Skip to main content
International Journal of Tryptophan Research : IJTR logoLink to International Journal of Tryptophan Research : IJTR
. 2009 Jan 8;2:1–19. doi: 10.4137/ijtr.s2097

Kynurenine Pathway Metabolites in Humans: Disease and Healthy States

Yiquan Chen 1, Gilles J Guillemin 1,2,
PMCID: PMC3195227  PMID: 22084578

Abstract

Tryptophan is an essential amino acid that can be metabolised through different pathways, a major route being the kynurenine pathway. The first enzyme of the pathway, indoleamine-2,3-dioxygenase, is strongly stimulated by inflammatory molecules, particularly interferon gamma. Thus, the kynurenine pathway is often systematically up-regulated when the immune response is activated. The biological significance is that 1) the depletion of tryptophan and generation of kynurenines play a key modulatory role in the immune response; and 2) some of the kynurenines, such as quinolinic acid, 3-hydroxykynurenine and kynurenic acid, are neuroactive. The kynurenine pathway has been demonstrated to be involved in many diseases and disorders, including Alzheimer’s disease, amyotrophic lateral sclerosis, Huntington’s disease, AIDS dementia complex, malaria, cancer, depression and schizophrenia, where imbalances in tryptophan and kynurenines have been found. This review compiles most of these studies and provides an overview of how the kynurenine pathway might be contributing to disease development, and the concentrations of tryptophan and kynurenines in the serum, cerebrospinal fluid and brain tissues in control and patient subjects.

Introduction

Tryptophan is one of the 9 essential amino acids that the human body is incapable of synthesizing and thus, has to be obtained through external sources. Once absorbed by the body, tryptophan travels around the periphery circulation either bound to albumin or in free form, the two states existing in equilibrium, with the former accounting for up to 90%.1 However, tryptophan can only be transported across the blood brain barrier in its free form by the competitive and non-specific L-type amino acid transporter.2 Once in the central nervous system (CNS), tryptophan acts as a precursor to various metabolic pathways. This versatility results in different end-products, such as protein, serotonin and kynurenines.3 In both the peripheral and central systems, the kynurenine pathway represents a major route for the metabolism of tryptophan.

Following the kynurenine pathway (Fig. 1), tryptophan is oxidized by cleavage of the indole-ring, initiated either by tryptophan 2,3-dioxygenase (TDO), indoleamine 2,3-dioxygenase 1 (IDO-1) or IDO-2, a newly discovered IDO related enzyme.47 TDO resides primarily in the liver and is induced by tryptophan or corticosteroids.4 IDO-1, on the other hand, is the predominant enzyme extra-hepatically and can be found in numerous cells, including macrophages, microglia, neurons and astrocytes.811 It is up-regulated by certain cytokines and inflammatory molecules, such as lipopolysaccharides, amyloid peptides and human immunodeficiency virus (HIV) proteins,5,12,13 but its most potent stimulant is interferon gamma (IFN-γ).14,15 IFN-γ is able to induce both the gene expression and enzymatic activity of IDO-1.16,17 Recently, an IDO related enzyme, IDO-2, was identified.7,6 The encoding genes for IDO-1 and IDO-2 are located next to each other and IDO-2 possesses similar structural and enzymatic activities as IDO-1. However, IDO-2 differs in its expression pattern and signalling pathway and is preferentially inhibited by D-1-methyl-tryptophan.7,6

Figure 1.

Figure 1.

A schematic diagram of the kynurenine pathway.

As tryptophan proceeds along the kynurenine pathway to achieve the final product, nicotinamide adenosine dinucleotide (NAD), kynurenine is the first stable intermediate formed. Subsequently, several neuroactive intermediates are generated. These comprise the free-radical generator, 3-hydroxyanthranilic acid,18 the excitotoxin and N-methyl-D-aspartic acid (NMDA) receptor agonist, quinolinic acid,19 the NMDA antagonist, kynurenic acid,20 and the neuroprotectant, picolinic acid.21

During an immune response, the release of IFN-γ by activated T cells and leukocytes leads to an accelerated and sustained degradation of tryptophan. This significance was first speculated to be a defence mechanism that starved tumour cells, pathogens and parasites of tryptophan.22,23 However, with the discovery that IDO-1 activity was necessary for the preservation of allogenic fetuses in mice, further in vitro research found that tryptophan depletion had an anti-proliferative and apoptotic effect on T cells.2426 In particular, the general control non-derepressible-2 (GCN2) kinase was identified as a key mediator in IDO-1 induced tryptophan depletion immunosuppression.27 The activation of GCN2 triggers a stress-response program that can result in cell-cycle arrest, differentiation, adaptation or apoptosis.2830 Furthermore, some of the kynurenines, such as quinolinic acid and 3-hydroxyanthranilic acid, can also effectively suppress T cell proliferation.31 This inhibition appears to selectively target immune cells undergoing activation 32 and these kynurenines may act in concert to produce an additive effect.33 Lastly, the production of the excitotoxin quinolinic acid is often significantly increased following inflammation and resulting immune activation.34

To date, the kynurenine pathway has been implicated in a variety of diseases and disorders, including acquired immune deficiency syndrome (AIDS) dementia complex, Alzheimer’s disease (AD), schizophrenia, Huntington’s disease, amyotrophic lateral sclerosis (ALS) and neoplasia,3543 and numerous studies have measured the levels of tryptophan and kynurenines under those conditions. Significant imbalances in tryptophan and its metabolites were frequently observed, which when brought back within normal ranges, often resulted in alleviation of symptoms. This review brings together most of these studies to provide a better idea of the expected differences in tryptophan and kynurenine levels in the serum, cerebrospinal fluid (CSF) and brain between disease and healthy states.

The Kynurenines

Kynurenic acid

Kynurenic acid is an endogenous neuroprotectant that is usually present in the brain at nanomolar concentrations.44 An antagonist to quinolinic acid, kynurenic acid acts on the glycine modulatory site of the NMDA receptor at low concentrations;45 and at higher concentrations, at the glutamate site of the NMDA receptors and also on the a-amino-3-hydroxy-5-methyl-4-isoxazolepropionate (AMPA) receptors.46 In addition, it also antagonizes the alpha 7 nicotinic acetylcholine receptors47 and selectively activates a G-protein coupled receptor, GPR35.48

Increases in brain kynurenic acid were first observed to have sedative and anticonvulsant effects.49 Later, it was found to be protective against brain ischemia.50 The elevation in CSF kynurenic acid in schizophrenic patients also provided a new insight into the possible effect of kynurenic acid on the glutamatergic and dopaminergic systems, and its potential role in the pathogenesis of schizophrenia.51,52 Although it is argued that the physiological levels of kynurenic acid may fall below that which is necessary for glutamate receptor antagonism, at specific sites within synapses, those levels may be sufficient.53 This hypothesis is supported by the significant reduction in glutamate release and extracellular levels of dopamine seen with kynurenic acid in rats in vivo.54,55 In addition, the use of kynurenine 3-hydroxylase inhibitor also led to a hyperactivity in dopamine neurons.56

In a septic shock mouse model, kynurenic acid was able to significantly decrease the release of tumour necrosis factor α (TNF-α), nitric oxide and high mobility group box 1 protein, a molecule likely to be involved in lipopolysaccharides mediated toxicity.57,58 Rather unexpectedly though, kynurenic acid inhibited the release of fibroblastic growth factor 1, a compound that supports growth and recovery of injured cells and enhances proliferation of glia cells.59 However, this does not necessarily challenge the concept of kynurenic acid being neuroprotective but definitely warrants more investigation.

3-hydroxyanthranilic acid

3-hydroxyanthranilic acid can be derived either from the hydrolysis of 3-hydroxykyurenine or the oxidation of anthranilic acid (Fig. 1). Besides playing a role in immunoregulation,6062 3-hydroxyanthranilic acid is also a neurotoxin. Intracerebral injection of 3-hydroxyanthranilic acid leads to a decrease in choline acetyltransferase activity similar to those seen with quinolinic acid, but to a lesser extent.21 In addition, it is a free radical (superoxide and hydrogen peroxide) generator in the presence of copper.18 However, 3-hydroxyanthranilic acid can also act as an antioxidant, scavenging peroxyl radicals more effectively than equimolar concentrations of either ascorbic acid or Trolox (a water soluble analogue of vitamin E).63

In murine macrophages, 3-hydroxyanthranilic acid at sub-millimolar concentrations can inhibit the activation of nuclear factor κB and likewise, the expression and activity of inducible nitric oxide synthase (iNOS).64 iNOS catalyses the formation of NO, which is strongly correlated with antimicrobial and antitumoral activities in mouse macrophages.65 Following along the lines of tumoregenesis, non-toxic concentrations of 3-hydroxyanthranilic acid has no effect on T cell receptor triggered CD8+ T lymphocyte proliferation, but greatly inhibits that induced by antigen-independent cytokine (particularly interleukin (IL)-2, IL-7 and IL-15) stimulation.66 Thus, in the context of cancer, tumour cells could severely arrest CD8+ T cell proliferation by driving cytokine production without effectively triggering T cell receptor response.66

Furthermore, 3-hydroxyanthranilic acid exerts a selective apoptotic effect on murine thymocytes and T helper 1 (Thl) cells via the activation of caspase-8 and release of cytochrome c from mitochondria, but independent of the Fas pathway.61 This action occurs at concentrations well below those resulting in neurotoxicity or apoptosis of macrophages and could represent an important role in peripheral immunoregulation.61 Adding to this, following antigen stimulation of myelin basic protein Acl-11 T cell receptor transgenic CD4+ T cells, 3-hydroxyanthranilic acid was associated with a Gi/S phase arrest in CD4+ T cells and a cytokine profile shift in favour of Th2 cells.67 This finding has important implications in the treatment of multiple sclerosis (MS).67

Picolinic acid

Picolinic acid, a monocarboxylic acid, is an endogenous neuroprotectant and a natural iron and zinc chelator.21 It controls cellular growth and has anti-tumoral, antifungal and antiviral activities. In vitro, picolinic acid arrests normal cells in G1 phase, possibly through the interactions with NAD+ as the inhibition can be overcome by nicotinamide.68 Recently, the characterization of the kynurenine pathway in human primary adult neurons and SK-N-SH neuroblastoma cell line found the former capable of synthesizing picolinic acid but not the latter.69 This variation in kynurenine pathway activation in neuroblastoma cells may provide a key to understanding tumour persistence and associated neurotoxicity.

In vivo, the antitumoral effect of picolinic acid was observed when treatment in mice inoculated with MBL-2 lymphoma cells altered their ribosomal ribonucleic acid (RNA) metabolism, augmenting the cytotoxic and tumoricidal activities of macrophages, resulting in increased survival rate.70,71 As an antifungal, picolinic acid acts synergistically with IFN-γ to amplify the inhibitory effect of neutrophils, inhibiting Candida albicans growth in vitro and in vivo.72,73 Although the mechanism of this co-stimulatory effect is unclear, it is known to be vulnerable to IL-4 suppression.74 In mouse, the synergy with IFN-γ is further extended to include NOS and TNF-α gene expression.75,76

At relatively high concentrations (1.5–3 mM), picolinic acid exerts antiviral, cytotoxic and apoptotic effects on HIV-1 and human herpes simplex virus-2,77 which is likely to be associated with an up-regulation in macrophage inflammatory protein (MIP)-1α and MIP-1β messenger RNA (mRNA) expression, as both compounds inhibit HIV-1 infection.7880 Interestingly, this stimulatory effect on MlP-lα and β is antagonized by IFN-γ.81 The complex interplay between picolinic acid and IFN-γ highlights the importance of these molecules on the regulation of macrophage activities and perhaps, the inflammatory response.81

Like kynurenic acid, picolinic acid blocks quinolinic acid induced neurotoxicity, but not the neuroexcitatory component.21,82 Compared to kynurenic acid though, picolinic acid is less potent and appears to act via a different mechanism, attenuating calcium dependent glutamate release and/or chelating endogenous zinc.83,84,85 This lower potency of picolinic acid may also be partly explained by the weak stimulatory action it has on glutamate release from the striatum.84

Quinolinic acid

Quinolinic acid is a heterocyclic amino acid that selectively activates the neuronal NMDA subtype of glutamate receptors.19 Within the brain, quinolinic acid concentrations are normally lower compared to blood and systemic tissues as tryptophan is metabolized to 5-hydroxytryptamine rather than to formylkynurenine.86 However, during an immune response, either systemic or central, IDO-1 activity and levels of quinolinic acid rise dramatically, the significance of which is still obscure.8789

Under inflammatory conditions in the brain, infiltrating macrophages, microglia and dendritic cells are major sources of quinolinic acid production.90,91,92 Astrocytes, in contrast, are incapable of synthesizing quinolinic acid due to the absence of the enzyme, kynurenine hydroxylase.93 Rather, both astrocytes and neurons,9 being neuroprotective, uptake quinolinic acid and catabolize it to NAD. However, this catabolic system is easily saturated in the presence of high amounts of quinolinic acid, produced under pathological conditions, resulting in the toxic accumulation of quinolinic acid within the cells.94

As an endogenous molecule of the mammalian CNS, the immune and neurotoxic properties of quinolinic acid are of special interest.95 In vitro, the synthesis of quinolinic acid by CD8” dendritic cells induced apoptosis in Thl target cells;96 and quinolinic acid can also selectively inhibit the proliferation of CD4+ and CD8+ T lymphocytes and natural killer cells undergoing activation, the effect of which is amplified in the absence of tryptophan.32

In direct intracerebral administration and neuronal cell cultures, quinolinic acid led to neuronal death.97,98 Similarly, the chronic exposure to sub-micromolar concentrations of quinolinic acid on neurons produced an adverse effect and the converse was true too.99,98 In vivo, injection of quinolinic acid into discrete regions of the rat brain caused axon-sparing lesions similar to those produced by kainic and ibotenic acid.97 Several studies have already provided strong evidence suggesting that quinolinic acid plays a significant pathological role in the development of neurodegenerative disorders, such as Huntington’s disease (HD),99 AD100,101,102 and AIDS dementia complex.103,104,105

The Kynurenine Pathway in Disease States

Under various pathological conditions, an accelerated degradation of tryptophan with an accompanying increase in kynurenines is often observed in the serum, CSF and/or brain tissue (Tables 1, 2 and 3). Moreover, the breakdown of tryptophan via the kynurenine pathway is often routed preferentially towards the production of quinolinic acid. The pathologies associated with the up-regulation of the kynurenine pathway include infectious diseases (e.g. HIV), neurological disorders (e.g. AD, HD and ALS), affective disorders (e.g. schizophrenia, depression and anxiety), autoimmune diseases (e.g. MS and rheumatoid arthritis), peripheral conditions (e.g. cardiovascular disease) and malignancy (e.g. haematological neoplasia and colorectal cancer). However, significant elevations in tryptophan levels in lung and breast cancer have also been reported.106

Table 1.

Studies investigating kynurenine metabolites in plasma/serum.

References Pathology Compound Patients Controls Comments
Werner et al. 1988165 HIV TRP (μM) 44.8 ± 8.4++ 91.0 ± 22.0 Neopterin levels were significantly increased in patients (39.1 ± 17.0 nM vs. 4.5 ± 1.5 nM).
KYN (μM) 3.53 ± 0.89*** 2.31 ± 0.77
T/K ratio 13.4 ± 3.7++ 42.5 ± 13.7
Larsson et al. 1989166 HIV TRP (μM) 28.4 39.7 Platelets bound serotonin (5-HT) (ng/109) significantly reduced in patients compared to controls (430 vs. 676).
Cascino et al. 1991106 Cancer TRP (μM) 10.9 ± 5.2* (L-pre) 4.7 ± 0.7 L-pre: Lung cancer, pre-operation;
B: Breast cancer; pt: post-operation;
A:P < 0.05 from pre-op.
TRP data here is that of free tryptophan. Total plasma TRP was similar between patients and controls, pre-operation and post-operation.
6.6 ± 3.2* (B-pre)
7.1 ± 2.6*A (L-pt) 5.4 ± 0.9
4.6 ± l.lA (B-pt)
Fuchs et al. 1991114 HIV TRP (μM) 57.0 ± 2.8** (+) 91.0 ± 6.63 IFN-γ (U/I): 259 ± 70** in seropositive patients compared to 23.5 ± 1.7 in seronegative patients.
KYN (μM) 3.45 ± 0.14** (+) 2.31 ± 0.23
Denz et al. 1993120 Hematological neoplasias TRP (μM) 56.4 ± 13.1 (HD) ≤65 HD: Hodgkin’s disease; NHL: non-Hodgkin’s lymphoma; M/M: multiple myeloma/monoclonal gammopathy of unknown significance.
An inverse correlation was found between TRP and weight loss in patients.
50.5 ± 16.9+ (NHL)
44.9 ± 12.9+ (MM)
KYN (μM) 2.3 ± 1.1 (HD) ≤3.5
2.8 ± 1.4 (NHL)
2.5 ± 1.0 (M/M)
Heyes et al. 1994167 Epilepsy (intractable complex partial seizure) TRP (μM) 85.2 ± 3.7* (I.I.) 76.7 ± 4.7 I.I.: inter-ictal; P.I.: post-ictal
Data are shown only when differences were significant.
Patients’ data are approximates as results were presented only with a bar graph.
KYN (μM) 68.5 ± 3.7+ (I.I.) 3.27 ± 0.3
70.4 ± 3.7+ (P.I.)
KYNA (nM) 55.6 ± 5.56+ (I.I.) 32.1 ± 3.6
60.2 ± 7.4l** (P.I.)
3-HK (nM) No difference 383 ± 24
QUIN (nM) 73.1 ± 3.7*** (I.I.) 432 ± 60
70.4 ± 3.7*** (P.I.)
Orlikov et al. 1994168 Anxiety (A) and Depression (D) KYN (μM) 9.32 ± 0.2*** (A) 4.32 ± 0.3 After treatment, the KYN concentrations returned back to normal.
A significant correlation exists between KYN concentrations and anxiety severity.
2.98 ± 0.01* (D)
Fujigaki et al. 1998169 None KYN (μM) 1.6 ± 0.1 Species (human, macaques, rabbit, guinea pig, rat, gerbil and mouse) differences present in KYN and AA.
AA (nM) 16.5 ± 0.7
Heyes et al. 1998170 HIV QUIN (nM) 16847 ± 3358** 451 ± 78
Huengsberg et al. 1998171 HIV TRP (μM) 33.2 56.3 Asym: asymptomatic patients.
KT ratio (×1000): 119.9 in patients;
50.5 in asymptomatic AIDS subjects;
34.9 in controls. K/T ratio had a reciprocal relationship with CD4+ count.
50.1 (asym)
KYN (μM) 3.98 1.98
2.55 (asym)
Look et al. 2000115 HIV TRP (μM) 44.6 (pre) 52.6 pre: pre-treatment with HAART.
Post-treatment saw a significant increase in TRP and a decrease in QUIN.
KYN (μM) 4.1*** (pre) 2.7
KYNA (μM) 27 (pre) 30.1
QUIN (nM) 848+ (pre) 303.3
K/T (×103) 108.2*** (pre) 51.4
Murr et al. 2001112 Streptococcus pyogenes TRP (μM) 25.3** (STSS) STSS: streptococcal toxic shock syndrome; data are median values. Neopterin levels: STSS (152 nM) vs. tonsillitis (12 nM).
Neopterin levels correlated with kynurenine, K/T and inversely with tryptophan significantly.
80.9 (tonsillitis)
KYN (μM) 12.8** (STSS)
2.7 (tonsillitis)
K/T (×103) 560** (STSS)
40 (tonsillitis)
Murray et al. 2001137 HIV TRP (μM) 49.4 ± 6.5 (pre) pre: pre-treatment; post: post-treatment. Treatment with 3 g of nicotinamide daily for 2 mths.
69.2 ± 6.3 (post)
(Zangerle et al. 2002)116 HIV TRP (μM) 44.1 ± 13.3 (pre) 65.8 ± 12.8 pre: pre-treatment with ART.
6 mths after ART, median increase in TRP was 20.2%, median decrease in KYN was 19.3% and median decrease in KT ratio was 28.1%. During ART, change in KT ratio significantly correlated with change in HIV RNA, CD4+ T cells and neopterin.
KYN (μM) 3.01 ± 0.91 (pre) 2.02 ± 0.66
K/T (×103) 79.2 ± 60.3 (pre) 30.7 ± 8.7
(Huang et al. 2002)172 Colorectal cancer TRP (μM) 53.5* (median) 63.7
KYN (μM) 2.1 (median) 2.0
K/T (×103) 42.9* 31.8
(Ilzecka et al. 2003)173 ALS KYNA (nM) 57.8 ± 35.0 59.6 ± 20.5 m/s.c.s.: mild/severe clinical status
a: significantly lower KYNA in s.c.s. compared to m.c.s. There was no difference in serum KYNA and type of ALS onset.
81.6 ± 41.2a (m.c.s.)
39.9 ± 14.7* (s.c.s.)
Schrocksnadel et al. 2003174 Rheumatoid arthritis TRP (μM) 44.95** (median) 62.62
KYN (μM) 1.86 (median) 2.06
K/T (×103) 42.39** 31.72
Wirleitner et al. 2003175 Coronary heart disease TRP (μM) 53.5 ± 9.26** 65.9 ± 12.7 Subdividing patients into 3 groups: 1, 2/3-artery disease and those with restenosis showed no significant difference in TRP or KYN between groups.
KYN (μM) l.88 ± 0.53 l.85 ± 0.51
K/T (×103) 36.3 ± 13.0** 28.l ± 5.15
Schrocksnadel et al. 2005176 Gynaecological cancer TRP (μM) 43.5* (median) 53.5 Subdivision of patients found only those with ovarian cancer had significantly lower TRP than control. TRP, KYN or K/T did not correlate with disease stage.
KYN (μM) 1.91 (median) 1.73
Stoy et al. 200542 HD Data in graphs: The comparisons here are for baseline values only. The paper also looked at values after TRP depletion and loading. Big variations in QUIN values were observed but overall, the concentrations were similar between patients and controls. Neopterin levels were significantly increased in patients (18.6 ± 1.7 nM vs. 12.7 ± 0.8 nM).
TRP (μM) No difference
KYN (μM) Higher**
KYNA (μM) No difference
3-HK (μM) Lower*
3-HAA (μM) Lower*
QUIN (μM) No difference
K/T (×103) Higher**
Forrest et al. 2006177 Osteoporosis TRP (μM) 36.69 ± 1.8 (pre) 42.08 ± 2.28 Patients were treated for 2 yrs with either raloxifene or disodium etidronate with calcium.
42.42 ± 1.65 (post)
KYN (μM) 1.87 ± 0.12 (pre) 1.96 ± 0.11
2.01 ± 0.14 (post)
KYNA (nM) 32.68 ± 2.98 (pre) 24.76 ± 2.46
34.09 ± 3.75 (post)
3-HAA (nM) 1.04 ± 0.13* (pre) 7.89 ± 1.15
AA (nM) 139 ± 14.7* (pre) 21.56 ± 2.25
Mackay et al. 2006178 Chronic brain injury Data in graphs: The comparisons here are for baseline values only. The paper also looked at values after TRP depletion and loading. Big variations in QUIN values were observed but overall, the concentrations were similar between patients and controls. Neopterin levels were significantly increased in patients (18.8 ± 2.4 nM vs. 12.7 ± 0.8 nM).
TRP (μM) No difference
KYN (μM) Higher*
KYNA (μM) Lower**
3-HK (uM) Lower**
3-HAA (μM) Lower*
QUIN (μM) No difference
K/T (×103) Higher**
Darlington et al. 2007179 Stroke Data in graphs: The comparisons were made at different time points after stroke and the values here are only baseline values. Various correlations between kynurenines, neopterin, peroxidation products and volume of brain damage were analysed and TRP metabolism may contribute to brain damage following stroke.
TRP (μM) Lower+
KYN (μM) Higher*
3-HAA (nM) Lower+
AA (nM) Higher**
K/T (×103) Higher+
Hartai et al. 2007180 AD KYN (μM) 2.5 ± 0.1 2.01 ± 0.2 In red blood cells, comparing patients to controls, KYNA (nM): 43.9 ± 5.9*vs. 67.4 ± 8.6; KYN (mM): 8.1 ± 0.5 vs. 9.3 ± 0.6. Activities of KAT I and II were similar in both instances in patients and controls.
KYNA (nM) 15.82.31 ± 1.1* 23.13 ± 2.2
Myint et al. 2007123 Major depressioin TRP (μM) 65.8 ± 15.57 69.71 ± 13.65
KYN (\M) 1.81 ± 0.56 1.87 ± 0.43
KYNA (nM) 24.29 ± 8.09** 35.95 ± 13.4
3-HAA (nM) 24.53 ± 11.91 24.12 ± 7.3
K/T (×103) 25 ± 12* 17 ± 14
Schrocksnadel et al. 2006181 Rheumatoid arthritis TRP (μM) 58.0 ± 19.3* There was an inverse relation between TRP and the disease stage (P < 0.01)
KYN (nM) 2.20 ± 0.82*
Chen et al. unpublished182 ALS TRP (μM) 143.28 ± 5.64++ 75.0 ± 10.5
KYN (μM) 4.02 ± 0–2++ 2.52 ± 0.19
QUIN (MM) 0.37 ± 0.018* 0.30 ± 0.026
PIC (MM) 1.42 ± 0.087* 2.38 ± 0.37
K/T (×103) 37 ± 2.5 39 ± 4
*

P < 0.05;

**

P < 0.01;

***

P < 0.005;

+

P < 0.001;

++

P < 0.0001.

Table 2.

Studies investigating kynurenine metabolites in CSF.

Ref. Pathology Compound Patients Controls Comments
Young et al. 1983183 Epilepsy TRP (μM) 1.58 ± 0.61 1.66 ± 0.64 CSF data shown here were from the lumbar region. Cisternal CSF showed no differences between patients and controls and there were no CSF gradient differences either.
KYN (nM) 28.4 ± 15.3* 43.9 ± 24.5
5-HIAA (nM) 96.7 ± 37.7 117.2 ± 62.7
Larsson et al. 1989166 HIV TRP (nM) 1518 2179 No significant change in 5-HIAA.
Baig et al. 1991125 MS and Cerebro-vascular disease (CVD) TRP (nM) 1.25 ± 0.14+ (MS) 2.02 ± 0.34 Metabolites of the noradrenergic and dopaminergic systems [3-methoxy-4-hydroxyphenylglyco (MHPG), 3,4-dihydroxyphenylacetic acid (DOPAC) and homovanillic acid (HVA)] were also found to be significantly different in MS and CVD patients compared to controls.
3.34 ± 0.54+ (CVD)
5-HT (pM) 5 ± 1* (MS) 7 ± 2
7 ± 2 (CVD)
5-HIAA (pM) 116 ± 15** (MS) 173 ± 20
299 ± 50** (CVD)
Gisslen et al. 1994184 HIV TRP (nM) 1097 (pre) 3–14 months treatment with zidovudin.
Decrease in neopterin correlated with increase in TRP. 5-hydroxyindoleacetic acid (5-HIAA).
1535.8 (post)
Heyes et al. 1994167 Epilepsy (intractable complex partial) TRP (μM) No difference 1.32 ± 0.13 I.I.: inter-ictal; P.I.: post-ictal
Data are shown only when differences were significant. Patients’ data are approximates as results were presented only with a bar graph.
QUIN:KYNA in patients vs. controls: 61.1 ± 11.1** (I.I.), 58.3 ± 5.55***(P.I.) vs. 86.1 ± 19.4
KYN (nM) 68.1 ± 2.78+ (I.I.) 42.2 ± 3.8
65.3 ± 2.78+ (P.I.)
KYNA (nM) No difference 2.32 ± 0.35
QUIN (nM) 72.2.1 ± 2.78+ (I.I.) 21.9 ± 2.8
68.1 ± 2.78+ (P.I.)
Demitrack et al. 1995117 Eating disorders (anorexia nervosa) TRP (nM) 1.9 ± 0.5 2.1 ± 0.3 In anorectics, weight normalized restored all compounds tested to within the control range. The relative amount of QUIN (QUIN: KYNA) was significantly higher during the underweight phase for anorectics.
Kynurenines were within control range for normal weight bulimics.
KYN (nM) 25.6 ± 9.9 34.4 ± 12.3
KYNA (nM) 1.5 ± 0.5* 2.8 ± 1.2
QUIN (nM) 13.4 ± 5.4 13.8 ± 4.3
5-HIAA (nM) 107.2 ± 31.4* 146.3 ± 30.2
Heyes et al. 1995185 CNS pathology QUIN (nM) 31 ± 5 (Hy) 20 ± 2 Hy: hydrocephalus; H: haemorrhage; T: tumour; C: CSF infection. Subjects were all children. Both TNF-a and IL-6 were increased, with a significant correlation between IL-6 and QUIN.
200 ± 113** (H)
282 ± 82** (T)
1084 ± 549** (C)
KYN (nM) 185 ± 40 (Hy) 54 ± 7
254 ± 128** (H)
1698 ± 589** (T)
2610 ± l 067** (C)
Fujigaki et al. 1998169 None KYN (nM) 29.1 ± 3.2 Species (human, macaques, rabbit, guinea pig, rat, gerbil and mouse) differences detected in levels of KYN and AA.
AA (nM) 16.3 ± 4.2
Heyes et al. 1998170 HIV QUIN (nM) 3789 ± 888** 22.1 ± 2.1
Erhardt et al. 2001a51 Schizophrenia KYN (nM) 1.67 ± 0.027* 0.97 ± 0.07 A correlation between age and KYN was found in schizophrenics.
Medana et al. 2002186 Malaria (severe) KYNA (μM) 0.06 0.07 None of the kynurenines were associated with convulsions or coma.
QUIN (μM) 0.80++ 0.07
PIC (μM) 0.19+ 0.08
Rejdak et al. 2002187 MS KYNA (nM) 0.41** (MS) MS: were patients with relapsing MS during remission or not progressing for at least 2 months; ID: infectious inflammatory disease; OND: non-inflammatory neurological disorders. MS had significantly lower KYNA than either ID or OND.
0.67 (OND)
1.7 (ID)
Ilzecka et al. 2003173 ALS KYNA (nM) 2.41 ± 1.7 (grp) 1.59 ± 0.9 Bul: bulbar onset; s.c.s: severe clinical status No significant difference between KYNA levels and gender and no correlation between KYNA and age.
3.61 ± 2.0** (bul)
3.26 ± 2.1*(s.c.s.)
Medana et al. 2003109 Cerebral Malaria (Malawian children) QUIN (μM) 0.09 For QUIN, KYNA and PIC, 72% (2%), 77% (43%) and 74% (38%) of Malawian children had higher levels than median (reference range) UK control levels respectively. Elevated levels of QUIN and PIC were associated with a fatal outcome. Other diseases tested include convulsions, sepsis and acute hepatitis.
KYNA (μM) 0.21
PIC (μM) 0.18
Nilsson et al. 2005188 Schizophrenia KYNA (nM) 1.45 ± 0.10* (grp) 1.06 ± 0.06 Grp: All patients; 1st: Drug naive, first episode patients; T: patients undergoing treatment with anti-psychotic drugs; no D: patients who had been treated but are now drug free. In patients, a positive correlation was found between KYNA levels and age.
1.53 ± 0.19* (1st)
1.53 ± 0.17*(T)
1.16 ± 0.06 (noD)
Atlas et al. 2007189 HIV KYNA (nM) median levels 4.54 (psy.) 3.02 (no psy.) 1.23 psy: psychotic symptoms In controls, KYNA levels were significantly higher in females (2.29 nM) than males (1.10 nM) (P < 0.05). However, this gender difference was absent in the patient population.
Chen et al. unpublished182 ALS TRP (μM) 5.02 ± 0.19 2.58 ± 0.16
KYN (μM) 0.23 ± 0.0016 0.027 ± 0.001
QUIN (μM) 0.053 ± 0.0054* 0.038 ± 0.006
PIC (μM) 0.36 ± 0.034 0.51 ± 0.11
K/T(×103) 43.7 ± 2 11.1 ± 0.8
*

P < 0.05;

**

P < 0.01;

***

P < 0.005;

+

P < 0.001;

++

P < 0.0001.

Table 3.

Studies investigating kynurenine metabolites in brain.

Ref. Pathology Compound Patients Controls Comments
Beal et al. 199040 HD KYNA (nM) 1.29 ± 0.14* (HD) 5.10 ± 1.04 PD: Parkinson’s disease; IS: ischemic stroke. 2-fold increase in KYN/KYNA in HD (P < 0.01). KYNA was found to be considerably lower in HD compared to controls and patients with other neurological disorders.
3.93 ± 0.71 (AD)
4.59 ± 1.75 (PD)
5.04 ± 1.66 (IS)
Beal et al. 1992190 HD TRP (ng/g) 4658 ± 442** (i.t.) 8053 ± 1120 p.g.:precentral gyrus; f.c: frontal cortex; i.t.: inferior temporal; m.t: middle temporal; s.t: superior temporal. Kynurenine metabolites, tryptophan, indoleamines and tyrosine and metabolites were anaylzed in 8 different regions of the brain. The data presented here are only for kynurenine metabolites that were significantly different in patients compared to controls.
KYN (ng/g) 2334 ± 33*** (i.t.) 5884 ± 129
KYNA (ng/g) 223 ± 33** (m.t.) 422 ± 83
3-HK(ng/g) 18.4 ± 5.3** (p.g.) 81.3 ± 18.1
17.9 ± 2.4* (f.c) 31.2 ± 5.6
16.0 ± 2.7*** (i.t.) 70.04 ± 17.2
17.0 ± 3.7** (m.t.) 39.0 ± 6.4
29.4 ± 9.7** (s.t.) 130.3 ± 60.4
26.8 ± 8.3** (i.t.) 67.7 ± 19.6
Pearson and Reynolds. 1992191 HDand AD 3-HK(ng/g) 110 ± 47**(HDt.c.) 65 ± 56 t.c: temporal cortex; f.c: frontal cortex; p: putamen In HD, a general increase in 3-HK was observed, rather than a region-specific one. In t.c. of AD cases, where neuronal loss was greater than in HD, suggested that 3-HK increases in HD is not due entirely to neuronal atrophy.
82 ± 41 (ADt.c.) 65 ± 33
93 ± 60**(HDf.c.) 33 ± 26
65 ± 47*** (HD p) 19 ± 14
Sardar et al. 1995192 HIV 3-HK(ng/g) 71.3 ± 12.7 (grp)** 19.95 ± 3.18 N-D: HIV without dementia; D: HIV with dementia. Tissues were taken from the frontal cortex. Higher levels of 3-HK in D was not significantly different from N-D. 3-HA formation was an indicator for 3-hydroxykinurease (3-HKase) activity, which was highest in N-D. Thus, increase in 3-HK reflected an overall increase in KP, instead of a decrease in 3-HKase activity.
3-HA formation (ng/h/g) 64.9 ± 11.4 (N-D)** 15.8 = t2.14
85.5 ± 32.8 (D)**
66.4 ± 11.5 (grp)**
61.6 ± 16.5 (N-D)**
75.5 ± 12.5 (D)**
Heyes et al. 1998170 HIV QUIN (pmol/g) 20942 ± 2959** (bg) 72 ± 26 bg: basal ganglia; wm: cortical white matter; gm: cortical grey matter.
25397 ± 11435** (wm) 75 ± 12
26292 ± 8615** (gm) 81 ± 20
Bara et al. 2000193 HIV KYN (pmol/mg) 22.66 ± 5.38 (f.c.) 12.08 ± 1.24 f.c: frontal cortex; cb: cerebellum KAT I activity rose significantly in both frontal cortex and cerebellum (341% and 262% of control, respectively), whereas KAT II activity increased only in the frontal cortex (141% of control).
KYNA (pmol/mg) 24.67 ± 2.62 (cb) 16.33 ± 2.00
7.31 ± 1.33(f.c.) 3.49 ± 0.55
4.54 ± 0.87 (cb) 2.77 ± 0.63
Schwarcz et al. 2001111 Schizophrenia KYN (ng/g) 35.2 ± 28.0* (b.a.9) 22.4 ± 14.3 b.a.: Brodmann area KYN, KYNA and 3-HK were tested in b.a 9,10 and 19. Only data that were significantly different from controls are presented here. Positive correlation found between KYN and KYNA but not KYN and 3-HK.
KYNA (ng/g) 40.3 ± 23.4* (b.a.l9) 30.9 ± 10.8
1.9 ± 1.3* (b.a.9) 2.9 ± 2.2
*

P < 0.05;

**

P < 0.01;

***

P < 0.005;

+

P < 0.001;

++

P < 0.0001.

We also observed an increase in tryptophan levels ALS patients’ samples (unpublished). At this stage, we speculate that this phenomenon might be associated with either a disturbance in albumin binding of tryptophan, an over-compensatory response to decreased tryptophan concentrations in the brain and/or a malfunctioning in the L-type amino acid transporter at the blood brain barrier in ALS. The elevation in tryptophan notwithstanding, ALS patients still exhibited a larger kynurenine/tryptophan (K/T) ratio, an index for IDO activity, than control subjects due to a significant concomitant rise in kynurenine.

The enhanced degradation of tryptophan and higher K/T ratio are also often associated with advanced stages of disease, more severe symptoms or a fatal outcome.107 108,109 However, it is important to note that a progressive increased in tryptophan catabolism is part of the “normal” ageing process.110 Nonetheless, the degree of tryptophan depletion is still far more substantial in neurodegenerative disorders compared to normal ageing and most of the studies on pathological conditions were performed using age matched control subjects.111,100

In some studies, neopterin concentrations were also measured. Neopterin is a marker for immune activation and show a correlation with the K/T ratio and kynurenine, and inversely with tryptophan.112,113,87 This suggests an increase in endogenous IFN-γ production and an up-regulation in the kynurenine pathway. Indeed, HIV patients exhibit a 10-fold increase in IFN-γ through direct measurements.114

When HIV patients are treated with highly active antiretroviral therapy (HAART) or antiretroviral treatment (ART), which significantly decreases immune activation through reduction in viral load, a repletion in tryptophan and reduction in kynurenine and quinolinic acid often follows.115,116 It is interesting to note that the alteration in tryptophan levels occurred in the absence of any dietary modification and that changes in K/T ratio correlated strongly with HIV mRNA and CD4+ T cell count.116

The most important consequences of dramatic decline in tryptophan, thus, are likely to be immunosuppression and immunodeficiency, particularly evident in HIV infection, but also in autoimmune diseases and cancer. Other effects include weight loss, mood disturbances and cognitive impairment.117,118

In anorexia nervosa, underweight anorexic patients had lower tryptophan levels which rose with weight normalization.117 The association of tryptophan levels and the development of cachexia and weight loss are also evident in neoplasia.119,120 This could be associated with the release of pro-inflammatory cytokines. TNF-α, for instance, is a known cachexia, featuring prominently in muscle pathophysiology.121 The heightened catabolism of tryptophan via the kynurenine pathway may also divert this essential amino acid away from protein synthesis, thus, contributing to weight loss and muscle wasting.119

Tryptophan also acts as a precursor for the synthesis of serotonin, which has a broad spectrum of action, two of which are in mood and cognitive functioning.118,122 Imbalances in kynurenines and significant decline in 5-hydroxyindoleacetic acid (5-HIAA), a serotonin metabolite, have been reported in major depression, MS and cardiovascular disease, among others.123,124,125 However, the activation of the immune response is also postulated as a cause of depression126,127 and a strong association exists between inflammatory diseases and depression.128,129

In normal subjects, the deliberate depletion of tryptophan selectively impaired long-term memory consolidation,130 opposed to the results observed with the administration of selective serotonin reuptake inhibitors.131 In AD and HD patients, the K/T ratio was also inversely correlated with cognitive performance;132,133 and in HIV-1 patients, treatment with HAART, which elevates tryptophan levels, markedly improved cognitive function.134,135

Potential Treatments Involving the Kynurenine Pathway

The involvement of the kynurenine pathway in a wide range of diseases suggests that research on treatment strategies targeting the kynurenine pathway (Fig. 2) may provide an alternative means of treatment or as a complement to what is already available.

Figure 2.

Figure 2.

Drugs targeting the kynurenine pathway—inhibitors and analogues.

Niacin supplementation

One of the consequences of accelerated degradation and depletion of tryptophan in the body is the suppression of T cell proliferation,136 which compromises the body’s immunity. Repletion of tryptophan could lead to improve immune response but may also inadvertently cause an increase in neurotoxins. Niacin supplementation, however, provides an indirect way to increase tryptophan and act as a feedback mechanism to suppress IDO-1 activity.137 In clinical studies, dietary supplementation of niacin to HIV-1 patients was associated with higher CD4 counts and improved survival rates.138,139

IDO inhibitors

The suppression of IDO-1 activity has been targeted directly in cancer research. Using transgenic mouse model of breast cancer, IDO-1 inhibitors, 1-methyl-DL-tryptophan and methyl-thiohydantoin-tryptophan, were able to potentiate the efficacy of chemotherapy drugs, promoting tumour regression without increasing the side effects.140 The discovery of the preferential inhibition by D-l-methyl-tryptophan on IDO-2 could also provide the key to understanding the mechanism behind the antitumoral action of 1-methyl-tryptophan and in designing future IDO inhibitors.7

Kynurenine analogues

Another approach to modifying the kynurenine pathway is to skew the balance of kynurenines towards neuroprotection and away from neurotoxicity. Currently, there are several therapeutic agents, either already on the market or undergoing clinical trials, which are either analogues of neuroprotective kynurenines or act to inhibit the production of quinolinic acid. They include 4-chlorokynurenine, laquinimod, leflunomide, tranilast, nicotinylalanine, meta-nitrobenzoylalanine and Ro61-8048.

7-chlorokynurenate, a synthetic derivative of kynurenic acid, attenuates the neurotoxic effect of quinolinic acid through blockade of the glycine modulatory site of the NMDA receptor.141,142 However, 7-chlorokynurenate crosses the blood brain barrier with great difficulty.143 4-chlorokynurenine, a precursor of 7-chlorokynurenate, on the other hand, is able to overcome this obstacle.144 When administered together with quinolinic acid in vivo, 4-chlorokynurenine was converted into the active 7-chlorokynurenate successfully, providing neuroprotection.145,146

Laquinimod (ABR-215062), a novel synthetic quinoline, has demonstrated immunomodulatory properties without immunosuppression in preclinical trials.147149 In MS animal model, experimental autoimmune encephalomyelitis (EAE), laquinimod delayed disease progression, inhibited infiltration of CD4+ T cells and macrophages into the CNS and modulated the immune response in favour of Th2/Th3 cytokines IL-4, IL-10 and transforming growth factor (TGF-P).150 Furthermore, in patients with relapsing MS, treatment with laquinimod successfully reduced the development of active lesions.151

Leflunomide (Avara®), an immunosuppressive and anti-inflammatory prodrug is converted into terflunomide in vivo (A771126). Terflunomide is an inhibitor of mitochondrial dihydroorotate dehydrogenase, an essential enzyme for de novo pyrimidine synthesis.152 In 1998, the Food and Drug Administration (FDA, U.S.A.) approved leflunomide for the treatment of rheumatoid arthritis. Furthermore, in a recent phase II trial with MS patients, terflunomide proved well tolerated and effective in reducing active lesions.153

Tranilast (Rizaben®), a synthetic anthranilic acid derivative drug, has the ability to inhibit the release of chemical mediators, TGF-É¿ and suppress angiogenesis.154,155 Tranilast has been effective against many diseases, such as allergic rhinitis, atopic dermatitis and bronchial asthma. Recently, when tested against EAE, tranilast inhibited the actions of Thl cells while enhancing those of Th2 cells, an action similar to that of natural tryptophan catabolites, 3-hydroxyanthranilic acid and 3-hydroxykynurenic acid.67

Finally, kynurenine hydroxylase inhibitors are also effective in diverting the kynurenine pathway away from the synthesis of quinolinic acid towards that of kynurenic acid. These compounds include nicotinylalanine, meta-nitrobenzoylalanine and Ro61-8048.156 Nicotinylalanine, an analogue of kynurenine, protects the brain from induced seizures157,158 and quinolinic acid induced striatal damage in the rat.159 With meta-nitrobenzoylalanine, sedation and anticonvulsant effects were achieved in rats,160 while reduced neuronal loss from brain ischemia were seen in gerbils.50 In immune activated mice, meta-nitronemzoylalanine also significantly decreased quinolinic acid production in the blood and brain.161 With Ro61-8048, there is an additional benefit of reducing glutamate levels in the extracellular spaces of the basal ganglia in rats, while maintaining the learning and memory process.162 In EAE rats, administration of Ro61-8048 significantly reduced the neurotoxic levels of 3-hydroxykynurenine and quinolinic acid in the CNS;163 and in a cerebral malaria mouse model, it significantly increased the neuroprotective levels of picolinic acid, prevented the development of neurological symptoms and prolonged survival by threefold.164 Like meta-nitrobenzoylalanine, Ro61-8048 too decreased neuronal loss due to brain ischemia.50

Conclusion

The kynurenine pathway is an effective mechanism in modulating the immune response and in inducing immune tolerance. This is achieved by accelerating the degradation of tryptophan and the generation of kynurenines. The metabolites of the pathway, with their different inherent properties, can also synergize or antagonize the effects of one another. By measuring the levels of tryptophan, kynurenines and the K/T ratio under various pathological conditions, the degree of immune activation and the relationship between the kynurenine pathway and disease states may be gleaned. However, much research is still needed to fully understand the complex interaction between tryptophan, IDO and kynurenines among themselves and within the CNS and in the periphery. With the seemingly prevalent involvement of the kynurenine pathway in a wide range of different diseases and disorders, the knowledge gained from research focusing on the kynurenine pathway may be translated into designing novel and more effective treatment strategies.

Acknowledgments

This study was funded by the Motor Neuron Disease Research Institute Association (Australia).

Footnotes

Disclosure

The authors report no conflicts of interest.

References

  • 1.Mcmenamy RH. Binding of indole analogues to human serum albumin. Effects of fatty acids. J Biol Chem. 1965;240:4235–43. [PubMed] [Google Scholar]
  • 2.Hargreaves KM, Pardridge WM. Neutral amino acid transport at the human blood-brain barrier. J Biol Chem. 1988;263:19392–7. [PubMed] [Google Scholar]
  • 3.Ruddick JP, Evans AK, Nutt DJ, Lightman SL, Rook GA, Lowry CA. Tryptophan metabolism in the central nervous system: medical implications. Expert Rev Mol Med. 2006;8:1–27. doi: 10.1017/S1462399406000068. [DOI] [PubMed] [Google Scholar]
  • 4.Salter M, Pogson CI. The role of tryptophan 2,3-dioxygenase in the hormonal control of tryptophan metabolism in isolated rat liver cells. Effects of glucocorticoids and experimental diabetes. Biochem J. 1985;229:499–504. doi: 10.1042/bj2290499. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Takikawa O. Biochemical and medical aspects of the indoleamine 2,3-dioxygenase-initiated L-tryptophan metabolism. Biochem Biophys Res Commun. 2005;338:12–9. doi: 10.1016/j.bbrc.2005.09.032. [DOI] [PubMed] [Google Scholar]
  • 6.Ball HJ, Sanchez-Perez A, Weiser S, et al. Characterization of an indoleamine 2,3-dioxygenase-like protein found in humans and mice. Gene. 2007;396:203–13. doi: 10.1016/j.gene.2007.04.010. [DOI] [PubMed] [Google Scholar]
  • 7.Metz R, Duhadaway JB, Kamasani U, Laury-kleintop L, Muller AJ, Prendergast GC. Novel tryptophan catabolic enzyme IDO2 is the preferred biochemical target of the antitumor indoleamine 2,3-dioxygenase inhibitory compound D-1-methyl-tryptophan. Cancer Res. 2007;67:7082–7. doi: 10.1158/0008-5472.CAN-07-1872. [DOI] [PubMed] [Google Scholar]
  • 8.Guillemin GJ, Smythe G, Takikawa O, Brew BJ. Expression of indoleamine 2,3-dioxygenase and production of quinolinic acid by human microglia, astrocytes, and neurons. Glia. 2005d;49:15–23. doi: 10.1002/glia.20090. [DOI] [PubMed] [Google Scholar]
  • 9.Guillemin GJ, Cullen KM, Lim CK, et al. Characterization of the Kynurenine Pathway in Human Neurons. J Neuro sci. 2007b;27:12884–92. doi: 10.1523/JNEUROSCI.4101-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Guillemin GJ, Kerr SJ, Smythe GA, et al. Kynurenine pathway metabolism in human astrocytes: a paradox for neuronal protection. J Neuro chem. 2001;78:1–13. doi: 10.1046/j.1471-4159.2001.00498.x. [DOI] [PubMed] [Google Scholar]
  • 11.Guillemin GJ, Smith DG, Smythe GA, Armati PJ, Brew BJ. Expression of the kynurenine pathway enzymes in human microglia and macrophages. Adv Exp Med Biol. 2003a;527:105–12. doi: 10.1007/978-1-4615-0135-0_12. [DOI] [PubMed] [Google Scholar]
  • 12.Fujigaki S, Saito K, Sekikawa K, et al. Lipopolysaccharide induction of indoleamine 2,3-dioxygenase is mediated dominantly by an IFN-gamma-independent mechanism. Eur J Immunol. 2001;31:2313–8. doi: 10.1002/1521-4141(200108)31:8<2313::aid-immu2313>3.0.co;2-s. [DOI] [PubMed] [Google Scholar]
  • 13.Guillemin GJ, Smythe GA, Veas LA, Takikawa O, Brew BJ. A beta 1–42 induces production of quinolinic acid by human macrophages and microglia. Neuroreport. 2003b;14:2311–5. doi: 10.1097/00001756-200312190-00005. [DOI] [PubMed] [Google Scholar]
  • 14.Hayaishi O, Yoshida R. Specific induction of pulmonary indoleamine 2,3-dioxygenase by bacterial lipopolysaccharide. Ciba Found Symp. 1978:199–203. doi: 10.1002/9780470715413.ch12. [DOI] [PubMed] [Google Scholar]
  • 15.Werner-felmayer G, Werner ER, Fuchs D, Hausen A, Reibnegger G, Wachter H. Characteristics of interferon induced tryptophan metabolism in human cells in vitro. Biochim Biophys Acta. 1989;1012:140–7. doi: 10.1016/0167-4889(89)90087-6. [DOI] [PubMed] [Google Scholar]
  • 16.Yasui H, Takai K, Yoshida R, Hayaishi O. Interferon enhances tryptophan metabolism by inducing pulmonary indoleamine 2,3-dioxygenase: its possible occurrence in cancer patients. Proc Natl Acad Sci U S A. 1986;83:6622–6. doi: 10.1073/pnas.83.17.6622. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Dai W, Gupta SL. Regulation of indoleamine 2,3-dioxygenase gene expression in human fibroblasts by interferon-gamma. Upstream control region discriminates between interferon-gamma and interferon-alpha. J Biol Chem. 1990;265:19871–7. [PubMed] [Google Scholar]
  • 18.Goldstein LE, Leopold MC, Huang X, et al. 3-Hydroxykynurenine and 3-hydroxyanthranilic acid generate hydrogen peroxide and promote alpha-cry stallin cross-linking by metal ion reduction. Biochemistry. 2000;39:7266–75. doi: 10.1021/bi992997s. [DOI] [PubMed] [Google Scholar]
  • 19.Stone TW, Perkins MN. Quinolinic acid: a potent endogenous excitant at amino acid receptors in CNS. Eur J Pharmacol. 1981;72:411–2. doi: 10.1016/0014-2999(81)90587-2. [DOI] [PubMed] [Google Scholar]
  • 20.Perkins MN, Stone TW. An iontophoretic investigation of the actions of convulsant kynurenines and their interaction with the endogenous excitant quinolinic acid. Brain Res. 1982;247:184–7. doi: 10.1016/0006-8993(82)91048-4. [DOI] [PubMed] [Google Scholar]
  • 21.Jhamandas K, Boegman RJ, Beninger RJ, Bialik M. Quinolinate-induced cortical cholinergic damage: modulation by tryptophan metabolites. Brain Res. 1990;529:185–91. doi: 10.1016/0006-8993(90)90826-w. [DOI] [PubMed] [Google Scholar]
  • 22.Brown RR, Ozaki Y, Datta SP, Borden EC, Sondel PM, Malone DG. Implications of interferon-induced tryptophan catabolism in cancer, auto-immune diseases and AIDS. Adv Exp Med Biol. 1991;94:425–35. doi: 10.1007/978-1-4684-5952-4_39. [DOI] [PubMed] [Google Scholar]
  • 23.Pfefferkorn ER. Interferon gamma blocks the growth of Toxoplasma gondii in human fibroblasts by inducing the host cells to degrade tryptophan. Proc Natl Acad Sci U S A. 1984;81:908–12. doi: 10.1073/pnas.81.3.908. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Lee GK, Park HJ, Macleod M, Chandler P, Munn DH, Mellor AL. Tryptophan deprivation sensitizes activated T cells to apoptosis prior to cell division. Immunology. 2002;107:452–60. doi: 10.1046/j.1365-2567.2002.01526.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Munn DH, Shafizadeh E, Attwood JT, Bondarev I, Pashine A, Mellor AL. Inhibition of T cell proliferation by macrophage tryptophan catabolism. J Exp Med. 1999;189:1363–72. doi: 10.1084/jem.189.9.1363. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Munn DH, Zhou M, Attwood JT, et al. Prevention of allogeneic fetal rejection by tryptophan catabolism. Science. 1998;281:1191–3. doi: 10.1126/science.281.5380.1191. [DOI] [PubMed] [Google Scholar]
  • 27.Munn DH, Sharma MD, Baban B, et al. GCN2 kinase in T cells mediates proliferative arrest and anergy induction in response to indoleamine 2,3-dioxygenase. Immunity. 2005;22:633–42. doi: 10.1016/j.immuni.2005.03.013. [DOI] [PubMed] [Google Scholar]
  • 28.Rao RV, Ellerby HM, Bredesen DE. Coupling endoplasmic reticulum stress to the cell death program. Cell Death Differ. 2004;11:372–80. doi: 10.1038/sj.cdd.4401378. [DOI] [PubMed] [Google Scholar]
  • 29.De haro C, Mendez R, Santoyo J. The eIF-2alpha kinases and the control of protein synthesis. Faseb J. 1996;10:1378–87. doi: 10.1096/fasebj.10.12.8903508. [DOI] [PubMed] [Google Scholar]
  • 30.Bi M, Naczki C, Koritzinsky M, et al. ER stress-regulated translation increases tolerance to extreme hypoxia and promotes tumor growth. Embo J. 2005;24:3470–81. doi: 10.1038/sj.emboj.7600777. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Fallarino F, Grohmann U, Vacca C, et al. T cell apoptosis by kynurenines. Adv Exp Med Biol. 2003;527:183–90. doi: 10.1007/978-1-4615-0135-0_21. [DOI] [PubMed] [Google Scholar]
  • 32.Frumento G, Rotondo R, Tonetti M, Damonte G, Benatti U, Ferrara GB. Tryptophan-derived catabolites are responsible for inhibition of T and natural killer cell proliferation induced by indoleamine 2,3-dioxygenase. J Exp Med. 2002;196:459–68. doi: 10.1084/jem.20020121. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Terness P, Bauer TM, Rose L, et al. Inhibition of allogeneic T cell proliferation by indoleamine 2,3-dioxygenase-expressing dendritic cells: mediation of suppression by tryptophan metabolites. J Exp Med. 2002;196:447–57. doi: 10.1084/jem.20020052. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Moffett JR, Els T, Espey MG, Walter SA, Streit WJ, Namboodiri MA. Quinolinate immunoreactivity in experimental rat brain tumors is present in macrophages but not in astrocytes. Exp Neurol. 1997;144:287–301. doi: 10.1006/exnr.1996.6365. [DOI] [PubMed] [Google Scholar]
  • 35.Guillemin G, Brew BJ. Implications of the kynurenine pathway and quinolinic acid in Alzheimer’s disease. Redox Rep. 2002;7:199–206. doi: 10.1179/135100002125000550. [DOI] [PubMed] [Google Scholar]
  • 36.Guillemin G, Meininger V, Brew B. Implications for the kynurenine pathway and quinolinic acid in amyotrophic lateral sclerosis. Neurodegenerative diseases. 2006;2:166–76. doi: 10.1159/000089622. [DOI] [PubMed] [Google Scholar]
  • 37.Guillemin GJ, Kerr SJ, Brew BJ. Involvement of quinolinic acid in AIDS dementia complex. Neurotox Res. 2005c;7:103–23. doi: 10.1007/BF03033781. [DOI] [PubMed] [Google Scholar]
  • 38.Ting K, Brew BJ, Guillemin GJ. Effect of quinolinic acid on gene expression in human astrocytes: Implications for Alzheimer’s disease. In: Takai K, editor. International Congress Series. 2007. [Google Scholar]
  • 39.Bruijn LI, Miller TM, Cleveland DW. Unraveling the mechanisms involved in motor neuron degeneration in ALS. Annu Rev Neurosci. 2004;27:723–49. doi: 10.1146/annurev.neuro.27.070203.144244. [DOI] [PubMed] [Google Scholar]
  • 40.Beal MF, Matson WR, Swartz KJ, Gamache PH, Bird ED. Kynurenine pathway measurements in Huntington’s disease striatum: evidence for reduced formation of kynurenic acid. J Neurochem. 1990;55:1327–39. doi: 10.1111/j.1471-4159.1990.tb03143.x. [DOI] [PubMed] [Google Scholar]
  • 41.LIM CK, Smythe GA, Stocker R, Brew BJ, Guillemin GJ. Characterization of the kynurenine pathway in human oligodendrocytes. In: Takai K, editor. International Congress Series. 2007. [Google Scholar]
  • 42.Stoy N, Mackay GM, Forrest CM, et al. Tryptophan metabolism and oxidative stress in patients with Huntington’s disease. J Neurochem. 2005;93:611–23. doi: 10.1111/j.1471-4159.2005.03070.x. [DOI] [PubMed] [Google Scholar]
  • 43.Zamanakou M, Germenis AE, Karanikas V. Tumor immune escape mediated by indoleamine 2,3-dioxygenase. Immunol Lett. 2007;111:69–75. doi: 10.1016/j.imlet.2007.06.001. [DOI] [PubMed] [Google Scholar]
  • 44.Moroni F, Russi P, Lombardi G, Beni M, Carla V. Presence of kynurenic acid in the mammalian brain. J Neurochem. 1988;51:177–80. doi: 10.1111/j.1471-4159.1988.tb04852.x. [DOI] [PubMed] [Google Scholar]
  • 45.Stone TW. Neuropharmacology of quinolinic and kynurenic acids. Pharmacol Rev. 1993;45:309–79. [PubMed] [Google Scholar]
  • 46.Stone TW, Addae JI. The pharmacological manipulation of glutamate receptors and neuroprotection. Eur J Pharmacol. 2002;447:285–96. doi: 10.1016/s0014-2999(02)01851-4. [DOI] [PubMed] [Google Scholar]
  • 47.Hilmas C, Pereira EF, Alkondon M, Rassoulpour A, Schwarcz R, Albuquerque EX. The brain metabolite kynurenic acid inhibits alpha7 nicotinic receptor activity and increases non-alpha7 nicotinic receptor expression: physiopathological implications. J Neuro Sci. 2001;21:7463–73. doi: 10.1523/JNEUROSCI.21-19-07463.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Wang H, Liao H, Ochani M, et al. Cholinergic agonists inhibit HMGB1 release and improve survival in experimental sepsis. Nat Med. 2004;10:1216–21. doi: 10.1038/nm1124. [DOI] [PubMed] [Google Scholar]
  • 49.Carpenedo R, Chiarugi A, Russi P, et al. Inhibitors of kynurenine hydroxylase and kynureninase increase cerebral formation of kynurenate and have sedative and anticonvulsant activities. Neuro Science. 1994;61:237–43. doi: 10.1016/0306-4522(94)90227-5. [DOI] [PubMed] [Google Scholar]
  • 50.Cozzi A, Carpenedo R, Moroni F. Kynurenine hydroxylase inhibitors reduce ischemic brain damage: studies with (m-nitrobenzoyl)-alanine (mNBA) and 3,4-dimethoxy-[-N-4-(nitrophenyl)thiazol-2yl]- benzene-sulfonamide (Ro 61–8048) in models of focal or global brain ischemia. J Cereb Blood Flow Metab. 1999;19:771–7. doi: 10.1097/00004647-199907000-00007. [DOI] [PubMed] [Google Scholar]
  • 51.Erhardt S, Blennow K, Nordin C, Skogh E, Lindstrom LH, Engberg G. Kynurenic acid levels are elevated in the cerebrospinal fluid of patients with schizophrenia. Neurosci Lett. 2001;33:96–8. doi: 10.1016/s0304-3940(01)02242-x. [DOI] [PubMed] [Google Scholar]
  • 52.Grace AA. Phasic versus tonic dopamine release and the modulation of dopamine system responsivity: a hypothesis for the etiology of schizophrenia. Neuroscience. 1991;41:1–24. doi: 10.1016/0306-4522(91)90196-u. [DOI] [PubMed] [Google Scholar]
  • 53.Scharfman HE, Goodman JH, Schwarcz R. Electrophysiological effects of exogenous and endogenous kynurenic acid in the rat brain: studies in vivo and in vitro. Amino Acids. 2000;19:283–97. doi: 10.1007/s007260070060. [DOI] [PubMed] [Google Scholar]
  • 54.Carpenedo R, Pittaluga A, Cozzi A, et al. Presynaptic kynurenate-sensitive receptors inhibit glutamate release. Eur J Neurosci. 2001;13:2141–7. doi: 10.1046/j.0953-816x.2001.01592.x. [DOI] [PubMed] [Google Scholar]
  • 55.Rassoulpour A, Wu HQ, Ferre S, Schwarcz R. Nanomolar concentrations of kynurenic acid reduce extracellular dopamine levels in the striatum. J Neurochem. 2005;93:762–5. doi: 10.1111/j.1471-4159.2005.03134.x. [DOI] [PubMed] [Google Scholar]
  • 56.Erhardt S, Oberg H, Mathe JM, Engberg G. Pharmacological elevation of endogenous kynurenic acid levels activates nigral dopamine neurons. Amino Acids. 2001b;20:353–62. doi: 10.1007/s007260170032. [DOI] [PubMed] [Google Scholar]
  • 57.Riedemann NC, Guo RF, Ward PA. Novel strategies for the treatment of sepsis. Nat Med. 2003;9:517–24. doi: 10.1038/nm0503-517. [DOI] [PubMed] [Google Scholar]
  • 58.Moroni F, Fossati S, Chiarugi A, Cozzi A. Kynurenic acid actions in brain and periphery. International Congress Series. 2007;1304:305–13. [Google Scholar]
  • 59.Serio CD, Cozzi A, Angeli I, et al. Kynurenic acid inhibits the release of the neurotrophic fibroblast growth factor (FGF)-l and enhances proliferation of glial cells, in vitro. Cell Mol Neurobiol. 2005 doi: 10.1007/s10571-005-8469-y. [DOI] [PubMed] [Google Scholar]
  • 60.Morita T, Saito K, Takemura M, et al. 3-Hydroxyanthranilic acid, an L-tryptophan metabolite, induces apoptosis in monocyte-derived cells stimulated by interferon-gamma. Ann Clin Biochem. 2001;38:242–51. doi: 10.1258/0004563011900461. [DOI] [PubMed] [Google Scholar]
  • 61.Fallarino F, Grohmann U, Vacca C, et al. T cell apoptosis by tryptophan catabolism. Cell Death Differ. 2002;9:1069–77. doi: 10.1038/sj.cdd.4401073. [DOI] [PubMed] [Google Scholar]
  • 62.Lopez AS, Alegre E, Diaz-lagares A, Garcia-giron C, Coma MJ, Gonzalez A. Effect of 3-hydroxyanthranilic acid in the immunosuppressive molecules indoleamine dioxygenase and HLA-G in macrophages. Immunol Lett. 2008;117:91–5. doi: 10.1016/j.imlet.2008.01.001. [DOI] [PubMed] [Google Scholar]
  • 63.Christen S, Peterhans E, Stocker R. Antioxidant activities of some tryptophan metabolites: possible implication for inflammatory diseases. Proc Natl Acad Sci U S A. 1990;87:2506–10. doi: 10.1073/pnas.87.7.2506. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Sekkai D, Guittet O, Lemaire G, Tenu JP, Lepoivre M. Inhibition of nitric oxide synthase expression and activity in macrophages by 3-hydroxyanthranilic acid, a tryptophan metabolite. Arch Biochem Biophys. 1997;340:117–23. doi: 10.1006/abbi.1997.9913. [DOI] [PubMed] [Google Scholar]
  • 65.Nathan CF, Hibbs JB., Jr Role of nitric oxide synthesis in macrophage antimicrobial activity. Curr Opin Immunol. 1991;3:65–70. doi: 10.1016/0952-7915(91)90079-g. [DOI] [PubMed] [Google Scholar]
  • 66.Weber WP, Feder-mengus C, Chiarugi A, et al. Differential effects of the tryptophan metabolite 3-hydroxyanthranilic acid on the proliferation of human CD8+ T cells induced by TCR triggering or homeostatic cytokines. Eur J Immunol. 2006;36:296–304. doi: 10.1002/eji.200535616. [DOI] [PubMed] [Google Scholar]
  • 67.Platten M, Ho PP, Youssef S, et al. Treatment of autoimmune neuroinflammation with a synthetic tryptophan metabolite. Science. 2005;310:850–5. doi: 10.1126/science.1117634. [DOI] [PubMed] [Google Scholar]
  • 68.Fernandez-pol JA, Bono VH, Johnson GS. Control of growth by picolinic acid: differential response of normal and transformed cells. Proc Natl Acad Sci U S A. 1977;74:2889–93. doi: 10.1073/pnas.74.7.2889. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Guillemin GJ, Cullen KM, Lim CK, et al. Characterization of the kynurenine pathway in human neurons. J Neurosci. 2007c;27:12884–92. doi: 10.1523/JNEUROSCI.4101-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Varesio L, Clayton M, Blasi E, Ruffman R, Radzioch D. Picolinic acid, a catabolite of tryptophan, as the second signal in the activation of IFN-gamma-primed macrophages. J Immunol. 1990;145:4265–71. [PubMed] [Google Scholar]
  • 71.Ruffmann R, Schlick R, Chirigos MA, Budzynsky W, Varesio L. Antiproliferative activity of picolinic acid due to macrophage activation. Drugs Exp Clin Res. 1987;13:607–14. [PubMed] [Google Scholar]
  • 72.Abe S, Hu W, Ishibashi H, Hasumi K, Yamaguchi H. Augmented inhibition of Candida albicans growth by murine neutrophils in the presence of a tryptophan metabolite, picolinic acid. J Infect Chemother. 2004;10:181–4. doi: 10.1007/s10156-004-0311-9. [DOI] [PubMed] [Google Scholar]
  • 73.Blasi E, Mazzolla R, Pitzurra L, Barluzzi R, Bistoni F. Protective effect of picolinic acid on mice intracerebrally infected with lethal doses of Candida albicans. Antimicrob Agents Chemother. 1993;37:2422–6. doi: 10.1128/aac.37.11.2422. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Cox GW, Chattopadhyay U, Oppenheim JJ, Varesio L. IL-4 inhibits the costimulatory activity of IL-2 or picolinic acid but not of lipopolysaccharide on IFN-gamma-treated macrophages. J Immunol. 1991;147:3809–14. [PubMed] [Google Scholar]
  • 75.Melillo G, Cox GW, Biragyn A, Sheffler LA, Varesio L. Regulation of nitric-oxide synthase mRNA expression by interferon-gamma and picolinic acid. J Biol Chem. 1994;269:8128–33. [PubMed] [Google Scholar]
  • 76.Pais TF, Appelberg R. Macrophage control of mycobacterial growth induced by picolinic acid is dependent on host cell apoptosis. J Immunol. 2000;164:389–97. doi: 10.4049/jimmunol.164.1.389. [DOI] [PubMed] [Google Scholar]
  • 77.Alkhatib G, Combadiere C, Broder CC, et al. CC CKR5: a Rantes MIP-1 alpha, MlP-lbeta receptor as a fusion cofactor for macrophage-tropic HIV-1. Science. 1996;272:1955–8. doi: 10.1126/science.272.5270.1955. [DOI] [PubMed] [Google Scholar]
  • 78.Fernandez-pol JA, Klos DJ, Hamilton PD. Antiviral cytotoxic and apoptotic activities of picolinic acid on human immunodeficiency virus-1 and human herpes simplex virus-2 infected cells. Anticancer Res. 2001;21:3773–6. [PubMed] [Google Scholar]
  • 79.Bosco MC, Rapisarda A, Massazza S, Melillo G, Young H, Varesio L. The tryptophan catabolite picolinic acid selectively induces the chemokines macrophage inflammatory protein-1 alpha and -1 beta in macrophages. J Immunol. 2000;164:3283–91. doi: 10.4049/jimmunol.164.6.3283. [DOI] [PubMed] [Google Scholar]
  • 80.Cocchi F, Devico AL, Garzino-demo A, Arya SK, Gallo RC, Lusso P. Identification of Rantes MIP-1 alpha, and MIP-1 beta as the major HIV-suppressive factors produced by CD8+ T cells. Science. 1995;270:1811–5. doi: 10.1126/science.270.5243.1811. [DOI] [PubMed] [Google Scholar]
  • 81.Rapisarda A, Pastorino S, Massazza S, Varesio L, Bosco MC. Antagonistic effect of picolinic acid and interferon-gamma on macrophage inflammatory protein-1 alpha/beta production. Cell Immunol. 2002;220:70–80. doi: 10.1016/s0008-8749(03)00008-x. [DOI] [PubMed] [Google Scholar]
  • 82.Beninger RJ, Colton AM, Ingles JL, Jhamandas K, Boegman RJ. Picolinic acid blocks the neurotoxic but not the neuroexcitant properties of quinolinic acid in the rat brain: evidence from turning behaviour and tyrosine hydroxylase immunohistochemistry. Neuroscience. 1994;61:603–12. doi: 10.1016/0306-4522(94)90438-3. [DOI] [PubMed] [Google Scholar]
  • 83.Jhamandas KH, Boegman RJ, Beninger RJ, Flesher S. Role of zinc in blockade of excitotoxic action of quinolinic acid by picolinic acid. Amino Acids. 1998;14:257–61. doi: 10.1007/BF01345272. [DOI] [PubMed] [Google Scholar]
  • 84.Vrooman L, Jhamandas K, Boegman RJ, Beninger RJ. Picolinic acid modulates kainic acid-evoked glutamate release from the striatum in vitro. Brain Res. 1993;627:193–8. doi: 10.1016/0006-8993(93)90320-m. [DOI] [PubMed] [Google Scholar]
  • 85.Cockhill J, Jhamandas K, Boegman RJ, Beninger RJ. Action of picolinic acid and structurally related pyridine carboxylic acids on quinolinic acid-induced cortical cholinergic damage. Brain Res. 1992;599:57–63. doi: 10.1016/0006-8993(92)90852-z. [DOI] [PubMed] [Google Scholar]
  • 86.Heyes MP, Chen CY, Major EO, Saito K. Different kynurenine pathway enzymes limit quinolinic acid formation by various human cell types. Biochem J. 1997;326(Pt 2):351–6. doi: 10.1042/bj3260351. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Heyes MP, Saito K, Crowley JS, et al. Quinolinic acid and kynurenine pathway metabolism in inflammatory and non-inflammatory neurological disease. Brain. 1992;115(Pt 5):1249–73. doi: 10.1093/brain/115.5.1249. [DOI] [PubMed] [Google Scholar]
  • 88.Flanagan EM, Erickson JB, Viveros OH, Chang SY, Reinhard JF., Jr Neurotoxin quinolinic acid is selectively elevated in spinal cords of rats with experimental allergic encephalomyelitis. J Neurochem. 1995;64:1192–6. doi: 10.1046/j.1471-4159.1995.64031192.x. [DOI] [PubMed] [Google Scholar]
  • 89.Espey MG, Tang Y, Morse HC, 3rd, Moffett JR, Namboodiri MA. Localization of quinolinic acid in the murine AIDS model of retrovirus-induced immunodeficiency: implications for neurotoxicity and dendritic cell immunopathogenesis. Aids. 1996;10:151–8. doi: 10.1097/00002030-199602000-00004. [DOI] [PubMed] [Google Scholar]
  • 90.Brew BJ, Corbeil J, Pemberton L, et al. Quinolinic acid production is related to macrophage tropic isolates of HIV-1. J Neurovirol. 1995;1:369–74. doi: 10.3109/13550289509111026. [DOI] [PubMed] [Google Scholar]
  • 91.Heyes MP, Achim CL, Wiley CA, Major EO, Saito K, Markey SP. Human microglia convert 1-tryptophan into the neurotoxin quinolinic acid. Biochem J. 1996;320(Pt 2):595–7. doi: 10.1042/bj3200595. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Moffett JR, Espey MG, Gaudet SJ, Namboodiri MA. Antibodies to quinolinic acid reveal localization in select immune cells rather than neurons or astroglia. Brain Res. 1993;623:337–40. doi: 10.1016/0006-8993(93)91450-7. [DOI] [PubMed] [Google Scholar]
  • 93.Guillemin GJ, Smith DG, Kerr SJ, et al. Characterisation of kynurenine pathway metabolism in human astrocytes and implications in neuropathogenesis. Redox Report. 2000;5:108–11. doi: 10.1179/135100000101535375. [DOI] [PubMed] [Google Scholar]
  • 94.Guillemin GJ, Brew BJ, Noonan CE, Takikawa O, Cullen KM. Indoleamine 2,3 dioxygenase and quinolinic acid immunoreactivity in Alzheimer’s disease hippocampus. Neuropathol Appl Neurobiol. 2005a;31:395–404. doi: 10.1111/j.1365-2990.2005.00655.x. [DOI] [PubMed] [Google Scholar]
  • 95.Wolfensberger M, Amsler U, Cuenod M, Foster AC, Whetsell WO, Jr, Schwarcz R. Identification of quinolinic acid in rat and human brain tissue. Neurosci Lett. 1983;41:247–52. doi: 10.1016/0304-3940(83)90458-5. [DOI] [PubMed] [Google Scholar]
  • 96.Belladonna ML, Grohmann U, Guidetti P, et al. Kynurenine pathway enzymes in dendritic cells initiate tolerogenesis in the absence of functional IDO. J Immunol. 2006;111:130–7. doi: 10.4049/jimmunol.177.1.130. [DOI] [PubMed] [Google Scholar]
  • 97.Schwarcz R, Whetsell WO, Jr, Mangano RM. Quinolinic acid: an endogenous metabolite that produces axon-sparing lesions in rat brain. Science. 1983;219:316–8. doi: 10.1126/science.6849138. [DOI] [PubMed] [Google Scholar]
  • 98.Kim JP, Choi DW. Quinolinate neurotoxicity in cortical cell culture. Neuroscience. 1987;23:423–32. doi: 10.1016/0306-4522(87)90066-2. [DOI] [PubMed] [Google Scholar]
  • 99.Whetsell WO, Jr, Schwarcz R. Prolonged exposure to submicromolar concentrations of quinolinic acid causes excitotoxic damage in organotypic cultures of rat corticostriatal system. Neurosci Lett. 1989;97:271–5. doi: 10.1016/0304-3940(89)90609-5. [DOI] [PubMed] [Google Scholar]
  • 100.Widner B, Leblhuber F, Fuchs D. Increased neopterin production and tryptophan degradation in advanced Parkinson’s disease. J Neural Transm. 2002;109:181–9. doi: 10.1007/s007020200014. [DOI] [PubMed] [Google Scholar]
  • 101.Guillemin GJ, Brew BJ, Noonan CE, Takikawa O, Cullen KM. Indoleamine 2,3 dioxygenase and quinolinic acid immunoreactivity in Alzheimer’s disease hippocampus. Neuropathol Appl Neurobiol. 2005b;31:395–404. doi: 10.1111/j.1365-2990.2005.00655.x. [DOI] [PubMed] [Google Scholar]
  • 102.Guillemin GJ, Brew BJ, Noonan CE, Knight TG, Smythe GA, Cullen KM. Mass spectrometric detection of quinolinic acid in microdissected Alzheimer’s disease plaques. In: Takai K, editor. International Congress Series. 2007a. [Google Scholar]
  • 103.Heyes MP, Brew BJ, Martin A, Markey SP, Price RW, Bhalla RB, Salazar A. Cerebrospinal fluid quinolinic acid concentrations are increased in acquired immune deficiency syndrome. Adv Exp Med Biol. 1991a;294:687–90. doi: 10.1007/978-1-4684-5952-4_94. [DOI] [PubMed] [Google Scholar]
  • 104.Heyes MP, Brew BJ, Martin A, et al. Quinolinic acid in cerebrospinal fluid and serum in HIV-1 infection: relationship to clinical and neurological status. Ann Neurol. 1991b;29:202–9. doi: 10.1002/ana.410290215. [DOI] [PubMed] [Google Scholar]
  • 105.Guillemin GJ, Kerr SJ, Brew BJ. Involvement of quinolinic acid in ADDS dementia complex. Neurotox Res. 2004;7:103–24. doi: 10.1007/BF03033781. [DOI] [PubMed] [Google Scholar]
  • 106.Cascino A, Cangiano C, Ceci F, et al. Increased plasma free tryptophan levels in human cancer: a tumor related effect. Anticancer Research. 1991;11:1313–6. [PubMed] [Google Scholar]
  • 107.Schroecksnadel K, Winkler C, Duftner C, Wirleitner B, Schirmer M, Fuchs D. Tryptophan degradation increases with stage in patients with rheumatoid arthritis. Clin Rheumatol. 2006;25:334–7. doi: 10.1007/s10067-005-0056-6. [DOI] [PubMed] [Google Scholar]
  • 108.Giusti RM, Maloney EM, Hanchard B, et al. Differential patterns of serum biomarkers of immune activation in human T-cell lymphotropic virus type I-associated myelopathy/tropical spastic paraparesis, and adult T-cell leukemia/lymphoma. Cancer Epidemiol Biomarkers Prev. 1996;5:699–704. [PubMed] [Google Scholar]
  • 109.Medana IM, Day NP, Salahifar-sabet H, et al. Metabolites of the kynurenine pathway of tryptophan metabolism in the cerebrospinal fluid of Malawian children with malaria. J Infect Dis. 2003;188:844–9. doi: 10.1086/377583. [DOI] [PubMed] [Google Scholar]
  • 110.Frick B, Schroecksnadel K, Neurauter G, Leblhuber F, Fuchs D. Increasing production of homocysteine and neopterin and degradation of tryptophan with older age. Clin Biochem. 2004;37:684–7. doi: 10.1016/j.clinbiochem.2004.02.007. [DOI] [PubMed] [Google Scholar]
  • 111.Schwarcz R, Rassoulpour A, Wu HQ, Medoff D, Tamminga CA, Roberts RC. Increased cortical kynurenate content in schizophrenia. Biol Psychiatry. 2001;50:521–30. doi: 10.1016/s0006-3223(01)01078-2. [DOI] [PubMed] [Google Scholar]
  • 112.Murr C, Gerlach D, Widner B, Dierich MP, Fuchs D. Neopterin production and tryptophan degradation in humans infected by Streptococcus pyogenes. Med Microbiol Immunol. 2001;189:161–3. doi: 10.1007/s430-001-8023-3. [DOI] [PubMed] [Google Scholar]
  • 113.Schroecksnadel K, Winkler C, Fuith LC, Fuchs D. Tryptophan degradation in patients with gynecological cancer correlates with immune activation. Cancer Lett. 2005;223:323–9. doi: 10.1016/j.canlet.2004.10.033. [DOI] [PubMed] [Google Scholar]
  • 114.Fuchs D, Moller AA, Reibnegger G, Werner ER, Werner-felmayer G, Dierich MP, Wachter H. Increased endogenous interferon-gamma and neopterin correlate with increased degradation of tryptophan in human immunodeficiency virus type 1 infection. Immunol Lett. 1991;28:207–11. doi: 10.1016/0165-2478(91)90005-u. [DOI] [PubMed] [Google Scholar]
  • 115.Look MP, Altfeld M, Kreuzer KA, et al. Parallel decrease in neurotoxin quinolinic acid and soluble tumor necrosis factor receptor p75 in serum during highly active antiretroviral therapy of HIV type 1 disease. AIDS Res Hum Retroviruses. 2000;16:1215–21. doi: 10.1089/08892220050116989. [DOI] [PubMed] [Google Scholar]
  • 116.Zangerle R, Widner B, Quirchmair G, Neurauter G, Sarcletti M, Fuchs D. Effective antiretroviral therapy reduces degradation of tryptophan in patients with HIV-1 infection. Clin Immunol. 2002;104:242–7. doi: 10.1006/clim.2002.5231. [DOI] [PubMed] [Google Scholar]
  • 117.Demitrack MA, Heyes MP, Altemus M, Pigott TA, Gold PW. Cerebrospinal fluid levels of kynurenine pathway metabolites in patients with eating disorders: relation to clinical and biochemical variable. Biol Psychiatry. 1995;37:512–20. doi: 10.1016/0006-3223(94)00173-Z. [DOI] [PubMed] [Google Scholar]
  • 118.Schmitt JA, Wingen M, Ramaekers JG, Evers EA, Riedel WJ. Serotonin and human cognitive performance. Curr Pharm Des. 2006;12:2473–86. doi: 10.2174/138161206777698909. [DOI] [PubMed] [Google Scholar]
  • 119.Iwagaki H, Hizuta A, Tanaka N, Orita K. Decreased serum tryptophan in patients with cancer cachexia correlates with increased serum neopterin. Immunol Invest. 1995;24:467–78. doi: 10.3109/08820139509066843. [DOI] [PubMed] [Google Scholar]
  • 120.Denz H, Orth B, Weiss G, Herrmann R, Huber P, Wachter H, Fuchs D. Weight loss in patients with hematological neoplasias is associated with immune system stimulation. Clin Investig. 1993;71:37–41. doi: 10.1007/BF00210961. [DOI] [PubMed] [Google Scholar]
  • 121.Beutler B, Greenwald D, Hulmes JD, et al. Identity of tumour necrosis factor and the macrophage-secreted factor cachectin. Nature. 1985;316:552–4. doi: 10.1038/316552a0. [DOI] [PubMed] [Google Scholar]
  • 122.Riedel WJ, Klaassen T, Schmitt JA. Tryptophan mood, and cognitive function. Brain Behav Immun. 2002;16:581–9. doi: 10.1016/s0889-1591(02)00013-2. [DOI] [PubMed] [Google Scholar]
  • 123.Myint AM, Kim YM, Verkerk R, Scharpe S, Steinbusch H, Leonard B. Kynurenine pathway in major depression: evidence of impaired neuroprotection. J Affect Disord. 2007;98:143–51. doi: 10.1016/j.jad.2006.07.013. [DOI] [PubMed] [Google Scholar]
  • 124.Maes M, Scharpe S, Meltzer HY, et al. Increased neopterin and interferon-gamma secretion and lower availability of L-tryptophan in major depression: further evidence for an immune response. Psychiatry Res. 1994;54:143–60. doi: 10.1016/0165-1781(94)90003-5. [DOI] [PubMed] [Google Scholar]
  • 125.Baig S, Halawa I, Qureshi GA. High performance liquid chromatography as a tool in the definition of abnormalities in monoamine and tryptophan metabolites in cerebrospinal fluid from patients with neurological disorders. Biomed Chromatogr. 1991;5:108–12. doi: 10.1002/bmc.1130050304. [DOI] [PubMed] [Google Scholar]
  • 126.Smith RS. The macrophage theory of depression. Med Hypotheses. 1991;35:298–306. doi: 10.1016/0306-9877(91)90272-z. [DOI] [PubMed] [Google Scholar]
  • 127.Wichers MC, Koek GH, Robaeys G, Verkerk R, Scharpe S, Maes M. IDO and interferon-alpha-induced depressive symptoms: a shift in hypothesis from tryptophan depletion to neurotoxicity. Mol Psychiatry. 2005;10:538–44. doi: 10.1038/sj.mp.4001600. [DOI] [PubMed] [Google Scholar]
  • 128.Minden SL, Orav J, Reich P. Depression in multiple sclerosis. Gen Hosp Psychiatry. 1987;9:426–34. doi: 10.1016/0163-8343(87)90052-1. [DOI] [PubMed] [Google Scholar]
  • 129.Katon W, Sullivan MD. Depression and chronic medical illness. J Clin Psychiatry. 1990;51(Suppl 3–11) discussion 12–4. [PubMed] [Google Scholar]
  • 130.Riedel WJ, Klaassen T, Deutz NE, Van someren A, Van praag HM. Tryptophan depletion in normal volunteers produces selective impairment in memory consolidation. Psychopharmacology (Berl) 1999;141:362–9. doi: 10.1007/s002130050845. [DOI] [PubMed] [Google Scholar]
  • 131.Harmer CJ, Bhagwagar Z, Cowen PJ, Goodwin GM. Acute administration of citalopram facilitates memory consolidation in healthy volunteers. Psychopharmacology (Berl) 2002;163:106–10. doi: 10.1007/s00213-002-1151-x. [DOI] [PubMed] [Google Scholar]
  • 132.Leblhuber F, Walli J, Jellinger K, et al. Activated immune system in patients with Huntington’s disease. Clin Chem Lab Med. 1998;36:747–50. doi: 10.1515/CCLM.1998.132. [DOI] [PubMed] [Google Scholar]
  • 133.Widner B, Leblhuber F, Walli J, Tilz GP, Demel U, Fuchs D. Tryptophan degradation and immune activation in Alzheimer’s disease. J Neural Transm. 2000;107:343–53. doi: 10.1007/s007020050029. [DOI] [PubMed] [Google Scholar]
  • 134.Gendelman HE, Zheng J, Coulter CL, et al. Suppression of inflammatory neurotoxins by highly active antiretroviral therapy in human immunodeficiency virus-associated dementia. J Infect Dis. 1998;178:1000–7. doi: 10.1086/515693. [DOI] [PubMed] [Google Scholar]
  • 135.Suarez S, Baril L, Stankoff B, et al. Outcome of patients with HIV-1-related cognitive impairment on highly active antiretroviral therapy. Aids. 2001;15:195–200. doi: 10.1097/00002030-200101260-00008. [DOI] [PubMed] [Google Scholar]
  • 136.Mellor AL, Munn DH. Tryptophan catabolism and T-cell tolerance: immunosuppression by starvation. Immunol Today. 1999;20:469–73. doi: 10.1016/s0167-5699(99)01520-0. [DOI] [PubMed] [Google Scholar]
  • 137.Murray MF, Langan M, Macgregor RR. Increased plasma tryptophan in HIV-infected patients treated with pharmacologic doses of nicotinamide. Nutrition. 2001;17:654–6. doi: 10.1016/s0899-9007(01)00568-8. [DOI] [PubMed] [Google Scholar]
  • 138.Abrams B, Duncan D, Hertz-Picciotto I. A prospective study of dietary intake and acquired immune deficiency syndrome in HIV-seropositive homosexual men. J Acquir Immune Defic Syndr. 1993;6:949–58. [PubMed] [Google Scholar]
  • 139.Tang AM, Graham NM, Saah AJ. Effects of micronutrient intake on survival in human immunodeficiency virus type 1 infection. Am J Epidemiol. 1996;143:1244–56. doi: 10.1093/oxfordjournals.aje.a008712. [DOI] [PubMed] [Google Scholar]
  • 140.Muller AJ, Duhadaway JB, Donover PS, Sutanto-Ward E, Prendergast GC. Inhibition of indoleamine 2,3-dioxygenase, an immunoregulatory target of the cancer suppression gene Binl potentiates cancer chemotherapy. Nat Med. 2005;11:312–9. doi: 10.1038/nm1196. [DOI] [PubMed] [Google Scholar]
  • 141.Foster AC, Willis CL, Tridgett R. Protection Against N-methyl-D-aspartate Receptor-Mediated Neuronal Degeneration In Rat Brain by 7-chlorokynurenate and 3-amino-l-hydroxypyrrolid-2-one, Antagonists at The Allosteric Site for Glycine. Eur J Neurosci. 1990;2:270–7. doi: 10.1111/j.1460-9568.1990.tb00418.x. [DOI] [PubMed] [Google Scholar]
  • 142.Hartley DM, Monyer H, Colamarino SA, Choi DW. 7-Chlorokynurenate Blocks NMDA Receptor-Mediated Neurotoxicity in Murine Cortical Culture. Eur J Neurosci. 1990;2:291–5. doi: 10.1111/j.1460-9568.1990.tb00420.x. [DOI] [PubMed] [Google Scholar]
  • 143.Rao TS, Gray NM, Dappen MS, et al. Indole-2-carboxylates, novel antagonists of the N-methyl-D-aspartate (NMDA)-associated glycine recognition sites: in vivo characterization. Neuropharmacology. 1993;32:139–47. doi: 10.1016/0028-3908(93)90094-j. [DOI] [PubMed] [Google Scholar]
  • 144.Hokari M, Wu HQ, Schwarcz R, Smith QR. Facilitated brain uptake of 4-chlorokynurenine and conversion to 7-chlorokynurenic acid. Neuroreport. 1996;8:15–8. doi: 10.1097/00001756-199612200-00004. [DOI] [PubMed] [Google Scholar]
  • 145.Wu HQ, Salituro FG, Schwarcz R. Enzyme-catalyzed production of the neuroprotective NMD A receptor antagonist 7-chlorokynurenic acid in the rat brain in vivo. Eur J Pharmacol. 1997;319:13–20. doi: 10.1016/s0014-2999(96)00829-1. [DOI] [PubMed] [Google Scholar]
  • 146.Wu HQ, Lee SC, Schwarcz R. Systemic administration of 4-chlorokynurenine prevents quinolinate neurotoxicity in the rat hippocampus. Eur J Pharmacol. 2000;390:261–74. doi: 10.1016/s0014-2999(00)00024-8. [DOI] [PubMed] [Google Scholar]
  • 147.Brunmark C, Runstrom A, Ohlsson L, et al. The new orally active immunoregulator laquinimod (ABR-215062) effectively inhibits development and relapses of experimental autoimmune encephalomyelitis. J Neuroimmunol. 2002;130:163–72. doi: 10.1016/s0165-5728(02)00225-4. [DOI] [PubMed] [Google Scholar]
  • 148.Zou LP, Abbas N, Volkmann I, et al. Suppression of experimental autoimmune neuritis by ABR-215062 is associated with altered Thl/Th2 balance and inhibited migration of inflammatory cells into the peripheral nerve tissue. Neuropharmacology. 2002;42:731–9. doi: 10.1016/s0028-3908(02)00015-1. [DOI] [PubMed] [Google Scholar]
  • 149.Yang JS, Xu LY, Xiao BG, Hedlund G, Link H. Laquinimod (ABR-215062) suppresses the development of experimental autoimmune encephalomyelitis, modulates the Thl/Th2 balance and induces the Th3 cytokine TGF-beta in Lewis rats. J Neuroimmunol. 2004;156:3–9. doi: 10.1016/j.jneuroim.2004.02.016. [DOI] [PubMed] [Google Scholar]
  • 150.Yang Y, Hentati A, Deng HX, et al. The gene encoding alsin, a protein with three guanine-nucleotide exchange factor domains, is mutated in a form of recessive amyotrophic lateral sclerosis. Nat Genet. 2001;29:160–5. doi: 10.1038/ng1001-160. [DOI] [PubMed] [Google Scholar]
  • 151.Polman C, Barkhof F, Sandberg-wollheim M, Linde A, Nordle O, Nederman T. Treatment with laquinimod reduces development of active MRI lesions in relapsing MS. Neurology. 2005;64:987–91. doi: 10.1212/01.WNL.0000154520.48391.69. [DOI] [PubMed] [Google Scholar]
  • 152.Williamson RA, Yea CM, Robson PA, et al. Dihydroorotate dehydrogenase is a high affinity binding protein for A77 1726 and mediator of a range of biological effects of the immunomodulatory compound. J Biol Chem. 1995;270:22467–72. doi: 10.1074/jbc.270.38.22467. [DOI] [PubMed] [Google Scholar]
  • 153.O’connor PW, Li D, Freedman MS, et al. A Phase II study of the safety and efficacy of teriflunomide in multiple sclerosis with relapses. Neurology. 2006;66:894–900. doi: 10.1212/01.wnl.0000203121.04509.31. [DOI] [PubMed] [Google Scholar]
  • 154.Suzawa H, Kikuchi S, Arai N, Koda A. The mechanism involved in the inhibitory action of tranilast on collagen biosynthesis of keloid fibroblasts. Jpn J Pharmacol. 1992;60:91–6. doi: 10.1254/jjp.60.91. [DOI] [PubMed] [Google Scholar]
  • 155.Isaji M, Miyata H, Ajisawa Y, Takehana Y, Yoshimura N. Tranilast inhibits the proliferation, chemotaxis and tube formation of human microvascular endothelial cells in vitro and angiogenesis in vivo. Br J Pharmacol. 1997;122:1061–6. doi: 10.1038/sj.bjp.0701493. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Pellicciari R, Natalini B, Costantino G, et al. Modulation of the kynurenine pathway in search for new neuroprotective agents. Synthesis and preliminary evaluation of (m-nitrobenzoyl)alanine, a potent inhibitor of kynurenine-3-hydroxylase. J Med Chem. 1994;37:647–55. doi: 10.1021/jm00031a015. [DOI] [PubMed] [Google Scholar]
  • 157.Connick JH, Heywood GC, Sills GJ, Thompson GG, Brodie MJ, Stone TW. Nicotinylalanine increases cerebral kynurenic acid content and has anticonvulsant activity. Gen Pharmacol. 1992;23:235–9. doi: 10.1016/0306-3623(92)90017-e. [DOI] [PubMed] [Google Scholar]
  • 158.Russi P, Alesiani M, Lombardi G, Davolio P, Pellicciari R, Moroni F. Nicotinylalanine increases the formation of kynurenic acid in the brain and antagonizes convulsions. J Neurochem. 1992;59:2076–80. doi: 10.1111/j.1471-4159.1992.tb10097.x. [DOI] [PubMed] [Google Scholar]
  • 159.Harris CA, Miranda AR, Tanguay JJ, Boegman RJ, Beninger RJ, Jhamandas K. Modulation of striatal quinolinate neurotoxicity by elevation of endogenous brain kynurenic acid. Br J Pharmacol. 1998;124:391–9. doi: 10.1038/sj.bjp.0701834. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Chiarugi A, Carpenedo R, Molina MT, Mattoli L, Pellicciari R, Moroni F. Comparison of the neurochemical and behavioral effects resulting from the inhibition of kynurenine hydroxylase and/or kynureninase. J Neurochem. 1995;65:1176–83. doi: 10.1046/j.1471-4159.1995.65031176.x. [DOI] [PubMed] [Google Scholar]
  • 161.Chiarugi A, Moroni F. Quinolinic acid formation in immune-activated mice: studies with (m-nitrobenzoyl)-alanine (mNBA) and 3,4-dimethoxy-[-N-4-(-3-nitrophenyl)thiazol-2yl]-benzenesul fonamide (Ro 61–8048), two potent and selective inhibitors of kynurenine hydroxylase. Neuropharmacology. 1999;38:1225–33. doi: 10.1016/s0028-3908(99)00048-9. [DOI] [PubMed] [Google Scholar]
  • 162.Moroni F, Cozzi A, Carpendo R, Cipriani G, Veneroni O, Izzo E. Kynurenine 3-mono-oxygenase inhibitors reduce glutamate concentration in the extracellular spaces of the basal ganglia but not in those of the cortex or hippocampus. Neuropharmacology. 2005;48:788–95. doi: 10.1016/j.neuropharm.2004.10.019. [DOI] [PubMed] [Google Scholar]
  • 163.Chiarugi A, Cozzi A, Ballerini C, Massacesi L, Moroni F. Kynurenine 3-mono-oxygenase activity and neurotoxic kynurenine metabolites increase in the spinal cord of rats with experimental allergic encephalomyelitis. Neuroscience. 2001;102:687–95. doi: 10.1016/s0306-4522(00)00504-2. [DOI] [PubMed] [Google Scholar]
  • 164.Clark CJ, Mackay GM, Smythe GA, Bustamante S, Stone TW, Phillips RS. Prolonged survival of a murine model of cerebral malaria by kynurenine pathway inhibition. Infect Immun. 2005;73:5249–51. doi: 10.1128/IAI.73.8.5249-5251.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165.Werner ER, Fuchs D, Hausen A, et al. Tryptophan degradation in patients infected by human immunodeficiency virus. Biol Chem Hoppe Seyler. 1988;369:337–40. doi: 10.1515/bchm3.1988.369.1.337. [DOI] [PubMed] [Google Scholar]
  • 166.Larsson M, Hagberg L, Norkrans G, Forsman A. Indole amine deficiency in blood and cerebrospinal fluid from patients with human immunodeficiency virus infection. J Neurosci Res. 1989;23:441–6. doi: 10.1002/jnr.490230410. [DOI] [PubMed] [Google Scholar]
  • 167.Heyes MP, Saito K, Devinsky O, Nadi NS. Kynurenine pathway metabolites in cerebrospinal fluid and serum in complex partial seizures. Epilepsia. 1994;35:251–7. doi: 10.1111/j.1528-1157.1994.tb02428.x. [DOI] [PubMed] [Google Scholar]
  • 168.Orlikov AB, Prakhye IB, Ryzov IV. Kynurenine in blood plasma and DST in patients with endogenous anxiety and endogenous depression. Biol Psychiatry. 1994;36:97–102. doi: 10.1016/0006-3223(94)91189-4. [DOI] [PubMed] [Google Scholar]
  • 169.Fujigaki S, Saito K, Takemura M, et al. Species differences in L-tryptophan-kynurenine pathway metabolism: quantification of anthranilic acid and its related enzymes. Arch Biochem Biophys. 1998;358:329–35. doi: 10.1006/abbi.1998.0861. [DOI] [PubMed] [Google Scholar]
  • 170.Heyes MP, Saito K, Lackner A, Wiley CA, Achim CL, Markey SP. Sources of the neurotoxin quinolinic acid in the brain of HIV1-infected patients and retrovirus-infected macaques. Faseb J. 1998;12:881–96. doi: 10.1096/fasebj.12.10.881. [DOI] [PubMed] [Google Scholar]
  • 171.Huengsberg M, Winer JB, Gompels M, Round R, Ross J, Shahmanesh M. Serum kynurenine-to-tryptophan ratio increases with progressive disease in HIV-infected patients. Clin Chem. 1998;44:858–62. [PubMed] [Google Scholar]
  • 172.Huang A, Fuchs D, Widner B, Glover C, Henderson DC, Allen-mersh TG. Serum tryptophan decrease correlates with immune activation and impaired quality of life in colorectal cancer. Br J Cancer. 2002;86:1691–6. doi: 10.1038/sj.bjc.6600336. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Ilzecka J, Kocki T, Stelmasiak Z, Turski WA. Endogenous protectant kynurenic acid in amyotrophic lateral sclerosis. Acta Neurol Scand. 2003;107:412–8. doi: 10.1034/j.1600-0404.2003.00076.x. [DOI] [PubMed] [Google Scholar]
  • 174.Schrocksnadel K, Widner B, Bergant A, et al. Longitudinal study of tryptophan degradation during and after pregnancy. Life Sci. 2003;72:785–93. doi: 10.1016/s0024-3205(02)02304-4. [DOI] [PubMed] [Google Scholar]
  • 175.Wirleitner B, Rudzite V, Neurauter G, et al. Immune activation and degradation of tryptophan in coronary heart disease. Eur J Clin Invest. 2003;33:550–4. doi: 10.1046/j.1365-2362.2003.01186.x. [DOI] [PubMed] [Google Scholar]
  • 176.Schrocksnadel K, Winkler C, Fuith LC, Fuchs D. Tryptophan degradation in patients with gynecological cancer correlates with immune activation. Cancer Lett. 2005;223:323–9. doi: 10.1016/j.canlet.2004.10.033. [DOI] [PubMed] [Google Scholar]
  • 177.Forrest CM, Mackay GM, Oxford L, Stoy N, Stone TW, Darlington LG. Kynurenine pathway metabolism in patients with osteoporosis after 2 years of drug treatment. Clin Exp Pharmacol Physiol. 2006;33:1078–87. doi: 10.1111/j.1440-1681.2006.04490.x. [DOI] [PubMed] [Google Scholar]
  • 178.Mackay GM, Forrest CM, Stoy N, et al. Tryptophan metabolism and oxidative stress in patients with chronic brain injury. Eur J Neurol. 2006;13:30–42. doi: 10.1111/j.1468-1331.2006.01220.x. [DOI] [PubMed] [Google Scholar]
  • 179.Darlington LG, Mackay GM, Forrest CM, Stoy N, George C, Stone TW. Altered kynurenine metabolism correlates with infarct volume in stroke. Eur J Neurosci. 2007;26:2211–21. doi: 10.1111/j.1460-9568.2007.05838.x. [DOI] [PubMed] [Google Scholar]
  • 180.Hartai Z, Juhasz A, Rimanoczy A, et al. Decreased serum and red blood cell kynurenic acid levels in Alzheimer’s disease. Neurochem Int. 2007;50:308–13. doi: 10.1016/j.neuint.2006.08.012. [DOI] [PubMed] [Google Scholar]
  • 181.Schrocksnadel K, Winkler C, Duftner C, Wirleitner B, Schirmer M, Fuchs D. Tryptophan degradation increases with stage in patients with rheumatoid arthritis. Clin Rheumatol. 2006;25:334–7. doi: 10.1007/s10067-005-0056-6. [DOI] [PubMed] [Google Scholar]
  • 182.Chen Y, Stankovic R, Cullen KA, et al. The kynurenine pathway and inflammation in amyotrophic lateral sclerosis. [Abstract] Amyotroph Lateral Scler. (Suppl. 1):81. doi: 10.1007/s12640-009-9129-7. [DOI] [PubMed] [Google Scholar]
  • 183.Young SN, Joseph MH, Gauthier S. Studies on kynurenine in human cerebrospinal fluid: lowered levels in epilepsy. J Neural Transm. 1983;58:193–204. doi: 10.1007/BF01252805. [DOI] [PubMed] [Google Scholar]
  • 184.Gisslen M, Larsson M, Norkrans G, Fuchs D, Wachter H, Hagberg L. Tryptophan concentrations increase in cerebrospinal fluid and blood after zidovudine treatment in patients with HIV type 1 infection. AIDS Res Hum Retroviruses. 1994;10:947–51. doi: 10.1089/aid.1994.10.947. [DOI] [PubMed] [Google Scholar]
  • 185.Heyes MP, Saito K, Milstien S, Schiff SJ. Quinolinic acid in tumors, hemorrhage and bacterial infections of the central nervous system in children. J Neurol Sci. 1995;133:112–8. doi: 10.1016/0022-510x(95)00164-w. [DOI] [PubMed] [Google Scholar]
  • 186.Medana IM, Hien TT, Day NP, et al. The clinical significance of cerebrospinal fluid levels of kynurenine pathway metabolites and lactate in severe malaria. J Infect Dis. 2002;185:650–6. doi: 10.1086/339009. [DOI] [PubMed] [Google Scholar]
  • 187.Rejdak K, Bartosik-psujek H, Dobosz B, et al. Decreased level of kynurenic acid in cerebrospinal fluid of relapsing-onset multiple sclerosis patients. Neurosci Lett. 2002;331:63–5. doi: 10.1016/s0304-3940(02)00710-3. [DOI] [PubMed] [Google Scholar]
  • 188.Nilsson LK, Linderholm KR, Engberg G, et al. Elevated levels of kynurenic acid in the cerebrospinal fluid of male patients with schizophrenia. Schizophr Res. 2005;80:315–22. doi: 10.1016/j.schres.2005.07.013. [DOI] [PubMed] [Google Scholar]
  • 189.Atlas A, Gisslen M, Nordin C, Lindstrom L, Schwieler L. Acute psychotic symptoms in HIV-1 infected patients are associated with increased levels of kynurenic acid in cerebrospinal fluid. Brain Behav Immun. 2007;21:86–91. doi: 10.1016/j.bbi.2006.02.005. [DOI] [PubMed] [Google Scholar]
  • 190.Beal MF, Matson WR, Storey E, et al. Kynurenic acid concentrations are reduced in Huntington’s disease cerebral cortex. J Neurol Sci. 1992;108:80–7. doi: 10.1016/0022-510x(92)90191-m. [DOI] [PubMed] [Google Scholar]
  • 191.Pearson SJ, Reynolds GP. Increased brain concentrations of a neurotoxin, 3-hydroxykynurenine, in Huntington’s disease. Neurosci Lett. 1992;144:199–201. doi: 10.1016/0304-3940(92)90749-w. [DOI] [PubMed] [Google Scholar]
  • 192.Sardar AM, Bell JE, Reynolds GP. Increased concentrations of the neurotoxin 3-hydroxykynurenine in the frontal cortex of HIV1-positive patients. J Neurochem. 1995;64:932–5. doi: 10.1046/j.1471-4159.1995.64020932.x. [DOI] [PubMed] [Google Scholar]
  • 193.Bara H, Hainfellner JA, Kepplinger B, Mazal PR, Shchmid H, Budka Kynurenic acid metabolism in the brain of HIV-1 infected patients. J Neural Transm. 2000;107:1127–38. doi: 10.1007/s007020070026. [DOI] [PubMed] [Google Scholar]

Articles from International Journal of Tryptophan Research : IJTR are provided here courtesy of SAGE Publications

RESOURCES