The proposed structure of this conjugate is 3-(␥-glutamyl-␥-glutamylcystein-Syl)acetaminophen (see Scheme 1). MS/MS of the pseudomolecular ion at m/z 529 produced a characteristic ion at m/z
466 that probably was formed by the losses of ammonia and
formic acid from the terminal glutamate moiety. Further
MS/MS experiments with the fragment ions at m/z 400 and
271 confirmed the postulated sequence of amino acids as
proposed in Scheme 1.
LC/MS Characterization of Metabolites Produced by
Cysteinylglycine Conjugate (M4) of Acetaminophen.
Analyses of bile samples from the rat dosed with M4 showed
that most of the metabolite was converted to N-acetylcysteine conjugate via the mercapturic acid pathway. Interestingly, when acetaminophen was given orally to the rats,
cysteine and not N-acetylcysteine conjugate was detected as
one of the major metabolites. The purpose of administering
M4 was to show that GSH could be formed from this metabolite via the glutamic acid pathway. Indeed, the GSH adduct
was detected as a minor metabolite in the bile. In addition to
these metabolites, the 3-(␥-glutamylcystein-S-yl)acetaminophen conjugate M8 was found in significant quantities in
bile. Overall, the conversion of M4 via the mercapturic acid
pathway was the dominant route of metabolism.
Incubation of 3-(␥-Glutathion-S-yl)acetaminophen
with Rat Liver/Kidney Subcellular Fractions and Bile.
It was found that the GSH adduct, M5, was converted to the
cysteinylglycine and cysteine conjugates, M4 and M3, respectively, in the presence of rat bile and liver/kidney subcellular
fractions. The formation of M6 or any of the peptide conjugates M7, M8, and M9 were not detected in any of these
experiments. The LC/MS profiles suggested that mercapturic
acid pathway predominated under the experimental conditions used during the incubations.
Discussion
Numerous reports have appeared in the literature describing the formation of GSH adducts and their subsequent catabolism to cysteinylglycine, cysteine, and N-acetylcysteine
conjugates via the mercapturic acid pathway (Commandeur
et al., 1995, and the literature cited therein). The GSH conjugates are formed by the interaction of an electrophilic compound, such as those produced by bioactivation of drugs by
cytochrome P450, with the endogenous pool of GSH. These
reactions are either mediated by specific GSH transferases or
are nonenzymatic in nature (Habig et al., 1974; Ketterer,
1986). The resulting GSH adducts are excreted in bile or are
Downloaded from jpet.aspetjournals.org at ASPET Journals on May 7, 2017
Fig. 4. 1H LC/NMR of the tetrapeptide conjugate, 3-(␥-glutamyl-glutathion-S-yl)acetaminophen (M6), isolated from rat bile.
742
Mutlib et al.
Vol. 294
further catabolized via the mercapturic acid pathway to cysteinylglycine, cysteine, or N-acetylcysteine adducts (Commandeur et al., 1995). The cleaved cysteine conjugates are
often reabsorbed through the gastrointestinal tract and are
converted to the N-acetylcysteine derivatives by N-acetyltransferases present in kidneys. The formation of GSH conjugates of electrophilic compounds has generally been regarded as a detoxification process. However, conjugation
with GSH has also been implicated in the activation of some
xenobiotics to mutagenic and carcinogenic electrophiles
(Lash and Anders, 1986; Vamvaka et al., 1987; Anders et al.,
1988; Monks and Lau, 1989; Koob and Dekant, 1991). Furthermore, evidence has been presented showing that GSH
adducts of a variety of compounds and/or their corresponding
cysteine conjugates are nephrotoxic (Elfarra and Anders,
1984; Nash et al., 1984; Monks et al., 1985; Anders et al.,
1987; Monks and Lau, 1989).
Acetaminophen, a commonly used analgesic, has been
shown to produce a GSH adduct via an interaction between
its metabolic intermediate, NAPQI, and endogenous GSH
(Dahlin et al., 1984). The GSH adduct of acetaminophen,
3-(glutathion-S-yl)acetaminophen (M5) has been characterized before (Hinson et al., 1982). In this study, the GSH
adduct of acetaminophen and the products resulting from
further processing of this conjugate via the mercapturic acid
pathway were characterized spectroscopically. In addition to
these metabolites, a number of peptide conjugates were iso-
lated from rat bile and characterized spectroscopically. The
MS and NMR analyses of these conjugates confirmed the
presence of additional glutamic acid moieties linked to the
intact GSH, cysteinylglycine, and cysteine adducts of acetaminophen. The addition of one and two glutamic acids to the
GSH adduct of acetaminophen produced metabolites M6 and
M7, respectively. The addition of glutamic acid was also
observed with the cysteinylglycine and cysteine conjugates of
acetaminophen to produce M5 (GSH adduct) and M8, respectively. Metabolites M8 and M9 were produced as a result of
conjugation of one and two glutamic acids to the cysteine
conjugate (M3) of acetaminophen, respectively. The enzyme
responsible for mediating the addition of glutamic acid to
cysteine, cysteinylglycine, and GSH adducts in the glutamic
acid pathway is currently unknown. It is postulated that
␥-glutamyltranspeptidase could be involved in the formation
of these conjugates. The natural occurrence of ␥-glutamylGSH, a polyanionic tetrapeptide, has been described previously (Abbott et al., 1986). The formation of ␥-Glu-GSH from
endogenous GSH was demonstrated to have been mediated
by ␥-glutamyltranspeptidase (Abbott et al., 1986). Studies
are ongoing to determine the nature of the enzyme or enzymes responsible for the “glutamic acid pathway”.
The coupling of glutamic acid to other amino acids was
demonstrated to occur via its ␥-carboxylic acid group. This
was confirmed by comparing the NMR and MS data of the
isolated metabolites with those of synthetic standards. The
Downloaded from jpet.aspetjournals.org at ASPET Journals on May 7, 2017
Fig. 5. TOCSY of the tetrapeptide conjugate, 3-(␥-glutamyl-glutathion-S-yl)acetaminophen (M6), isolated from rat bile.
2000
Disposition of Glutathione Adducts by Glutamic Acid Pathway
743
identification of metabolites M8 and M9 was important because it confirmed that the enzymatic addition of glutamate
was taking place on amino acids other than just the glutamate of the GSH adducts. Furthermore, it was demonstrated
from in vivo studies that the cysteinylglycine conjugate itself
could be converted to the GSH conjugate via the aforementioned glutamic acid pathway. However, it was shown that
mercapturic acid pathway is more dominant than the glutamic acid pathway.
The discovery of these tetrapeptide and pentapeptide conjugates of acetaminophen represents a novel disposition of
GSH adducts of compounds. The degradation of GSH adducts
via the mercapturic acid pathway is the only one that has
been described in the literature (Commandeur et al., 1995,
and the literature cited therein). GSH adducts undergoing
catabolism via mercapturic pathway lead to smaller molecules such as cysteinylglycine and cysteine conjugates, which
are normally excreted in bile. In this study, we have uncovered an alternate disposition pathway for GSH adducts (and
for their breakdown products formed via the mercapturic
acid pathway) in which the conjugates were taken as substrates by an unknown enzyme that added glutamic acids in
an ordered manner (see Scheme 1). This pathway seems to
act in the opposite manner compared with the mercapturic
acid pathway. Although the mercapturic acid pathway represented essentially a degradation route for GSH adducts,
the newly discovered glutamic acid pathway appears to build
on the GSH adducts (and their breakdown products from
mercapturic acid pathway) by adding elements of glutamic
acid. As shown in Scheme 1, glutamic acid was added sequen-
tially to the GSH adducts of acetaminophen to give metabolites M6 and M7. The addition of glutamic acid was not only
confined to the glutamic acid moiety of the GSH adduct;
instead, products were found in which the glutamic acid was
added sequentially to the cysteine to give metabolites M8 and
M9 or to the cysteinylglycine adduct to give M5. The structures of these dipeptides, tripeptides, tetrapeptides, and pentapeptides suggested that the ␥-carboxylic acid of glutamic
acid was activated and subsequently added to the ␣-amino
group of the conjugates. This type of reaction has gone unnoticed for a number of reasons. Analytical techniques, such
as LC/MS and LC/NMR, were not available in the past to
characterize these products that were once considered trivial
and nonidentifiable. In addition, because of their low abundance in bile and their close proximity to the GSH adducts
during chromatography, these peptides were easily overlooked (see Fig. 2). Numerous metabolism studies in the past
with radiolabeled compounds have shown the existence of
radioactive peaks in radiochromatograms that were not characterized. It is quite possible that some of the radioactivity
could have been due to products resulting from glutamic acid
pathway. With the advent of LC/MS, LC/NMR, and high field
NMR, it has become much easier to characterize compounds
that were once considered unidentifiable by previously existing techniques. LC/MS has been found to be useful tool in
characterizing polar metabolites such as the GSH adducts
(Pallante et al., 1986; Haroldsen et al., 1988, Draper et al.,
1989; Muck and Henion, 1990; Burlingame et al., 1992, and
the references cited therein; Baillie and Davis, 1993). However, there are limitations with MS in obtaining complete
Downloaded from jpet.aspetjournals.org at ASPET Journals on May 7, 2017
Fig. 6. 1H NMR of the synthetic dipeptide conjugate, 3-(␥-glutamylcystein-S-yl)acetaminophen. The LC/MS retention time and mass spectral
fragmentation pattern of this standard matched those of the metabolite M8.
744
Mutlib et al.
Vol. 294
structural information on these metabolites. More frequently, NMR has been found to be a complementary tool in
elucidating structures of complex metabolites such as those
described in this report. Metabolites formed as very polar
diconjugates have recently been confirmed by using a combination of such techniques (Spraul et al., 1993; Mutlib et al.,
1995, 1999; Shockcor et al., 1996; Tang and Abbott, 1996;
Lindon et al., 1997).
The characterization of these peptides provides very convincing evidence of the existence of a competing pathway
(“glutamic acid pathway”) for GSH conjugates. Although the
levels of these peptides appeared low in bile compared with
the GSH adducts and the mercapturic acid products, the
amount of these products formed intracellularly is not
known. The formation of tetrapeptides and pentapeptides
suggests the possibility that larger peptides were made by
cells (probably elongation of the peptide chain by the sequential addition of glutamic acid via the ␥-carboxylic acid linkage) that were not detected in this study. These probably
were not identified because they were confined to cells (i.e.,
could not be transported out) and hence were not eliminated
in bile. In vivo studies with M4 and in vitro studies with M5
incubated with rat bile or rat kidney/liver S9 fractions suggested that the mercapturic acid pathway was the predominant route for the disposition of GSH adducts. The contribution toward disposition of the GSH adducts by the glutamic
acid pathway appears to be minor in comparison with the
mercapturic acid pathway. A number of in-house discovery
compounds that were structurally different from acetamino-
phen and were capable of producing GSH adducts were found
to produce metabolites via the glutamic acid pathway.
The formation of such conjugates may represent yet another pathway by which drugs could produce covalent binding via their reactive metabolic intermediates. Previous reports have suggested the direct interaction of reactive
metabolic intermediate of compounds such as acetaminophen
with proteins and cellular components leading to covalent
binding and possibly toxicity (Jollow et al., 1973; Streeter et
al., 1984). The results from these studies suggest an alternate route for the interaction of reactive metabolites with
cellular components may exist. Because the cellular concentration of endogenous GSH is fairly high (approximately 5
mM in hepatocytes), there is a good possibility that a reactive
metabolic intermediate will interact with the thiolate of GSH
to a greater extent than with other protein components. This
in turn leads to the GSH adducts such as metabolites M5 of
acetaminophen. Subsequent processing of these adducts via
this novel metabolic pathway could lead to longer peptides
(e.g., M6 and M7) in which the incorporated moiety includes
the trapped reactive metabolic intermediate already linked
to cysteine or GSH. The consequence of having peptides such
as the ones described in this report is not known. It is possible that these peptides could disrupt cellular functions with
toxicological consequences. Furthermore, the retention of exogenous compounds linked to peptides could potentially have
immunological consequences such as the production of allergies or idiosyncratic reactions. The metabolic and physiological significance of existence of such drug-peptide conjugates
Downloaded from jpet.aspetjournals.org at ASPET Journals on May 7, 2017
Fig. 7. LC/MS/MS selected ion monitoring chromatogram (top) and MS/MS spectrum of the protonated parent ion (MH⫹ 400) of the 3-(␥glutamylcystein-S-yl)acetaminophen (M8) present in rat bile. Metabolite M8 had a retention time of 11.2 min. The LC/MS/MS data for the synthetic
standard were found to be identical with those produced by the metabolite.
2000
Disposition of Glutathione Adducts by Glutamic Acid Pathway
requires additional study. One must also consider the possibility that these conjugates have no function and that their
formation represents a side reaction catalyzed by the ␥-glutamyltranspeptidase.
The formation of peptide conjugates of acetaminophen was
confirmed by MS and NMR analyses of isolated metabolites.
It was shown that these peptides were formed by the addition
of (activated) glutamic acid via the ␥-carboxylic acid functional group to the ␣-amino group of the adducts. The glutamic acid pathway appears to function in the opposite manner to mercapturic acid pathway. Longer peptide conjugates
are produced via this metabolic pathway compared with the
mercapturic acid pathway, which essentially degrades the
GSH adducts to smaller molecules.
References
Ketterer B (1986) Detoxification reactions of glutathione and glutathione transferases. Xenobiotica 16:957–973.
King DS, Fields CG and Fields GB (1990) A cleavage method which minimizes side
reactions following Fmoc solid phase peptide synthesis. Int J Pept Prot Res 36:
255–266.
Koob M and Dekant W (1991) Bioactivation of xenobiotics by formation of toxic
glutathione conjugates. Chem-Biol Interact 77:107–136.
Lash LH and Anders MW (1986) Bioactivation and cytotoxicity of amino acid and
glutathione S-conjugates. Comments Toxicol 1:87–112.
Lindon JC, Nicholson JK, Sidelmann UG and Wilson ID (1997) Directly coupled
HPLC NMR and its application to drug metabolism. Drug Metab Rev 29:705–747.
Monks TJ and Lau SS (1989) Sulphur conjugate-mediated toxicity. Rev Biochem
Toxicol 10:41–90.
Monks TJ, Lau SS, Highet RJ and Gillette JR (1985) Glutathione conjugates of
2-bromohydroquinone are nephrotoxic. Drug Metab Dispos 13:553–559.
Muck WM and Henion JD (1990) High-performance liquid chromatography/tandem
mass spectrometry: Its use for the identification of stanozolol and its major metabolites in human and equine urine. Biomed Environ Mass Spectrom 19:37–51.
Mutlib AE, Chen H, Nemeth G, Gan L and Christ DD (1999) LC/MS and high field
NMR characterization of novel mixed diconjugates of the non-nucleoside HIV-1
reverse transcriptase inhibitor, efavirenz. Drug Metab Dispos 27:1045–1056.
Mutlib AE, Strupczewski J and Chesson S (1995) Application of hyphenated LC/
NMR and LC/MS techniques in rapid identification of in vitro and in vivo metabolites of iloperidone. Drug Metab Dispos 23:951–964.
Nash JA, King LJ, Lock EA and Green T (1984) The metabolism and disposition of
hexachloro-1:3-butadiene in the rat and its relevance to nephrotoxicity. Toxicol
Appl Pharmacol 73:124 –137.
Pallante SL, Lisek CA, Dulik DM and Fenselau C (1986) Glutathione conjugates:
Immobilized enzyme synthesis and characterization by fast bombardment mass
spectrometry. Drug Metab Dispos 14:313–318.
Shockcor JP, Wurm RM, Frick LW, Sanderson PN, Farrant RD, Sweatman BC and
Lindon JC (1996) HPLC NMR identification of the human urinary metabolites of
(⫺)-cis-5-fluoro-1-[2-(hydroxymethyl)-1,3-oxathiolan-5-yl]cytosine, a nucleoside
analogue active against human immunodeficiency virus (HIV). Xenobiotica 26:
189 –199.
Spraul M, Hofmann M, Wilson ID, Lenz E, Nicholson JK and Lindon JC (1993)
Coupling of HPLC with 19F- and 1H-NMR spectroscopy to investigate the human
urinary excretion of flurbiprofen metabolites. J Pharm Biomed Appl 11:1009 –
1015.
Streeter AJ, Dahlin DC, Nelson SD and Baillie TA (1984) The covalent binding of
acetaminophen to protein: Evidence for cysteine residues as major sites of arylation in vitro. Chem-Biol Interact 48:349 –366.
Tang W and Abbott FS (1996) Bioactivation of a toxic metabolite of valproic acid,
(E)-2-propyl-2,4-pentadienoic acid, via glucuronidation: LC/MS/MS characterization of the GSH-glucuronide diconjugates. Chem Res Toxicol 9:517–526.
Vamvaka S, Dekant W, Berthold K, Schmidt S, Wild D and Henschler D (1987)
Enzymic transformation of mercapturic acids derived from halogenated alkenes to
reactive and mutagenic intermediates. Biochem Pharmacol 36:2741–2748.
Weidolf LO, Lee ED and Henion JD (1988) Determination of boldenone sulfoconjugates and related steroid sulfates in equine urine by high performance liquid
chromatography/mass spectrometry. Biomed Environ Mass Spectrom 15:283–289.
Send reprint requests to: Dr. A. E. Mutlib, Drug Metabolism and Pharmacokinetics Section, DuPont Pharmaceuticals Company, P.O. Box 30, 1094 Elkton
Rd., Newark, DE 19714-0030. E-mail: [email protected]
Downloaded from jpet.aspetjournals.org at ASPET Journals on May 7, 2017
Abbott WA, Griffith OW and Meister A (1986) Glutamyl-glutathione: Natural occurrence and enzymology. J Biol Chem 261:13657–13661.
Anders MW, Elfarra AA and Lash LH (1987) Cellular effects of reactive intermediates: Nephrotoxicity of S-conjugates of amino acids. Arch Toxicol 60:103–108.
Anders MW, Lash L, Dekant W, Elfarra AA and Dohn DR (1988) Biosynthesis and
biotransformation of glutathione S-conjugates to toxic metabolites. CRC Crit Rev
Toxicol 18:311–342.
Baillie TA and Davis MR (1993) Mass spectrometry in the analysis of glutathione
conjugates. Biol Mass Spectrom 22:319 –325.
Burlingame AL, Baillie TA and Russell DH (1992) Mass spectrometry. Anal Chem
64:467R–502R (and references cited therein).
Commandeur JNM, Stijntjes GJ and Vermeulen NPE (1995) Enzymes and transport
systems involved in the formation and disposition of glutathione S-conjugates:
Role in bioactivation and detoxification mechanisms of xenobiotics. Pharmacol Rev
47:271–330.
Dahlin DC, Miwa GT, Lu AYH and Nelson SD (1984) N-Acetyl-P-benzoquinone
imine: A cytochrome P-450 mediated oxidation product of acetaminophen. Proc
Natl Acad Sci USA 81:1327–1331.
Draper WM, Brown FR, Bethem R and Miller MJ (1989) Thermospray mass spectrometry and tandem mass spectrometry of polar, urinary metabolites and metabolic conjugates. Biomed Environ Mass Spectrom 18:767–774.
Elfarra AA and Anders MW (1984) Renal processing of glutathione conjugates: Role
in nephrotoxicity. Biochem Pharmacol 33:3729 –3732.
Habig WH, Pabst MJ and Jakoby WB (1974) Glutathione S-transferases: The first
enzymatic step in mercapturic acid formation. J Biol Chem 249:7130 –7139.
Haroldsen PE, Reilly MH, Hughes H, Gaskell SJ and Porter CJ (1988) Characterization of glutathione conjugates by fast atom bombardment/tandem mass spectrometry. Biomed Environ Mass Spectrom 15:615– 621.
Hinchman CA andd Ballatori (1990) Glutathione-degrading capacities of liver and
kidney in different species. Biochem Pharmacol 40:1131–11135.
Hinson JA, Monks TJ, Hong M, Highet RJ and Pohl LR (1982) 3-(Glutathion-Syl)acetaminophen: A biliary metabolite of acetaminophen. Drug Metab Dispos
10:47–50.
Jollow DJ, Mitchell JR, Potter WZ, Davis DC, Gillette JR and Brodie BB (1973)
Acetaminophen-induced hepatic necrosis. II. Role of covalent binding in vivo.
J Pharmacol Exp Ther 187:195–202.
745