Next Article in Journal
A Preliminary Study Indicating Improvement in the Median Survival Time of Glioblastoma Multiforme Patients by the Application of Deuterium Depletion in Combination with Conventional Therapy
Next Article in Special Issue
The Prognostic and Predictive Value of Human Gastrointestinal Microbiome and Exosomal mRNA Expression of PD-L1 and IFNγ for Immune Checkpoint Inhibitors Response in Metastatic Melanoma Patients: PROTOCOL TRIAL
Previous Article in Journal
Immunohistological Analysis of Lichen Sclerosus of the Foreskin in Pediatric Age: Could It Be Considered a Premalignant Lesion?
Previous Article in Special Issue
Updates on Lymphovascular Invasion in Breast Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Indoleamine 2,3-Dioxygenase (IDO) Activity: A Perspective Biomarker for Laboratory Determination in Tumor Immunotherapy

1
Department of Laboratory Medicine, Beijing Hospital, National Center of Gerontology, Institute of Geriatric Medicine, Chinese Academy of Medical Sciences, Beijing 100730, China
2
The Key Laboratory of Geriatrics, Beijing Institute of Geriatrics, Institute of Geriatric Medicine, Chinese Academy of Medical Sciences, Beijing Hospital/National Center of Gerontology of National Health Commission, Beijing 100730, China
*
Author to whom correspondence should be addressed.
Biomedicines 2023, 11(7), 1988; https://doi.org/10.3390/biomedicines11071988
Submission received: 14 June 2023 / Revised: 5 July 2023 / Accepted: 12 July 2023 / Published: 13 July 2023

Abstract

:
Indoleamine 2,3-dioxygenase 1 (IDO1) is a heme enzyme involved in catalyzing the conversion of tryptophan (Trp) into kynurenine (Kyn) at the first rate-limiting step in the kynurenine pathway of L-tryptophan metabolism. It has been found to be involved in several biological functions such as aging, immune microorganism, neurodegenerative and infectious diseases, and cancer. IDO1 plays an important role in immune tolerance by depleting tryptophan in the tumor microenvironment and inhibiting the proliferation of effector T cells, which makes it an important emerging biomarker for cancer immunotherapy. Therefore, the research and development of IDO1 inhibitors are of great importance for tumor therapy. Of interest, IDO activity assays are of great value in the screening and evaluation of inhibitors. Herein, we mainly review the biological functions of IDO1, immune regulation, key signaling molecules in the response pathway, and the development of IDO1 inhibitors in clinical trials. Furthermore, this review provides a comprehensive overview and, in particular, a discussion of currently available IDO activity assays for use in the evaluation of IDO inhibitors in human blood. We believe that the IDO activity is a promising biomarker for the immune escape and laboratory evaluation of tumor immunotherapy.

1. Introduction

L-Tryptophan (Trp) is the least abundant essential amino acid with an indole ring backbone, which is required for a variety of cellular activities and normal physiological functions [1]. Trp is metabolized via four major pathways: decarboxylation for producing tryptamine, protein synthesis, the serotonin pathway, and the kynurenine (Kyn) pathway (KP), where ~95% of Trp is degraded through the KP as the substrate for indoleamine-2,3-dioxygenase 1 (IDO1), generating N-Formyl kynurenine (NFK) [2]. NFK is rapidly converted to Kyn, catalyzed via formamidase (FAMID). Kyn can then be metabolized to kynurenic acid through the activation of kynurenine aminotransferases (KATs), participating in the aryl hydrocarbon receptor (AhR) signaling pathway. Meanwhile, kynurenine 3-monooxygenase (KMO) catalyzes Kyn to 3-hydroxykynurenine, which is subsequently converted to 3-hydroxyanthranilic acid by Kynureninase (KYNU). And then 3-hydroxyanthranilic acid is metabolized to quinolinic acid (QUIN). QUIN has toxic effects but are metabolized to the de novo nicotinamide adenine dinucleotide (NAD+) biosynthesis pathway [3,4] (Figure 1).
In addition to the IDO, KATs, and KMO described above, the kynurenine pathway involves other important enzymes, such as tryptophan dioxygenase (TDO) and 3-hydroxyanthranilic acid dioxygenase (3HAAO) [5,6]. In mammals, the three rate-limiting catabolic enzymes, i.e., IDO1, IDO2, and TDO2, can break the indole ring, each of which has specialized tissue-specific and regulated expression [7,8,9]. They play an important role in regulating immune response and tumor immune escape. Although IDO1, IDO2, and TDO2 have similar biochemical properties, they are different in structure, substrate conversion rate, substrate specificity, tissue distribution, and expression regulation [9,10,11]. IDO1 can be expressed in many types of biological tissues, and it increases with age in specific tissues [12,13]. IDO2 is a genetic paralog of IDO1 and is directly adjacent to IDO1 on the same chromosome. The expression level of IDO2 is low in a more restricted set of tissues, including liver, kidney, lymph nodes, and placenta [14]. Similarly, the expression of TDO2 is relatively small, predominantly in the liver where it plays a key role in maintaining the systemic homeostasis of the tryptophan levels [15]. Here, we will focus on the biology of IDO1 because, of the three enzymes, it has the highest capacity to catalyze the Trp metabolism, while simultaneously being the tryptophan-metabolizing enzyme most closely linked to immune function [16].
IDO works along the tryptophan-depleting kynurenine pathway and might trigger an immunosuppressive microenvironment in tumors, which in turn allows tumor cells to escape immune recognition and cytotoxicity [17]. Therefore, the IDO activity involved in KP plays an important role in the laboratory diagnosis of multiple types of human cancers. An interesting study performed by Botticelli et al. shows that higher kynurenine/tryptophan (Kyn/Trp) ratio representing IDO1 activity could predict resistance to anti-PD-1 treatment in NSCLC, which suggest the possibility of using anti-PD-1 plus IDO inhibitors in those patients with a high level of Kyn/Trp ratio. Consequently, IDO activity has been suggested as a possible mechanism of resistance to anti-PD-1 treatment, leading to an immunosuppressive microenvironment [18,19]. Recently, it was also reported that platinum-resistant lung tumors utilized both IDO1 and TDO2 enzymes for survival in order to escape immune surveillance [20]. In conclusion, high IDO activity can serve as a prognostic and predictive biomarker to optimize the immunotherapy of patients with lung tumors [21]. In this review, we summarized the characteristics of IDO as a biomarker in clinical trials, focusing on various methods to determine IDO activity so far.

2. Immune Regulation of IDO Pathway

IDO1 (EC 1.13.11.42), a heme-containing dioxygenase, assumes a key role in catalyzing the catabolism of mammalian L-tryptophan [22,23]. Studies have confirmed that human IDO1 is a monomeric enzyme with high enzymatic activity for the conversion of Trp to Kyn and its metabolites [24]. IDO pathways have important effects on T cells in response to antigenic stimulation (Figure 2). It activates the tryptophan metabolic pathway, causing a large depletion of Trp. T cells stagnate in G1 phase due to severe lack of Trp, resulting in the suppression of their proliferation. Meanwhile, metabolites such as pyridine carboxylic acid and kynurenine induce T cell apoptosis due to their obvious toxic effects. The suppression of effector T cells allows the tumor to escape host immune surveillance, thus creating an immunosuppressive microenvironment [25,26]. Furthermore, IDO acts on the dendritic cells (DCs) surface and interacts with regulatory T cells (Tregs), producing immunosuppressive effects due to their high expression, through the direct contact or production of suppressive cytokines. In turn, increased Tregs can bind to DC surface-associated receptors and induce IDO expression in DCs [27].
IDO is involved in immune regulation as an important negative immune regulator [28,29,30], and its regulatory mechanism mainly includes the following: (i) upon tryptophan depletion by IDO, the general control nonderepressible-2 (GCN2) kinase is activated, which then binds to tryptophan tRNA and inhibits the mammalian target of rapamycin (mTOR) kinase pathway to modulate immune responses; (ii) the AhR signaling pathway is regulated by kynurenine, the product catalyzed by IDO, resulting in the arrest of T cell activation, the differentiation of Treg cells, the alteration of DC functional immunogenicity, and other immunological responses; and (iii) IDO initiates downstream signaling effectors, including the noncanonical nuclear factor-κB (NF-κB) pathway, which result in sustained tumor growth factor (TGF)-β production and the downregulation of proinflammatory cytokine production in DCs, inducing type I interferons and plasmacytoid dendritic cell (pDC) phenotype. These results promote IDO1 expression, which in turn induces (TGF)-β secretion, thus creating a positive feedback loop. Noninflammatory DCs are generated by activating the noncanonical NF-κB pathway, thereby inhibiting T cell activation and promoting the development of Tregs [31]. In this context, these immunosuppressive effects allow IDO to play important roles in cancers, organ transplantations, and pathogenic infectious diseases [6,27,32,33,34,35,36].

3. Biochemical Properties of IDO in the KP

Studies of IDO1 provided considerable insight into the properties of the enzyme, including substrate specificity, reactivity with dioxygen, and spectroscopic characteristics. Suzuki and co-workers reported that IDO1 was evolutionarily related to myoglobin with a molecular weight comparable with that of IDO1 [37,38,39]. An electron source is required to maintain IDO1 in the active ferrous form to overcome autoxidation to the ferric derivative [40,41]. Tryptophan catabolism begins with the break of the pyrrole ring by IDO1 to produce NFK. The cofactor of this enzyme is ferriporphyrin, and vitamin C is the activator of this enzyme, which has the function of protecting the Fe2+ in the cofactor from oxidation. It was found that IDO1 can oxidize NADH under aerobic conditions [42]. In experiments under aerobic conditions, IDOFe3+ was added to NADH in the absence of any other reactants, resulting in the formation of IDOFe2+-O2 with an increase in absorbance at both 577 nm and 544 nm (Figure 3). Subsequently, the reaction of IDOFe2+-O2 with L-Trp resulted in the production of NFK, as indicated by the increase in absorbance at 321 nm, with the generation of NAD+. A recent study demonstrated for the first time that an increase in immune-mediated IDO activity increases not only NAD+ biosynthesis but also NAD+ catabolism in primary human macrophages [43].

4. A Perspective Biomarker in a Variety of Human Diseases

Many studies have shown that IDO1 is closely associated with a variety of human diseases, including inflammatory diseases, microbial infections, autoimmunity, tumors and psychiatric disorders [27,44,45,46,47,48]. Trp depletion via IDO1 will disable the protein synthesis of T cells and inhibit the proliferation of T cells, resulting in the activation of GCN2 pathway and the suppression of mTOR pathway. Kyn will also activate AhR and then promote immune tolerance [32]. Immune response is the most common mechanism of most kidney diseases, the inhibition of which can delay the progress of kidney diseases. IDO activity was found to be associated with chronic kidney disease (CKD)-related indicators, such as calcium, phosphorus, uric acid, hemoglobin, albumin, coagulation indicators, and hypertension. Ischemic kidney injury can be protected by intervention in the IDO1/kynurenine pathway. In addition, increased IDO1 abundance and stress gene expression was shown in kidney tissue from patients with Ab-driven nephropathy. Therefore, IDO1 is supposed to be a promising biomarker to predict CKD and assess kidney function [49,50,51].
Recently, researchers used microarray data for the first time for immune cell infiltration analysis to identify IDO1 as a diagnostic and prognostic biomarker for diabetic nephropathy (DN) [52]. In addition, IDO expression is significantly induced by intrahepatic bacterial and viral infections as well as by non-pathogenic inflammatory conditions, such as liver fibrosis, cirrhosis, liver tumors, and certain liver parasites, which explains the role of IDO in liver-related diseases [53,54,55]. HBV and HCV infection can cause the upregulation of IDO and induce immunosuppression, while hepatitis B and C are closely associated with the development of hepatocellular carcinoma, so it can be further speculated that IDO may increase the risk of HBV and HCV-induced hepatocellular carcinoma [56,57]. Recent experiments showed that low IDO expression in asthmatic mice weakened immune tolerance, thus making it easier for bronchial asthma to develop or exacerbate [58].
On the one hand, IDO exerts immunosuppressive effects on the cellular microenvironment, leading to infections and immune escape of tumor cells; on the other hand, it also exerts suppressive effects on pathogens, such as bacteria and parasites, protecting the organism from pathogens to some extent. At present, it is found that IDO genes from parasites are related to the immunosuppressive effect of parasites and participate in the immune escape process of certain parasites, but the mechanism is still unclear [59,60]. Dai et al. confirmed that the activation of IDO gene in vivo was partially responsible for the anti-Toxoplasma effect mediated by IFN-γ [61]. In Helicobacter pylori-infected gastric mucosa, the activation of IFN-γ and the regulation of TGFB1 and CTLA4 together induced the upregulation of IDO1 expression [62]. This synergistic mechanism can promote immune tolerance and help bacteria evade host surveillance mechanisms and may reduce the effectiveness of bacterial eradication, promoting the progression of chronic gastritis. Therefore, for the above diseases involving abnormal IDO expression, the study of IDO inhibitors can be valuable in reversing the development and treatment of related diseases.

5. IDO Inhibitors in Clinical Trials for Cancer Immunotherapy

Thus, IDO has become an attractive target in cancer immunotherapy. A large number of IDO inhibitors have been reported, but most of the currently reported IDO inhibitors have still not reached clinical trials [63,64]. To date, six small molecule inhibitors of IDO1 have entered clinical studies, including indoximod (1-D-MT), epacadostat (INCB024360), BMS-986205, navoximod (GDC-0919, NLG-919), PF-06840003, and HTI-1090 (SHR9146) (Figure 1). The major clinical trials evaluating efficacy of these IDO inhibitors are shown in Table 1.
Indoximod (1-methyl-D-tryptophan, 1MT, NLG-8189) is the most representative tryptophan analogue inhibitor developed by the NewLink Genetics (NLNK) company [65]. Trp depletion can lead to the inhibition of mTORC1 signaling pathway. Indoximod, a Trp analogue, can reduce this inhibition and restore the activity of mTORC1, which is a central regulator for cell growth. Therefore, indoximod does not directly inhibit the enzymatic activity of IDO or TDO, but instead resists the effects elicited by these enzymes [66]. Indoximod has good oral bioavailability, safety, and tolerability in treating prostate cancer, breast cancer, non-small cell lung cancer, and brain tumors. It has been shown that the antitumor efficacy of indoximod is significantly enhanced when used in combination with other therapies [67,68].
Epacadostat (INCB024360), a competitive inhibition of IDO, effectively interferes with tryptophan metabolism by competing with tryptophan to bind to the catalytic domain of IDO [69]. Studies showed that epacadostat could reduce tumor growth and promote the proliferation of T cells and NK cells [70]. It is mainly used in ovarian cancer, melanoma, myelodysplastic syndrome, and myelodysplastic syndromes with a high specificity inhibition of IDO1 [71,72]. Moreover, HTI-1090 (SHR9146) is a novel and efficient small molecule IDO1/TDO dual inhibitor with a similar structure to epacadostat and has good oral bioavailability and safety. A Phase I trial of HTRI-1090 in advanced solid tumors was completed in January 2019.
Inrodostat mesylate (BMS-986205), a potent and selective oral IDO1 inhibitor, can specifically bind to IDO1 but not IDO2 or TDO [73]. BMS-986205 occupies the heme cofactor binding site to prevent the further activation of IDO1, thereby reversing the immunosuppression system in cancer patients, potentially improving cancer prognosis, especially when used in combination with other immunotherapies. BMS-986205 combined with Nivolumab in the treatment of liver cancer, melanoma, and solid tumors has been shown to be well tolerated in patients [74,75]. Navoximod (GDC-0919, NLG-919) is an orally active IDO1 inhibitor, contains the 4-phenylimidazole structure of tryptophan IDO inhibitor, which has stronger anti-IDO activity than epacadostat and BMS-986205. Studies showed that navoximod was able to restore the T cell function in vitro, and Kyn levels decreased after navoximod treatment in vivo. It is a promising novel immune checkpoint inhibitor for the treatment of recurrent solid tumors [76,77,78]. PF-06840003 is a non-competitive, non-hemoglobin-binding IDO inhibitor with high bioavailability, which was reported to restore the function of T cells in vitro and to reduce the Kyn levels in vivo. It can cross the blood–brain barrier and, therefore, may interfere with and inhibit tumor brain metastasis for the treatment of glioblastoma [79,80].

6. Applications of IDO Activity Assays

IDO activity induced by cellular immune activation plays a crucial role in tumor-promoting inflammatory processes, cancer formation and immunosuppressive therapy [81]. Recent studies have reported that IDO activity, as reflected by Kyn/Trp ratio, shows significant differences in different types of mental disorders. However, the results of studies on Kyn/Trp calculations for severe mental disorders are inconsistent [82]. Therefore, with the important involvement of IDO in various vital metabolic activities, it is necessary to summarize the currently available methods for the determination of IDO activity, hoping to find a suitable method for the determination of IDO activity to be applied in clinical diagnosis and treatment.

6.1. IDO Enzyme Activity Assays Based on Different Intermediates

In the kynurenine pathway, the formation of certain intermediates is accompanied by the changes of absorption value at different nanometer wavelengths. Therefore, we here briefly summarized the determination of different intermediates. First of all, IDO activity can be determined through the concentration of N-Formyl kynurenine, which is the first intermediate where tryptophan is catalyzed. The assay of DO1 enzyme activity was performed under normal atmospheric oxygen partial pressure (approximately 250 μM O2 in buffer at 24 °C) based on the method described by Shimizu et al. [83]. The experiment mixture contained potassium phosphate buffer, methylene blue, ascorbic acid, catalase, L-tryptophan, and the determined IDO1 enzyme [84]. The increase in absorbance at 321 nm due to the formation of N-Formyl kynurenine was continuously observed and recorded at 24 °C. IDO activity was then calculated according to the change of absorbance value in unit time (Figure 3).
Likewise, IDO activity can also be determined through the concentration of kynurenine produced. For example, the IDO activity can be determined in the cell homogenate via a slight modification of the method (Hara et al.) [85]. Briefly, 30% trichloroacetic acid (TCA) was added to the culture supernatant, vortexed, and spun. The supernatant was then added to an equal amount of Ehrlich reagent (100 mg/5 mL glacial acetic acid). The absorbance at 492 nm was recorded in multifunctional microplate reader (Figure 3). The change in kynurenine concentration was obtained by subtracting the kynurenine content of fresh uncultured media from the value derived from the cultured supernatant [43]. Currently, the ratio of kynurenine/tryptophan is also considered to determine the enzymatic activity of IDO [86]. In addition, the Kyn in the supernatant can also be analyzed by high performance liquid chromatography (HPLC).
Moreover, recent studies have shown that IDO1 can oxidize extra NADH to NAD+ under aerobic conditions. Through binding to NADH, IDO1 was confirmed in the oxidation of L-Trp with the conversion of IDOFe3+ to IDOFe2+-O2 (Figure 3). The NADH-supported oxidation of the L-Trp assay can be detected at three different absorption values as follows. First, the solutions of IDO1 and NADH were mixed in MOPS buffer and incubated at 20 °C. When the spectrum of the sample was completely converted to that of IDOFe2+-O2, the sample was mixed with an equal volume of L-Trp solution with the increasing absorption values at 577 nm and 544 nm that are characteristic of the α and β transitions in the spectrum of IDOFe2+-O2 [41]. Second, the reaction of the NADH-generated IDOFe2+-O2 with L-Trp resulted in the formation of NFK, as indicated by the increasing absorbance at 321 nm. Finally, the consumption of NADH could be detected in the decreasing absorbance maxima at 340 nm with the formation of NAD+ [42]. Moreover, the IDO activity in vivo can also be determined by NAD+, which is the end product of KP metabolism. The IFN-γ activation of mononuclear phagocytes significantly increases IDO and flux through the KP. It was first shown that an immune-mediated increase in IDO activity increased NAD+ production in IFN-γ-stimulated human primary mononuclear cells. Therefore, the determination of NAD+ is able to reflect the activity of IDO in vivo. The cellular pyridine nucleotide content (NAD++NADH) can be measured by the Thiazolyl blue microcycling assay of Bernofsky and Swan adapted to a 96-well plate format [43].

6.2. IDO Activity Assays in Inhibitors In Vitro

In recent years, since IDO inhibitors have been effective in the treatment of a variety of tumors, the establishment of a stable IDO activity assay system is essential for the screening of novel and efficient IDO inhibitors and the evaluation of the effects of IDO mechanisms. The screening process for IDO inhibitors is generally as follows. The purified recombinant human IDO was constructed by genetic engineering. Taking a compound as the object, IDO activity can be initially determined by the several methods described above to obtain the inhibition type, half-inhibition concentration, and inhibition constant. Next, a high expression-human IDO plasmid was constructed and transfected with HEK 293 cells to evaluate the IDO inhibitory activity of the compound at the cellular level. MTT colorimetric assay was used to investigate the growth inhibitory effect of the compound on a specific tumor cell. Then, the animal experiments were performed to be verified by further studies. In addition, the yeast-based screening assay was also regarded as a powerful tool for the early stages of the identification of inhibitors of IDO [87]. Judging from these assays, the activity assay of IDO is crucial and central in the screening and evaluation of IDO inhibitors in vitro.

6.3. IDO Activity Assays in Human Plasma/Serum

Since IDO is a potential biomarker for many solid tumors, the detection of IDO activity in human blood is a direct indicator of the efficacy of tumor therapy. Currently, laboratories can detect the transcript level and expression level of IDO in serum via real-time PCR and Western blotting methods, respectively. Alternatively, the expression of IDO in cancer and paraneoplastic tissues can be determined through semi-quantitative immunohistochemistry. For example, HPLC is used to determine the concentration of Trp and Kyn in human plasma/serum, such as assessing the activity of IDO before and after treatment in patients with early stage NSCLC via Kyn and the kynurenine to tryptophan ratio (Kyn/Trp) in human blood [88]. This ratio is often used instead of the measurement of IDO activity. However, IDO is not the only determinant of the increase in the plasma (Kyn/Trp) ratio in vivo. This ratio is not accurate in vivo due to various factors such as the availability of nutrients, the presence of multiple cell types, complex structural and functional tissue arrangements, and the interactions of extracellular matrix, hormones, cytokines, and paracrine [86]. Moreover, HPLC has not been used clinically due to the inconvenient experiment and the high price. Therefore, the universal simple assay for direct detection of IDO activity in clinical tests is urgently needed.

7. Conclusions

IDO is an evolutionarily ancient enzyme that catalyzes the amino acid tryptophan metabolism, which is also an endogenous mechanism of immune-regulation and tolerance in the immune system. It suppresses the proliferation and differentiation of effector T cells, and significantly enhances the inhibitory activity of Tregs. Thus, this immune escape causes the immune system to fail to respond effectively to tumor antigens, which induces abnormal levels of IDO in host cells [64,89]. Based on these properties, IDO has become both a promising therapeutic target and an unfavorable prognostic biomarker with clinical potential for the diagnosis and treatment of tumor diseases [90,91]. At present, a number of clinical trials are actively underway that some chemotherapy drugs or vaccines against IDO to potentiate anti-Tregs can be combined with IDO inhibitor drugs in some cases to treat solid malignant tumors [92,93]. The current status of the study shows that this combined immunotherapy has shown positive therapeutic effects during the window of opportunity for conventional treatment of multiple tumor diseases [94].
As a consequence, the determination of IDO activity is particularly important both in the diagnosis of oncological diseases and in the evaluation of the effect of IDO inhibitors in immunotherapy. In addition to the assessment of the efficacy of solid malignant tumor therapy, the determination of IDO activity can also be applied to the diagnosis and therapeutic monitoring of various diseases, such as kidney disease, intestinal disease, cardiovascular disease, reproductive system, parasitic infections, mental illness, etc. At present, the determination of IDO activity in human plasma/serum has great application prospects in the diagnosis and treatment of a variety of tumor-related diseases. The calculation method of the Kyn/Trp ratio, which is widely used in experimental studies, cannot be applied to clinical practice because of its high price, inconvenience, and susceptibility to many other factors in vivo [86]. Therefore, the current experimental methods for the determination of IDO activity were summarized in this review. We expect that, based on these activity assays, a method for the direct determination of IDO activity will be available for clinical use in the near future.

Author Contributions

All authors conceptualized and designed the work. P.Y. carried out the literature review, data collection, and the drafting of the article. J.Z. performed data analysis, critically revised the manuscript, and gave final approval. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Beijing Hospital Project (Grant BJ-2020-134) and the National Natural Science Foundation of China (Grant 22274108).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing is not applicable to this article as no datasets were generated or analyzed in the current study.

Acknowledgments

We thank Fei Xiao for his valuable suggestions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Barik, S. The Uniqueness of Tryptophan in Biology: Properties, Metabolism, Interactions and Localization in Proteins. Int. J. Mol. Sci. 2020, 21, 8776. [Google Scholar] [CrossRef]
  2. Badawy, A.A. Kynurenine Pathway of Tryptophan Metabolism: Regulatory and Functional Aspects. Int. J. Tryptophan Res. IJTR 2017, 10, 1178646917691938. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Badawy, A.A. Kynurenine pathway and human systems. Exp. Gerontol. 2020, 129, 110770. [Google Scholar] [CrossRef]
  4. González Esquivel, D.; Ramírez-Ortega, D.; Pineda, B.; Castro, N.; Ríos, C.; Pérez de la Cruz, V. Kynurenine pathway metabolites and enzymes involved in redox reactions. Neuropharmacology 2017, 112, 331–345. [Google Scholar] [CrossRef] [PubMed]
  5. Pires, A.S.; Gupta, S.; Barton, S.A.; Vander Wall, R.; Tan, V.; Heng, B.; Phillips, J.K.; Guillemin, G.J. Temporal Profile of Kynurenine Pathway Metabolites in a Rodent Model of Autosomal Recessive Polycystic Kidney Disease. Int. J. Tryptophan Res. IJTR 2022, 15, 11786469221126063. [Google Scholar] [CrossRef] [PubMed]
  6. Song, P.; Ramprasath, T.; Wang, H.; Zou, M.H. Abnormal kynurenine pathway of tryptophan catabolism in cardiovascular diseases. Cell. Mol. Life Sci. CMLS 2017, 74, 2899–2916. [Google Scholar] [CrossRef] [PubMed]
  7. Platten, M.; Nollen, E.A.A.; Röhrig, U.F.; Fallarino, F.; Opitz, C.A. Tryptophan metabolism as a common therapeutic target in cancer, neurodegeneration and beyond. Nat. Rev. Drug Discov. 2019, 18, 379–401. [Google Scholar] [CrossRef]
  8. Too, L.K.; Li, K.M.; Suarna, C.; Maghzal, G.J.; Stocker, R.; McGregor, I.S.; Hunt, N.H. Deletion of TDO2, IDO-1 and IDO-2 differentially affects mouse behavior and cognitive function. Behav. Brain Res. 2016, 312, 102–117. [Google Scholar] [CrossRef]
  9. Peyraud, F.; Guegan, J.P.; Bodet, D.; Cousin, S.; Bessede, A.; Italiano, A. Targeting Tryptophan Catabolism in Cancer Immunotherapy Era: Challenges and Perspectives. Front. Immunol. 2022, 13, 807271. [Google Scholar] [CrossRef]
  10. Dostal, C.R.; Carson Sulzer, M.; Kelley, K.W.; Freund, G.G.; McCusker, R.H. Glial and tissue-specific regulation of Kynurenine Pathway dioxygenases by acute stress of mice. Neurobiol. Stress 2017, 7, 1–15. [Google Scholar] [CrossRef]
  11. Bakmiwewa, S.M.; Fatokun, A.A.; Tran, A.; Payne, R.J.; Hunt, N.H.; Ball, H.J. Identification of selective inhibitors of indoleamine 2,3-dioxygenase 2. Bioorganic Med. Chem. Lett. 2012, 22, 7641–7646. [Google Scholar] [CrossRef]
  12. Théate, I.; van Baren, N.; Pilotte, L.; Moulin, P.; Larrieu, P.; Renauld, J.C.; Hervé, C.; Gutierrez-Roelens, I.; Marbaix, E.; Sempoux, C.; et al. Extensive profiling of the expression of the indoleamine 2,3-dioxygenase 1 protein in normal and tumoral human tissues. Cancer Immunol. Res. 2015, 3, 161–172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Carvajal-Hausdorf, D.E.; Mani, N.; Velcheti, V.; Schalper, K.A.; Rimm, D.L. Objective measurement and clinical significance of IDO1 protein in hormone receptor-positive breast cancer. J. Immunother. Cancer 2017, 5, 81. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Merlo, L.M.; Mandik-Nayak, L. IDO2: A Pathogenic Mediator of Inflammatory Autoimmunity. Clin. Med. Insights. Pathol. 2016, 9, 21–28. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Kozlova, A.; Frédérick, R. Current state on tryptophan 2,3-dioxygenase inhibitors: A patent review. Expert. Opin. Ther. Pat. 2019, 29, 11–23. [Google Scholar] [CrossRef]
  16. Meininger, D.; Zalameda, L.; Liu, Y.; Stepan, L.P.; Borges, L.; McCarter, J.D.; Sutherland, C.L. Purification and kinetic characterization of human indoleamine 2,3-dioxygenases 1 and 2 (IDO1 and IDO2) and discovery of selective IDO1 inhibitors. Biochim. Et. Biophys. Acta 2011, 1814, 1947–1954. [Google Scholar] [CrossRef]
  17. Takkenkamp, T.J.; Jalving, M.; Hoogwater, F.J.H.; Walenkamp, A.M.E. The immune tumour microenvironment of neuroendocrine tumours and its implications for immune checkpoint inhibitors. Endocr. Relat. Cancer 2020, 27, R329–R343. [Google Scholar] [CrossRef]
  18. Botticelli, A.; Cerbelli, B.; Lionetto, L.; Zizzari, I.; Salati, M.; Pisano, A.; Federica, M.; Simmaco, M.; Nuti, M.; Marchetti, P. Can IDO activity predict primary resistance to anti-PD-1 treatment in NSCLC? J. Transl. Med. 2018, 16, 219. [Google Scholar] [CrossRef] [Green Version]
  19. Botticelli, A.; Mezi, S.; Pomati, G.; Cerbelli, B.; Cerbelli, E.; Roberto, M.; Giusti, R.; Cortellini, A.; Lionetto, L.; Scagnoli, S.; et al. Tryptophan Catabolism as Immune Mechanism of Primary Resistance to Anti-PD-1. Front. Immunol. 2020, 11, 1243. [Google Scholar] [CrossRef]
  20. Wu, C.; Spector, S.A.; Theodoropoulos, G.; Nguyen, D.J.M.; Kim, E.Y.; Garcia, A.; Savaraj, N.; Lim, D.C.; Paul, A.; Feun, L.G.; et al. Dual inhibition of IDO1/TDO2 enhances anti-tumor immunity in platinum-resistant non-small cell lung cancer. Cancer Metab. 2023, 11, 7. [Google Scholar] [CrossRef]
  21. Mandarano, M.; Orecchini, E.; Bellezza, G.; Vannucci, J.; Ludovini, V.; Baglivo, S.; Tofanetti, F.R.; Chiari, R.; Loreti, E.; Puma, F.; et al. Kynurenine/Tryptophan Ratio as a Potential Blood-Based Biomarker in Non-Small Cell Lung Cancer. Int. J. Mol. Sci. 2021, 22, 4403. [Google Scholar] [CrossRef] [PubMed]
  22. Takikawa, O. Biochemical and medical aspects of the indoleamine 2,3-dioxygenase-initiated L-tryptophan metabolism. Biochem. Biophys. Res. Commun. 2005, 338, 12–19. [Google Scholar] [CrossRef] [PubMed]
  23. Lancellotti, S.; Novarese, L.; De Cristofaro, R. Biochemical properties of indoleamine 2,3-dioxygenase: From structure to optimized design of inhibitors. Curr. Med. Chem. 2011, 18, 2205–2214. [Google Scholar] [CrossRef]
  24. Najfeld, V.; Menninger, J.; Muhleman, D.; Comings, D.E.; Gupta, S.L. Localization of indoleamine 2,3-dioxygenase gene (INDO) to chromosome 8p12-->p11 by fluorescent in situ hybridization. Cytogenet. Cell Genet. 1993, 64, 231–232. [Google Scholar] [CrossRef] [PubMed]
  25. Puccetti, P.; Grohmann, U. IDO and regulatory T cells: A role for reverse signalling and non-canonical NF-kappaB activation. Nat. Rev. Immunol. 2007, 7, 817–823. [Google Scholar] [CrossRef] [PubMed]
  26. Azimnasab-Sorkhabi, P.; Soltani-Asl, M.; Yoshinaga, T.T.; Zaidan Dagli, M.L.; Massoco, C.O.; Kfoury Junior, J.R. Indoleamine-2,3 dioxygenase: A fate-changer of the tumor microenvironment. Mol. Biol. Rep. 2023, 50, 6133–6145. [Google Scholar] [CrossRef]
  27. Munn, D.H.; Mellor, A.L. IDO in the Tumor Microenvironment: Inflammation, Counter-Regulation, and Tolerance. Trends Immunol. 2016, 37, 193–207. [Google Scholar] [CrossRef] [Green Version]
  28. Wang, Q.; Liu, D.; Song, P.; Zou, M.H. Tryptophan-kynurenine pathway is dysregulated in inflammation, and immune activation. Front. Biosci. (Landmark Ed.) 2015, 20, 1116–1143. [Google Scholar] [CrossRef] [Green Version]
  29. Heitger, A. Regulation of expression and function of IDO in human dendritic cells. Curr. Med. Chem. 2011, 18, 2222–2233. [Google Scholar] [CrossRef]
  30. Stone, T.W.; Williams, R.O. Modulation of T cells by tryptophan metabolites in the kynurenine pathway. Trends Pharmacol. Sci. 2023, 44, 442–456. [Google Scholar] [CrossRef]
  31. Wu, H.; Gong, J.; Liu, Y. Indoleamine 2, 3-dioxygenase regulation of immune response (Review). Mol. Med. Rep. 2018, 17, 4867–4873. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Salminen, A. Role of indoleamine 2,3-dioxygenase 1 (IDO1) and kynurenine pathway in the regulation of the aging process. Ageing Res. Rev. 2022, 75, 101573. [Google Scholar] [CrossRef]
  33. Zhai, L.; Bell, A.; Ladomersky, E.; Lauing, K.L.; Bollu, L.; Sosman, J.A.; Zhang, B.; Wu, J.D.; Miller, S.D.; Meeks, J.J.; et al. Immunosuppressive IDO in Cancer: Mechanisms of Action, Animal Models, and Targeting Strategies. Front. Immunol. 2020, 11, 1185. [Google Scholar] [CrossRef] [PubMed]
  34. Löb, S.; Königsrainer, A. Role of IDO in organ transplantation: Promises and difficulties. Int. Rev. Immunol. 2009, 28, 185–206. [Google Scholar] [CrossRef] [PubMed]
  35. Hainz, U.; Jürgens, B.; Heitger, A. The role of indoleamine 2,3-dioxygenase in transplantation. Transpl. Int. Off. J. Eur. Soc. Organ. Transplant. 2007, 20, 118–127. [Google Scholar] [CrossRef]
  36. Barth, H.; Raghuraman, S. Persistent infectious diseases say—IDO. Role of indoleamine-2,3-dioxygenase in disease pathogenesis and implications for therapy. Crit. Rev. Microbiol. 2014, 40, 360–368. [Google Scholar] [CrossRef]
  37. Suzuki, T.; Kawamichi, H.; Imai, K. A myoglobin evolved from indoleamine 2,3-dioxygenase, a tryptophan-degrading enzyme. Comp. Biochem. Physiol. Part B Biochem. Mol. Biol. 1998, 121, 117–128. [Google Scholar] [CrossRef]
  38. Suzuki, T.; Imai, K. Evolution of myoglobin. Cell. Mol. Life Sci. CMLS 1998, 54, 979–1004. [Google Scholar] [CrossRef]
  39. Suzuki, T.; Yokouchi, K.; Kawamichi, H.; Yamamoto, Y.; Uda, K.; Yuasa, H.J. Comparison of the sequences of Turbo and Sulculus indoleamine dioxygenase-like myoglobin genes. Gene 2003, 308, 89–94. [Google Scholar] [CrossRef]
  40. Hirata, F.; Ohnishi, T.; Hayaishi, O. Indoleamine 2,3-dioxygenase. Characterization and properties of enzyme. O2- complex. J. Biol. Chem. 1977, 252, 4637–4642. [Google Scholar] [CrossRef]
  41. Sono, M. Spectroscopic and equilibrium properties of the indoleamine 2,3-dioxygenase-tryptophan-O2 ternary complex and of analogous enzyme derivatives. Tryptophan binding to ferrous enzyme adducts with dioxygen, nitric oxide, and carbon monoxide. Biochemistry 1986, 25, 6089–6097. [Google Scholar] [CrossRef]
  42. Rosell, F.I.; Kuo, H.H.; Mauk, A.G. NADH oxidase activity of indoleamine 2,3-dioxygenase. J. Biol. Chem. 2011, 286, 29273–29283. [Google Scholar] [CrossRef] [Green Version]
  43. Grant, R.S. Indoleamine 2,3-Dioxygenase Activity Increases NAD+ Production in IFN-γ-Stimulated Human Primary Mononuclear Cells. Int. J. Tryptophan Res. IJTR 2018, 11, 1178646917751636. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Mehraj, V.; Routy, J.P. Tryptophan Catabolism in Chronic Viral Infections: Handling Uninvited Guests. Int. J. Tryptophan Res. IJTR 2015, 8, 41–48. [Google Scholar] [CrossRef] [PubMed]
  45. Mbongue, J.C.; Nicholas, D.A.; Torrez, T.W.; Kim, N.S.; Firek, A.F.; Langridge, W.H. The Role of Indoleamine 2, 3-Dioxygenase in Immune Suppression and Autoimmunity. Vaccines 2015, 3, 703–729. [Google Scholar] [CrossRef] [Green Version]
  46. Myint, A.M.; Kim, Y.K. Network beyond IDO in psychiatric disorders: Revisiting neurodegeneration hypothesis. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2014, 48, 304–313. [Google Scholar] [CrossRef] [PubMed]
  47. Zhang, L.; Jiang, T.; Yang, Y.; Deng, W.; Lu, H.; Wang, S.; Liu, R.; Chang, M.; Wu, S.; Gao, Y.; et al. Postpartum hepatitis and host immunity in pregnant women with chronic HBV infection. Front. Immunol. 2022, 13, 1112234. [Google Scholar] [CrossRef] [PubMed]
  48. Lashgari, N.A.; Roudsari, N.M.; Shayan, M.; Niazi Shahraki, F.; Hosseini, Y.; Momtaz, S.; Abdolghaffari, A.H. IDO/Kynurenine; novel insight for treatment of inflammatory diseases. Cytokine 2023, 166, 156206. [Google Scholar] [CrossRef]
  49. Pan, B.; Zhang, F.; Sun, J.; Chen, D.; Huang, W.; Zhang, H.; Cao, C.; Wan, X. Correlation of Indoleamine-2,3-Dioxygenase and Chronic Kidney Disease: A Pilot Study. J. Immunol. Res. 2021, 2021, 8132569. [Google Scholar] [CrossRef]
  50. Torosyan, R.; Huang, S.; Bommi, P.V.; Tiwari, R.; An, S.Y.; Schonfeld, M.; Rajendran, G.; Kavanaugh, M.A.; Gibbs, B.; Truax, A.D.; et al. Hypoxic preconditioning protects against ischemic kidney injury through the IDO1/kynurenine pathway. Cell Rep. 2021, 36, 109547. [Google Scholar] [CrossRef] [PubMed]
  51. Chaudhary, K.; Shinde, R.; Liu, H.; Gnana-Prakasam, J.P.; Veeranan-Karmegam, R.; Huang, L.; Ravishankar, B.; Bradley, J.; Kvirkvelia, N.; McMenamin, M.; et al. Amino acid metabolism inhibits antibody-driven kidney injury by inducing autophagy. J. Immunol. 2015, 194, 5713–5724. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Yu, K.; Li, D.; Xu, F.; Guo, H.; Feng, F.; Ding, Y.; Wan, X.; Sun, N.; Zhang, Y.; Fan, J.; et al. IDO1 as a new immune biomarker for diabetic nephropathy and its correlation with immune cell infiltration. Int. Immunopharmacol. 2021, 94, 107446. [Google Scholar] [CrossRef]
  53. Zhou, Q.; Shi, Y.; Chen, C.; Wu, F.; Chen, Z. A narrative review of the roles of indoleamine 2,3-dioxygenase and tryptophan-2,3-dioxygenase in liver diseases. Ann. Transl. Med. 2021, 9, 174. [Google Scholar] [CrossRef]
  54. Xu, L.; Ling, J.; Su, C.; Su, Y.W.; Xu, Y.; Jiang, Z. Emerging Roles on Immunological Effect of Indoleamine 2,3-Dioxygenase in Liver Injuries. Front. Med. 2021, 8, 756435. [Google Scholar] [CrossRef]
  55. Fujiwara, N.; Kubota, N.; Crouchet, E.; Koneru, B.; Marquez, C.A.; Jajoriya, A.K.; Panda, G.; Qian, T.; Zhu, S.; Goossens, N.; et al. Molecular signatures of long-term hepatocellular carcinoma risk in nonalcoholic fatty liver disease. Sci. Transl. Med. 2022, 14, eabo4474. [Google Scholar] [CrossRef] [PubMed]
  56. Yoshio, S.; Sugiyama, M.; Shoji, H.; Mano, Y.; Mita, E.; Okamoto, T.; Matsuura, Y.; Okuno, A.; Takikawa, O.; Mizokami, M.; et al. Indoleamine-2,3-dioxygenase as an effector and an indicator of protective immune responses in patients with acute hepatitis B. Hepatol. 2016, 63, 83–94. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Lepiller, Q.; Soulier, E.; Li, Q.; Lambotin, M.; Barths, J.; Fuchs, D.; Stoll-Keller, F.; Liang, T.J.; Barth, H. Antiviral and Immunoregulatory Effects of Indoleamine-2,3-Dioxygenase in Hepatitis C Virus Infection. J. Innate Immun. 2015, 7, 530–544. [Google Scholar] [CrossRef]
  58. Chunna, X.; Xiao, L.; Yu, T.; Lei, Z.; Xiufang, W. IDO expression in lung tissue and its relationship with IFN-γ in asthmatic mice. Chin. Pediatr. Integr. Tradit. West. Med. 2021, 13, 103–107. (In Chinese) [Google Scholar]
  59. Makala, L.H.; Baban, B.; Lemos, H.; El-Awady, A.R.; Chandler, P.R.; Hou, D.Y.; Munn, D.H.; Mellor, A.L. Leishmania major attenuates host immunity by stimulating local indoleamine 2,3-dioxygenase expression. J. Infect. Dis. 2011, 203, 715–725. [Google Scholar] [CrossRef] [Green Version]
  60. Wang, L.J.; Ma, P.Y.; Liu, H.; Cao, J.P.; Li, H.M.; Zheng, M.H. Progress of researches on the involvement of indoleamine 2, 3-dioxygenase in regulation of parasite-host immune interactions. Chin. J. Schistosomiasis Control 2020, 33, 209–212. [Google Scholar] [CrossRef]
  61. Dai, W.; Pan, H.; Kwok, O.; Dubey, J.P. Human indoleamine 2,3-dioxygenase inhibits Toxoplasma gondii growth in fibroblast cells. J. Interferon Res. 1994, 14, 313–317. [Google Scholar] [CrossRef] [PubMed]
  62. Raitala, A.; Karjalainen, J.; Oja, S.S.; Kosunen, T.U.; Hurme, M. Helicobacter pylori-induced indoleamine 2,3-dioxygenase activity in vivo is regulated by TGFB1 and CTLA4 polymorphisms. Mol. Immunol. 2007, 44, 1011–1014. [Google Scholar] [CrossRef] [PubMed]
  63. Dolšak, A.; Gobec, S.; Sova, M. Indoleamine and tryptophan 2,3-dioxygenases as important future therapeutic targets. Pharmacol. Ther. 2021, 221, 107746. [Google Scholar] [CrossRef]
  64. Le Naour, J.; Galluzzi, L.; Zitvogel, L.; Kroemer, G.; Vacchelli, E. Trial watch: IDO inhibitors in cancer therapy. Oncoimmunology 2020, 9, 1777625. [Google Scholar] [CrossRef]
  65. Brincks, E.L.; Adams, J.; Wang, L.; Turner, B.; Marcinowicz, A.; Ke, J.; Essmann, M.; Mautino, L.M.; Allen, C.V.; Kumar, S.; et al. Indoximod opposes the immunosuppressive effects mediated by IDO and TDO via modulation of AhR function and activation of mTORC1. Oncotarget 2020, 11, 2438–2461. [Google Scholar] [CrossRef] [PubMed]
  66. Metz, R.; Rust, S.; Duhadaway, J.B.; Mautino, M.R.; Munn, D.H.; Vahanian, N.N.; Link, C.J.; Prendergast, G.C. IDO inhibits a tryptophan sufficiency signal that stimulates mTOR: A novel IDO effector pathway targeted by D-1-methyl-tryptophan. Oncoimmunology 2012, 1, 1460–1468. [Google Scholar] [CrossRef] [Green Version]
  67. Soliman, H.H.; Minton, S.E.; Han, H.S.; Ismail-Khan, R.; Neuger, A.; Khambati, F.; Noyes, D.; Lush, R.; Chiappori, A.A.; Roberts, J.D.; et al. A phase I study of indoximod in patients with advanced malignancies. Oncotarget 2016, 7, 22928–22938. [Google Scholar] [CrossRef] [Green Version]
  68. Fox, E.; Oliver, T.; Rowe, M.; Thomas, S.; Zakharia, Y.; Gilman, P.B.; Muller, A.J.; Prendergast, G.C. Indoximod: An Immunometabolic Adjuvant That Empowers T Cell Activity in Cancer. Front. Oncol. 2018, 8, 370. [Google Scholar] [CrossRef] [Green Version]
  69. Labadie, B.W.; Bao, R.; Luke, J.J. Reimagining IDO Pathway Inhibition in Cancer Immunotherapy via Downstream Focus on the Tryptophan-Kynurenine-Aryl Hydrocarbon Axis. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2019, 25, 1462–1471. [Google Scholar] [CrossRef] [Green Version]
  70. Yue, E.W.; Sparks, R.; Polam, P.; Modi, D.; Douty, B.; Wayland, B.; Glass, B.; Takvorian, A.; Glenn, J.; Zhu, W.; et al. INCB24360 (Epacadostat), a Highly Potent and Selective Indoleamine-2,3-dioxygenase 1 (IDO1) Inhibitor for Immuno-oncology. ACS Med. Chem. Lett. 2017, 8, 486–491. [Google Scholar] [CrossRef] [Green Version]
  71. Odunsi, K.; Qian, F.; Lugade, A.A.; Yu, H.; Geller, M.A.; Fling, S.P.; Kaiser, J.C.; Lacroix, A.M.; D’Amico, L.; Ramchurren, N.; et al. Metabolic adaptation of ovarian tumors in patients treated with an IDO1 inhibitor constrains antitumor immune responses. Sci. Transl. Med. 2022, 14, eabg8402. [Google Scholar] [CrossRef] [PubMed]
  72. Long, G.V.; Dummer, R.; Hamid, O.; Gajewski, T.F.; Caglevic, C.; Dalle, S.; Arance, A.; Carlino, M.S.; Grob, J.J.; Kim, T.M.; et al. Epacadostat plus pembrolizumab versus placebo plus pembrolizumab in patients with unresectable or metastatic melanoma (ECHO-301/KEYNOTE-252): A phase 3, randomised, double-blind study. Lancet. Oncol. 2019, 20, 1083–1097. [Google Scholar] [CrossRef] [PubMed]
  73. Cherney, E.C.; Zhang, L.; Nara, S.; Zhu, X.; Gullo-Brown, J.; Maley, D.; Lin, T.A.; Hunt, J.T.; Huang, C.; Yang, Z.; et al. Discovery and Preclinical Evaluation of BMS-986242, a Potent, Selective Inhibitor of Indoleamine-2,3-dioxygenase 1. ACS Med. Chem. Lett. 2021, 12, 288–294. [Google Scholar] [CrossRef] [PubMed]
  74. Balog, A.; Lin, T.A.; Maley, D.; Gullo-Brown, J.; Kandoussi, E.H.; Zeng, J.; Hunt, J.T. Preclinical Characterization of Linrodostat Mesylate, a Novel, Potent, and Selective Oral Indoleamine 2,3-Dioxygenase 1 Inhibitor. Mol. Cancer Ther. 2021, 20, 467–476. [Google Scholar] [CrossRef] [PubMed]
  75. Sonpavde, G.; Necchi, A.; Gupta, S.; Steinberg, G.D.; Gschwend, J.E.; Van Der Heijden, M.S.; Garzon, N.; Ibrahim, M.; Raybold, B.; Liaw, D.; et al. ENERGIZE: A Phase III study of neoadjuvant chemotherapy alone or with nivolumab with/without linrodostat mesylate for muscle-invasive bladder cancer. Future Oncol. 2020, 16, 4359–4368. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Ma, S.; Suchomel, J.; Yanez, E.; Yost, E.; Liang, X.; Zhu, R.; Le, H.; Siebers, N.; Joas, L.; Morley, R.; et al. Investigation of the absolute bioavailability and human mass balance of navoximod, a novel IDO1 inhibitor. Br. J. Clin. Pharmacol. 2019, 85, 1751–1760. [Google Scholar] [CrossRef]
  77. Cheong, J.E.; Ekkati, A.; Sun, L. A patent review of IDO1 inhibitors for cancer. Expert. Opin. Ther. Pat. 2018, 28, 317–330. [Google Scholar] [CrossRef]
  78. Kumar, S.; Waldo, J.P.; Jaipuri, F.A.; Marcinowicz, A.; Van Allen, C.; Adams, J.; Kesharwani, T.; Zhang, X.; Metz, R.; Oh, A.J.; et al. Discovery of Clinical Candidate (1R,4r)-4-((R)-2-((S)-6-Fluoro-5H-imidazo[5,1-a]isoindol-5-yl)-1-hydroxyethyl)cyclohexan-1-ol (Navoximod), a Potent and Selective Inhibitor of Indoleamine 2,3-Dioxygenase 1. J. Med. Chem. 2019, 62, 6705–6733. [Google Scholar] [CrossRef]
  79. Gomes, B.; Driessens, G.; Bartlett, D.; Cai, D.; Cauwenberghs, S.; Crosignani, S.; Dalvie, D.; Denies, S.; Dillon, C.P.; Fantin, V.R.; et al. Characterization of the Selective Indoleamine 2,3-Dioxygenase-1 (IDO1) Catalytic Inhibitor EOS200271/PF-06840003 Supports IDO1 as a Critical Resistance Mechanism to PD-(L)1 Blockade Therapy. Mol. Cancer Ther. 2018, 17, 2530–2542. [Google Scholar] [CrossRef] [Green Version]
  80. Reardon, D.A.; Desjardins, A.; Rixe, O.; Cloughesy, T.; Alekar, S.; Williams, J.H.; Li, R.; Taylor, C.T.; Lassman, A.B. A phase 1 study of PF-06840003, an oral indoleamine 2,3-dioxygenase 1 (IDO1) inhibitor in patients with recurrent malignant glioma. Investig. New Drugs 2020, 38, 1784–1795. [Google Scholar] [CrossRef]
  81. Muller, A.J.; Mondal, A.; Dey, S.; Prendergast, G.C. IDO1 and inflammatory neovascularization: Bringing new blood to tumor-promoting inflammation. Front. Oncol. 2023, 13, 1165298. [Google Scholar] [CrossRef] [PubMed]
  82. Fellendorf, F.T.; Bonkat, N.; Dalkner, N.; Schönthaler, E.M.D.; Manchia, M.; Fuchs, D.; Reininghaus, E.Z. Indoleamine 2,3-dioxygenase (IDO)-activity in Severe Psychiatric Disorders: A Systemic Review. Curr. Top. Med. Chem. 2022, 22, 2107–2118. [Google Scholar] [CrossRef] [PubMed]
  83. Shimizu, T.; Nomiyama, S.; Hirata, F.; Hayaishi, O. Indoleamine 2,3-dioxygenase. Purification and some properties. J. Biol. Chem. 1978, 253, 4700–4706. [Google Scholar] [CrossRef]
  84. Sono, M.; Taniguchi, T.; Watanabe, Y.; Hayaishi, O. Indoleamine 2,3-dioxygenase. Equilibrium studies of the tryptophan binding to the ferric, ferrous, and CO-bound enzymes. J. Biol. Chem. 1980, 255, 1339–1345. [Google Scholar] [CrossRef]
  85. Hara, T.; Yamakura, F.; Takikawa, O.; Hiramatsu, R.; Kawabe, T.; Isobe, K.; Nagase, F. Diazotization of kynurenine by acidified nitrite secreted from indoleamine 2,3-dioxygenase-expressing myeloid dendritic cells. J. Immunol. Methods 2008, 332, 162–169. [Google Scholar] [CrossRef] [PubMed]
  86. Badawy, A.A.; Guillemin, G. The Plasma [Kynurenine]/[Tryptophan] Ratio and Indoleamine 2,3-Dioxygenase: Time for Appraisal. Int. J. Tryptophan Res. IJTR 2019, 12, 1178646919868978. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Cerejo, M.; Andrade, G.; Roca, C.; Sousa, J.; Rodrigues, C.; Pinheiro, R.; Chatterjee, S.; Vieira, H.; Calado, P. A powerful yeast-based screening assay for the identification of inhibitors of indoleamine 2,3-dioxygenase. J. Biomol. Screen. 2012, 17, 1362–1371. [Google Scholar] [CrossRef] [Green Version]
  88. Lanser, L.; Kink, P.; Egger, E.M.; Willenbacher, W.; Fuchs, D.; Weiss, G.; Kurz, K. Inflammation-Induced Tryptophan Breakdown is Related With Anemia, Fatigue, and Depression in Cancer. Front. Immunol. 2020, 11, 249. [Google Scholar] [CrossRef] [Green Version]
  89. Andersen, M.H. Tumor microenvironment antigens. Semin. Immunopathol. 2023, 45, 253–264. [Google Scholar] [CrossRef]
  90. Bello, C.; Heinisch, P.P.; Mihalj, M.; Carrel, T.; Luedi, M.M. Indoleamine-2,3-Dioxygenase as a Perioperative Marker of the Immune System. Front. Physiol. 2021, 12, 766511. [Google Scholar] [CrossRef]
  91. Passarelli, A.; Pisano, C.; Cecere, S.C.; Di Napoli, M.; Rossetti, S.; Tambaro, R.; Ventriglia, J.; Gherardi, F.; Iannacone, E.; Venanzio, S.S.; et al. Targeting immunometabolism mediated by the IDO1 Pathway: A new mechanism of immune resistance in endometrial cancer. Front. Immunol. 2022, 13, 953115. [Google Scholar] [CrossRef] [PubMed]
  92. Andersen, M.H. The T-win® technology: Immune-modulating vaccines. Semin. Immunopathol. 2019, 41, 87–95. [Google Scholar] [CrossRef] [PubMed]
  93. Ala, M. The footprint of kynurenine pathway in every cancer: A new target for chemotherapy. Eur. J. Pharmacol. 2021, 896, 173921. [Google Scholar] [CrossRef] [PubMed]
  94. Ascierto, P.A.; Flaherty, K.; Goff, S. Emerging Strategies in Systemic Therapy for the Treatment of Melanoma. Am. Soc. Clin. Oncol. Educ. Book. Am. Soc. Clin. Oncol. Annu. Meet. 2018, 38, 751–758. [Google Scholar] [CrossRef]
Figure 1. Metabolism of tryptophan and the action of IDO (indoleamine 2,3-dioxygenase) inhibitors in the tryptophan metabolic pathway. In vivo, ~95% of L-tryptophan is catabolized via IDO in the form of kynurenine degradation products. Additional enzymatic metabolites in the pathway promote immune suppression. IDO inhibitors primarily compete with tryptophan by binding to the catalytic domain of IDO or non-competitive mechanisms to affect tryptophan metabolism, thereby exerting an inhibitory effect on IDO. IDO1, indoleamine 2,3-dioxygenase 1; IDO2, indoleamine 2,3-dioxygenase 2; TDO, tryptophan dioxygenase; FAMID, formamidase; KATs, kynurenine aminotransferases; KMO, kynurenine 3-monooxygenase; KYNU, kynureninase; 3HAAO, 3-hydroxyanthranilic acid dioxygenase; QPRT, quinolinic-acid phosphoribosyl transferase; NAD+, nicotinamide adenine dinucleotide; GCN2, general control nondepressible-2; mTORC1, mammalian target of rapamycin complex 1; AhR, aryl hydrocarbon receptor; Treg cells, regulatory T cells.
Figure 1. Metabolism of tryptophan and the action of IDO (indoleamine 2,3-dioxygenase) inhibitors in the tryptophan metabolic pathway. In vivo, ~95% of L-tryptophan is catabolized via IDO in the form of kynurenine degradation products. Additional enzymatic metabolites in the pathway promote immune suppression. IDO inhibitors primarily compete with tryptophan by binding to the catalytic domain of IDO or non-competitive mechanisms to affect tryptophan metabolism, thereby exerting an inhibitory effect on IDO. IDO1, indoleamine 2,3-dioxygenase 1; IDO2, indoleamine 2,3-dioxygenase 2; TDO, tryptophan dioxygenase; FAMID, formamidase; KATs, kynurenine aminotransferases; KMO, kynurenine 3-monooxygenase; KYNU, kynureninase; 3HAAO, 3-hydroxyanthranilic acid dioxygenase; QPRT, quinolinic-acid phosphoribosyl transferase; NAD+, nicotinamide adenine dinucleotide; GCN2, general control nondepressible-2; mTORC1, mammalian target of rapamycin complex 1; AhR, aryl hydrocarbon receptor; Treg cells, regulatory T cells.
Biomedicines 11 01988 g001
Figure 2. Regulatory effect of IDO on T cells and Tregs in immune microenvironment. IDO depletes tryptophan and exerts important immunosuppressive functions via the regulation of T effector cells and the induction of T regulatory cell proliferation. This immune response eventually results in clinical conditions, such as cancer, infection, and transplantation. Tregs, regulatory T cells; GCN2, general control nondepressible-2; mTOR, mammalian target of rapamycin; AhR, aryl hydrocarbon receptor; TGF-β, tumor growth factor β. Red arrow, an increase; green arrow, a decrease.
Figure 2. Regulatory effect of IDO on T cells and Tregs in immune microenvironment. IDO depletes tryptophan and exerts important immunosuppressive functions via the regulation of T effector cells and the induction of T regulatory cell proliferation. This immune response eventually results in clinical conditions, such as cancer, infection, and transplantation. Tregs, regulatory T cells; GCN2, general control nondepressible-2; mTOR, mammalian target of rapamycin; AhR, aryl hydrocarbon receptor; TGF-β, tumor growth factor β. Red arrow, an increase; green arrow, a decrease.
Biomedicines 11 01988 g002
Figure 3. Strategy for the determination of indoleamine 2,3-dioxygenase (IDO) activity. IDOFe3+ can be converted to IDOFe2+-O2 in addition to β-NADH under aerobic conditions, resulting in increasing absorbance at 577 and 544 nm of the α and β transitions. The absorbance at 340 nm depicts the consumption of NADH. The incubation of NADH-generated IDOFe2+-O2 with L-Trp resulted in the production of NFK, as indicated by the increasing absorbance at 321 nm. NFK is rapidly converted to Kyn catalyzed by FAMID along with the increasing absorbance at 492 nm. NFK, N-Formyl kynurenine; FAMID, formamidase; Kyn, kynurenine.
Figure 3. Strategy for the determination of indoleamine 2,3-dioxygenase (IDO) activity. IDOFe3+ can be converted to IDOFe2+-O2 in addition to β-NADH under aerobic conditions, resulting in increasing absorbance at 577 and 544 nm of the α and β transitions. The absorbance at 340 nm depicts the consumption of NADH. The incubation of NADH-generated IDOFe2+-O2 with L-Trp resulted in the production of NFK, as indicated by the increasing absorbance at 321 nm. NFK is rapidly converted to Kyn catalyzed by FAMID along with the increasing absorbance at 492 nm. NFK, N-Formyl kynurenine; FAMID, formamidase; Kyn, kynurenine.
Biomedicines 11 01988 g003
Table 1. IDO inhibitors in clinical trials.
Table 1. IDO inhibitors in clinical trials.
InhibitorPhaseMedicationTumor TypeIDStatus
Indoximod (1-D-MT)ICombination with idarubicin and cytarabineAcute myeloid leukemiaNCT028357292019-12 completed
IICombination with docetaxel or paclitaxelMetastatic breast cancerNCT017920502017-07 completed
I/IICombination with ipilimumab, pembrolizumab and nivolumabMetastatic melanomaNCT020731232019-07 completed
IICombination with sipuleucel-TMetastatic prostate cancerNCT015609232018-12 completed
ICombination with ibrutinib, cyclophosphamide and etoposidePediatric brain cancerNCT05106296Recruiting
I/IICombination with temozolomide and bevacizumabNewly diagnosed glioblastomaNCT020526482019-06 completed
ISingle agentAdvanced malignancyNCT005679312012-09 completed
ICombination with temozolomidePediatric brain tumorNCT025027082020-02 completed
I/IICombination with p53 DC vaccineMetastatic breast cancerNCT010425352018-02 completed
ICombination with docetaxelMetastatic solid tumorNCT011912162013-08 completed
IICombination with partial radiationProgressive brain cancerNCT04049669Recruiting
Epacadostat (INCB024360)IICombination with pembrolizumabGastrointestinal stromal tumorNCT032910542020-08 completed
ICombination with sirolimusAdvanced malignancyNCT03217669Active, not recruiting
IICombination with pembrolizumabGastroesophageal junction or gastric cancerNCT031962322018-05 completed
IISingle agentMyelodysplastic syndromeNCT018226912015-02 completed
IICombination with pembrolizumabLocally advanced/metastatic sarcomaNCT03414229Active, not recruiting
IICombination with pembrolizumabovarian clear cell carcinomaNCT03602586Active, not recruiting
ISingle agentAdvanced malignancyNCT011953112013-07 completed
I/IICombination with durvalumabAdvanced solid tumorNCT023182772020-10 completed
IIICombination with pembrolizumabCisplatin-ineligible urothelial carcinomaNCT033618652020-08 completed
IICombination with pembrolizumabMetastatic non-small cell lung cancerNCT033225402020-11 completed
IIICombination with pembrolizumabMetastatic urothelial carcinomaNCT033744882020-07 completed
I/IICombination with MK-3475High colorectal cancer, Endometrial cancer, Head and neck cancer and Hepatocellular carcinomaNCT021787222020-11 completed
I/IICombination with intralesional SD101Advanced/refractory solid tumors and lymphomaNCT033223842020-04 completed
I/IICombination with nivolumabAdvanced solid tumors and lymphomas; squamous cell carcinoma of head and neck and non-small cell lung cancerNCT023270782020-06 completed
IICombination with MELITAC 12.1 peptide vaccineStage III-IV melanomaNCT019611152017-05 completed
IIICombination with pembrolizumabLocally advanced/metastatic renal cell carcinomaNCT03260894Active, not recruiting
IIICombination with pembrolizumabUnresectable or metastatic melanomaNCT027520742019-08 completed
IICombination with retifanlimab, INCAGN02385 and INCAGN02390Neoadjuvant urothelial carcinomaNCT04586244Recruiting
I/IISingle agentLocally advanced rectal cancerNCT03516708Recruiting
I/IICombination with DPX-Survivac and cyclophosphamideRecurrent ovarian cancerNCT02785250Active, not recruiting
ISingle agentEpithelial ovarian, fallopian tube, or primary peritoneal cancerNCT02042430Active, not recruiting
IICombination with pembrolizumabMetastatic pancreas cancerNCT03006302Active, not recruiting
IICombination with pembrolizumab and platinum-based chemotherapyMetastatic non-small cell lung cancerNCT033225662020-10 completed
IICombination with retifanlimab and pemigatinibAdvanced or metastatic endometrial cancerNCT04463771Recruiting
IIICombination with pembrolizumab and cetuximabRecurrent or metastatic head and neck squamous cell carcinomaNCT03358472Active, not recruiting
BMS-986205IICombination with nivolumabEndometrial cancerNCT04106414Active, not recruiting
I/IICombination with nivolumabLiver cancerNCT03695250Active, not recruiting
I/IICombination with nivolumabAdvanced malignant solid tumorNCT037927502020-12 completed
IIICombination with nivolumabAdvanced melanomaNCT033298462020-07 completed
IICombination with nivolumabSquamous cell cancer of the head and neckNCT03854032Active, not recruiting
ICombination with nivolumabNewly diagnosed glioblastomaNCT04047706Active, not recruiting
I/IICombination with nivolumab, relatlimab and ipilimumabAdvanced tumorNCT03459222Active, not recruiting
IICombination with nivolumab, relatlimab and ipilimumabAdvanced renal cell carcinomaNCT029961102021-11 completed
ICombination with nivolumabAdvanced malignant tumorNCT031929432018-12 completed
I/IICombination with nivolumab, and ipilimumabAdvanced malignant tumorNCT026588902021-10 completed
IICombination with nivolumab, relatlimab and ipilimumabAdvanced gastric cancerNCT029356342022-05 completed
IIICombination with nivolumab, gemcitabine and cisplatinMuscle invasive bladder cancerNCT03661320Active, not recruiting
ICombination with nivolumab, relatlimab and cabiralizumabSolid tumorNCT033355402021-08 completed
Navoximod (GDC-0919, NLG-919)ISingle agentAdvanced solid tumorNCT020487092016-02 completed
ICombination with atezolizumabLocally advanced or metastatic solid tumorsNCT024718462019-10 completed
PF-06840003ISingle agentMalignant gliomasNCT02764151Terminated
HTI-1090 (SHR9146)ISingle agentAdvanced solid tumorNCT032089592019-01 completed
ICombination with SHR-1210 and apatinibAdvanced solid tumorNCT03491631Active, not recruiting
Data is available at ClinicalTrials.gov, https://www.clinicaltrials.gov.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yang, P.; Zhang, J. Indoleamine 2,3-Dioxygenase (IDO) Activity: A Perspective Biomarker for Laboratory Determination in Tumor Immunotherapy. Biomedicines 2023, 11, 1988. https://doi.org/10.3390/biomedicines11071988

AMA Style

Yang P, Zhang J. Indoleamine 2,3-Dioxygenase (IDO) Activity: A Perspective Biomarker for Laboratory Determination in Tumor Immunotherapy. Biomedicines. 2023; 11(7):1988. https://doi.org/10.3390/biomedicines11071988

Chicago/Turabian Style

Yang, Pengbo, and Junhua Zhang. 2023. "Indoleamine 2,3-Dioxygenase (IDO) Activity: A Perspective Biomarker for Laboratory Determination in Tumor Immunotherapy" Biomedicines 11, no. 7: 1988. https://doi.org/10.3390/biomedicines11071988

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

Article Metrics

Back to TopTop