ISSN 0006-2979, Biochemistry (Moscow), 2024, Vol. 89, No. 10, pp. 1660-1680 © The Author(s) 2024. This article is an open access publication.
1660
REVIEW
ω-Amidase and Its Substrate α-Ketoglutaramate
(the α-Keto Acid Analogue of Glutamine)
as Biomarkers in Health and Disease
Arthur J. L. Cooper
1,a#
and Travis T. Denton
2,3,4,5,b,c
*
1
Department of Biochemistry and Molecular Biology, New York Medical College, Valhalla, NY 10595, USA
2
LiT Biosciences, 120 N Pine ST, Ste 242, Spokane, WA 99202-5029, USA
3
Department of Pharmaceutical Sciences, College of Pharmacy and Pharmaceutical Sciences,
Washington State University Health Sciences Spokane, Spokane, WA, USA
4
Department of Translational Medicine and Physiology, Elson S. Floyd College of Medicine,
Washington State University Health Sciences Spokane, Spokane, WA, USA
5
Steve Gleason Institute for Neuroscience, Washington State University Health Sciences Spokane, Spokane, WA, USA
a
e-mail: arthur_cooper@nymc.edu; ORCID 0000-0002-9143-8504
b
e-mail: travis.denton@litbiosciencesllc.com; ORCID 0000-0002-1222-2538 
c
e-mail: travis.denton@wsu.edu
Received July 12, 2024
Revised September 10, 2024
Accepted September 15, 2024
AbstractA large literature exists on the biochemistry, chemistry, metabolism, and clinical importance of the
α-keto acid analogues of many amino acids. However, although glutamine is the most abundant amino acid
in human tissues, and transamination of glutamine to its α-keto acid analogue (α-ketoglutaramate; KGM) was
described more than seventy years ago, little information is available on the biological importance of KGM.
Herein, we summarize the metabolic importance of KGM as an intermediate in the glutamine transaminase
ω-amidase (GTωA) pathway for the conversion of glutamine to anaplerotic α-ketoglutarate. We describe some
properties ofKGM, notably its occurrence as a lactam (2-hydroxy-5-oxoproline; 99.7% at pH7.2), and its presence
in normal tissues and body fluids. We note that the concentration of KGM is elevated in the cerebrospinal fluid
of liver disease patients and that the urinary KGM/creatinine ratio is elevated in patients with an inborn error
of the urea cycle and in patients with citrin deficiency. Recently, of the 607 urinary metabolites measured in a
kidney disease study, KGM was noted to be one of five metabolites that was most significantly associated with
uromodulin (a potential biomarker for tubular functional mass). Finally, we note that KGM is an intermediate
in the breakdown of nicotine in certain organisms and is an important factor in nitrogen homeostasis in some
microorganisms and plants. In conclusion, we suggest that biochemists and clinicians should consider KGM as
(i) akey intermediate in nitrogen metabolism in all branches of life, and (ii) a biomarker, along with ω-amidase,
inseveral diseases.
DOI: 10.1134/S000629792410002X
Keywords: ω-amidase, ammonia, anaplerosis, glutamine, glutaminase 1, glutaminase 2, glutamine addiction, glu-
tamine transaminases, 2-hydroxy-5-oxoproline, hyperammonemia, α-ketoglutarate, α-ketoglutaramate, methi-
onine, methionine salvage pathway, transamination
* To whom correspondence should be addressed.
# Deceased.
DEDICATION
This article is dedicated to my long-time friend
and mentor, Dr. Arthur J. L. Cooper. We suddenly lost
Dr. Cooper on May 30, 2024, at the age of 78, during
the time he and I were in the process of submitting
this article together. Arthur was, undeniably, one of
the world’s foremost experts in glutamine metabo-
lism, as well as countless other areas of biochemistry
and neuroscience. Arthur spent much time in the final
years of his life seeking to raise scientific awareness
ω-AMIDASE AND α-KETOGLUTARAMATE AS BIOMARKERS 1661
BIOCHEMISTRY (Moscow) Vol. 89 No. 10 2024
about the Glutamine Transaminase-omega-Amidase
(GTωA) pathway. It is in Arthurs memory that I pres-
ent this final publication highlighting the GTωA path-
way, co-authored with Arthur J. L. Cooper. You will
never be forgotten, my friend!
INTRODUCTION. THE DISCOVERY
OF ENZYME-CATALYZED TRANSAMINATION
A discussion of the discovery of enzyme-catalyzed
transamination provides a background for the discov-
ery of glutamine transamination. Enzyme-catalyzed
transamination [Eq.(1)] was discovered by Alexander E.
Braunstein and coworkers in the 1930s. For a review of
this work and its metabolic importance see [1]. Braun-
stein coined the word “umaminierung” (i.e., transami-
nation) to describe this process. Enzymes that catalyze
transamination reactions were originally referred to as
transaminases. However, currently, the more common
term for transaminases is aminotransferases. In this re-
view we will use the original term transaminase where
appropriate. The first transamination reactions discov-
ered by Braunstein and colleagues were between
L
-glu-
tamate and pyruvate [Eq.(2)] and between
L
-glutamate
and oxaloacetate [Eq. (3)]. The enzymes that catalyze
these reactions are often referred to as glutamate py-
ruvate transaminase (GPT) and glutamate oxaloacetate
transaminase (GOT), respectively, especially among cli-
nicians. The recommended Enzyme Commission terms
for these enzymes are alanine aminotransferase and
aspartate aminotransferase, respectively, where it is
understood that L-glutamate is the amino acid substrate
in the forward direction. Here, we use the abbrevia-
tions AlaAT and AspAT, respectively, for these enzymes.
L-Amino acid (1) + α-keto acid (2) ⇆
α-keto acid (1) +
L-amino acid (2)
(1)
L-Glutamate + pyruvate ⇆
α-ketoglutarate +
L-alanine
(2)
L-Glutamate + oxaloacetate ⇆
α-ketoglutarate +
L-aspartate
(3)
Mammalian tissues contain two isozymes of
AspAT – a mitochondrial form (mitAspAT) and a cy-
tosolic form (cytAspAT). Each is a homodimer. The pig
heart cytosolic and mitochondrial isozyme monomers
contain 412 and 401 residues, respectively ([1] and ref-
erences cited therein). In 1973, after enormous effort,
Braunstein and coworkers sequenced pig heart cytoso-
lic AspAT monomer [2] – at that time, it was the third
largest protein to have been sequenced! Braunstein
was the first to emphasize the importance of gluta-
mate/α-ketoglutarate-linked aminotransferases [Eq.(4)]
coupled to the glutamate dehydrogenase (GDH) reac-
tion [Eq.(5)] in generating ammonia from many amino
acids, or for assimilating ammonia nitrogen into many
amino acids in mammalian issues. He termed these
processes transdeamination and transreamination, re-
spectively [3]. The transdeamination reaction [forward
direction of Eq.(6)] is especially important in directing
nitrogen from the catabolism of many of the common
amino acids toward ammonia to be used for urea syn-
thesis in the liver [4]. Esmond Snell (e.g.,[5, 6]), Alton
Meister (e.g., [7, 8]) and their colleagues established
pyridoxal 5′-phosphate (PLP) as the cofactor involved
in transamination reactions.
L-Amino acid + α-ketoglutarate ⇆
α-keto acid +
L-glutamate
(4)
L-Glutamate + H
2
O + NAD
+
α-ketoglutarate + NADH +
+
NH
4
(5)
Net:
L-Amino acid + H
2
O + NAD
+
α-keto acid + NADH +
+
NH
4
(6)
Following the pioneering work of Braunstein and
colleagues it was soon established that transamination
in mammals is a major step in the metabolism of, for
example, aspartate, branched chain amino acids, γ-am-
inobutyric acid (GABA), tyrosine and phosphoserine.
However, discussion of transaminases involved in the
metabolism of these amino acids is beyond the scope
of the present review.
DISCOVERY AND PROPERTIES
OF L-GLUTAMINE TRANSAMINASES
Discovery. In the late 1940s, Alton Meister and col-
leagues at the National Institutes of Health discovered
transamination of L-glutamine to α-ketoglutaramate
(KGM) [Eq.(7)]. They also discovered an enzyme that
they named ω-amidase, which catalyzes the hydrolysis
of KGM to α-ketoglutarate and ammonia [Eq.(8)]. The
net reaction resulting from the combination of gluta-
mine transaminase and ω-amidase [Eq.(9)] was named
the glutaminaseII pathway [9-12].
L-Glutamine + α-keto acid ⇆
α-ketoglutaramate (KGM) +
L-amino acid
(7)
KGM + H
2
O → α-ketoglutarate +
+
NH
4
(8)
Net:
L-Glutamine + α-keto acid + H
2
O →
α-ketoglutarate +
L-amino acid +
+
NH
4
(9)
For a description of the background leading to
the discovery of the glutaminase II pathway see [13].
COOPER, DENTON1662
BIOCHEMISTRY (Moscow) Vol. 89 No. 10 2024
Fig.1. The glutamine transaminase-ω-amidase (GTωA) pathway. Glutamine transamination with a suitable α-keto acid sub-
strate is catalyzed by at least two glutamine transaminases in mammalian tissues– glutamine transaminaseK (GTK) and glu-
tamine transaminase L (GTL). The resulting α-keto acid (α-ketoglutaramate; KGM) is in equilibrium with a cyclic lactam form
(2-hydroxy- 5-oxoproline) [16, 17]. Under physiological conditions, the lactam predominates (~99.7%) over the open chain form
(~0.3%; ω-amidase substrate) [16, 17]. [Throughout this review, the term KGM is understood to be the sum of open-chain form
plus lactam, unless specified otherwise.] Note that although the transamination of glutamine is freely reversible (see[13]),
the pathway is irreversibly pulled in the direction of glutamine transamination by cyclization of KGM and by the action
ofω-amidase. Note also that not all enzyme reactants and products are depicted.
In order to prevent confusion between the name glu-
taminaseII and the “true” glutaminase (GLS2) we have
recently renamed the glutaminase II pathway as the
GTωA pathway [13]. The GTωA pathway is depicted
inFig.1.
Purification of glutamine transaminases. Meis-
ter and colleagues partially purified a glutamine trans-
aminase from rat liver and showed that the enzyme
exhibits wide specificity toward α-keto acids [9, 10].
Subsequently, Cooper and Meister purified the rat
liver enzyme to homogeneity and showed that it is
a homodimer, and PLP dependent [14]. The enzyme
was also shown to exhibit broad
L-amino acid speci-
ficity [14]. Cooper and Meister later showed that rat
tissues contain an additional glutamine transaminase
that is prominent in kidney and that they named gluta-
mine transaminaseK, to distinguish it from glutamine
transaminase L (GTL) – the predominant form in rat
liver [15]. Like the K-isozyme the
L-isozyme is a ho-
modimer and also has broad amino acid and α-keto
acid specificity.
These authors also noted that the activities are
present in cytosolic and mitochondrial fractions [15].
This was verified for GTK by Malherbe etal. [18] who
showed that the N-terminus of the rat enzyme possess-
es a 32 amino acid mitochondrial targeting sequence
that allows entry into the mitochondria; removal of
this sequence ensures a competing cytosolic location
[18]. The biological basis for the evolution of two gluta-
mine transaminases with such overlapping specificity
is currently unknown.
Nomenclature of glutamine transaminases and
subunit composition. As noted in [13], owing to the
broad substrate specificity of the
L
-glutamine transam-
inases, it is probable that some preparations of phe-
nylalanine, serine, histidine and aminoadipate trans-
aminases reported in the literature can be ascribed
to
L-glutamine transaminases. Of special note is the
finding that two kynurenine aminotransferase (KAT)
isozymes that have been described in the scientific lit-
erature, namely KAT1 and KAT3, are identical to GTK,
and GTL, respectively [13, 19-21]. In addition, KAT2
(aminoadipate aminotransferase) has some glutamine
transaminase activity [22].
Human GTK, which is annotated as KYAT1 in
UniProt, is a homodimer [23]. Three isoforms of hu-
man GTK/KYAT1, produced by alternative splicing, are
predicted in this database, namely Q16773-1 (canoni-
cal form; 422 amino acid residues; subunit molecular
mass 47,875 Da), Q16773-2 (canonical form missing res-
idues 68-117), and Q16773-3 (different sequence from
the canonical form at residues 230-250 and missing
ω-AMIDASE AND α-KETOGLUTARAMATE AS BIOMARKERS 1663
BIOCHEMISTRY (Moscow) Vol. 89 No. 10 2024
residues 251-422). The biological roles of Q16773-2
and3 are unknown.
Human GTL, which is annotated as KYAT3 in
UniProt, is also a homodimer [24] and is predicted to
exist as three alternatively spliced isoforms. Isoform1
(Q6YP21-1) is the canonical isoform (length 454 ami-
no acid residues; subunit molecular mass 51,400 Da)
and contains a 34 amino acid mitochondrial targeting
sequence. Isoform 3 (Q6YP21-3) lacks residues 1-34
of the canonical isoform. Isoforms 1 and 3 presum-
ably account for the presence of the enzyme in the
mitochondria and cytosol, respectively. In addition, a
third isoform (Q6YP21-2) is greatly truncated (missing
residues 168-454) and contains an altered sequence
152-167. The function of this polypeptide is unknown.
A relatively large amount of work has been car-
ried out on the four KATs described in the scientific
literature (i.e., KATs 1, 2, 3, and 4) in part because
the product of kynurenine transamination (i.e., ky-
nurenate) is a neuromodulator (e.g.,[25] and referenc-
es cited therein). According to[25], kynurenate is “an
endogenous antagonist of α7 nicotinic acetylcholine
and excitatory amino acid receptors, regulates gluta-
matergic, GABAergic, cholinergic and dopaminergic
neurotransmission in several regions of the rodent
brain”. The contribution of GTK/KAT1 and GTL/KAT3 to
the formation of cerebral neuromodulator kynurenate
is debatable and KAT2 is now regarded as the probable
main contributor to brain kynurenate formation [26].
Moreover, among the amino acids tested as substrates
for both GTK/KAT1 and GTL/KAT3, the highest catalytic
efficiency (V
max
/K
m
) occurs with L-glutamine. The cata-
lytic efficiency is much lower for kynurenine [19, 20].
In addition, L-glutamine is the most abundant amino
acid in human tissues and its concentration is estimat-
ed to be 70-80 g per 70 kg individual ([27] and refer-
ences cited therein). Given the molecular mass of
L
-glu-
tamine [147.13 AMU] and assuming 80% water content
in the body, the average concentration of
L
-glutamine
in human tissues is about 9 mM. On the other hand,
the concentration of kynurenine in mammalian tissues
is much lower. The concentration of kynurenine is re-
ported to be 20nmol/g (~22μM) in rat liver, with much
lower concentrations in the brain, lung, and spleen
[28]. Thus, the transamination of
L
-glutamine catalyzed
by GTK/KAT1 and GTL/KAT3 in vivo in human/mamma-
lian tissues is quantitatively likely to be orders of mag-
nitude greater than that of
L-kynurenine[13].
In addition to catalyzing transamination reactions,
both GTK and GTL catalyze competing β-elimination
reactions with many cysteine S-conjugates [Eq. (10)],
where -SR is a good leaving group [e.g., [13, 21, 29-32].
Because GTK was the first enzyme recognized to have
strong cysteine S-conjugate β-lyase activity ([32] and
references cited therein) the alternative abbreviation
CCBL1 (cysteine S-conjugate beta-lyase 1) is given in
UniProt for GTK/KAT1. This name is also used by some
commercial vendors. Inasmuch as GTL/KAT3 also has
strong cysteine S-conjugate β-lyase activity [30]. This
enzyme has been given the alternative name CCBL2 in
the UniProt and by some commercial vendors.
In addition, Commandeur et al. [29] have shown
that selenocysteine Se-conjugates are excellent trans-
aminase and β-lyase substrates of highly purified rat
liver GTK [Eq.(11)].
RSCH
2
CH(CO
2
)(NH
3
+
) + H
2
O ⇆
CH
3
C(O)CO
2
+
+
NH
4
+ RSH
(10)
RSeCH
2
CH(CO
2
)(NH
3
+
) + H
2
O ⇆
CH
3
C(O)CO
2
+
+
NH
4
+ RSeH
(11)
Substrate specificity of GTK and GTL. The wide
L-amino acid and α-keto acid specificity noted by Coo-
per and Meister for rat kidney GTK and rat liver GTL
[14, 15] was also noted for human KAT1 (GTK) [19] and
mouse KAT3 (GTL) [20]. In general, amino acid sub-
strates have the general structure RCH(CO
2
)NH
3
+
and
α-keto acid substrates have the structure RC(O)(CO
2
)
where R is a relatively hydrophobic, non-charged
group. Thus, for example, methionine and leucine (and
their respective α-keto acids) in addition to glutamine,
are good substrates of both enzymes [14, 15]. Valine
and isoleucine and their corresponding α-keto acids,
however, are poor substrates [14, 15, 19, 20], presum-
ably due to steric hindrance at the active site as a re-
sult of branching at the β position. Interestingly, gly-
oxylate [HC(O)(CO
2
)] (but not glycine [CH
2
(CO
2
)NH
3
+
])
is a good substrate of both GTK and GTL, presumably
because its small size allows ready entry at the ac-
tivesite.
METABOLIC IMPORTANCE
OF THE GTωA PATHWAY
Here we summarize the suggested metabolic role
of this pathway as recently reviewed in [13].
Closure of the methionine salvage pathway
[Eq. (12)]. Strong evidence suggests that the GTωA
pathway is largely responsible for closure of the methi-
onine salvage pathway in mammals ([13] and referenc-
es cited therein). Moreover, glutamine transaminase
and ω-amidase have been shown to act in tandem to
close the methionine salvage pathway in plants and
bacteria [33].
L-Glutamine + α-keto-γ-methiolbutyrate (KMB) +
H
2
O → α-ketoglutarate + L-methionine +
+
NH
4
(12)
Salvage of α-keto acids. We have proposed that
the glutamine transaminases may act as repair enzymes
COOPER, DENTON1664
BIOCHEMISTRY (Moscow) Vol. 89 No. 10 2024
to correct the “mistakes” of other transaminases. Such
mistakes may generate α-keto acids that are potential-
ly toxic/neurotoxic at elevated levels (e.g., phenylpy-
ruvate, p-hydroxyphenylpyruvate). Owing to the irre-
versibility of the GTωA pathway, the α-keto acid will be
favorably converted back to the less toxic amino acid,
while at the same time providing anaplerotic α-keto-
glutarate ([13] and references cited therein). Caligiore
etal. suggest that GTK may be especially important in
salvaging tryptophan from its corresponding α-keto
acid [34].
Detoxification of α-keto acids. Excess glyoxylate
may be toxic due to its potential conversion to oxalate.
Highly insoluble calcium oxalate is a major contributor
to kidney stones. It has been suggested that glutamine
transaminases/KATs may divert glyoxylate to glycine
in the kidney thereby lowering the possibility of the
conversion of glyoxylate to oxalate [35].
Possible antioxidant role of the GTωA pathway.
Many α-keto acids are oxidatively decarboxylated by
H
2
O
2
which, in the process, is converted to H
2
O  ([36]
and references quoted therein). For example, α-ke-
toglutarate is oxidized by H
2
O
2
to succinate and CO
2
[Eq.  (13)]. We have suggested that the GTωA path-
way, as a source of α-ketoglutarate, may play a role
in antioxidant defenses and upregulation of defense
systems under hypoxic conditions [13]. Moreover, it
has been recently shown that oxidation of specific
cysteine residues in ωA by H
2
O
2
inhibit ωAs catalytic
activity which, in turn, would decrease the produc-
tion of the α-ketoglutarate, making it less available to
neutralize H
2
O
2
[37]. In light of these new findings,
although α-ketoglutarate itself does react with H
2
O
2
,
it now seems less likely that the α-ketoglutarate,
which has come directly from the GTωA pathway,
would be the α-ketoglutarate acting as an intracellular
antioxidant.
O
2
CCH
2
CH
2
C(O)CO
2
+ H
2
O
2
O
2
CCH
2
CH
2
CO
2
+ CO
2
+H
2
O
(13)
Possible role of the GTωA pathway in the trans-
port of α-keto acids/-amino acids. A large number
of cellular and subcellular transporters have been de-
scribed for L-amino acids and α-keto acids. We have
suggested that the GTωA pathway, coupled to an α-ke-
toglutarate-linked aminotransferase, may assist in the
transport of
L-amino acids [13]. For example, consider
transamination of an
L
-amino acid with α-ketoglutarate
to the corresponding α-keto acid in compartment 1
and uptake of that α-keto acid into compartment 2.
TheGTωA pathway will ensure stoichiometric appear-
ance of the corresponding L-amino acid in compart-
ment 2 [Eqs. (14), (15)]. Removal of the α-keto acid in
compartment 2 via the GTωA pathway will ensure uni-
directional movement of the α-keto acid from compart-
ment 1 into compartment 2. The compartments may be
at the cellular or subcellular level.
Compartment 1:
L-Amino acid +
α-ketoglutarate ⇆ α-keto acid +
L-glutamate
(14)
Compartment 2:
L-Glutamine + α-keto acid +
H
2
O → α-ketoglutarate + L-amino acid +
+
NH
4
(15)
SUBSTRATE SPECIFICITY, DISTRIBUTION
AND STRUCTURE OF ω-AMIDASE
Substrate specificity. Meister and colleagues
showed that partially purified rat liver ω-amidase hy-
drolyzes the ω-monoamides of 4- and 5-C dicarboxyl-
ates such as KGM, α-ketosuccinamate (KSM; the α-keto
acid analogue of asparagine), glutaramate, succinamate
[9, 10]. However, neither glutamine nor asparagine was
found to be a substrate of this preparation of ω-ami-
dase [9, 10]. Thus, the -C(O)- in the α position may be
replaced by a -CH
2
- but not by a -CH(NH
3
+
)-. Subse-
quently, Hersh [16, 17] showed that rat liver ω-amidase
catalyzes hydroxaminolysis reactions with a number of
carboxamide substrates, in which water is replaced by
hydroxylamine as the attacking nucleophile, generating
the corresponding hydroxamate [Eq.  (16)]. In addition,
rat liver ω-amidase was shown to catalyze the hydro-
lysis of the terminal carboxylate esters of a number
of 5C- and 4C-dicarboxylates and some transamidation
reactions at the ω-carboxamide [16, 17]. For a detailed
discussion of these reactions see [38].
RC(O)NH
2
+ NH
2
OH → RC(O)NHOH +
+
NH
4
(16)
Assay procedures. Meister showed that partially
purified rat liver ω-amidase exhibits a relatively sharp
pH optimum at ~9.0 for KGM, yet a much broader pH
optimum (~5.0-9.0) with KSM as substrate, despite the
apparent close similarity in structure between these
two α-keto acids [11]. An explanation was provided by
Hersh using a continuous coupled enzyme assay pro-
cedure in which the formation of α-ketoglutarate from
KGM, catalyzed by ω-amidase [Eq. (8)], is reductively
aminated by the action of GDH in the presence of am-
monia and NADH– the reverse direction of Eq.(5) [13].
The disappearance of NADH is continuously monitored
spectrophotometrically at 340 nm (ε = 6220 M
–1
cm
–1
).
Hersh showed that, at the enzyme concentration used
in the assay, the rate of change of NAD
+
production at
pH values below 8.0 is biphasic [16]. From this, Hersh
was able to calculate the ratio of open-chain form
(0.3%; substrate) to the lactam form (99.7%; non sub-
strate). [Nevertheless, throughout this review, unless
stated otherwise, it is understood that the term KGM
refers to the sum of open-chain form plus lactam.]
ω-AMIDASE AND α-KETOGLUTARAMATE AS BIOMARKERS 1665
BIOCHEMISTRY (Moscow) Vol. 89 No. 10 2024
Hersh concluded that the rate of interconversion
between open-chain and cyclized forms of KGM is
OH-catalyzed. As a result, assays of ω-amidase with
KGM as substrate are typically carried out in buffers at
pH8.5-9-0 where rate of ring opening is unlikely to be
rate limiting for amounts of enzyme used in atypical
assay [39].
An alternative assay procedure uses succinamate
[R =
O
2
CH
2
CH
2
- in Eq. (16)] as substrate. ω-Amidase
catalyzes the conversion of succinamate to the cor-
responding hydroxamate in the presence of hydrox-
ylamine. The hydroxamate exhibits a brown color
(ε=920M
–1
cm
–1
at 535nm) when acidic ferric chloride
is added to the reaction mixture  [39]. The advantages
of this assay are that succinamate, unlike KGM, is
commercially available and the assay can be carried
out at physiological pH values (i.e., ~7.0-7.4). A third
assay procedure involves the use of KGM as a sub-
strate dissolved in buffers at pH8.5-9.0. The reaction is
quenched by addition of acidic 5mM 2,4-dinitrophenyl-
hydrazine. After incubation for a few minutes, the pH
of the reaction is adjusted by the addition of 1  M KOH
and the absorbance of the 2,4-dinitrophenylhydrazone
is measured at 430 nm (ε ~ 16,000 M
–1
cm
–1
) [39].
Distribution of ω-amidase. The enzyme activ-
ity is ubiquitously expressed in nature, attesting to
its central importance in nitrogen metabolism in the
biosphere. For example, Meister showed that ω-ami-
dase activity (KGM and KSM as substrates) is present
in all eight rat tissues investigated [11]. It was also
shown to be present in a Novikoff tumor, lettuce leaves,
spinach leaves, Escherichia coli and Streptococcus fae-
calis [11]. In another study, ω-amidase activity was
shown to be present in all ten rat organs investigated,
with highest specific activity in liver and kidney [40].
Prostate was not included in these studies, but lat-
er work showed that glutamine transaminase and
ω-amidase specific activities are exceptionally high in
that organ [41]. Lin et al noted the presence of Nit2
(nitrilase-like protein  2) protein in all sixteen human
tissues/cells investigated with highest levels in liver
and kidney [42]. [Nit2 is identical to ω-amidase (see be-
low)]. GeneCards indicates that the message for ω-ami-
dase/Nit2 is present in all 37 human tissues investi-
gated [43].
ω-Amidase is well represented in Neuropora cras-
sa where it is thought to be an important intermediate
in the cycling of glutamine nitrogen [44, 45]. In plants,
transamination of asparagine to KSM with glyoxylate
as co-substrate [Eq. (17)] followed by hydrolysis of
KSM, catalyzed by ω-amidase [Eq.(18)], is thought to
play a key role in photorespiration– the net reaction
is shown in Eq.(19) [46, 47].
L-Asparagine + glyoxylate →
α-ketosuccinamate (KSM) + glycine
(17)
KSM + H
2
O → oxaloacetate +
+
NH
4
(18)
Net:
L-Asparagine + glyoxylate + H
2
O →
oxaloacetate + glycine +
+
NH
4
(19)
Special role of an ω-amidase in microbial nico-
tine metabolism. Gram-positive soil bacteria Arthrobac-
ter nicotinovorans, Nocardioides sp. JS614 and Rhodo-
coccus opacus were shown to contain a similarly
organized gene cluster for the catabolism of nicotine
[48]. An intermediate in the catabolism of nicotine by
A. nicotinovorans was shown to be KGM and one of
the genes in the gene cluster was found to code for
an enzyme that hydrolyzes KGM to α-ketoglutarate and
ammonia [48]. The enzyme was found to have a mono-
mer mass of 32.3  kDa – only slightly larger than that
of the mammalian ω-amidase monomer (30.6  kDa; see
below)– and possess a characteristic EKC triad at the
active site [48]. Interestingly, the mammalian ω-ami-
dase is a homodimer (see below), whereas preliminary
data suggested that the A.  nicotinovorans enzyme is a
homotetramer [48].
Structure and location of mammalian ω-ami-
dase/Nit2. The enzyme is annotated as Nit2 in human
and mammalian gene data banks – see [49, 50]. The
human enzyme (Q9NQR4 (UniProt) [51]) is a homodi-
mer; each monomer consists of 276 amino acids with
a molecular mass of 30,608Da. Predicted three dimen-
sional structures for the human enzyme are report-
ed in [52] and [53] – the AlphaFold predicted struc-
ture [AF-Q9NQR4] is classified as “very high, with a
pLDDT>90. The enzyme active site on each monomer
possesses a reactive cysteine residue [16, 17], part of a
glutamate, lysine, cysteine triad (E43; K112; C153 in the
human enzyme) [52, 53].
Cellular and extracellular location of ω-Ami-
dase/Nit2. The enzyme is present in rat liver cytosol
[10,  40] and mitochondria [40,  54]. The mechanism
for the import of ω-amidase into mitochondria re-
quires a detailed study. Interestingly, UniProt lists a
protein– epididymis secretory sperm binding protein
Li  8a (Q9NQR4-1) that has a sequence identical to that
of ω-amidase and is present in the centrosome [51].
However, we are unable to find any additional informa-
tion on the apparent identity of ω-amidase/Nit2 with
this binding protein. Finally, it has been shown that
Nit2 (i.e., ω-amidase) is one of 1160 proteins identified
in urinary exosomes by the NHLBI Epithelial Systems
Biology Laboratory [55].
Possible post translational modification sites.
PhosphSitePlus® indicates about thirty potential sites
for post translational modifications of Nit2/ω-amidase
of which Y49 (phosphorylation), Y145 (phosphorylation)
and K249 (acetylation) are the most prominent [56].
In this context, it was previously noted that folate
deficiency in human colonocytes in culture resulted
COOPER, DENTON1666
BIOCHEMISTRY (Moscow) Vol. 89 No. 10 2024
in a remarkable, ~98% loss of Nit2 (i.e., ω-amidase)
protein, compared to that of control cultures as as-
sessed by staining intensity of the protein spot on a
2D gel [57]. Possibly, Nit2 is phosphorylated under low
folate conditions and moves to a different position on
the 2D gel. Moreover, in a proteomic study of MCF7
breast cancer cells, a shift in isoelectric point of Nit2
(i.e., ω-amidase) compared to normal breast cells was
noted, but only in those cancer cells overexpressing
ERBB2 (Erb-B2 receptor tyrosine kinase) [58]. Whether
phosphorylation of the enzyme under normal physi-
ological conditions occurs in vivo, and whether this
plays a role in regulating enzyme activity and location
requires a detailed study.
ω-AMIDASE/Nit2 IS A MEMBER
OF THE NITRILASE SUPERFAMILLY
The nitrilase superfamily of enzymes, which in-
cludes nitrilases, amidases and amidotransferases, has
been classified into 13 families [59,  60]. ω-Amidase/
Nit2 [along with Nit  1 (deaminated glutathione (dGSH)
amidase [61])] are the sole mammalian members of
family 10 of the nitrilase superfamily. All members of
the superfamily, including ω-amidase, contain a canon-
ical glutamate  (E), lysine  (K), cysteine  (C) triad at the
active site [60, 62]. Interestingly, this superfamily does
not include the two mammalian glutaminase isozymes.
The architecture of the active site in the two mamma-
lian glutaminase isozymes is different from those of
the nitrilase superfamily and the active site contains
a reactive serine residue rather than a reactive cys-
teine residue – see below. The importance of the EKC
triad (i.e., E43, K112, C153) for the catalytic activity in
human ω-amidase is shown by mutation studies. Sepa-
rate mutations of each of these residues to an alanine
residue results in an inactive enzyme  [52]. Moreover,
removal of the loop that forms a lid over the active
site (residues 116-128) also results in loss of activity
[52]. In addition to the EKC triad, according to Weber
et al. “All known nitrilase superfamily amidase and
carbamoylase structures have an additional glutamate
that is hydrogen bonded to the catalytic lysine” [62].
Geobacillus pallidus amidase catalyzes the hydrolysis
of simple aliphatic amides [62]. Both an E(142)L mu-
tation and an E(142)D mutation result in an inactive
enzyme [62]; see also [63]. Interestingly, one of the
substrates studied in the latter study was glutaramate,
which is also a substrate of ω-amidase [11, 17].
Yeast Nit2 (yNit2) is an ortholog of mamma-
lian Nit1 but NOT of mammalian Nit2. The crystal
structure of yeast Nit2 (yNit2) has been reported by
Liu et al. [64]. Confusingly, yNit2 is more related to
mammalian Nit1 than to mammalian Nit2 ([61, 64] and
references cited therein). A compound that is present
in the active site of yNit2 was tentatively identified as
GSH-like by Liu etal. [64]. Subsequently, Peracchi etal.
[61] showed that dGSH is a major substrate of both
yNit2 and mouse Nit1 (mNit1) [61]. KGM is a relatively
poor substrate of both yNit2 [61, 64] and mNit1 [64].
As noted by Perrachi etal. for mouse Nit2 (mNit2)[61]
“Sequence comparisons and structural data indicate
that the residues of the Nit2 catalytic pocket that sur-
round the substrate α-KGM are conserved in Nit1.”
A major difference is, however, found in the subsite
where the amido group of α-KGM is predicted to bind,
which is a quite small niche in the structure of Nit2,
but a much larger cavity in Nit1 – see also [65,  66].
These considerations are elaborated further in the
next section which discusses the mechanism of human
ω-amidase/Nit2.
Catalytic mechanism of ω-amidase/Nit2. Based
on a QM/MM study of hNit2 (i.e., human ω-amidase)
a theoretical four-step process that includes two tran-
sition states (TSs) at the active site during turnover
of KSM to oxaloacetate and ammonia has been sug-
gested [53]. This mechanism, depicted using KGM as
the substrate and the numbering of the human EKC
triad (i.e., Glu
43
, Lys
112
, Cys
153
), is depicted in Fig.  2.
In the first step, the amide group of the substrate is
activated by coordination with Lys
112
of the catalytic
triad, and the supporting Glu
128
residue, followed by
nucleophilic attack by Cys
153
. The stabilized tetrahedral
intermediate collapses to form the thioester interme-
diate with concomitant loss of ammonia/ammonium.
Subsequently, a water molecule, activated by coor-
dination to Glu
43
, attacks the thioester intermediate
releasing KG followed by redistribution of active site
residues for a subsequent round of catalysis. Curiously,
the authors mention that the enzyme has 276 resi-
dues (as also noted above for human ω-amidase and
in UniProt [51]) yet show the active site residues to be
Glu81, Lys150, and Cys191, which is shifted by 38 ami-
no acids. We assume that the authors recapitulate the
residues mentioned in the work of Barglowetal. [66]
and also assume that the 38 additional residues are a
product of the expression system used for production
of the enzyme (N-terminal hexahistidine fusion with
the gene 10 leader sequence).
ω-AMIDASE/Nit2 – CLINICAL ASPECTS
Differentiation of Crohn’s disease (CD) from
ulcerative colitis (UC). Burakoff et al. investigated
whether blood-based biomarkers can differentiate UC
from CD and noninflammatory diarrhea [67]. These
authors generated whole blood gene expression pro-
files for 21 patients with UC, 24 patients with CD, and
10 control patients with diarrhea, but without colon-
ic pathology. According to the authors: “A supervised
ω-AMIDASE AND α-KETOGLUTARAMATE AS BIOMARKERS 1667
BIOCHEMISTRY (Moscow) Vol. 89 No. 10 2024
Fig. 2. Proposed reaction mechanism for the hydrolysis of KGM by human ω-amidase. The catalytic triad residue numbers
reflect that on the human enzyme correlating to that of Chienetal. [52]. Molecular modeling of the mechanism was performed
by Silva Teixeiraetal. [53] and are the basis of the mechanism proposed.
learning method (logistic regression) was used to iden-
tify specific panels of probe sets which were able to
discriminate between UC and CD and from controls.
The UC panel consisted of the four genes, CD300A,
KPNA4, IL1R2, and ELAVL1; the CD panel comprised
the four genes CAP1, BID, NIT2, and NPL. These panels
clearly differentiated between CD and UC” [67]. Note
the inclusion of Nit2 (i.e., ω-amidase) in the CD panel.
ω-Amidase/Nit2 as a possible diagnostic biomarker in
CD requires further study.
Nit2 and Down syndrome. Myung et al. carried
out MALDI-TOF mass-spectrometric identification and
quantification of spots on 2D gel electrophoresis after
in-gel digestion of proteins using cortical brain sam-
ples from 7 controls and 7 samples from fetal Down
syndrome at the early second trimester  [68]. Three
proteins tentatively identified as reduced by 50% in
Down syndrome fetal brain versus normal fetal brain
were identified as Rik protein, mitochondrial inner
membrane protein and Nit2 [68]. The significance of
this finding also requires further study.
Role of ω-amidase/Nit2 in tumor progression.
Nit1, as noted above, is now known to be a dGSH de-
aminase. The enzyme catalyzes the hydrolysis of dGSH
to α-ketoglutarate and cysteinylglycine [61]. Consider-
able evidence suggests that Nit1 is a tumor suppres-
sor (e.g., [69-71]). Because Nit2 has moderate sequence
similarity to Nit1, Lin et al. considered the possibili-
ty that Nit2 (i.e., ω-amidase) is also a tumor suppres-
sor [42]. Ectopic expression of Nit2 in HeLa cells was
found to inhibit cell growth through G(2) arrest rather
than by apoptosis [42]. The authors also showed, using
proteomic and RT-PCR analysis, that Nit2 up-regulated
the protein and mRNA levels of 14-3-3sigma, an inhib-
itor of both G(2)/M progression and protein kinase B
(Akt)-activated cell growth, and down-regulated 14-3-
3beta, a potential oncogenic protein [42]. The authors
concluded that Nit2 may be a tumor suppressor can-
didate [42]. However, other workers have concluded
that Nit2 may be a cancer promoter in colorectal can-
cer [72]. Moreover, the authors suggested that Nit2
may be a promising target for the treatment of colon
cancer [72]. In another study, it was shown that the
expression of Nit2 in tongue squamous cell carcino-
ma is significantly higher than that in normal tongue
tissues [73]. The authors also suggested that Nit2 may
be a therapeutic target [73]. As noted above, GTK and
ω-amidase are well represented in rat prostate tissue
[41]. Furthermore, it was shown that protein levels
of these enzymes increased in human prostate cells
in culture in tandem with the aggressiveness of the
cancer [41]. We suggest that, on balance, ω-amidase
COOPER, DENTON1668
BIOCHEMISTRY (Moscow) Vol. 89 No. 10 2024
is a tumor promoter by providing anaplerotic α-ketoglu-
tarate to rapidly dividing cancer cells (see the next sec-
tion).
Glutamine addiction in tumors – the canonical
pathway. It is well known that many cancers utilize
glutamine as a major source of anaplerotic α-ketoglu-
tarate and nitrogen for DNA synthesis. This has been
termed “glutamine addiction” ([13] and references cited
therein). Most researchers in the field only consider one
pathway for the conversion of glutamine to α-ketoglu-
tarate. This pathway (the canonical pathway) consists
of the conversion of glutamine to glutamate via the
action of a glutaminase, followed by conversion of glu-
tamate to α-ketoglutarate by a transamination reaction
[Eqs. (20)-(22)]. Note that Eq.(21) depicts the reverse
direction of Eq.(4). Alternatively, glutamate may be con-
verted to α-ketoglutarate by the action of GDH [Eq.(5)].
L-Glutamine + H
2
O → L-glutamate +
+
NH
4
(20)
L-Glutamate + α-keto acid ⇆
α-ketoglutarate +
L-amino acid (21)
Net:
L-Glutamine + H
2
O + α-keto acid →
α-ketoglutarate +
L-amino acid +
+
NH
4
(22)
Human cells contain two glutaminase isozymes
encoded by GLS and GLS2. GLS encodes a kidney-type
glutaminase (KGA) – also known as GLS1 (or simply
GLS). GLS also encodes its active, shorter splice variant
(glutaminase C, GAC). GAC is present in many cancers
([74] and references quoted therein) and is thought
to be a major enzyme contributing to their glutamine
addiction. GLS2 encodes an enzyme (GLS2) that is
normally predominant in the liver and also exists as
splice variants– LGA and GAB (glutaminaseB) [75-77].
In addition, LGA exists as two isoforms, one of which
is more active than the other [76]. There appears to be
some confusion in the literature on the naming of the
more active isoform of GLS2 [75-77]. In their recent
article, Feng et al., use the term LGA to describe the
more active GLS2 isoform [74]. Unlike GLS1, which is a
tumor promoter in many cancers, GSL2 may be either
a tumor promoter or tumor suppressor [74, 78].
An Aside – Glutaminases involved in glutamine
addiction – can their catalytic properties reveal
insights into ω-amidase catalysis and regulation?
Many enzymes are activated by forming polymers
[79,  80]. Perhaps the most well studied such enzyme
is acetyl-CoA carboxylase ([80] and references cited
therein). The enzyme is active as a homodimer, but
its activity is greatly increased upon polymerization.
Inaddition, the activity is regulated in a complex fash-
ion by phosphate binding and by allosteric regulators
[80]. It is now becoming apparent that the mamma-
lian glutaminases are activated in a similar fashion.
GLS1 can exist as an inactive dimer as well as an
active tetramer [80-82]. Moreover, both glutaminase
isozymes are activated by phosphate and contain al-
losteric binding sites ([74] and references cited therein).
Inaddition, it is known that GLS1 (i.e., KGA), GAC (i.e.,
active shortened form of KGA) and GLS2 can exist
in active polymeric forms [74,  81-83]. Recently, Feng
et  al. demonstrated that the ability of GAC and GLS2
to form filaments is directly coupled to their catalytic
activity  [74]. The authors further noted that “Filament
formation guides an ‘activation loop’ to assume a spe-
cific conformation that works together with a ‘lid’ to
close over the active site and position glutamine for
nucleophilic attack by an essential serine.” The active
site of GLS1 is different from that of ω-amidase. For
example, as noted above, the active site of GLS/GAC
contains a reactive serine residue [74], whereas that of
ω-amidase contains a reactive cysteine residue [16,  52,
53]. However, a point of similarity is that both enzymes
require a lid to cover the active site for catalysis to
proceed [52,  53,  74]. Nothing is currently known about
possible regulation of ω-amidase activity. Nevertheless,
based on the fact that KGM and glutamine are almost
identical in size (differing in molecular mass by just
1  Da) and are monoamides of 5-C dicarboxylic acids,
it is possible that ω-amidase shares some regulatory
features with GLS1 and GLS2. This is another area that
requires detailed study.
Glutamine addiction in tumors – the alterna-
tive GTωA pathway for the formation of anaple-
rotic α-ketoglutarate. Several clinical studies have
been completed or are ongoing for the treatment of
various cancers with the allosteric glutaminase inhib-
itor CB-839 monotherapy or in combination therapy
with other drugs (e.g.,[84-87]). Whereas some of these
treatments seem to be well tolerated, they have only
been moderately successful. We have suggested that
this may be due, in part, to an alternative pathway for
the conversion of glutamine to α-ketoglutarate – i.e.,
the GTωA pathway (originally referred to as the gluta-
minase  II pathway), that can provide anaplerotic α-ke-
toglutarate to cancer cells when GLS1 is inhibited [13,
88]. Although this pathway has mostly been ignored
(or unrecognized) by cancer researchers, strong ev-
idence has recently been presented that the GTωA
pathway is a major source of anaplerotic α-ketogluta-
rate in pancreatic cancer. Thus, Udupa et al. showed
that genetic suppression of GTK in pancreatic tumors
(P198 shGTK-KD cancer cells injected in the back of
nude mice) led to complete suppression of pancreatic
tumorigenesis [89]. The authors suggested that a GTK
inhibitor may be useful, either alone or in combina-
tion with a GLS1 inhibitor, for the treatment of cancer
[89]. In another study, Pham et al., investigated the
isotopomer patterns in
L-glutamate generated from
L-[U-
13
C]glutamine, in orthotopic human Myc-amplified
ω-AMIDASE AND α-KETOGLUTARAMATE AS BIOMARKERS 1669
BIOCHEMISTRY (Moscow) Vol. 89 No. 10 2024
D425MED medulloblastoma tumor and showed that
this tumor preferentially uses the GTωA pathway over
the canonical GLS1 pathway to convert L-glutamine
to L-glutamate (in this case, via α-ketoglutarate) [90].
The authors noted that KYAT1 (i.e., the gene for GTK)
and its mRNA are upregulated in medulloblastoma
compared to other pediatric brain tumors in the Chil-
dren’s Brain Tumor Network/KidsFirst Pediatric Brain
Tumor Atlas RNAseq dataset [90]. In another study, as
noted above, not only does GLS1 protein increase in
prostate cancer cells in culture as the aggressiveness
of the tumor derived therefrom increases, but GTK
and ω-amidase proteins also increase concomitantly
[41]. It has been suggested that inhibitors of gluta-
mine transaminases may be useful anti-cancer agents,
perhaps in combination with GLS1 inhibitors [41, 89,
90]. Glutamine transaminase inhibitors are discussed
in the next section.
Glutamine transaminase inhibitors. Transition
state (TS) mimetics that inhibit KMB transamination
in cells in culture have been developed, including
L-methionine ethyl ester pyridoxal (MEEP) [91, 92].
Inasmuch as KMB is a substrate of both GTL and GTK,
it is reasonable to assume that these TS mimetic inhib-
itors will block L-glutamine transaminases. Interesting-
ly, MEEP was shown to induce DNA strand breaks in
HeLa cells typical of apoptotic cell death [92]. Some
inhibitors of KAT1 (i.e., GTK) have been synthesized
with K
i
values ranging from ~20 µM to ~1 mM [93] but,
to our knowledge, have not been tested as anti-can-
cer agents. The anti-cancer agent cisplatin is known
to be nephrotoxic at least in part due to formation of
the corresponding cysteine S-conjugate [94]. Mitochon-
drial AspAT is known to catalyze cysteine S-conjugate
β-lyase reactions [95]. Thus, it was reasoned that this
enzyme might catalyze a β-lyase reaction with the
cysteine S-conjugate of cisplatin. Overexpression of
mitAspAT in LLC-PK
1
cells led to increased toxicity of
cisplatin, increased platination in mitochondria and
increased inhibition of α-ketoglutarate dehydrogenase
complex, possibly due to close juxtapositioning of mi-
tAspAT to the enzyme complex [96].
In that study (i.e., [96]), GTK was not studied as
a possible β-lyase contributing to the nephrotoxicity.
Nevertheless, in a recent study, Sukeda etal. purified
human recombinant CCBL1 (i.e., GTK/KAT1) and used
a high-throughput screening assay (transamination be-
tween kynurenine and pyruvate) to screen chemical
libraries (not specified) for potential inhibitors [97].
2′,4′,6′-Trihydroxyacetophenone (THA; a naturally oc-
curring metabolite in Curcuma comosa) was identified
as a dose-dependent inhibitor of the human recombi-
nant GTK/KAT1 with an IC
50
of 13.2 μM [97]. The au-
thors stated that THA inhibited the β-lyase activity,
but inspection of their data shows that the inhibition
was actually of the transaminase activity [97]. Never-
theless, the authors noted some protective effects of
THA against cisplatin-induced toxicity toward mouse
kidney in vivo and LL-C-PK1 cells in culture. However,
the compound did not prevent the proliferation of can-
cer cell lines– LLC and MDA-MB-231 [97]. Sukeda etal.
also recognized other potential caveats. For example,
THA may have been protective by acting as an antiox-
idant. Nevertheless, THA could be the lead compound
in the search for compounds that will result in more
powerful glutamine transaminase inhibitors.
In conclusion, much work remains to be carried
out in the search for a selective and potent inhibitor
of GTK (and GTL). Such an inhibitor will be useful in
studies of the enzyme mechanism, but it may have
limitations as an anti-cancer agent due to the possi-
ble involvement of the glutamine transaminases in so
many physiologically relevant processes (see the above
discussion). For this reason, compounds designed to
disrupt the GTωA pathway may be better directed to-
ward inhibition of ω-amidase rather than toward in-
hibition of glutamine transaminases. We will return to
this point later.
ENDOGENOUS CONCENTRATIONS
OF α-KETOGLUTARAMATE (KGM) IN NORMAL
RAT TISSUES AND HUMAN BODY FLUIDS
Despite the fact that ω-amidase is inherently of
high specific activity in tissues, it is now well estab-
lished that KGM is an endogenous, natural metabolite.
This is presumably due to the cyclization of open-chain
(substrate) form to a lactam (inactive as a substrate),
and that the rate of ring opening is relatively slow at
neutral pH [16]. Because the rate of ring opening is hy-
droxide ion (
OH)-dependent, any pathological condi-
tion that will result in lowering of the cellular pH (e.g.,
hypercarbia [98]) will impede the ω-amidase reaction
with KGM even further.
The occurrence of KGM as a natural metabolite
was first observed in human cerebrospinal fluid (CSF).
In the procedure used by the investigators, endogenous
KGM was converted to α-ketoglutarate with ω-amidase
purified from rat liver. The generated α-ketoglutarate
was then measured fluorometrically with GDH in the
presence of NADH and ammonia [99-101]. The pres-
ence of KGM in human CSF was later verified by us-
ing a gas chromatography-mass spectrometry (GC-MS)
procedure [102]. KGM has also been measured in hu-
man urine by a GC-MS procedure [103, 104]– see also
below. Both GC-MS procedures measure the cyclized
lactam form of KGM. In another procedure, KGM was
measured in human CSF rat liver, rat kidney and rat
brain by an isotope dilution method [105]. Finally, a
high-performance liquid chromatography (HPLC) pro-
cedure has been developed to measure underivatized
COOPER, DENTON1670
BIOCHEMISTRY (Moscow) Vol. 89 No. 10 2024
KGM in rat tissues and plasma [106]. Taken together,
the data indicate that the concentration of KGM in nor-
mal human CSF is ~5 μM, whereas the concentration
in rat liver, kidney, brain, and plasma is ~8-216, 5-13,
6-11, and 19 μM, respectively [99, 101, 105, 106]. The
concentration in normal human urine is ~1-3 μmol
KGM/mmol creatinine [103]. Halámková and colleagues
showed that two separate spectrophotometric coupled
enzyme assays could be used to determine the con-
centration of KGM estimated to be present in human
serum [107]. The authors estimated that the concen-
tration of KGM in normal human plasma is ~5-10 μM
[107]– a concentration in the same order of magnitude
as that noted by Shurubor etal. of 19μM [103].
CLINICAL STUDIES IN WHICH KGM
WAS MEASURED/DETECTED IN BODY FLUIDS
Increased concentrations of KGM in the CSF
of patients with liver disease. Duffy and colleagues
showed that KGM occurs in human CSF and that the
concentration is increased in hyperammonemic pa-
tients with hepatic encephalopathy (HE) [99, 101, 105].
Moreover, the concentration was found to increase in
proportion to the severity of the disease. In some se-
vere cases of HE, CSF KGM concentrations exceeded
50  μM [99,  105]. A strong correlation was also noted
between concentrations of KGM and glutamine in the
CSF [99,  105]. Most likely, increased ammonia in the
HE patients resulted in increased tissue glutamine
concentrations, which in turn contributed to increased
transamination of glutamine. Mild neurotoxic effects
were noted upon infusion of KGM into rat CSF [99,
105]. This finding led Duffy etal. to suggest that KGM
may be a neurotoxic agent contributing to the enceph-
alopathy noted in HE patients [99,  101,  105]. However,
the concentration of KGM in the infusate (10  mM) is
>2000-fold greater than that expected in normal CSF
and 100-fold greater than the highest concentration
found in the CSF of an HE patient.
KGM exhibits a red color when spotted onto fil-
ter paper (or onto paper chromatograms) and then
sprayed with an ethanolic solution of Ehrlich’s reagent
[(4-dimethylamino)benzaldehyde] [108]. The procedure
is sensitive enough to detect KGM at a concentration
of 20  μM in 20  μL of CSF, spotted directly onto paper
without prior treatment [108]. The chemistry behind
this reaction is unknown, but the spot test may pos-
sibly serve as a starting point for the development of
a clinical test for KGM.
The KGM used in the studies of Duffy et al. was
synthesized by the method of Meister [11] in which
a solution of glutamine is incubated with snake ven-
om
L-amino acid oxidase in the presence of catalase,
followed by purification of KGM by cation exchange
chromatography. However, it has been known for more
than 75 years that glutamine spontaneously cyclizes
in solution to 5-oxoproline (5-OP) and formation of
ammonia [109]. Thus, even the best preparations of
KGM prepared by this enzymatic procedure contain a
few percent 5-OP. This compound is known to be en-
zymatically converted to
L-glutamate [110] – an amino
acid in the CSF suggested to be an excitotoxin even
at μM concentrations (e.g., [111,  112] and references
cited therein).
Increased urinary KGM/creatinine ratios in
primary and secondary hyperammonemia. Prima-
ry hyperammonemia is due to a defect that directly
affects enzymes or transporters of the urea cycle [113].
Secondary hyperammonemia occurs when the func-
tion of the urea cycle is inhibited by toxic metabolites
or by substrate deficiencies [113, 114]. The possibility
that increased production of KGM also occurs in pa-
tients who do not have overt liver disease but have
an inborn error of metabolism resulting in primary
hyperammonemia was investigated by Kuhara and col-
leagues [103]. The KGM/creatinine ratio was shown to
be markedly elevated in urine obtained from patients
with primary hyperammonemia due to an inherited
metabolic defect in any one of the five enzymes of
the urea cycle [103]. Kuhara etal. also noted increased
urinary KGM in three patients with a defect result-
ing in lysinuric protein intolerance and one of two
patients with a defect in the ornithine transporter  I
[103]. It was suggested that the increase in urinary
KGM concentrations in patients with primary hyper-
ammonemia is related to concomitant increase of glu-
tamine concentration [103]. On the other hand, urinary
KGM levels were not well correlated with secondary
hyperammonemia in patients with propionic acidemia
or methylmalonic acidemia, possibly as a result, in
part, of decreased glutamine levels [103].
In another study, Kuhara and colleagues noted an
increase in urinary KGM in most patients with second-
ary hyperammonemia resulting from citrin deficiency,
despite normal levels of urinary glutamine [104]. Citrin
is a hepatic mitochondrial aspartate-glutamate carrier
coded by the SLC25A13 gene and a defect in this gene
can lead to episodic hyperammonemia and disturbed
consciousness ([115] and references cited therein).
Frainay et al. have devised a network-based rec-
ommendation system to interpret and enrich metab-
olomics results that they have named MetaboRank
[116]. Interestingly, the authors state “...MetaboRank
recommended the overlooked α-ketoglutaramate as a
metabolite which should be added to the metabolic
fingerprint of HE…” [116].
In summary, KGM may be a useful biomarker for
many hyperammonemic diseases including hepatic en-
cephalopathy, inborn errors of the urea cycle, citrin de-
ficiency and lysinuric protein intolerance.
ω-AMIDASE AND α-KETOGLUTARAMATE AS BIOMARKERS 1671
BIOCHEMISTRY (Moscow) Vol. 89 No. 10 2024
Strong association of urinary KGM with uro-
modulin in chronic kidney disease patients. In a
recent study, it was shown that KGM is one of five me-
tabolites (out of 607) that strongly (P 2.11e
–440
) correlat-
ed with the protein uromodulin (UMOD) in a cohort
of kidney disease patients (n =  462) in Germany [117].
The kidney-specific UMOD has been suggested to be a
potential marker for tubular functional mass in pop-
ulation-based studies [117-121]. According to Pruijm
et al. [118] the associations of UMOD excretion with
markers of tubular functions and kidney dimensions
may reflect tubule activity in the general population.
It has also been suggested to be a marker for structur-
al integrity of the distal nephron, and renal function
[120, 121]. The reason for the association of KGM with
uromodulin is not obvious and requires further study.
KGM as a potential biomarker in serum ob-
tained from stroke victims. Sidorov etal. used me-
tabolomic analysis to prospectively analyze potential
serum biomarkers in acute ischemic (7 h) and chronic
(3-6 months) stages of 60 stroke victims [122]. The au-
thors noted the presence of four unknown metabolites
at the acute stage that were significantly associated
with infarct volume (all p< 0.01) [122]. Nine metabo-
lites at the chronic stage were found to significantly
associate with infarct volume, namely 3-indolepropio-
nate, α-ketoglutaramate, picolinate, and six unknowns
(all p<0.048) [122].
OTHER STUDIES INVOLVING
BIOLOGICALLY OCCURING KGM
Plasma and tissue KGM in rats treated with the
hepatotoxin thioacetamide. Shurobor et al. used an
HPLC method to show that the concentration of KGM
in normal rat plasma is ~19  μM [106,  123]. The au-
thors treated rats with various doses (200, 400, and
600mg/kg) of hepatotoxic thioacetamide and, after six
days of recovery, measured the concentration of sever-
al TCA cycle metabolites, and that of KGM, in the plas-
ma [123]. Despite the apparent physiological recovery
of the rats at this time, a metabolic imbalance was still
apparent [123]. Notably, the concentration of plasma
KGM was decreased by about 15-20% in rats treated
with all three doses, relative to that of a control [123].
The authors noted that the concentration of KGM in
liver and kidney, but not in brain, tended to decrease
in proportion to the amount of thioacetamide adminis-
tered [123]. The α-ketoglutarate/KGM ratio was signifi-
cantly increased in brain and kidney as the concentra-
tion of administered thioacetamide increased, but the
ratios in the kidney were not significantly different
from that of the control [123].
Increased serum KGM in dogs with exocrine
pancreatis insufficiency (EPI). EPI results from in-
sufficient secretion of pancreatic digestive enzymes
([124] and references cited therein). Barko etal., used
UPLC-MS/MS to carry out untargeted serum metabolo-
mics to identify metabolic disturbances associated with
EPI [124]. Fasted serum samples were collected from
dogs with EPI (n =  20) and healthy controls (n =  10),
all receiving pancreatic replacement therapy [122].
759 serum metabolites were detected, of which the con-
centration of 114 varied significantly (p<0.05, q <  0.2)
between dogs with EPI and healthy controls. Of note
was a marked increase in the concentration of serum
KGM [124]. The reason for this increase requires fur-
ther study.
ADDIONAL ASPECTS
OF ω-AMIDASE/KGM BIOLOGY
Wing and eye development in Drosophila mela-
nogaster. Knockdown of the gene CG813 in D.  mela-
nogaster, that encodes a homologue of mammalian
ω-amidase, results in severe eye- [125] and wing de-
fects [126]. It is not clear whether the defects are due
to a buildup of KGM that is specifically toxic to eye
and wing development, decreased synthesis of anaple-
rotic α-ketoglutarate that specifically affects wings and
eyes, or to an as yet unidentified moonlighting prop-
erty of ω-amidase required for correct eye and wing
development.
Role of KGM in stimulating nitrogen assimila-
tion in plants. Unkefer etal. have recently shown that
KGM (also known as 2-hydroxy-5-oxoproline) stimu-
lates nitrate uptake by plants and is a “likely signal for
ammonium assimilation” [127]. These authors suggest
that application of KGM to agricultural plants “would
potentially reduce the use of nitrate fertilizer per unit
of yield, leading to a further reduction in agriculture’s
carbon footprint” [127].
THE NEED FOR LARGE SCALE
SYNTHESIS OF KGM CHEAPLY
AND ON AN INDUSTRIAL SCALE
One of the reasons why ω-amidase has been so
little studied is that the substrate KGM is not available
commercially and, until recently, could only be made
enzymatically by oxidation of
L-glutamine with snake
venom L-amino acid oxidase in the presence of catalase
[11]. However, as noted above, such preparations are
invariably contaminated with small amounts of 5-oxop-
roline and trace amounts of α-ketoglutarate. Recently,
one of us (TTD) has devised a laboratory scale organic
synthesis of KGM that is not contaminated with 5-OP
and/or α-ketoglutarate [128]. Martinez and Unkefer have
patented another organic synthesis procedure [127].
COOPER, DENTON1672
BIOCHEMISTRY (Moscow) Vol. 89 No. 10 2024
Fig. 3. The importance of ω-amidase as a clinical marker; the central role of KGM as a clinical marker and as a marker of
nitrogen metabolism; the GTωA pathway as a source of anaplerotic α-ketoglutarate to cancer cells.
This procedure relies on the oxidation of glutamine or
5-oxoproline with Fremy’s salt. The yield is reported to
be high [129]. The possible advantages and disadvan-
tages of the procedure are discussed in [130]. In con-
clusion, (i)  given the potentially enormous agricultural
benefit of using KGM to stimulate nitrate assimilation
and (ii)  the need to study the biological and clinical
significance of ω-amidase/KGM, it is of utmost impor-
tance that methods be devised for the cheap, large-
scale synthesis of KGM.
ω-AMIDASE AND α-KETOGLUTARAMATE AS BIOMARKERS 1673
BIOCHEMISTRY (Moscow) Vol. 89 No. 10 2024
THE NEED FOR KGM LABLED
WITH HEAVY ISOTOPES
Most of the metabolomic studies mentioned above
that recognized KGM as an important intermediate
were carried out by Metabolon Inc. The company
does not currently have access to authentic KGM or
a suitable internal standard. Nevertheless, Metabolon
recognizes the KGM peak in the MS profile with a high
degree of certainty based on fragmentation pattern.
In other studies, Kuhara et al. measured KGM peak
heights relative to creatinine by using authentic KGM
but, in this case, γ-methyl KGM was used as an internal
standard. Given the growing recognition of the meta-
bolic and clinical importance of KGM, it is important
that KGM, labeled with one or more heavy isotopes, be
synthesized as an appropriate internal standard and
for quantitation.
THE NEED FOR SELECTIVE
INHIBITORS OF ω-AMIDASE
Given the many roles of the pathway in interme-
diary metabolism and especially in providing anaple-
rotic α-ketoglutarate to cancer cells, it will be very
important to develop inhibitors of the enzymes of the
GTωA pathway. As noted above, some researchers have
suggested that inhibitors of GTK may be useful anti-
cancer agents, perhaps in combination with a GLS1/
GAC inhibitor. Some inhibitors of GTK have been de-
scribed in the literature. These have not been tested
in a clinical setting but can perhaps be used as lead
compounds in the development of more potent inhib-
itors. However, as also noted, selective inhibitors of
ω-amidase may be even more useful as anticancer
agents, perhaps in combination with a GLS1/GAC in-
hibitor. Unfortunately, no selective and potent inhibi-
tor of ω-amidase is currently available, although some
strategies for the development of such inhibitors have
been suggested [130].
CONCLUSIONS
ω-Amidase is ubiquitously expressed in nature,
often in high levels, yet its biological roles remain
largely unappreciated by biomedical scientists. This
review emphasizes (i)  the central role of KGM in nitro-
gen metabolism, (ii)  ω-amidase and its substrate KGM
as clinical markers and (iii)the proposed importance of
the GTωA pathway as a source of anaplerotic α-ketoglu-
tarate to cancer cells. This is schematically shown in
Fig.  3. Finally, we emphasize the need for the devel-
opment of selective inhibitors of GTK and especially
of ω-amidase.
Abbreviations.
OH, Hydroxide ion; 2D, Two di-
mensional; 5-OP, 5-Oxoproline (synonym: pyrogluta-
mate); ω-Amidase,ω-Amidodicarboxylate amidohydro-
lase; AlaAT,alanine aminotransferase; AspAT,aspartate
aminotransferase; BID, BH3 interacting domain death
agonist is a member of the BCL-2 family; CAP1,adeny-
lyl cyclase-associated protein-1; CCBL1,cysteine S-con-
jugate β-lyase1; CCBL2,cysteine S-conjugate β-lyase2;
CD, Crohn disease; CD300A, Cluster of differentiation
300A; CSF,cerebrospinal fluid; dGSH, deaminated glu-
tathione; EKC, Glutamate (E), lysine (K), cysteine (C);
ELAVL1,Embryonic lethal abnormal vision [synonym:
or HuR (Drosophila-like Hu antigen R)]; EPI,exocrine
pancreatic insufficiency; GABA, γ-aminobutyric acid;
GAB,glutaminaseB; GAC,glutaminaseC; dGSH,deam-
inated glutathione; GC-MS, Gas chromatography-mass
spectrometry; GDH, glutamate dehydrogenase; GLS
or GLS1, glutaminase isozyme 1 (synonyms: kid-
ney type glutaminase); GLS2, glutaminase isozyme 2
(synonyms: liver type glutaminase); GTK, glutamine
transaminase K [synonyms: kynurenine aminotrans-
ferase 1; GTL, glutamine transaminase L [synonyms:
kynurenine aminotransferase 3; GTωA, glutamine
transaminase ω-amidase; GOT,glutamate oxaloacetate
transaminase; GPT, glutamate pyruvate transaminase;
GSH, glutathione; HE, hepatic encephalopathy; HPLC,
high performance liquid chromatography; IL1R2,
interleukin 1 receptor, type  II; KAT/KYAT, kynure-
nine aminotransferase; KGA, kidney-type glutaminase;
KGM, α-Ketoglutaramate (synonyms: 2-oxoglutara-
mate, 5-amino-2,5 dioxopentanoate); KMB, α-Keto-γ-
methiolbutyrate (synonyms: 4-methylthio-2-oxobu-
tanoate, 4-methylthio2-oxobutyrate); KOH, potassium
hydroxide; KPNA4, karyo pherin alpha 4 (synonym:
importin alpha 3); KSM, α-ketosuccinamate (syn-
onyms: 2-oxosuccinamate, 4-amino-2,4-oxobutanoate);
LGA, liver type glutaminase; MALDI-TOF, matrix as-
sisted laser desorption ionization-time of flight; MEEP,
L-methionine ethyl ester pyridoxal; NADH, nicoti-
namide adenine dinucleotide; mNit2, mouse nitri-
lase like protein 2; NPL, N-Acetylneuraminate pyru-
vate lyase; mRNA, messenger ribonucleic acid; Nit2,
Nitrilase like protein 2 (synonyms: ω-amidase); PLP,
pyridoxal 5′-phosphate; RT-PCR, reverse transcriptase-
polymerase chain reaction; TCA, tricarboxylic acid;
THA, 2′,A′,6′-Trihydroxyacetophenone; TS, transi-
tion state; UC, ulcerative colitis; UMOD, uromodulin;
UPLC-MS/MS, ultraperformance liquid chromatography-
tandem mass spectrometry; yNit2, yeast nitrilase like
protein2.
Acknowledgments. We thank Dr. Niklas Müller
and Dr. Flávia Rezende (Institute for Cardiovascular
Physiology, Goethe University, Frankfurt am Main,
Germany and German Center of Cardiovascular Re-
search (DZHK), Partner Site Rhein Main, Germany) for
helpful discussions.
COOPER, DENTON1674
BIOCHEMISTRY (Moscow) Vol. 89 No. 10 2024
Contributions. AJLC and TTD carried out a con-
siderable amount of work on the GTωA pathway de-
scribed in this review. AJLC wrote the first draft of
the manuscript. TTD drew the figures. Both authors
contributed to and approved the final version of the
manuscript.
Funding. This work was supported by ongoing in-
stitutional funding. No additional grants to carry out
or direct this particular research were obtained.
Ethics declarations. TTD, in association with
Dr.Albert Li, has created a Limited Liability Company
(LiT Biosciences, Spokane, WA, USA) to commercialize
products related to the GTωA pathway. However, LiT
Biosciences did not contribute funding to the writ-
ing of the current review or to any work previously
carried out by the authors. Thus, the authors declare
noconflict of interest.
Open Access. This article is licensed under a Cre-
ative Commons Attribution 4.0 International License,
which permits use, sharing, adaptation, distribution,
and reproduction in any medium or format, as long
as you give appropriate credit to the original author(s)
and the source, provide a link to the Creative Com-
mons license, and indicate if changes were made.
Theimages or other third party material in this article
are included in the article’s Creative Commons license,
unless indicated otherwise in a credit line to the mate-
rial. If material is not included in the article’s Creative
Commons license and your intended use is not permit-
ted by statutory regulation or exceeds the permitted
use, you will need to obtain permission directly from
the copyright holder. To view a copy of this license,
visit http://creativecommons.org/licenses/by/4.0/.
REFERENCES
1. Cooper, A.J., and Meister,A. (1989) An appreciation
of Professor Alexander E. Braunstein. The discovery
and scope of enzymatic transamination, Biochimie, 71,
387-404, https://doi.org/10.1016/0300-9084(89)90169-7.
2. Ovchinnikov, Y. A., Egorov, C. A., Aldanova, N. A.,
Feigina, M. Y., Lipkin, V. M., Abdulaev, N.G., Grishin,
E.V., Kiselev, A.P., Modyanov, N.N., Braunstein, A.E.,
Polyanovsky, O.L., and Nosikov, V.V. (1973) The com-
plete amino acid sequence of cytoplasmic aspartate
aminotransferase from pig heart, FEBS Lett., 29, 31-34,
https://doi.org/10.1016/0014-5793(73)80008-0.
3. Braunstein, A. E. (1957) Les Voies Principales De L’as-
similation Et Dissimilation De LAzote Chez Les Ani-
maux, in Advances in Enzymology and Related Areas
of Molecular Biology, pp. 335-389, https://doi.org/
10.1002/9780470122648.ch5.
4. Brosnan, J.T. (2000) Glutamate, at the interface between
amino acid and carbohydrate metabolism, J.Nutr., 130,
988S-990S, https://doi.org/10.1093/jn/130.4.988S.
5. Snell, E. E. (1993) From bacterial nutrition to en-
zyme structure: a personal odyssey, Annu. Rev. Bio-
chem., 62, 1-26, https://doi.org/10.1146/annurev.bi.
62.070193.000245.
6. Snell, E. E., and Jenkins, W. T. (1959) The mecha-
nism of the transamination reaction, J. Cell. Com-
par. Physiol., 54, 161-177, https://doi.org/10.1002/jcp.
1030540413.
7. Meister, A. (1955) Transamination, Adv. Enzymol.
Relat. Subj. Biochem., 16, 185-246, https://doi.org/
10.1002/9780470122617.ch4.
8. Meister, A. (1990) On the transamination of en-
zymes, Ann. NY Acad. Sci., 585, 13-31, https://doi.org/
10.1111/j.1749-6632.1990.tb28038.x.
9. Meister,A., and Tice, S.V. (1950) Transamination from
glutamine to alpha-keto acids, J.Biol. Chem., 187, 173-
187, https://doi.org/10.1016/S0021-9258(19)50942-5.
10. Meister, A., Sober, H. A., Tice, S. V., and Fraser, P. E.
(1952) Transamination and associated deamidation of
asparagine and glutamine, J.Biol. Chem., 197, 319-330,
https://doi.org/10.1016/S0021-9258(18)55681-7.
11. Meister,A. (1953) Preparation of enzymatic reactions
of the keto analogues of asparagine and glutamine,
J. Biol. Chem., 200, 571-589, https://doi.org/10.1016/
S0021-9258(18)71403-8.
12. Meister, A., and Otani, T. T. (1957) Omega-Amide and
omega-amino acid derivatives of alpha-ketoglutar-
ic and oxalacetic acids, J. Biol. Chem., 224, 137-148,
https://doi.org/10.1016/S0021-9258(18)65016-1.
13. Cooper, A.J.L., Dorai,T., Pinto, J.T., and Denton, T.T.
(2023) Metabolic heterogeneity, plasticity, and adapta-
tion to “glutamine addiction” in cancer cells: the role
of glutaminase and the GTωA [glutamine transami-
nase-ω-amidase (glutaminase II)] pathway, Biology,
12, 1131, https://doi.org/10.3390/biology12081131.
14. Copper, A. J. L., and Meister, A. (1972) Isolation and
properties of highly purified glutamine transaminase,
Biochemistry, 11, 661-671, https://doi.org/10.1021/
bi00755a001.
15. Cooper, A. J. L., and Meister, A. (1974) Isolation and
properties of a new glutamine transaminase from
rat kidney, J. Biol. Chem., 249, 2554-2561, https://
doi.org/10.1016/S0021-9258(19)42765-8.
16. Hersh, L. B. (1971) Rat liver omega-amidase. Purifi-
cation and properties, Biochemistry, 10, 2884-2891,
https://doi.org/10.1021/bi00791a014.
17. Hersh, L.B. (1972) Rat liver-amidase. Kinetic evidence
for an acyl-enzyme intermediate, Biochemistry, 11,
2251-2256, https://doi.org/10.1021/bi00762a007.
18. Malherbe,P., Alberati-Giani,D., Köhler,C., and Cesura,
A.M. (1995) Identification of a mitochondrial form of
kynurenine aminotransferase/glutamine transami-
naseK from rat brain, FEBS Lett., 367, 141-144, https://
doi.org/10.1016/0014-5793(95)00546-l.
19. Han, Q., Li,J., and Li, J. (2004) pH dependence, sub-
strate specificity and inhibition of human kynurenine
ω-AMIDASE AND α-KETOGLUTARAMATE AS BIOMARKERS 1675
BIOCHEMISTRY (Moscow) Vol. 89 No. 10 2024
aminotransferase I, Eur. J. Biochem., 271, 4804-4814,
https://doi.org/10.1111/j.1432-1033.2004.04446.x.
20. Han, Q., Robinson,H., Cai, T., Tagle, D.A., and Li, J.
(2009) Biochemical and structural properties of mouse
kynurenine aminotransferase III, Mol. Cell. Biol., 29,
784-793, https://doi.org/10.1128/mcb.01272-08.
21. Pinto, J. T., Krasnikov, B. F., Alcutt, S., Jones, M. E.,
Dorai,T., Villar, M. T., Artigues, A., Li, J., and Cooper,
A.J. (2014) Kynurenine aminotransferase III and glu-
tamine transaminase L are identical enzymes that
have cysteine S-conjugate β-lyase activity and can
transaminate L-selenomethionine, J. Biol. Chem., 289,
30950-30961, https://doi.org/10.1074/jbc.M114.591461.
22. Han, Q., Robinson,H., Cai, T., Tagle, D.A., and Li, J.
(2011) Biochemical and structural characterization
of mouse mitochondrial aspartate aminotransferase,
a newly identified kynurenine aminotransferase-IV,
Biosci. Rep., 31, 323-332, https://doi.org/10.1042/
bsr20100117.
23. URL: https://www.uniprot.org/uniprotkb/Q16773/entry.
24. URL: https://www.uniprot.org/uniprotkb/Q6YP21/entry.
25. Blanco Ayala, T., Lugo Huitrón, R., Carmona Apari-
cio, L., Ramírez Ortega, D., González Esquivel, D.,
Pedraza Chaverrí, J., Pérez de la Cruz, G., Ríos, C.,
Schwarcz, R., and Pérez de la Cruz, V. (2015) Alter-
native kynurenic acid synthesis routes studied in the
rat cerebellum, Front. Cell. Neurosci., 9, 178, https://
doi.org/10.3389/fncel.2015.00178.
26. Yu,P., Di Prospero, N.A., Sapko, M.T., Cai,T., Chen,A.,
Melendez-Ferro,M., Du,F., Whetsell, W.O.,Jr., Guidet-
ti, P., Schwarcz, R., and Tagle, D. A. (2004) Biochem-
ical and phenotypic abnormalities in kynurenine
aminotransferase II-deficient mice, Mol. Cell. Biol.,
24, 6919-6930, https://doi.org/10.1128/mcb.24.16.6919-
6930.2004.
27. Cruzat,V., Macedo Rogero,M., Noel Keane,K., Curi,R.,
and Newsholme, P. (2018) Glutamine: metabolism
and immune function, supplementation and clinical
translation, Nutrients, 10, 1564, https://doi.org/10.3390/
nu10111564.
28. Ohta,Y., Kubo,H., Yashiro,K., Ohashi,K., Tsuzuki,Y.,
Wada,N., Yamamoto,Y., and Saito,K. (2017) Effect of
water-immersion restraint stress on tryptophan catab-
olism through the kynurenine pathway in rat tissues,
J. Physiol. Sci., 67, 361-372, https://doi.org/10.1007/
s12576-016-0467-y.
29. Commandeur, J.N.M., Andreadou,I., Rooseboom,M.,
Out, M., de Leur, L. J., Groot, E., and Vermeulen,
N.P.E. (2000) Bioactivation of selenocysteine se-conju-
gates by a highly purified rat renal cysteine conjugate
β-lyase/glutamine transaminaseK, J.Pharmacol. Exp.
Ther., 294, 753-761.
30. Cooper, A.J., Krasnikov, B.F., Niatsetskaya, Z.V., Pinto,
J.T., Callery, P. S., Villar, M.T., Artigues, A., and Brus-
chi, S.A. (2011) Cysteine S-conjugate β-lyases: import-
ant roles in the metabolism of naturally occurring sul-
fur and selenium-containing compounds, xenobiotics
and anticancer agents, Amino Acids, 41, 7-27, https://
doi.org/10.1007/s00726-010-0552-0.
31. Cooper, A.J., Pinto, J. T., Krasnikov, B. F., Niatsetska-
ya, Z.V., Han,Q., Li,J., Vauzour,D., and Spencer, J. P.
(2008) Substrate specificity of human glutamine trans-
aminase K as an aminotransferase and as a cysteine
S-conjugate beta-lyase, Arch. Biochem. Biophys., 474,
72-81, https://doi.org/10.1016/j.abb.2008.02.038.
32. Lash, L.H., Nelson, R.M., Van Dyke, R.A., and Anders,
M.W. (1990) Purification and characterization of hu-
man kidney cytosolic cysteine conjugate beta-lyase ac-
tivity, Drug Metab. Dispos. Biol. Fate Chem., 18, 50-54.
33. Ellens, K.W., Richardson, L.G.L., Frelin,O., Collins,J.,
Ribeiro, C. L., Hsieh, Y.-f., Mullen, R.T., and Hanson,
A. D. (2015) Evidence that glutamine transaminase
and omega-amidase potentially act in tandem to close
the methionine salvage cycle in bacteria and plants,
Phytochemistry, 113, 160-169, https://doi.org/10.1016/
j.phytochem.2014.04.012.
34. Caligiore, F., Zangelmi, E., Vetro, C., Kentache, T.,
Dewulf, J.P., Veiga-da-Cunha,M., Van Schaftingen, E.,
Bommer,G., and Peracchi,A. (2022) Human cytosolic
transaminases: side activities and patterns of discrim-
ination towards physiologically available alternative
substrates, Cell. Mol. Life Sci., 79, 421, https://doi.org/
10.1007/s00018-022-04439-3.
35. Han, Q., Yang, C., Lu, J., Zhang, Y., and Li, J. (2019)
Metabolism of oxalate in humans: a potential role ky-
nurenine aminotransferase/glutamine transaminase/
cysteine conjugate betalyase plays in hyperoxaluria,
Curr. Med. Chem., 26, 4944-4963, https://doi.org/10.217
4/0929867326666190325095223.
36. Cooper,A., Ginos, J.Z., and Meister, A. (1983) Synthe-
sis and properties of the α-keto acids, Chem. Rev., 83,
321-358, https://doi.org/10.1021/cr00055a004.
37. Herrle, N., Malacarne, P. F., Warwick, T., Cabrera-
Orefice, A., Chen, Y., Gheisari, M., Chatterjee, S.,
Leisegang, M. S., Sarakpi, T., Wionski, S., Lopez, M.,
Koch, I., Keßler, M., Klein, S., Uschner, F. E., Trebic-
ka, J., Brunst, S., Proschak, E., Günther, S., Rosas-
Lemus, M., et al. (2024) The transaminase-ω-amidase
pathway is a redox switch in glutamine metabo-
lism that generates α-ketoglutarate, bioRxiv, https://
doi.org/10.1101/2024.08.28.610061.
38. Cooper, A. J., Shurubor, Y. I., Dorai, T., Pinto, J. T.,
Isakova, E.P., Deryabina, Y.I., Denton, T.T., and Kras-
nikov, B. F. (2016) ω-Amidase: an underappreciated,
but important enzyme in L-glutamine and L-aspar-
agine metabolism; relevance to sulfur and nitrogen
metabolism, tumor biology and hyperammonemic
diseases, Amino Acids, 48, 1-20, https://doi.org/10.1007/
s00726-015-2061-7.
39. Krasnikov, B.F., Nostramo,R., Pinto, J.T., and Cooper,
A.J. (2009) Assay and purification of omega- amidase/
Nit2, a ubiquitously expressed putative tumor
COOPER, DENTON1676
BIOCHEMISTRY (Moscow) Vol. 89 No. 10 2024
suppressor, that catalyzes the deamidation of the
alpha-keto acid analogues of glutamine and aspar-
agine, Anal. Biochem., 391, 144-150, https://doi.org/
10.1016/j.ab.2009.05.025.
40. Cooper, A. J. L. (1988) Glutamine aminotransferases
and ω-amidases, in Glutamine and Glutamate in Mam-
mals, CRC Press, Inc., Boca Raton, Florida, pp.33-52,
https://doi.org/10.1201/9781351072298-4.
41. Dorai,T., Dorai,B., Pinto, J.T., Grasso,M., and Cooper,
A. J. L. (2019) High levels of glutaminase II pathway
enzymes in normal and cancerous prostate suggest
a role in ‘glutamine addiction’, Biomolecules, 10, 2,
https://doi.org/10.3390/biom10010002.
42. Lin, C.-H., Chung, M.-Y., Chen, W.-B., and Chien, C.-H.
(2007) Growth inhibitory effect of the human NIT2
gene and its allelic imbalance in cancers, FEBS J.,
274, 2946-2956, https://doi.org/10.1111/j.1742-4658.
2007.05828.x.
43. GeneCards (2024) NIT2 Gene – Nitrilase Family
Member 2, URL: https://www.genecards.org/cgi-bin/
carddisp.pl?gene=NIT2.
44. Mora,J. (1990) Glutamine metabolism and cycling in
Neurospora crassa, Microbiol. Rev., 54, 293-304, https://
doi.org/10.1128/mr.54.3.293-304.1990.
45. Calderón,J., Morett,E., and Mora,J. (1985) Omega-ami-
dase pathway in the degradation of glutamine in
Neurospora crassa, J.Bacteriol., 161, 807-809, https://
doi.org/10.1128/jb.161.2.807-809.1985.
46. Ta, T. C., Joy, K. W., and Ireland, R. J. (1985) Role of
asparagine in the photorespiratory nitrogen metabo-
lism of pea leaves, Plant Physiol., 78, 334-337, https://
doi.org/10.1104/pp.78.2.334.
47. Zhang, Q., and Marsolais,F. (2014) Identification and
characterization of omega-amidase as an enzyme
metabolically linked to asparagine transamination
in Arabidopsis, Phytochemistry, 99, 36-43, https://
doi.org/10.1016/j.phytochem.2013.12.020.
48. Cobzaru, C., Ganas, P., Mihasan, M., Schleberger, P.,
and Brandsch, R. (2011) Homologous gene clusters
of nicotine catabolism, including a new ω-amidase
for α-ketoglutaramate, in species of three genera of
Gram-positive bacteria, Res. Microbiol., 162, 285-291,
https://doi.org/10.1016/j.resmic.2011.01.001.
49. Jaisson, S., Veiga-da-Cunha, M., and Van Schaftin-
gen, E. (2009) Molecular identification of ω-amidase,
the enzyme that is functionally coupled with gluta-
mine transaminases, as the putative tumor suppressor
Nit2, Biochimie, 91, 1066-1071, https://doi.org/10.1016/
j.biochi.2009.07.002.
50. Krasnikov, B. F., Chien, C. H., Nostramo, R., Pinto,
J.T., Nieves,E., Callaway,M., Sun,J., Huebner,K., and
Cooper, A. J. (2009) Identification of the putative tu-
mor suppressor Nit2 as omega-amidase, an enzyme
metabolically linked to glutamine and asparagine
transamination, Biochimie, 91, 1072-1080, https://
doi.org/10.1016/j.biochi.2009.07.003.
51. Q9NQR4 NIT2_HUMAN, URL: https://www.uniprot.org/
uniprotkb/Q9NQR4/entry#structure.
52. Chien, C.-H., Gao, Q.-Z., Cooper, A.J.L., Lyu, J.-H., and
Sheu, S.-Y. (2012) Structural insights into the catalyt-
ic active site and activity of human Nit2/ω-amidase:
kinetic assay and molecular dynamics simulation,
J.Biol. Chem., 287, 25715-25726, https://doi.org/10.1074/
jbc.M111.259119.
53. Silva Teixeira, C. S., Sousa, S. F., and Cerqueira, N.
(2021) An unsual Cys-Glu-Lys catalytic triad is respon-
sible for the catalytic mechanism of the nitrilase su-
perfamily: a QM/MM study on Nit2, Chemphyschem,
22, 796-804, https://doi.org/10.1002/cphc.202000751.
54. Cooper, A. J. L., and Meister, A. (1981) Comparative
studies of glutamine transaminases from rat tissues,
Compar. Biochem. Physiol. B, 69, 137-145, https://
doi.org/10.1016/0305-0491(81)90223-6.
55. Urinary Exosome Protein Database, URL: https://
esbl.nhlbi.nih.gov/UrinaryExosomes/.
56. PhosphoSitePlus, URL : https://www.phosphosite.org/
proteinAction?id=4207&showAllSites=true.
57. Duthie, S.J., Mavrommatis,Y., Rucklidge,G., Reid,M.,
Duncan, G., Moyer, M. P., Pirie, L. P., and Bestwick,
C. S. (2008) The response of human colonocytes to
folate deficiency in vitro: functional and proteom-
ic analyses, J. Proteome Res., 7, 3254-3266, https://
doi.org/10.1021/pr700751y.
58. Wang,D., Jensen, R.H., Williams, K.E., and Pallavicini,
M. G. (2004) Differential protein expression in MCF7
breast cancer cells transfected with ErbB2, neomycin
resistance and luciferase plus yellow fluorescent pro-
tein, Proteomics, 4, 2175-2183, https://doi.org/10.1002/
pmic.200300728.
59. Pace, H. C., and Brenner, C. (2001) The nitrilase su-
perfamily: classification, structure and function, Ge-
nome Biol., 2, Reviews0001, https://doi.org/10.1186/
gb-2001-2-1-reviews0001.
60. Brenner,C. (2002) Catalysis in the nitrilase superfam-
ily, Curr. Opin. Struct. Biol., 12, 775-782, https://doi.org/
10.1016/s0959-440x(02)00387-1.
61. Peracchi, A., Veiga-da-Cunha, M., Kuhara, T., Ellens,
K.W., Paczia,N., Stroobant,V., Seliga, A.K., Marlaire,S.,
Jaisson,S., Bommer, G.T., Sun,J., Huebner,K., Linster,
C. L., Cooper, A. J. L., and Van Schaftingen, E. (2017)
Nit1 is a metabolite repair enzyme that hydrolyzes de-
aminated glutathione, Proc. Natl. Acad. Sci. USA, 114,
E3233-E3242, https://doi.org/10.1073/pnas.1613736114.
62. Weber, B.W., Kimani, S.W., Varsani,A., Cowan, D.A.,
Hunter, R., Venter, G. A., Gumbart, J. C., and Sewell,
B. T. (2013) The mechanism of the amidases: mutat-
ing the glutamate adjacent to the catalytic triad in-
activates the enzyme due to substrate mispositioning,
J.Biol. Chem., 288, 28514-28523, https://doi.org/10.1074/
jbc.M113.503284.
63. Makumire, S., Su, S., Weber, B. W., Woodward, J. D.,
Wangari Kimani,S., Hunter,R., and Sewell, B.T. (2022)
ω-AMIDASE AND α-KETOGLUTARAMATE AS BIOMARKERS 1677
BIOCHEMISTRY (Moscow) Vol. 89 No. 10 2024
The structures of the C146A variant of the amidase
from Pyrococcus horikoshii bound to glutaramide
and acetamide suggest the basis of amide recognition,
J. Struct. Biol., 214, 107859, https://doi.org/10.1016/
j.jsb.2022.107859.
64. Liu, H., Gao, Y., Zhang, M., Qiu, X., Cooper, A. J. L.,
Niu, L., and Teng, M. (2013) Structures of enzyme-
intermediate complexes of yeast Nit2: insights into its
catalytic mechanism and different substrate specifici-
ty compared with mammalian Nit2, Acta Crystallogr. D
Biol. Crystallogr., 69, 1470-1481, https://doi.org/10.1107/
s0907444913009347.
65. Pace, H. C., Hodawadekar, S. C., Draganescu, A.,
Huang, J., Bieganowski, P., Pekarsky,Y., Croce, C. M.,
and Brenner, C. (2000) Crystal structure of the worm
NitFhit Rosetta Stone protein reveals a Nit tetram-
er binding two Fhit dimers, Curr. Biol., 10, 907-917,
https://doi.org/10.1016/s0960-9822(00)00621-7.
66. Barglow, K. T., Saikatendu, K. S., Bracey, M. H.,
Huey,R., Morris, G.M., Olson, A.J., Stevens, R.C., and
Cravatt, B.F. (2008) Functional proteomic and struc-
tural insights into molecular recognition in the ni-
trilase family enzymes, Biochemistry, 47, 13514-13523,
https://doi.org/10.1021/bi801786y.
67. Burakoff,R., Chao,S., Perencevich,M., Ying,J., Fried-
man,S., Makrauer,F., Odze,R., Khurana,H., and Liew,
C.-C. (2011) Blood-based biomarkers can differentiate
ulcerative colitis from Crohn’s disease and noninflam-
matory diarrhea, Inflamm. Bowel Dis., 17, 1719-1725,
https://doi.org/10.1002/ibd.21574.
68. Myung, J. K., Gulesserian, T., Fountoulakis, M., and
Lubec, G. (2003) Deranged hypothetical proteins Rik
protein, Nit protein 2 and mitochondrial inner mem-
brane protein, Mitofilin, in fetal Down syndrome
brain, Cell Mol. Biol., 49, 739-746.
69. Semba,S., Han, S.Y., Qin, H.R., McCorkell, K.A., Iliop-
oulos, D., Pekarsky, Y., Druck, T., Trapasso,F., Croce,
C. M., and Huebner, K. (2006) Biological functions
of mammalian Nit1, the counterpart of the inverte-
brate NitFhit Rosetta stone protein, a possible tumor
suppressor, J. Biol. Chem., 281, 28244-28253, https://
doi.org/10.1074/jbc.M603590200.
70. Lin, C., Zhang,J., Lu, Y., Li, X., Zhang, W., Zhang, W.,
Lin, W., Zheng, L., and Li, X. (2018) NIT1 suppresses
tumour proliferation by activating the TGFβ1–Smad2/3
signalling pathway in colorectal cancer, Cell Death &
Disease, 9, 263, https://doi.org/10.1038/s41419-018-0333-3.
71. Mittag,S., Wetzel,F., Müller, S.Y., and Huber,O. (2023)
The Rosetta Stone hypothesis-based interaction of the
tumor suppressor proteins Nit1 and Fhit, Cells, 12,
353, https://doi.org/10.3390/cells12030353.
72. Zheng, B. a., Chai,R., and Yu, X. (2015) Downregula-
tion of NIT2 inhibits colon cancer cell proliferation
and induces cell cycle arrest through the caspase-3
and PARP pathways, Int.J. Mol. Med., 35, 1317-1322,
https://doi.org/10.3892/ijmm.2015.2125.
73. Chen, S., Wang, Z., and Feng,C. (2020) NIT2 overex-
pression predicts poor prognosis in tongue squamous
cell carcinoma patients, Mol. Biol. Rep., 47, 1553-1561,
https://doi.org/10.1007/s11033-019-05197-5.
74. Feng,S., Aplin, C., Nguyen, T.-T. T., Milano, S. K., and
Cerione, R. A. (2024) Filament formation drives ca-
talysis by glutaminase enzymes important in cancer
progression, Nat. Commun., 15, 1971, https://doi.org/
10.1038/s41467-024-46351-3.
75. Buczkowska, J., and Szeliga, M. (2023) Two faces of
glutaminase GLS2 in carcinogenesis, Cancers, 15,
5566, https://doi.org/10.3390/cancers15235566.
76. Ferreira, I. M., Quesñay, J. E. N., Bastos, A. C. S.,
Rodrigues, C. T., Vollmar, M., Krojer, T., Strain-
Damerell,C., Burgess-Brown, N.A., von Delft,F., Yue,
W. W., Dias, S. M. G., and Ambrosio, A. L. B. (2021)
Structure and activation mechanism of the human
liver-type glutaminase GLS2, Biochimie, 185, 96-104,
https://doi.org/10.1016/j.biochi.2021.03.009.
77. Martín-Rufián,M., Tosina,M., Campos-Sandoval, J.A.,
Manzanares, E., Lobo, C., Segura, J. A., Alonso, F. J.,
Matés, J. M., and Márquez, J. (2012) Mammalian glu-
taminase Gls2 gene encodes two functional alterna-
tive transcripts by a surrogate promoter usage mech-
anism, PLoS One, 7, e38380, https://doi.org/10.1371/
journal.pone.0038380.
78. Katt, W. P., Lukey, M. J., and Cerione, R. A. (2017) A
tale of two glutaminases: homologous enzymes with
distinct roles in tumorigenesis, Fut. Med. Chem., 9,
223-243, https://doi.org/10.4155/fmc-2016-0190.
79. Park, C.K., and Horton, N.C. (2019) Structures, func-
tions, and mechanisms of filament forming enzymes:
a renaissance of enzyme filamentation, Biophys.
Rev., 11, 927-994, https://doi.org/10.1007/s12551-019-
00602-6.
80. Park, C.K., and Horton, N.C. (2020) Novel insights into
filament-forming enzymes, Nat. Rev. Mol. Cell Biol.,
21, 1-2, https://doi.org/10.1038/s41580-019-0188-1.
81. Nguyen, T.-T. T., Ramachandran, S., Hill, M. J., and
Cerione, R. A. (2022) High-resolution structures of
mitochondrial glutaminase C tetramers indicate con-
formational changes upon phosphate binding, J.Biol.
Chem., 298, 101564, https://doi.org/10.1016/j.jbc.2022.
101564.
82. Jiang, B., Zhang, J., Zhao, G., Liu, M., Hu, J., Lin, F.,
Wang, J., Zhao, W., Ma,H., Zhang, C., Wu, C., Yao, L.,
Liu,Q., Chen,X., Cao,Y., Zheng,Y., Zhang,C., Han,A.,
Lin,D., and Li,Q. (2022) Filamentous GLS1 promotes
ROS-induced apoptosis upon glutamine depriva-
tion via insufficient asparagine synthesis, Mol. Cell,
82, 1821-1835.e1826, https://doi.org/10.1016/j.molcel.
2022.03.016.
83. Adamoski,D., Dias, M. M., Quesñay, J. E.N., Yang, Z.,
Zagoriy, I., Steyer, A. M., Rodrigues, C. T., da Silva
Bastos, A.C., da Silva, B.N., Costa, R.K.E., de Abreu,
F. M. O., Islam, Z., Cassago, A., van Heel, M. G.,
COOPER, DENTON1678
BIOCHEMISTRY (Moscow) Vol. 89 No. 10 2024
Consonni, S.R., Mattei,S., Mahamid,J., Portugal, R.V.,
Ambrosio, A.L.B., and Dias, S.M.G. (2023) Molecular
mechanism of glutaminase activation through fila-
mentation and the role of filaments in mitophagy pro-
tection, Nat. Struct. Mol. Biol., 30, 1902-1912, https://
doi.org/10.1038/s41594-023-01118-0.
84. Harding, J.J., Telli,M., Munster,P., Voss, M.H., Infan-
te, J. R., DeMichele, A., Dunphy,M., Le, M. H., Molin-
eaux,C., Orford,K., Parlati,F., Whiting, S.H., Bennett,
M. K., Tannir, N. M., and Meric-Bernstam, F. (2021)
A phase I dose-escalation and expansion study of
telaglenastat in patients with advanced or metastatic
solid tumors, Clin. Cancer Res., 27, 4994-5003, https://
doi.org/10.1158/1078-0432.ccr-21-1204.
85. Meric-Bernstam, F., Tannir, N. M., Iliopoulos, O.,
Lee, R. J., Telli, M.L., Fan, A. C., DeMichele,A., Haas,
N. B., Patel, M. R., Harding, J. J., Voss, M. H., Owon-
ikoko, T. K., Carthon,B., Srinivasan, R., Bendell, J. C.,
Jenkins, Y., Whiting, S.H., Orford, K., Bennett, M. K.,
and Bauer, T. M. (2022) Telaglenastat plus cabozan-
tinib or everolimus for advanced or metastatic renal
cell carcinoma: an open-label phaseI trial, Clin. Can-
cer Res., 28, 1540-1548, https://doi.org/10.1158/1078-
0432.ccr-21-2972.
86. Lee, C. H., Motzer, R., Emamekhoo, H., Matrana, M.,
Percent,I., Hsieh, J.J., Hussain,A., Vaishampayan,U.,
Liu, S., McCune, S., Patel, V., Shaheen, M., Bendell, J.,
Fan, A. C., Gartrell, B. A., Goodman, O. B., Nikolina-
kos, P.G., Kalebasty, A.R., Zakharia,Y., Zhang,Z., etal.
(2022) Telaglenastat plus everolimus in advanced re-
nal cell carcinoma: a randomized, double-blinded,
placebo-controlled, phaseII entrata trial, Clin. Cancer
Res., 28, 3248-3255, https://doi.org/10.1158/1078-0432.
ccr-22-0061.
87. Tannir, N. M., Agarwal, N., Porta, C., Lawrence, N. J.,
Motzer,R., McGregor,B., Lee, R.J., Jain, R.K., Davis,N.,
Appleman, L. J., Goodman,O., Jr., Stadler, W.M., Gand-
hi, S., Geynisman, D. M., Iacovelli, R., Mellado, B.,
Sepúlveda Sánchez, J. M., Figlin, R., Powles, T., Akel-
la, L., et al. (2022) Efficacy and safety of telaglenas-
tat plus cabozantinib vs placebo plus cabozantinib
in patients with advanced renal cell carcinoma: the
CANTATA randomized clinical trial, JAMA Oncol., 8,
1411-1418, https://doi.org/10.1001/jamaoncol.2022.3511.
88. Cooper, A. J. L., Dorai, T., Pinto, J. T., and Denton,
T. T. (2022) α-Ketoglutaramate – a key metabolite
contributing to glutamine addiction in cancer cells,
Front. Med., 13, 1035335, https://doi.org/10.3389/
fmed.2022.1035335.
89. Udupa, S., Nguyen, S., Hoang, G., Nguyen, T.,
Quinones, A., Pham, K., Asaka, R., Nguyen, K.,
Zhang, C., Elgogary, A., Jung, J. G., Xu, Q., Fu, J.,
Thomas, A.G., Tsukamoto,T., Hanes,J., Slusher, B.S.,
Cooper, A.J. L., and Le, A. (2019) Upregulation of the
glutaminase II pathway contributes to glutamate
production upon glutaminase I inhibition in pancre-
atic cancer, Proteomics, 19, e1800451, https://doi.org/
10.1002/pmic.201800451.
90. Pham, K., Hanaford, A. R., Poore, B. A., Maxwell,
M. J., Sweeney, H., Parthasarathy, A., Alt, J., Rais, R.,
Slusher, B.S., Eberhart, C.G., and Raabe, E. H. (2022)
Comprehensive metabolic profiling of MYC-amplified
medulloblastoma tumors reveals key dependencies on
amino acid, tricarboxylic acid and hexosamine path-
ways, Cancers (Basel), 14, 1311, https://doi.org/10.3390/
cancers14051311.
91. Ogier, G., Chantepie, J., Deshayes, C., Chantegrel, B.,
Charlot, C., Doutheau, A., and Quash, G. (1993) Con-
tribution of 4-methylthio-2-oxobutanoate and its
transaminase to the growth of methionine-depen-
dent cells in culture. Effect of transaminase inhib-
itors, Biochem. Pharmacol., 45, 1631-1644, https://
doi.org/10.1016/0006-2952(93)90304-f.
92. Quash,G., Roch, A.M., Charlot,C., Chantepie,J., Thom-
as,V., Hamedi-Sangsari,F., and Vila,J. (2004) 4-meth-
ylthio 2-oxobutanoate transaminase: a specific target
for antiproliferative agents, Bull. Cancer, 91, E61-E79.
93. Nematollahi, A., Sun, G., Jayawickrama, G. S., and
Church, W. B. (2016) Kynurenine aminotransferase
isozyme inhibitors: a review, Int. J. Mol. Sci., 17, 946,
https://doi.org/10.3390/ijms17060946.
94. Townsend, D. M., Tew, K. D., He, L., King, J. B., and
Hanigan, M. H. (2009) Role of glutathione S-transfer-
ase Pi in cisplatin-induced nephrotoxicity, Biomed.
Pharmacother., 63, 79-85, https://doi.org/10.1016/
j.biopha.2008.08.004.
95. Cooper, A. J., Bruschi, S. A., Iriarte, A., and Martinez-
Carrion, M. (2002) Mitochondrial aspartate amino-
transferase catalyses cysteine S-conjugate beta-lyase
reactions, Biochem. J., 368, 253-261, https://doi.org/
10.1042/bj20020531.
96. Zhang, L., Cooper, A. J., Krasnikov, B. F., Xu,H., Bub-
ber, P., Pinto, J. T., Gibson, G. E., and Hanigan, M.H.
(2006) Cisplatin-induced toxicity is associated with plat-
inum deposition in mouse kidney mitochondria in vivo
and with selective inactivation of the alpha-ketogluta-
rate dehydrogenase complex in LLC-PK1 cells, Biochem-
istry, 45, 8959-8971, https://doi.org/10.1021/bi060027g.
97. Sukeda,N., Fujigaki, H., Ando,T., Ando, H., Yamamo-
to,Y., and Saito,K. (2023) Identification of 2′,4′,6′-trihy-
droxyacetophenone as promising cysteine conjugate
beta-lyase inhibitor for preventing cisplatin-induced
nephrotoxicity, Mol. Cancer Ther., 22, 873-881, https://
doi.org/10.1158/1535-7163.mct-22-0564.
98. Behar, K. L., Fitzpatrick, S. M., Hetherington, H. P.,
and Shulman, R.G. (1993) Cerebral metabolic studies
in vivo by combined 1H/31P and 1H/13C NMR spec-
troscopic methods, Acta Neurochir. Suppl., 57, 9-20,
https://doi.org/10.1007/978-3-7091-9266-5_2.
99. Vergara,F., Duffy, T.E., and Plum, F. (1973) Alpha-ke-
toglutaramate, a neurotoxic agent in hepatic coma,
Transact. Assoc. Am. Phys., 86, 255-263.
ω-AMIDASE AND α-KETOGLUTARAMATE AS BIOMARKERS 1679
BIOCHEMISTRY (Moscow) Vol. 89 No. 10 2024
100. Vemuganti, R., and Hazell, A. S. (2014) Mechanisms
of hepatic encephalopathy and thiamine deficiency,
Metab. Brain Dis., 29, 889-890, https://doi.org/10.1007/
s11011-014-9579-3.
101. Vergara,F., Plum,F., and Duffy, T.E. (1974) α-ketoglu-
taramate: increased concentrations in the cerebrospi-
nal fluid of patients in hepatic coma, Science, 183, 81-
83, https://doi.org/10.1126/science.183.4120.81.
102. Cooper, A. J. L., Dhar, A.K., Kutt,H., and Duffy, T. E.
(1980) Determination of 2-pyrrolidone-5-carboxyl-
ic and α-ketoglutaramic acids in human cerebro-
spinal fluid by gas chromatography, Anal. Biochem.,
103, 118-126, https://doi.org/10.1016/0003-2697(80)
90245-6.
103. Kuhara, T., Inoue, Y., Ohse, M., Krasnikov, B. F., and
Cooper, A. J. (2011) Urinary 2-hydroxy-5-oxoproline,
the lactam form of α-ketoglutaramate, is marked-
ly increased in urea cycle disorders, Anal. Bioanal.
Chem., 400, 1843-1851, https://doi.org/10.1007/s00216-
011-4688-x.
104. Kuhara,T., Ohse,M., Inoue,Y., and Cooper, A.J. (2011)
A GC/MS-based metabolomic approach for diagnosing
citrin deficiency, Anal. Bioanal. Chem., 400, 1881-1894,
https://doi.org/10.1007/s00216-011-4766-0.
105. Duffy, T.E., Cooper, A.J.L., and Meister,A. (1974) Iden-
tification of α-ketoglutaramate in rat liver, kidney, and
brain: relationship to glutamine transaminase and
ω-amidase activities, J. Biol. Chem., 249, 7603-7606,
https://doi.org/10.1016/S0021-9258(19)81280-2.
106. Shurubor, Y.I., Cooper, A.J., Isakova, E.P., Deryabina,
Y.I., Beal, M. F., and Krasnikov, B.F. (2016) HPLC de-
termination of α-ketoglutaramate [5-amino-2,5-dioxo-
pentanoate] in biological samples, Anal. Biochem., 494,
52-54, https://doi.org/10.1016/j.ab.2015.11.003.
107. Halámková, L., Mailloux, S., Halámek, J., Cooper,
A. J., and Katz, E. (2012) Enzymatic analysis of α-ke-
toglutaramate – a biomarker for hyperammonemia,
Talanta, 100, 7-11, https://doi.org/10.1016/j.talanta.
2012.08.022.
108. Cooper, A. J. L. (1978) Spot test for the detection of
α-ketoglutaramic acid in human cerebrospinal fluid,
Anal. Biochem., 90, 444-446, https://doi.org/10.1016/
0003-2697(78)90049-0.
109. Bray, H.G., James, S.P., Raffan, I.M., and Thorpe, W.V.
(1949) The enzymic hydrolysis of glutamine and its
spontaneous decomposition in buffer solutions, Bio-
chem.J., 44, 625-627.
110. Van der Werf,P., Orlowski,M., and Meister,A. (1971)
Enzymatic conversion of 5-oxo-L-proline (L-pyrroli-
done carboxylate) to L-glutamate coupled with cleav-
age of adenosine triphosphate to adenosine diphos-
phate, a reaction in the -glutamyl cycle, Proc. Natl.
Acad. Sci. USA, 68, 2982-2985, https://doi.org/10.1073/
pnas.68.12.2982.
111. Vannucci, R.C., Brucklacher, R.M., and Vannucci, S.J.
(1999) CSF glutamate during hypoxia-ischemia in the
immature rat, Brain Res. Dev. Brain Res., 118, 147-151,
https://doi.org/10.1016/s0165-3806(99)00142-x.
112. Hagberg,H., Thornberg, E., Blennow, M., Kjellmer, I.,
Lagercrantz, H., Thiringer, K., Hamberger, A., and
Sandberg,M. (1993) Excitatory amino acids in the ce-
rebrospinal fluid of asphyxiated infants: relationship
to hypoxic-ischemic encephalopathy, Acta Paediatrica,
82, 925-929, https://doi.org/10.1111/j.1651-2227.1993.
tb12601.x.
113. Häberle,J. (2013) Clinical and biochemical aspects of
primary and secondary hyperammonemic disorders,
Arch. Biochem. Biophys., 536, 101-108, https://doi.org/
10.1016/j.abb.2013.04.009.
114. Cooper, A. J., and Kuhara, T. (2014) α-Ketoglutara-
mate: an overlooked metabolite of glutamine and a
biomarker for hepatic encephalopathy and inborn er-
rors of the urea cycle, Metab. Brain Dis., 29, 991-1006,
https://doi.org/10.1007/s11011-013-9444-9.
115. Komatsu, M., Tanaka, N., Kimura,T., and Yazaki, M.
(2023) Citrin deficiency: clinical and nutritional fea-
tures, Nutrients, 15, 2284, https://doi.org/10.3390/
nu15102284.
116. Frainay, C., Aros, S., Chazalviel, M., Garcia, T., Vin-
son, F., Weiss, N., Colsch, B., Sedel, F., Thabut, D.,
Junot, C., and Jourdan, F. (2019) MetaboRank: net-
work-based recommendation system to interpret
and enrich metabolomics results, Bioinformatics,
35, 274-283, https://doi.org/10.1093/bioinformatics/
bty577.
117. Bächle, H., Sekula, P., Schlosser, P., Steinbrenner, I.,
Cheng, Y., Kotsis, F., Meiselbach, H., Stockmann, H.,
Schönherr, S., Eckardt, K. U., Devuyst, O., Scher-
berich,J., Köttgen,A., and Schultheiss, U.T. (2023) Uro-
modulin and its association with urinary metabolites:
the German Chronic Kidney Disease Study, Nephrol.
Dial. Transplant., 38, 70-79, https://doi.org/10.1093/
ndt/gfac187.
118. Pruijm, M., Ponte, B., Ackermann, D., Paccaud, F.,
Guessous, I., Ehret, G., Pechère-Bertschi, A., Vogt,B.,
Mohaupt, M. G., Martin, P.Y., Youhanna, S. C., Näge-
le, N., Vollenweider, P., Waeber, G., Burnier, M.,
Devuyst, O., and Bochud, M. (2016) Associations of
urinary uromodulin with clinical characteristics
and markers of tubular function in the general pop-
ulation, Clin. J. Am. Soc. Nephrol., 11, 70-80, https://
doi.org/10.2215/cjn.04230415.
119. Devuyst,O., Olinger, E., and Rampoldi,L. (2017) Uro-
modulin: from physiology to rare and complex kid-
ney disorders, Nat. Rev. Nephrol., 13, 525-544, https://
doi.org/10.1038/nrneph.2017.101.
120. Scherberich, J. E., Gruber, R., Nockher, W. A., Chris-
tensen, E.I., Schmitt,H., Herbst,V., Block,M., Kaden,J.,
and Schlumberger, W. (2018) Serum uromodulin-
a marker of kidney function and renal parenchy-
mal integrity, Nephrol. Dial. Transplant., 33, 284-295,
https://doi.org/10.1093/ndt/gfw422.
COOPER, DENTON1680
BIOCHEMISTRY (Moscow) Vol. 89 No. 10 2024
121. Schaeffer, C., Devuyst, O., and Rampoldi, L. (2021)
Uromodulin: roles in health and disease, Annu. Rev.
Physiol., 83, 477-501, https://doi.org/10.1146/annurev-
physiol-031620-092817.
122. Sidorov, E. V., Bejar, C., Xu, C., Ray, B., Gordon, D.,
Chainakul,J., and Sanghera, D.K. (2021) Novel metab-
olites as potential indicators of ischemic infarction
volume: a pilot study, Translat. Stroke Res., 12, 778-
784, https://doi.org/10.1007/s12975-020-00876-z.
123. Shurubor, Y. I., Rogozhin, A. E., Isakova, E. P., Dery-
abina, Y. I., and Krasnikov, B. F. (2023) Tricarboxylic
acid metabolite imbalance in rats with acute thioac-
etamide-induced hepatic encephalopathy indicates
incomplete recovery, Int.J. Mol. Sci., 24, 1384, https://
doi.org/10.3390/ijms24021384.
124. Barko, P. C., Rubin, S. I., Swanson, K. S., McMichael,
M. A., Ridgway, M. D., and Williams, D.A. (2023) Un-
targeted analysis of serum metabolomes in dogs with
exocrine pancreatic insufficiency, Animals, 14, 2313,
https://doi.org/10.3390/ani13142313.
125. Pletcher, R.C., Hardman, S.L., Intagliata, S.F., Lawson,
R. L., Page, A., and Tennessen, J. M. (2019) A genetic
screen using the drosophila melanogaster trip rnai
collection to identify metabolic enzymes required
for eye development, G3, 9, 2061-2070, https://doi.org/
10.1534/g3.119.400193.
126. Rotelli, M. D., Bolling, A. M., Killion, A. W., Weinberg,
A.J., Dixon, M.J., and Calvi, B.R. (2019) An RNAi screen
for genes required for growth of drosophila wing tissue,
G3, 9, 3087-3100, https://doi.org/10.1534/g3.119.400581.
127. Unkefer, P.J., Knight, T. J., and Martinez, R. A. (2023)
The intermediate in a nitrate-responsive ω-amidase
pathway in plants may signal ammonium assimila-
tion status, Plant Physiol., 191, 715-728, https://doi.org/
10.1093/plphys/kiac501.
128. Shen,D., Kruger, L., Deatherage,T., and Denton, T. T.
(2020) Synthesis of α-ketoglutaramic acid, Anal. Bio-
chem., 607, 113862, https://doi.org/10.1016/j.ab.2020.
113862.
129. Martinez, R. A., and Unkefer, P. J. (2001) Preparation of
2-Hydroxy-5-Oxoproline and Analogs Thereof, Patent,
Los Alamos National Security LLC, United States.
130. Denton, T. T., and Cooper, A. J. L. (2023) Chemistry,
biochemistry and clinical relevance of the glutamine
metabolite α-ketoglutaramate/2-hydroxy-5-oxoproline,
Austr. J. Chem., 76, 361-371, https://doi.org/10.1071/
CH22264.
Publishers Note. Pleiades Publishing remains
neutral with regard to jurisdictional claims in published
maps and institutional affiliations. AI tools may have
been used in the translation or editing of this article.