Abstract
The heme enzyme indoleamine 2,3-dioxygenase (IDO) was found to catalyze the oxidation of indole by H2O2, with generation of 2- and 3-oxoindole as the major products. This reaction occurred in the absence of O2 and reducing agents and was not inhibited by superoxide dismutase or hydroxyl radical scavengers, although it was strongly inhibited by l-Trp. The stoichiometry of the reaction indicated a one-to-one correspondence for the consumption of indole and H2O2. The 18O-labeling experiments indicated that the oxygen incorporated into the monooxygenated products was derived almost exclusively from H218O2, suggesting that electron transfer was coupled to the transfer of oxygen from a ferryl intermediate of IDO. These results demonstrate that IDO oxidizes indole by means of a previously unrecognized peroxygenase activity. We conclude that IDO inserts oxygen into indole in a reaction that is mechanistically analogous to the “peroxide shunt” pathway of cytochrome P450.
Keywords: monooxygenase, oxygen isotopic labeling
The heme enzyme indoleamine 2,3-dioxygenase (IDO) catalyzes the first and rate-determining step of l-tryptophan metabolism in nonhepatic mammalian tissues by inserting both atoms of O2 into the indole ring to form N-formylkynurenine (N-FK) (1). This dioxygenase activity of IDO requires O2 and the reduced (Fe2+) enzyme or O2•− and the oxidized (Fe3+) enzyme. In recent years, an increasing number of physiological consequences of IDO activity have been identified, including links to neural, ocular, and immunological disorders (2). Of these roles, the ability of IDO expressed by tumors to suppress the normal response of T lymphocytes and enable tumor survival and growth has attracted the most intense interest, owing to its implications for the development of new therapeutic approaches in cancer treatment (3).
Aside from tryptophan (Trp), the dioxygenase activity of IDO extends to oxidation of other indoleamines, such as 5-hydroxy-l-Trp, serotonin, and tryptamine, whereas indole, 3-methylindole, and indoleacetic acid are not substrates (4–6). Although indole can apparently bind to the oxygenated enzyme (IDOFe3+–O2•−) or to the IDOFe2+–CO complex (7), autoxidation to IDOFe3+ is unaffected by the presence of indole, and oxidation of indole does not occur (5). Notably, IDO does not catalyze oxidation of l-Trp by H2O2 (4, 8, 9). Consequently, the reactivity of IDO with H2O2 received little attention for more than 30 years. During that time, IDO was noted to have peroxidase activity (6) and to catalyze the H2O2-supported N-demethylation of benzphetamine and hydroxylation of aniline (10), but these activities were not studied further.
Recently, reactivity of the enzyme with H2O2 has received considerable attention in light of spectroscopic detection of a compound II-like ferryl species (11, 12) and quantum mechanics/molecular mechanics simulations implicating a role for this intermediate in the catalytic cycle (13). Subsequently, the peroxidase activity of IDO has received renewed attention (9).
Although the source of the oxygen incorporated into the products of IDO-catalyzed aniline hydroxylation remains unclear, the limited literature that is available implies that IDO may be capable of processing alternative substrates in the presence of hydrogen peroxide, presumably owing to the relatively accessible nature of the active site (14). In the present work, we have evaluated the ability of IDO to catalyze the oxidation of indole and methylated derivatives of indole by H2O2, identified the products formed in these reactions, and identified the source of the oxygen that is incorporated into the reaction products. These results provide evidence that IDO exhibits peroxygenase activity similar to the so-called peroxide shunt of cytochromes P450 (P450).
Results
IDO-Catalyzed Indole Oxidation by H2O2.
Addition of H2O2 to a solution containing IDOFe3+ and indole resulted in decreased absorbance at 276 nm and corresponding increases at ∼233 nm, 390 nm, and 610 nm (Fig. 1A) that are consistent with consumption of indole. The lack of isosbestic points in this family of spectra indicated that multiple products and/or intermediates were formed during reaction. Accordingly, at least seven products were identified by HPLC, and a MeOH/H2O-insoluble pigment was identified by TLC to be indigo blue (Fig. 1B and Tables S1 and S2). Only the peaks with retention times corresponding to Tris buffer and/or enzyme (7.5 min) and indole (35.5 min) were detected when either H2O2 or the enzyme were omitted or when the enzyme was thermally denatured before reaction.
Fig. 1.
IDO-catalyzed oxidation of indole by H2O2. (A) Electronic spectra following aerobic addition of H2O2 (1 mM) to IDOFe3+ (750 nM) and indole (220 μM). (B) HPLC analysis of reaction components (I–VII) after addition of indole (2 mM) to IDOFe3+ (200 μM) and H2O2 (4 mM). Data supporting identification of products are provided in SI Materials and Methods.
Stoichiometry of Indole Oxidation and Inactivation of IDO.
IDO-catalyzed oxidation of indole by H2O2 initiated by aerobic addition of equimolar amounts of both substrates was monitored fluorimetrically to determine the stoichiometry of the reaction (Fig. 2A). Under these conditions (substrates, 26–416 μM; [IDO], 10 μM), indole was consumed nearly quantitatively (∼95%), consistent with one-to-one consumption of the two cosubstrates. Some peroxide was presumably consumed in downstream formation of dioxygenated indoles. Addition of more H2O2 on completion of reaction led to oxidation of the remaining indole and confirmed that depletion of H2O2 accounts for premature termination of the reaction. At lower enzyme concentration (e.g., 1 μM), less indole was oxidized, and reaction could be reestablished only by addition of fresh enzyme; however, diminishing amounts of indole were oxidized after each addition (Fig. 2B). In addition, rapid decay of the Soret band was evident with addition of modest concentrations of H2O2 to the free enzyme (Fig. S1A); thus, we attribute the loss in activity at lower [IDO] to oxidative damage to the enzyme by peroxide.
Fig. 2.
Stoichiometry of the oxidation of indole catalyzed by IDO and H2O2. (A) [Indole] oxidized after addition of equal [H2O2], as monitored by fluorescence emission, at 1 and 10 μM IDOFe3+ (dashed and solid lines, respectively). (B) Stalling and restoration of the reaction of indole (422 μM) with H2O2 (1 mM) and multiple additions of IDOFe3+ (500 nmol each) as indicated by the arrows.
Dependence of the Distribution of Oxidation Products on H2O2 Concentration.
The distribution of indole oxidation products after reaction (1 min) as a function of [H2O2] was analyzed by HPLC (Fig. 3). With initial [indole] > [H2O2], the major products were 2- and 3-oxoindole in a ∼3:1 ratio (I and V, respectively; Fig. 1B), with trace amounts of o-formylaminobenzaldehyde (II). Indole-2,3-dione (III) and 3-hydroxy-2-oxoindole (IV) were detected only when [H2O2] > [indole]. The yield of these latter products increased with [H2O2], but formation of V was maximal when [indole] > [H2O2], whereas I formed maximally when [indole] ∼ 2[H2O2]. In contrast, the yield of dioxygenated derivatives (II, III, and IV) increased with [H2O2], as also observed for VII, although the latter product was not quantified. The pattern of product accumulation suggested involvement of additional oxidation reactions of the initial products driven by excess peroxide.
Fig. 3.
[H2O2]-dependence of product distribution. Reactant concentrations and conditions are the same as for Fig. 1B, except for [H2O2] (0.5–6 mM): ○, indole; ■, 2-oxoindole; □, 3-oxoindole; ▲, o-formylaminobenzaldehyde; △, indole-2,3-dione; ◇, 3-hydroxy-2-oxoindole. Reaction time: 60 s.
Oxidation of 2-Oxoindole (I) and 3-Oxoindole (V).
The IDO-catalyzed oxidations of I and V by H2O2 were examined to determine whether one or both of these monooxygenated products are intermediates in formation of the dioxygenated products. Aerobic reaction of IDO with HPLC-purified I and H2O2 (1 min) yielded the dioxygenated products III and IV (Fig. 4A). In contrast to indole, conditions resulting in quantitative oxidation of I could not be identified. The basis for this observation remains unclear, but it is consistent with the accumulation of I during IDO catalysis of indole oxidation by H2O2.
Fig. 4.
Oxidation of 2- and 3-oxoindole by IDO. HPLC analyses of products from incubation (60 s except where noted) of 2- and 3-oxoindole with IDO and H2O2. (A) Products from aerobic incubation of 2-oxoindole with IDOFe3+ and H2O2 (2, 0.1, and 2 mM, respectively; bold line) or with no additions (dashed line). (B) Products from anaerobic incubation of 3-oxoindole with IDOFe3+ and H2O2 (1, 0.05, and 1 mM, respectively; bold line) or with no additions (dashed line), and products from aerobic incubation (2 h) of 3-oxoindole with no additions (thin line). (Inset) (i and ii) Kinetics of anaerobic oxidation of 3-oxoindole (100 μM) by with IDOFe3+ (2.5 μM) and H2O2 (100 μM), or H2O2 alone, respectively. (iii) Kinetics of aerobic incubation (2 h) of 3-oxoindole with no additions.
Preparation of V was performed in situ by deesterification of 3-acetoxyindole with porcine liver esterase, and all work with this preparation before HPLC was performed anaerobically to prevent spontaneous oxidation of V to indigo blue (15, 16). Under anaerobic conditions, V was stable for hours if either H2O2 or IDOFe3+ were excluded from solution, but with both enzyme and peroxide present, it was oxidized rapidly (Fig. 4B, Inset), and indigo blue resulted. 3-Oxoindolenine (VI) was detected among the reaction products, but no significant dioxygenated indole derivatives were formed (Fig. 4B). Nevertheless, reaction of 3-oxoindole with O2 (∼2 h) in the absence of enzyme did afford III as a by-product, as reported previously (16). This result suggests that at least two pathways may account for this product in the IDO-catalyzed oxidation of indole: enzymatic oxidation of I and nonenzymatic oxidation of V by O2. Intriguingly, neither II nor VII was detected when the initial substrate was either I or V, so it is most likely that these products are produced by a different pathway than that resulting in III or IV.
Source of Oxygen in the IDO-Catalyzed Oxidation of Indoles by H2O2.
Anaerobic oxidation of indole by H2O2 catalyzed by IDO resulted in products similar to those observed under aerobic conditions (Table S3 and Fig. S2). These reactions were then performed with H218O2 to determine whether this oxidant is a source of oxygen in the reaction (Tables 1–3). Electrospray ionization (ESI)-MS analysis of I and V produced under aerobic conditions (16O2-saturated) revealed that both exhibit a new [M+H]+ ion of m/z 136 in addition to the previously observed [M+H]+ ion (m/z 134) that was consistent with the incorporation of one atom of 18O into the monooxygenated products. The percent incorporation of 18O into I and V reflected the ∼90% 18O enrichment of H218O2 used in these experiments. This result unambiguously demonstrates that the oxidation of indole by IDO and H2O2 occurred with essentially stoichiometric insertion of oxygen from H2O2 into the monooxygenated products. No loss of the 18O label in I occurred after 3 h at room temperature, whereas V oxidized spontaneously in this time.
Table 1.
Incorporation of 18O into indole oxidation products I and V
| [M+H]+ ion relative abundance, % |
|||||
| Product | Oxidant | m/z 134 | m/z 136 | m/z 138 | 18O incorporated, %* |
| I | H216O2 | 100 | 0.8 | — | — |
| H218O2 | 8.3 | 100 | 0.4 | 92 [102] | |
| V | H216O2 | 100 | 0.7 | — | — |
| H218O2 | 10.6 | 100 | 0.7 | 90 [100] | |
—, below detection limit.
*Numbers in brackets are normalized based on the enrichment level of H218O2.
Table 3.
Incorporation of 18O in indole oxidation product III
| [M+H]+ ion relative abundance, %* |
18O incorporated, %† |
|||||
| Product | Oxidant | m/z 148 | m/z 150 | m/z 152 | 16O18O | 18O18O |
| III | H216O2 | 100 | 1.2 | 0.3 | — | — |
| H218O2 | 28 (66) | 100 (100) | 2.2 (2.9) | 78 [87] | 0 | |
*Numbers in parentheses denote relative abundances from a repeat analysis after a 3-h incubation.
†Numbers in brackets are normalized based on the enrichment level of H218O2.
Table 2.
Incorporation of 18O in indole oxidation products II and IV
| [M+H]+ ion relative abundance, %* |
18O incorporated, %† |
|||||
| Product | Oxidant | m/z 150 | m/z 152 | m/z 154 | 16O18O | 18O18O |
| II | H216O2 | 100 | 1.0 | — | — | — |
| H218O2 | 20 (100) | 100 (75) | 58 (9.1) | 56 [62] | 32 [40] | |
| IV | H216O2 | 100 | 0.9 | — | — | — |
| H218O2 | 7 (9) | 100 (100) | 45 (35) | 65 [72] | 30 [37] | |
—, below detection limit.
*Numbers in parentheses denote relative abundances from a repeat analysis after a 3-h incubation.
†Numbers in brackets are normalized based on the enrichment level of H218O2.
ESI-MS analysis of the dioxygenated product II obtained in the same aerobic reaction resulted in two new [M+H]+ ions (m/z 154 and 152) and trace amounts of the previously detected ion (m/z 150). The relative abundance of these ions corresponded to 56% of singly 18O-labeled product and 32% of doubly 18O-labeled product. Over ∼3 h, 16O/18O exchange was evident and resulted in nearly complete disappearance of the m/z 154 ion and dominance of the fully 16O species. Both the lower than theoretical level of 18O incorporation in II seen initially and the subsequent, time-dependent 16O/18O exchange were consistent with behavior reported for the related compound N-FK (17) that exchanges oxygen with H216O (18).
Incorporation of 18O into III and IV was also evident. However, whereas at most one of the two oxygen atoms in III originated from H218O2 (78% incorporation of a single 18O atom), a second 18O was also found in IV (65% and 30% incorporation of one and two atoms of 18O, respectively). Time-dependent loss of the 18O label from both products was evident, but this loss was slower for IV than for II or III, with just one oxygen atom (presumably the C3 oxygen) exchanging with H2O over 3 h.
Methylated indoles were also readily oxidized by IDOFe3+ and H2O2 to products listed in Table S1. As with indole, these reactions were also performed with H218O2 (Table S4–S7). The aerobic reaction of 3-methylindole and 2-methylindole with IDOFe3+ and H2O2 proceeded with stoichiometric (92%) and partial (75%) incorporation of 18O from H218O2 into their respective monooxygenated products. In contrast, no 18O incorporation was evident in the monooxygenated derivative of 2,3-dimethylindole. Dioxygenated products were also produced in the oxidation of 3-methylindole and 2,3-dimethylindole, but the mechanism of their formation evidently differs from that of indole, because neither product incorporated 18O.
Kinetics of IDO-Catalyzed Indole Oxidation as Supported by H2O2.
In general, the use of low [IDO] (1 μM) in indole oxidation resulted in ∼100 turnovers before reaction stopped. Inactivation of IDO under these conditions presumably results from peroxidative damage to the protein. Consequently, the kinetics of indole oxidation by H2O2 catalyzed by IDO has been described in terms of the rate of indole consumption as monitored fluorometrically. The rate of disappearance of indole as a function of [H2O2] in the IDOFe3+-catalyzed reaction exhibited Michaelis–Menten behavior, with Km ∼1.1 mM for peroxide ([indole], 25–300 μM) (Fig. S3 and Table 4). In contrast, corresponding measurements as a function of [indole] exhibited biphasic kinetic behavior with increasing concentrations of substrate. This kinetic pattern resembled the behavior of several P450-catalyzed reactions for which it has been suggested that the availability a second substrate within the active site results in continuously increasing velocity at high [substrate] (19–21). A multisite kinetic model is also consistent with the conclusion that indole could bind to both IDOFe3+ and the IDOFe3+–Trp complex (7). Therefore, the concentration dependence of the initial rates was fitted to a single substrate, two-site model (19),
Table 4.
Kinetic parameters of the IDO-catalyzed oxidation of indole
| [H2O2] |
[Indole] |
|||
| kcat, s−1 | Km, mM | kcat1, s−1 | Km1, μM | kcat2/Km2, min−1 μM−1 |
| 2.6 ± 0.1* | 0.90 ± 0.08* | 1.3 ± 0.2§ | 9 ± 4§ | 0.96§ |
| 7.6 ± 0.7† | 1.3 ± 0.2† | 0.34 ± 0.07¶ | 3 ± 4¶ | 0.48¶ |
| 20 ± 3‡ | 1.2 ± 0.4‡ | —|| | —|| | 0.18|| |
—, values of kcat1 and Km1 were too small to obtain reliable estimates.
Error estimates reported as SE.
*Data obtained as a function of [H2O2] at 25 μM indole.
†Data obtained as a function of [H2O2] at 90 μM indole.
‡Data obtained as a function of [H2O2] at 300 μM indole.
§Data obtained as a function of [indole] at 500 μM H2O2.
¶Data obtained as a function of [indole] at 500 μM H2O2 and 2 μM l-Trp.
||Data obtained as a function of [indole] at 500 μM H2O2 and 5 μM l-Trp.
![]() |
where Km1 and Vmax1 are the kinetic constants for the first site and Km2 and Vmax2 are kinetic constants for the second site. This analysis resulted in an apparent kcat1 of 1.3 ± 0.2 s−1, Km1 of 9 ± 4 μM, and kcat2/Km2 of 0.96 min−1 μM−1 at 500 μM H2O2 (Table 4). Saturation of the second site was not observed in the range of [indole] examined. As little as 5 μM l-Trp strongly inhibited the reaction and changed the kinetics from biphasic to monophasic as the result of near abolition of kcat1, presumably owing to the high affinity of Trp for the ferryl enzyme [Kd = 0.3 μM for the high-affinity binding site (9)]. Nevertheless, even [l-Trp] as great as 30 μM was insufficient to completely inhibit activity. l-trypotophan was not oxidized by IDOFe3+ and H2O2 to any significant degree.
Enzyme Reduction in the Reaction with Indole and H2O2.
IDO-catalyzed oxidation of indole was monitored anaerobically and with substoichiometric [H2O2] (∼0.8 molar equivalent relative to IDOFe3+) to determine the oxidation state of IDO after turnover. Electronic spectra of the enzyme under these conditions revealed formation of a minor fraction of IDOFe2+ (∼12%) in ∼15 min. An accompanying increase in absorbance at 608 nm indicated that indigo blue accumulated during this time. At higher [H2O2] (5 molar equivalents), the fraction of IDOFe2+ formed increased to ∼50% (Fig. S1B). Immediately after addition of H2O2, [indole] decreased briefly, but then stopped changing even though IDO reduction continued. Introduction of catalase after H2O2 addition did not influence this decrease in the enzyme. These results argue against the involvement of a coupled process, but the anaerobic conversion of V to indigo blue by IDOFe3+ under anaerobic conditions in the absence of H2O2 indicates that reduction of IDOFe3+ is sufficient to permit formation of this product. In the absence of H2O2 and O2, V led to nearly quantitative reduction of the enzyme to IDOFe2+ (λmax = 559 nm; isosbestic points, 461.5 nm and 523 nm) (Fig. 5). No other product of indole oxidation was able to reduce IDOFe3+, so we attribute the formation of IDOFe2+ observed during oxidation of indole by IDOFe3+ and H2O2 to reduction of IDOFe3+ by V after its formation.
Fig. 5.
Anaerobic reduction of IDOFe3+ by 3-oxoindole. IDOFe3+ (7 μM) was incubated with 3-acetoxyindole (70 μM) under N2. Reaction was initiated by addition of esterase (2 units/mL) to generate 3-oxoindole. Spectra were recorded at 1-min intervals. (Inset) kinetics of indigo blue formation (608 nm) (●) and of enzyme reduction [ε559nm = 5.23 mM-1cm−1 (IDOFe3+) and 15.03 mM-1cm−1 (IDOFe2+)] (○).
Discussion
Our results demonstrate that IDO has peroxygenase activity toward indole but not toward Trp. This reaction proceeds under aerobic and anaerobic conditions and is not inhibited by superoxide dismutase (SOD), d-mannitol, or lidocaine (SI Materials and Methods). Thus, IDO-catalyzed oxidation of indole in this reaction does not require O2 or involve O2•−, •OH, or 1O2 outside the heme cavity. The consumption of indole and H2O2 proceeds at nearly stoichiometric proportions at moderate [H2O2], and the oxygen incorporated into the monooxygenated products originates from H2O2. In these experiments, the enzyme was initially present as IDOFe3+, and this state of the enzyme was regenerated when reaction was complete, thereby accounting for two oxidizing equivalents from H2O2, as expected for a peroxygenase. However, 3-oxoindole (V) formed initially can reduce IDOFe3+ to IDOFe2+ under specific conditions of [O2] and [H2O2]. Thus, the characteristics of this IDO reaction closely resemble those of the “H2O2 shunt” reported for P450 (22–26) and lead us to propose that a reactive compound I intermediate (Fe4+=O, with a porphyrin or protein radical center) transfers the ferryl oxygen to indole to generate monooxygenated products.
Although oxidation of indole by IDO has not been reported previously, oxidation of indole by other heme enzymes [e.g., HRP (27), chloroperoxidases (28), P450s (29–31)] is well known; however, the mechanisms and products of these reactions differ. Nevertheless, catalysis by both oxygenases and peroxidases involves conversion of the ferric enzyme to the principal reactive intermediate compound I (32–35).
No direct evidence for the formation of IDO compound I is currently available, although evidence of a ferryl species with a protein-based radical has been reported for the related enzyme Trp 2,3-dioxygenase (36). Resonance Raman experiments indicating formation of IDO compound II after reaction of IDOFe3+ with H2O2 provide the only evidence for formation of a higher oxidation state by this enzyme (9, 11, 12). Oxygenation of the neutral indole by IDO compound II is unlikely, however, given that transfer of the ferryl oxygen would generate IDOFe2+ and lead to IDOFe3+-O2•−, which is unreactive with indole. Similarly, single electron oxidation of indole by IDO compound II followed by oxygen transfer to the ensuing indolyl radical by a second IDO compound II is unlikely in view of the near-stoichiometric consumption of indole and H2O2. Although we propose involvement of a compound I intermediate in the oxidation of indole by IDO, the location of the radical on the porphyrin or the protein is speculative. Because only compound II has been detected at present, dissipation of the putative radical and/or transfer of the oxygen from IDO compound I to indole must be prompt. Such tightly coupled transfer of oxygen to indole presumably involves formation of a transient compound I–indole complex in which the substrate can access ferryl oxygen. The inhibition of indole oxidation by l-Trp, which binds to IDO compound II without reacting (9), is consistent with formation of such a complex.
The transfer of oxygen to indole by IDO compound I (Fig. 6) to produce I and V could occur by hydroxylation or epoxidation, as known to occur for many reactions of P450 (25, 26). Hydroxylation of either C2 or C3 is presumably a consequence of the large active site of IDO, which would allow binding of indole so that either carbon atom can access the ferryl center. Once formed, V can react with O2 to form VI, leading to indigo blue (15), or to form III (16). Two equivalents of IDOFe3+ alone can oxidize V to indigo blue, with the accumulation of IDOFe2+, presumably as the result of two one-electron oxidations. In the presence of H2O2, the ferryl/Fe3+ couple can promote rapid and extensive oxidation of V, again by two one-electron transfer reactions, consistent with the near absence of oxygenated products derived from anaerobic oxidation of V by H2O2 and IDOFe3+. In contrast, dioxygenated products III and IV can be derived by aerobic oxidation of I by IDOFe3+ in the presence of H2O2. The amount of 18O incorporated into IV by catalysis of H218O2 oxidation of indole by IDO was consistent with near-stoichiometric incorporation of 18O at C2 and partial (30%) incorporation of 18O at C3. This isotopic distribution could result from the time-dependent exchange of 18O with H216O, but alternative sources for the second oxygen atom are possible.
Fig. 6.
Proposed reactions for the IDOFe3+ catalyzed of indole oxidation by H2O2.
Formation of II (an analogue of N–FK) is notable, because this compound is the only product that we have identified that has not been reported to result from P450-catalyzed oxidation of indole (29–31, 37–39). In contrast, oxidation of indoles (40) and some indoleamines (41) by HRP can lead to oxidative indole ring opening by reaction of indolyl radical with O2 or O2•− (40), as reported for the photooxidation of l-Trp to N-FK (42). In principle, a peroxidatic mechanism could account for IDO oxidation of indole, as has been proposed for IDO oxidation of melatonin (8), but this possibility is ruled out for indole oxidation by incorporation of 18O from H218O2. On the other hand, peroxidation could account for IDO-catalyzed ring opening of 3-methylindole and 2,3-dimethyl indole, neither of which incorporated 18O from H218O2. Given that IDO lacks catalase activity (6), it is unlikely that sufficient 18O2 forms by disproportionation of H218O2 to react with a presumed indolyl radical and account for the substantial 18O content of the product. Instead, it is more likely that oxidation of indole to II and other indole products occurs along a common pathway initiated by the transfer of oxygen from IDO compound I.
Failure to observe II when I or V was used as substrate implies that neither of these compounds is an intermediate in formation of this product from indole. One possible intermediate is the highly labile indoline-2,3-epoxide (43–45), which decays to I- and V-type species (44, 45), consistent with the nearly identical incorporation of 18O into these two products. Recently, l-Trp epoxide has been proposed to accept an oxygen atom from IDO compound II during catalysis to regenerate IDOFe2+ (46, 47). A similar mechanism could convert the epoxide (before rearrangement) to II. The substoichiometric incorporation of 18O as the second atom of oxygen in II (and IV) is not consistent with this mechanism, however. One alternative mechanism involves nucleophilic attack on the epoxide by H2O, which generally leads to a vicinal diol (48–50), in this case indoline-2,3-diol. This mechanism could permit incorporation of one atom of 16O from bulk solvent or 18O from H218O (resulting from H218O2 after initial compound I formation). Indoline-2,3-diol may be unstable (51), but it could be oxidized by H2O2 and IDO-Fe3+ to form the stable products II and/or IV. The mechanism of this subsequent oxidation is speculative but may involve P450-type ferryl chemistry relevant to C-C bond cleavage of diols (26, 52) and alcohol oxidation (53, 54) (Fig. S4).
The remarkable selectivity that allows IDO to act on indole as a peroxygenase but on indoleamines exclusively as a dioxygenase reinforces the importance of the derivatized alkyl substituent in orientation of the substrate and possible stabilization of catalytic intermediates at the active site of this enzyme. As a result, electron transfer from indole to IDOFe3+-O2•− to produce a more highly oxidized form of IDO heme iron, as believed to occur with Trp (9, 11–13, 46, 55) does not occur. Similarly, no appreciable electron transfer from Trp bound near the heme iron to the presumed IDO compound I intermediate occurs to initiate substrate oxidation in the peroxygenase reaction. Without more detailed structural or spectroscopic information, it seems likely that the proposed ability of Trp to bind to more than one location at the active site of IDO is relevant to these observations (7, 56). Specifically, differences in the binding of Trp and indole to secondary sites at the active center of IDO may influence the binding orientations of both substrates in proximity to the heme iron and contribute significantly to the catalytic selectivity and the product distribution that we have observed.
Materials and Methods
Proteins and Reagents.
Bovine erythrocyte SOD, porcine liver esterase, glucose oxidase (Aspergillus niger), 30% H2O2, indole, 3-methylindole, 2-methylindole, 2,3-dimethylindole, 2-oxoindole, 3-methyl-2-oxoindole, 3-acetoxyindole, indole-2,3-dione (isatin), d-glucose, lidocaine, d-mannitol, and Trizma base were obtained from Sigma-Aldrich. Reagent concentrations were determined spectrophotometrically [H2O2, ε240 = 39.4 M−1cm−1 (57); indole, ε276 = 5,670 M−1cm−1 (58); 3-methylindole, ε280 = 5,160 M−1cm−1 (58); 2-methylindole, ε270 = 6,500 M−1cm−1 (58); 2,3-dimethylindole, ε280 = 6,360 M−1cm−1 (58); 3-acetoxyindole, ε278 = 5,500 M−1cm−1 (this study)]. Porcine liver esterase, glucose oxidase, and SOD were quantified from the activity reported by the manufacturer. IDO was expressed in Escherichia coli and purified (A404/A280 = 2.05) by a method culminating with elution over a Cibacron 3GA column (Sigma-Aldrich) (59). IDO concentration was determined spectrophotometrically (IDOFe3+ ε406 = 175,000 M−1cm−1) (6).
Spectroscopy.
Electronic absorption spectra (20 °C) were recorded with a Cary 4000 or 6000i spectrophotometer. Fluorescence emission spectra (20 °C) were recorded with a Cary Eclipse spectrofluorimeter [slit widths, 20 (emission) and 1.5 nm (excitation)].
IDO-Catalyzed Oxidation of Indoles.
Typically, reactions (30 μL) were performed in Tris buffer (20 mM, pH 7.5, 20 °C), initiated by the addition of H2O2 [0.5–6 mM; natural abundance H2O2 or 90% (vol/vol) 18O-enriched H218O2 in H216O; ICON Isotopes] to solutions containing IDO and indole (0.2 and 2 mM, respectively). Reactions were stopped after 60 s by addition of ice-cold methanol (50 μL). After centrifugation to remove inactivated enzyme, the supernatant fluid (20 μL) was analyzed immediately by HPLC. The lower aqueous solubilities of methylindoles required that they be dissolved in methanol (∼1–2%), which limited the concentrations of these compounds that could be assayed (1, mM for 3-methylindole, 1 mM for 2-methylindole, and 0.2 mM for 2,3-dimethylindole). Other derivatives assayed included 2-oxoindole (2 mM) and 3-oxoindole (1 mM). 3-Oxoindole was prepared from 3-acetoxyindole by anaerobic deesterification with porcine liver esterase (2 U/mL) immediately before use (60). Reactions were performed under aerobic conditions unless indicated otherwise. Anaerobic reactions were performed in a N2-filled glove box (O2 ≤1 ppm) using indole and protein solutions that had been purged with N2 for ∼30 min. H2O2 solutions were deoxygenated by repeated freeze–pump–thaw cycles. In some cases, d-glucose (10–100 mM) and glucose oxidase (24–47 U/mL) were included to scavenge traces of O2 (61). To monitor the progress of anaerobic reactions, reaction mixtures were prepared in Thunberg cuvettetes in a glove box. Anaerobically generated reaction products were collected in the glove box after the protein was precipitated with N2-purged methanol, and the resulting reaction mixtures were clarified by centrifugation and analyzed by aerobic HPLC and ESI-MS (SI Materials and Methods and Figs. S5 and S6).
Kinetics of Indole Oxidation.
Solutions of IDO (1 μM) and indole (12.5–400 μM) in Tris buffer (20 mM; pH 7.5) were mixed with H2O2 (25 μM to 4 mM) (final volume, 300 μL) in a dual-path length masked cuvette (1 cm × 1 cm), and indole consumption was monitored (triplicate or greater) by fluorescence at 345 nm (λex = 280 nm). The contribution of the primary product, 2-oxoindole, to fluorescence was negligible at this concentration. Nonlinearity of the fluorescence response was corrected with a standard curve ([indole] vs. λem = 345 nm). Initial rates were determined with the logarithmic approximation method (62), and kinetic parameters were obtained by nonlinear regression fitting of the data (Origin 8.0; OriginLab).
Supplementary Material
Acknowledgments
We thank Drs. Federico I. Rosell and Marcia R. Mauk for developing the original IDO purification protocol, and Drs. Emily Seo and David Williams for providing access to the mass spectrometers. This work was supported by a grant from the Canadian Cancer Research Institute.
Footnotes
The authors declare no conflict of interest.
This article is a PNAS Direct Submission.
This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1207191109/-/DCSupplemental.
References
- 1.Hayaishi O, et al. Indoleamine 2,3-dioxygenase: Incorporation of 18O2− and 18O2 into the reaction products. J Biol Chem. 1977;252:3548–3550. [PubMed] [Google Scholar]
- 2.Takikawa O. Biochemical and medical aspects of the indoleamine 2,3-dioxygenase-initiated L-tryptophan metabolism. Biochem Biophys Res Commun. 2005;338:12–19. doi: 10.1016/j.bbrc.2005.09.032. [DOI] [PubMed] [Google Scholar]
- 3.Katz JB, Muller AJ, Prendergast GC. Indoleamine 2,3-dioxygenase in T-cell tolerance and tumoral immune escape. Immunol Rev. 2008;222:206–221. doi: 10.1111/j.1600-065X.2008.00610.x. [DOI] [PubMed] [Google Scholar]
- 4.Hayaishi O. Properties and function of indoleamine 2,3-dioxygenase. J Biochem. 1976;79:13P–21P. doi: 10.1093/oxfordjournals.jbchem.a131115. [DOI] [PubMed] [Google Scholar]
- 5.Hirata F, Ohnishi T, Hayaishi O. Indoleamine 2,3-dioxygenase: Characterization and properties of enzyme O2- complex. J Biol Chem. 1977;252:4637–4642. [PubMed] [Google Scholar]
- 6.Shimizu T, Nomiyama S, Hirata F, Hayaishi O. Indoleamine 2,3-dioxygenase: Purification and some properties. J Biol Chem. 1978;253:4700–4706. [PubMed] [Google Scholar]
- 7.Sono M. Enzyme kinetic and spectroscopic studies of inhibitor and effector interactions with indoleamine 2,3-dioxygenase. 2. Evidence for the existence of another binding site in the enzyme for indole derivative effectors. Biochemistry. 1989;28:5400–5407. doi: 10.1021/bi00439a013. [DOI] [PubMed] [Google Scholar]
- 8.Ferry G, et al. Molecular evidence that melatonin is enzymatically oxidized in a different manner than tryptophan: Investigations with both indoleamine 2,3-dioxygenase and myeloperoxidase. Biochem J. 2005;388:205–215. doi: 10.1042/BJ20042075. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Lu C, Yeh SR. Ferryl derivatives of human indoleamine 2,3-dioxygenase. J Biol Chem. 2011;286:21220–21230. doi: 10.1074/jbc.M111.221507. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Takikawa O, Yoshida R, Hayaishi O. Monooxygenase activities of dioxygenases: Benzphetamine demethylation and aniline hydroxylation reactions catalyzed by indoleamine 2,3-dioxygenase. J Biol Chem. 1983;258:6808–6815. [PubMed] [Google Scholar]
- 11.Lewis-Ballester A, et al. Evidence for a ferryl intermediate in a heme-based dioxygenase. Proc Natl Acad Sci USA. 2009;106:17371–17376. doi: 10.1073/pnas.0906655106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Yanagisawa S, et al. Resonance Raman study on the oxygenated and the ferryl-oxo species of indoleamine 2,3-dioxygenase during catalytic turnover. Faraday Discuss. 2011;148:239–247, discussion 299–314. doi: 10.1039/c004552g. [DOI] [PubMed] [Google Scholar]
- 13.Capece L, et al. The first step of the dioxygenation reaction carried out by tryptophan dioxygenase and indoleamine 2,3-dioxygenase as revealed by quantum mechanical/molecular mechanical studies. J Biol Inorg Chem. 2010;15:811–823. doi: 10.1007/s00775-010-0646-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Sugimoto H, et al. Crystal structure of human indoleamine 2,3-dioxygenase: Catalytic mechanism of O2 incorporation by a heme-containing dioxygenase. Proc Natl Acad Sci USA. 2006;103:2611–2616. doi: 10.1073/pnas.0508996103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Russell GA, Kaupp G. Reactions of resonance stabilized carbanions. XXXI. Oxidation of carbanions: Oxidation of indoxyl to indigo in basic solution. J Am Chem Soc. 1969;91:3851–3859. [Google Scholar]
- 16.Cotson S, Holt SJ. Studies in enzyme cytochemistry. IV. Kinetics of aerial oxidation of indoxyl and some of its halogen derivatives. Proc R Soc Lond B Biol Sci. 1958;148:506–519. doi: 10.1098/rspb.1958.0042. [DOI] [PubMed] [Google Scholar]
- 17.Hayaishi O, Rothberg S, Mehler AH, Saito Y. Studies on oxygenases; enzymatic formation of kynurenine from tryptophan. J Biol Chem. 1957;229:889–896. [PubMed] [Google Scholar]
- 18.Ronsein GE, de Oliveira MC, de Medeiros MH, Di Mascio P. Characterization of O(2) ((1)delta(g))-derived oxidation products of tryptophan: A combination of tandem mass spectrometry analyses and isotopic labeling studies. J Am Soc Mass Spectrom. 2009;20:188–197. doi: 10.1016/j.jasms.2008.08.016. [DOI] [PubMed] [Google Scholar]
- 19.Korzekwa KR, et al. Evaluation of atypical cytochrome P450 kinetics with two-substrate models: Evidence that multiple substrates can simultaneously bind to cytochrome P450 active sites. Biochemistry. 1998;37:4137–4147. doi: 10.1021/bi9715627. [DOI] [PubMed] [Google Scholar]
- 20.Atkins WM. Non–Michaelis-Menten kinetics in cytochrome P450-catalyzed reactions. Annu Rev Pharmacol Toxicol. 2005;45:291–310. doi: 10.1146/annurev.pharmtox.45.120403.100004. [DOI] [PubMed] [Google Scholar]
- 21.Hutzler JM, Hauer MJ, Tracy TS. Dapsone activation of CYP2C9-mediated metabolism: Evidence for activation of multiple substrates and a two-site model. Drug Metab Dispos. 2001;29:1029–1034. [PubMed] [Google Scholar]
- 22.Hrycay EG, Gustafsson JA, Ingelman-Sundberg M, Ernster L. Sodium periodate, sodium chloride, organic hydroperoxides, and H2O2 as hydroxylating agents in steroid hydroxylation reactions catalyzed by partially purified cytochrome P-450. Biochem Biophys Res Commun. 1975;66:209–216. doi: 10.1016/s0006-291x(75)80315-9. [DOI] [PubMed] [Google Scholar]
- 23.Nordblom GD, White RE, Coon MJ. Studies on hydroperoxide-dependent substrate hydroxylation by purified liver microsomal cytochrome P-450. Arch Biochem Biophys. 1976;175:524–533. doi: 10.1016/0003-9861(76)90541-5. [DOI] [PubMed] [Google Scholar]
- 24.Denisov IG, Makris TM, Sligar SG, Schlichting I. Structure and chemistry of cytochrome P450. Chem Rev. 2005;105:2253–2277. doi: 10.1021/cr0307143. [DOI] [PubMed] [Google Scholar]
- 25.Sono M, Roach MP, Coulter ED, Dawson JH. Heme-containing oxygenases. Chem Rev. 1996;96:2841–2888. doi: 10.1021/cr9500500. [DOI] [PubMed] [Google Scholar]
- 26.Meunier B, de Visser SP, Shaik S. Mechanism of oxidation reactions catalyzed by cytochrome p450 enzymes. Chem Rev. 2004;104:3947–3980. doi: 10.1021/cr020443g. [DOI] [PubMed] [Google Scholar]
- 27.Holmes-Siedle AG, Saunders BC. Peroxidase-catalyzed oxidation of indole. Chem Ind. 1957:265–266. [Google Scholar]
- 28.Corbett MD, Chipko BR. Peroxide oxidation of indole to oxindole by chloroperoxidase catalysis. Biochem J. 1979;183:269–276. doi: 10.1042/bj1830269. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Gillam EM, Notley LM, Cai H, De Voss JJ, Guengerich FP. Oxidation of indole by cytochrome P450 enzymes. Biochemistry. 2000;39:13817–13824. doi: 10.1021/bi001229u. [DOI] [PubMed] [Google Scholar]
- 30.Nakamura K, Martin MV, Guengerich FP. Random mutagenesis of human cytochrome p450 2A6 and screening with indole oxidation products. Arch Biochem Biophys. 2001;395:25–31. doi: 10.1006/abbi.2001.2569. [DOI] [PubMed] [Google Scholar]
- 31.Li HM, Mei LH, Urlacher VB, Schmid RD. Cytochrome P450 BM-3 evolved by random and saturation mutagenesis as an effective indole-hydroxylating catalyst. Appl Biochem Biotechnol. 2008;144:27–36. doi: 10.1007/s12010-007-8002-5. [DOI] [PubMed] [Google Scholar]
- 32.Makris TM, von Koenig K, Schlichting I, Sligar SG. The status of high-valent metal oxo complexes in the P450 cytochromes. J Inorg Biochem. 2006;100:507–518. doi: 10.1016/j.jinorgbio.2006.01.025. [DOI] [PubMed] [Google Scholar]
- 33.Sligar SG, Makris TM, Denisov IG. Thirty years of microbial P450 monooxygenase research: Peroxo-heme intermediates—the central bus station in heme oxygenase catalysis. Biochem Biophys Res Commun. 2005;338:346–354. doi: 10.1016/j.bbrc.2005.08.094. [DOI] [PubMed] [Google Scholar]
- 34.Dunford HB. One-electron oxidations by peroxidases. Xenobiotica. 1995;25:725–733. doi: 10.3109/00498259509061888. [DOI] [PubMed] [Google Scholar]
- 35.Rittle J, Green MT. Cytochrome P450 compound I: Capture, characterization, and C-H bond activation kinetics. Science. 2010;330:933–937. doi: 10.1126/science.1193478. [DOI] [PubMed] [Google Scholar]
- 36.Fu R, et al. Enzyme reactivation by hydrogen peroxide in heme-based tryptophan dioxygenase. J Biol Chem. 2011;286:26541–26554. doi: 10.1074/jbc.M111.253237. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Gillam EM, et al. Formation of indigo by recombinant mammalian cytochrome P450. Biochem Biophys Res Commun. 1999;265:469–472. doi: 10.1006/bbrc.1999.1702. [DOI] [PubMed] [Google Scholar]
- 38.Gillam EMJ, Guengerich FP. Exploiting the versatility of human cytochrome P450 enzymes: The promise of blue roses from biotechnology. IUBMB Life. 2001;52:271–277. doi: 10.1080/152165401317291110. [DOI] [PubMed] [Google Scholar]
- 39.Huang WC, et al. Filling a hole in cytochrome P450 BM3 improves substrate binding and catalytic efficiency. J Mol Biol. 2007;373:633–651. doi: 10.1016/j.jmb.2007.08.015. [DOI] [PubMed] [Google Scholar]
- 40.Ximenes VF, Campa A, Catalani LH. The oxidation of indole derivatives catalyzed by horseradish peroxidase is highly chemiluminescent. Arch Biochem Biophys. 2001;387:173–179. doi: 10.1006/abbi.2000.2228. [DOI] [PubMed] [Google Scholar]
- 41.Ximenes VF, Catalani LH, Campa A. Oxidation of melatonin and tryptophan by an HRP cycle involving compound III. Biochem Biophys Res Commun. 2001;287:130–134. doi: 10.1006/bbrc.2001.5557. [DOI] [PubMed] [Google Scholar]
- 42.Davies MJ, Truscott RJ. Photo-oxidation of proteins and its role in cataractogenesis. J Photochem Photobiol B. 2001;63:114–125. doi: 10.1016/s1011-1344(01)00208-1. [DOI] [PubMed] [Google Scholar]
- 43.Adam W, Reinhardt D. Indole dioxetanes and epoxides by oxidation of N-acylated indoles with singlet oxygen and dimethyldioxirane: Kinetics and chemiluminescence yields of the thermal dioxetane decomposition and fluoride ion-induced CIEEL emission. J Chem Soc Perkin Trans. 1994;2:1503–1507. [Google Scholar]
- 44.Adam W, Ahrweiler M, Sauter M, Schmiedeskamp B. Oxidation of indoles by singlet oxygen and dimethyldioxirane: Isolation of indole dioxetanes and epoxides by stabilization through nitrogen acylation. Tetrahedron Lett. 1993;34:5247–5250. [Google Scholar]
- 45.Skordos KW, Skiles GL, Laycock JD, Lanza DL, Yost GS. Evidence supporting the formation of 2,3-epoxy-3-methylindoline: A reactive intermediate of the pneumotoxin 3-methylindole. Chem Res Toxicol. 1998;11:741–749. doi: 10.1021/tx9702087. [DOI] [PubMed] [Google Scholar]
- 46.Capece L, Lewis-Ballester A, Yeh SR, Estrin DA, Marti MA. Complete reaction mechanism of indoleamine 2,3-dioxygenase as revealed by QM/MM simulations. J Phys Chem B. 2012;116:1401–1413. doi: 10.1021/jp2082825. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Basran J, et al. The mechanism of formation of N-formylkynurenine by heme dioxygenases. J Am Chem Soc. 2011;133:16251–16257. doi: 10.1021/ja207066z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Whalen DL. Buffer catalysis in epoxide hydrolyses. J Am Chem Soc. 1973;95:3432–3434. [Google Scholar]
- 49.Whalen DL, Ross AM. Specific effects of chloride ion on epoxide hydrolysis: The pH-dependence of the rates and mechanisms for the hydrolysis of indene oxide. J Am Chem Soc. 1976;98:7859–7861. doi: 10.1021/ja00440a086. [DOI] [PubMed] [Google Scholar]
- 50.Whalen DL, Montemarano JA, Thakker DR, Yagi H, Jerina DM. Changes of mechanisms and product distributions in the hydrolysis of benzo[a]pyrene-7,8-diol 9,10-epoxide metabolites induced by changes in pH. J Am Chem Soc. 1977;99:5522–5524. doi: 10.1021/ja00458a069. [DOI] [PubMed] [Google Scholar]
- 51.Ensley BD, et al. Expression of naphthalene oxidation genes in Escherichia coli results in the biosynthesis of indigo. Science. 1983;222:167–169. doi: 10.1126/science.6353574. [DOI] [PubMed] [Google Scholar]
- 52.Ortiz de Montellano PR, editor. Cytochrome P450: Structure, Mechanism, and Biochemistry. 3rd Ed. New York: Plenum; 2004. [Google Scholar]
- 53.Vaz AD, Coon MJ. On the mechanism of action of cytochrome P450: Evaluation of hydrogen abstraction in oxygen-dependent alcohol oxidation. Biochemistry. 1994;33:6442–6449. doi: 10.1021/bi00187a008. [DOI] [PubMed] [Google Scholar]
- 54.Bell-Parikh LC, Guengerich FP. Kinetics of cytochrome P450 2E1-catalyzed oxidation of ethanol to acetic acid via acetaldehyde. J Biol Chem. 1999;274:23833–23840. doi: 10.1074/jbc.274.34.23833. [DOI] [PubMed] [Google Scholar]
- 55.Yanagisawa S, et al. Identification of the Fe-O2 and the Fe=O heme species for indoleamine 2,3-dioxygenase during catalytic turnover. Chem Lett. 2010;39:36–37. [Google Scholar]
- 56.Lu C, Lin Y, Yeh SR. Inhibitory substrate binding site of human indoleamine 2,3-dioxygenase. J Am Chem Soc. 2009;131:12866–12867. doi: 10.1021/ja9029768. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Noble RW, Gibson QH. The reaction of ferrous horseradish peroxidase with hydrogen peroxide. J Biol Chem. 1970;245:2409–2413. [PubMed] [Google Scholar]
- 58.Ramachandran LK, Witkop B. The interaction of mercuric acetate with indoles, tryptophan, and proteins. Biochemistry. 1964;3:1603–1611. doi: 10.1021/bi00899a001. [DOI] [PubMed] [Google Scholar]
- 59.Rosell FI, Kuo HH, Mauk AG. NADH oxidase activity of indoleamine 2,3-dioxygenase. J Biol Chem. 2011;286:29273–29283. doi: 10.1074/jbc.M111.262139. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Barrnett RJ, Seligman AM. Histochemical demonstration of esterases by production of indigo. Science. 1951;114:579–582. doi: 10.1126/science.114.2970.579. [DOI] [PubMed] [Google Scholar]
- 61.Benesch RE, Benesch R. Enzymatic removal of oxygen for polarography and related methods. Science. 1953;118:447–448. doi: 10.1126/science.118.3068.447. [DOI] [PubMed] [Google Scholar]
- 62.Lu WP, Fei L. A logarithmic approximation to initial rates of enzyme reactions. Anal Biochem. 2003;316:58–65. doi: 10.1016/s0003-2697(03)00034-4. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.







