Abstract
Iodotyrosine deiodinase (IYD) is unusual in its reliance on flavin to promote reductive dehalogenation of halotyrosines under aerobic conditions. Applications of this activity can be envisioned for bioremediation, but expansion of its specificity requires an understanding of the mechanistic steps that limit the rate of turnover. Key processes capable of controlling steady-state turnover have now been evaluated and described in this study. While proton transfer is necessary for converting the electron-rich substrate into an electrophilic intermediate suitable for reduction, kinetic solvent deuterium isotope effects suggest that this process does not contribute to the overall efficiency of catalysis under neutral conditions. Similarly, reconstituting IYD with flavin analogues demonstrates that a change in reduction potential by as much as 132 mV affects kcat by less than 3-fold. Furthermore, kcat/Km does not correlate with reduction potential and indicates that electron transfer is also not rate determining. Catalytic efficiency is most sensitive to the electronic nature of its substrates. Electron-donating substituents on the ortho position of iodotyrosine stimulate catalysis and conversely electron-withdrawing substituents suppress catalysis. Effects on kcat and kcat/Km range from 22- to 100-fold and fit a linear free-energy correlation with a ρ ranging from −2.1 to −2.8 for human and bacterial IYD. These values are consistent with a rate-determining process of stabilizing the electrophilic and nonaromatic intermediate poised for reduction. Future engineering can now focus on efforts to stabilize this electrophilic intermediate over a broad series of phenolic substrates that are targeted for removal from our environment.
Graphical Abstract
INTRODUCTION
Halophenols such as pentachlorophenol, bromoxynil, triclosan, and tetrabromobisphenol A have been widely dispersed in the environment as pesticides, herbicides, antimicrobial agents, and flame retardants. Unfortunately, their beneficial properties are now overshadowed by significant health risks caused by their accumulation in our surroundings. Natural pathways for their degradation are inefficient and waste treatment is problematic.1-3 A variety of enzyme-based strategies have been proposed for their detoxification but few have yet been applied successfully in the environment.4-6 A reductive dehalogenation catalyzed by the flavoprotein iodotyrosine deiodinase (IYD) offers a fresh alternative for bioremediation. Homologues of this enzyme are encoded by most metazoa and some bacteria. All variants examined to date prefer halotyrosines over halophenols as substrates.7-9 Efforts to broaden substrate tolerance by engineering regions of the active site have provided only modest success.10 High affinity between IYD and its substrates is not sufficient for rapid turnover,7 and ligands capable of promoting a disorder-to-order transition of the active site lid do not guarantee effective catalysis.10 Future manipulation of its substrate specificity remains stymied until the processes controlling catalytic efficiency are identified.
Successful engineering of a multistep enzyme requires a knowledge of the catalytic mechanism, the structural components contributing to efficient turnover, and the kinetic barriers encountered during turnover. Current data on IYD are consistent with a series of one electron transfers from the reduced flavin (FMNhq) to a dearomatized form of the substrate (Scheme 1).11 These transfers in turn depend on the ability of the substrate to reposition an active site Thr hydroxyl group to form a hydrogen bond with the N(5) position of flavin.8,12 Otherwise, IYD exhibits a promiscuous ability to promote hydride transfer as evident from a nitroreductase activity that is shared by other enzymes within its nitroreductase structural superfamily.8,13,14 The desired expansion of substrate specificity for dehalogenation is further complicated by the numerous bond reorganizations suggested in the mechanism of catalysis. For example, substrate aromaticity is disrupted by carbon protonation to create a highly electrophilic intermediate poised for electron transfer and subsequent halide elimination (Scheme 1). Equivalent loss of aromaticity has been proposed in mechanisms such as tyrosine phenol lyase for which an equivalent dienone contributes to the total energy barrier of reaction (Scheme 2).15 Subsequent steps of IYD such as product release do not likely control the rate of turnover since the steady-state kcat/Km for iodotyrosine (I-Tyr) deiodination is similar to the second-order rate constant (kox) for I-Tyr-dependent oxidation of the reduced (FMNhq) IYD.16 Cleavage of the carbon–halide bond also does not likely control the reaction since the kcat/Km values for I-Tyr and bromotyrosine are nearly identical despite the differences in their carbon–halogen bond energy.17,18 The key steps of substrate protonation and electron transfer are considered most likely to control turnover. These alternatives have now been examined in this report by use of a solvent isotope effect and a series of flavin and substrate analogues. Together, the data suggest that the extent rather than the rate of substrate protonation controls the catalytic rate under neutral conditions.
Scheme 1.
(A) Mechanism for Reductive Dehalogenation by IYD. (B) Isoalloxazine Component of Reduced Flavin Mononucleotide (FMNhq)
Scheme 2.
Protonation of Tyrosine during Catalysis by Tyrosine Phenol Lyase15
MATERIALS AND METHODS
General Materials.
Commercial sources for all starting materials and reagents are described in the Supporting Materials. Iodotyrosine deiodinase (Human, HsIYD) lacking its N-terminal transmembrane sequence (residues 2–31) was expressed as a SUMO fusion, purified by Ni2+ affinity and size exclusion chromatography and stored as described previously.12 IYD from Haliscomenobacter hydrossis (HhIYD) was expressed with a C-terminal His6-tag, purified by Ni2+ affinity and stored as described previously.7
General Methods.
NMR spectra were recorded on Bruker Avance-300 and Avance-400 spectrometers. UV measurements were performed with an Agilent 8453 spectrophotometer, and fluorescence measurements were performed with a FluoroMax-4 fluorescence spectrophotometer.
FMN and Substrate Analogues.
Detailed preparations are described in the Supporting Information. Briefly, FMN analogues were synthesized from the appropriately substituted nitroaniline as described previously.19 These were conjugated with ribose, reduced and condensed with alloxan. Phosphorylation was accomplished with ATP and riboflavin kinase.20 Tyrosine analogues were prepared by a Negishi cross-coupling of the appropriate para-iodophenol with N-Boc-3-iodo-l-alanine methyl ester.21 Iodination of the tyrosine analogues typically used iodine in ammonium hydroxide,22 but the cyano derivative (I-CN-Tyr) was prepared with N-iodosuccinimide.23 The transition-state analogue (N-MePyr) was generated by a similar Negishi cross-coupling using iodomethoxypyridine and N-Boc-3-iodo-l-alanine methyl ester.21
Reconstitution of HsIYD with FMN Analogues.
Cofactor exchange was performed as described previously.20 Briefly, a SUMO–HsIYD fusion was expressed and bound to a Ni-affinity column where it was treated with increasing concentrations of guanidinium hydrochloride from 0.6 to 1.5 M in 100 mM NaCl, 50 mM sodium phosphate pH 7.4, 10% glycerol, 20 mM imidazole, and 0.05 mM tris(2-carboxyethyl)-phosphine (TCEP) to release FMN. The gradient of guanidinium hydrochloride was then reversed and the column was equilibrated with 500 mM NaCl, 50 mM sodium phosphate pH 7.9, 10% glycerol, and 0.05 mM TCEP. The column was next washed with the FMN analogue of choice (200 μM) in 100 mM NaCl, 50 mM sodium phosphate pH 7.4, 10% glycerol, and 0.05 mM TCEP for 4 h at 1 mL/min (4 °C). Excess cofactor was removed by an additional wash with the same buffer lacking the FMN analogue. Finally, the reconstituted SUMO–HsIYD fusion was eluted by addition of 250 mM imidazole in 100 mM NaCl, 50 mM sodium phosphate pH 7.4, 10% glycerol, 20 mM imidazole, and 0.05 mM TCEP. The SUMO tag was removed by proteolysis and HsIYD was isolated by size exclusion chromatography.12
Catalytic Deiodination.
The product of deiodination was monitored by reverse-phase C18 HPLC after incubating IYD (70–80 nM final) in 111 mM potassium phosphate pH 7.4 (900 μL) with 5% sodium dithionite in 5% sodium bicarbonate (100 μL) at 25 °C as described previously.17
Binding Affinity to IYD Containing Oxidized FMN (FMNox).
Ligand was titrated into a solution of IYD (3 μM) in potassium phosphate (110 mM, pH 7.4) with gentle stirring (25 °C). Fluorescence (λex = 450 nm, λem = 516 nm) was measured 2 min after each addition. Dissociation constants were calculated by plotting remaining fluorescence relative to the initial fluorescence () with respect to total enzyme () and total ligand () using eq 1 (Origin 9.1) as described previously.24
(1) |
Reduction of Reconstituted HsIYD Using Xanthine/Xanthine Oxidase.
All assays were performed at 25 °C using xanthine/xanthin e oxidase as described previously.12,25,26 Briefly, a solution of 1 mM xanthine, 12 μM methyl viologen, 0.5 mM F-Tyr, and 110 mM potassium phosphate pH 7.4 was placed in a sealable quartz cuvette. Molecular oxygen was purged with argon for at least 20 min prior and 2 min subsequent to addition of HsIYD (20 μM). The headspace was then purged with argon for 30 min and reduction was initiated by addition of xanthine oxidase (140 μg/mL). The UV/vis spectral changes were recorded at room temperature every 2 min for over 4 h.
RESULTS AND DISCUSSION
Proton Transfer as a Potential Rate-Determining Step in IYD Catalysis.
Protonation of the aromatic ring of I-Tyr is an obvious process that can limit turnover since the loss of aromaticity is highly endergonic even for the electron-rich systems such as a phenol. Since no proton donor has yet been identified in the active sites of various IYDs,7,9,12,27 site-directed mutagenesis is ineffective for examining such protonation. However, the contribution of proton transfer to the overall rate of turnover can be measured by solvent kinetic isotope effects irrespective of the specific donor. IYD from H. hydrossis (HhIYD) was chosen for this analysis since it is both robust and easily purified.7 As usual, excess dithionite was used to reduce IYD since its own native reductase has not yet been identified.11 Reduction by dithionite does not limit turnover based on the similarity of the kcat/Km and kox values described above.16
A decrease in turnover is expected after replacing all exchangeable protons with deuterons if proton/deuteron transfer contributes to the overall rate of catalysis. Under neutral conditions used for the standard assays (pL 7.4), the effect of substituting H2O with D2O was quite modest on kcat and generated a solvent kinetic isotope effect (DODkcat) of 1.6 (Table 1 and Figure S1). For context, proton transfer was not thought to be rate limiting based on a DODkcat value of 1.7 for tyrosine phenol lyase that requires a carbon protonation most similar to that proposed for IYD (Scheme 2).15 Alternatively, a corresponding value of 2.3 for DODkox was considered sufficient to indicate that proton transfer contributes to the rate of flavin oxidation in a monooxygenase SidA.28 An exchange of H2O with D2O may also affect catalysis by increasing solvent viscosity and concomitantly decreasing rates of diffusion and conformational dynamics in proteins. The viscosity of D2O is equivalent to an aqueous solution of 9% glycerol.29,30 To test if the minimal effect of D2O on HhIYD was based on viscosity rather than proton transfer, steady-state kinetics were repeated in presence of 9% glycerol (Table 1 and Figure S1). The response to D2O and 9% glycerol were experimentally equivalent and thus proton transfer appears to contribute little to the overall rate of dehalogenation by HhIYD under neutral conditions. The limited effects of viscosity are not easily interpreted since only kcat, and not kcat/Km, appears sensitive to the glycerol. This result may reflect a viscosity-dependent partitioning of intermediates during catalysis, but its origins will remain ambiguous until extensive studies can sort through the many subtleties of IYD conformation and dynamics that may respond to solvent. A common explanation based on rate-determining product release would not be consistent with the similarity between steady-state rates of catalysis and the second-order rate of FMNhq oxidation by I-Tyr that is independent of product release.16,31
Table 1.
Characterization of I-Tyr Turnover by HhIYD in the Alternative Presence of D2O and Glycerola
conditionsb (pL 7.4) | kcat (s−1) | Km μM) | kcat/Km (M−1 s−1) × 102 |
---|---|---|---|
H2O | 0.64 ± 0.08 | 23 ± 4 | 280 ± 60 |
D2O | 0.41 ± 0.05 | 12 ± 4 | 340 ± 120 |
H2O, 9% glycerol | 0.39 ± 0.05 | 13 ± 4 | 300 ± 100 |
Data represent an average of two independent measurements and error corresponds to the range of these measurements. For details, see Figure S1.
pL maintained by 100 mM potassium phosphate.
A caveat associated with solvent isotope studies arises from the potential for functional group ionization to change after replacing H2O with D2O and hence measurements are typically performed under conditions in which catalysis is independent of pH.32 A pH of 7.4 has been the standard for the large majority of IYD assays and was consequently selected for the initial studies above. However, a pH rate profile of diiodotyrosine (I2-Tyr) deiodination by human IYD (HsIYD) previously indicated a pH sensitivity under these conditions.12 HhIYD was then subject to an equivalent examination over pH values from 5.5 to 8.5 (Figure 1). The responses of HhIYD and HsIYD are very similar, and both become relatively independent of pH at 6.0 and below. A fit of data from Figure 1 to eq 2 suggests that maximum turnover requires protonation of a group with an apparent pKa between 7.6 and 7.9 based on the response of V and V/K. The similarity in the pKa values from these plots is consistent with a common pH-dependent process controlling the two kinetic parameters. This is typically interpreted as a proton-dependent step in the chemical transformation, rather than binding, of substrate. A corresponding pKa of 7.5–7.6 was previously determined for HsIYD.12 Together, these values are similar to the pKa of the α-ammonium groups of I-Tyr (7.9) and I2-Tyr (7.8).33 A dependence on these groups is easily rationalized by their coordination to both a fully conserved Glu in the active site lid and the C(4) carbonyl of flavin. While the α-ammonium is not likely responsible for protonating the phenolic ring, it has the potential to influence the chemistry of the flavin very strongly.34,35
(2) |
Figure 1.
pH dependence of HhIYD catalysis. The pH profile of (A) log V and (B) log(V/K) were fit by nonlinear least-square regression using eq 2. Reductive deiodination with 0.5% dithionite was measured using standard protocols in the presence of 2-(N-morpholino)-ethanesulfonic acid (100 mM, MES) (■), potassium phosphate (100 mM) (○), and Tris-HCl (100 mM) (▲) at the indicated pH values (Figure S2). Error bars represent the range of values of two independent measurements.
Solvent deuterium isotope effects were repeated at a pL of 6.0 to check turnover under conditions that are independent of pL. In this case, a solvent kinetic isotope effect (DODkcat) of 2.3 was observed and represents a perturbation well beyond the 14% suppression of kcat caused by an increase in viscosity by 9% glycerol at pL of 6.0 (Figure S3). However, kcat/Km values measured at this pL in D2O, H2O, and 9% aqueous glycerol differed by no more than 12% from an average value of 250 × 102 M−1s−1. Thus, the overall energetics for one complete cycle of IYD catalysis remains relatively unperturbed by the change of solvent. In contrast, the sensitivity of kcat may suggest the presence of an internal partitioning that is altered by D2O. The different responses at pL of 6.0 and 7.4 could result from a change in the balance of processes contributing to the multistep catalysis of reductive dehalogenation promoted by HhIYD. Ultimately, the relative significance of the solvent effects depends on the complementary effects of FMN redox potential and substrate electronics as described below. These further investigations returned to the standard pH of 7.4 to evaluate each parameter when solvent isotope effects are minimal.
Electron Transfer as a Potential Rate-Determining Step in IYD Catalysis Using FMN Analogues.
Substitution of the methyl groups on the flavin C7 and C8 influences its redox potential as well as the basicity and nucleophilicity of its N5 group as predicted by inductive effects (Scheme 1).19,36-38 Linear free-energy correlations between such perturbations and enzyme turnover typically reveal key mechanistic constraints limiting catalysis.39,40 If electron transfer from reduced flavin to the protonated substrate was rate determining, then catalytic efficiency should increase as the reducing power increases.36 Accordingly, the 7-chloro- and 8-methoxy-desmethylFMN (7-Cl-FMN, 8-MeO-FMN) were prepared by literature procedures19,41 and used to reconstitute HhIYD with an on-column method described previously (Supporting Information).20 7-Cl-FMN and 8-MeO-FMN were selected as representatives with higher and lower potentials relative to the native FMN (Table 2). Unfortunately, HhIYD was not amenable to the reconstitution protocol optimized for HsIYD. Various concentrations of guanidinium HCl and KBr were later examined but none reconstituted HhIYD in reasonable yields. A switch to HsIYD seemed reasonable since HhIYD and HsIYD share similar substrate specificities, pH profiles, and active site structures.7,8,12 Additionally, HsIYD has been reconstituted with cofactor occupancies of 100% and enzyme recoveries of 97%.20 Steady-state characterization of these enzymes by I-Tyr dehalogenation indicated that catalytic efficiency (kcat/Km) remained greatest for IYD reconstituted with the native FMN (Table 2). In contrast, the kcat value did trend with reducing power and an enhancement of 2.8-fold was observed between analogues that differ by 132 mV. Electron transfer may then control the dehalogenation chemistry in part but the response is mild compared to the precedence of bacterial luciferase that induced a 6.6-fold acceleration of rate with a change in reduction potential of 108 mV44 and adrenodoxin reductase that induced a 35-fold acceleration with a change in potential of 136 mV.36
Table 2.
Steady-State Dehalogenation of I-Tyr by HsIYD Reconstituted with FMN and Its Analoguesa
cofactor | redox potential (mV)b | kcat (s−1 × 10−2) | Km (μM) | kcat/Km (M−1 × s−1) × 102 | pKae (FMNhq N5) |
---|---|---|---|---|---|
8-MeO-FMN | −260c | 9.0 ± 0.3 | 37 ± 3 | 25 ± 2 | 5.32 |
FMN | −205d | 5.2 ± 0.2 | 10 ± 1 | 50 ± 7 | 5.17 |
7-Cl-FMN | −128d | 3.3 ± 0.2 | 23 ± 2 | 15 ± 0.2 | 3.42 |
Data represent an average of two independent measurements under standard assay conditions (pH 7.4) and the error represents their range (see Figure S4 for details).
Redox potentials are reported for these derivatives when free in neutral aqueous solution at 25 °C.
From reference 42.
From reference 43.
From reference 19.
Insights Provided by FMN Analogues.
Additional FMN analogues were not pursued based on the weak correlation between catalytic efficiency and redox potential, but the properties of HsIYD alternatively containing 7-Cl-FMN and 8-MeO-FMN were examined for their ability to form their one-electron-reduced flavin semiquinone (FMNsq). This species does not form in detectable quantities in the native HsIYD containing FMN in the absence of a halotyrosine derivative bound in its active site.8,12 In presence of the inert substrate analogue fluorotyrosine, however, FMNsq accumulates as an intermediate during reductive titration of HsIYD and this feature correlates with the dehalogenase activity of the flavoprotein.8,11,45 No information was found in the literature by these authors on 7-Cl-FMNsq and 8-MeO-FMNsq and thus their ability to form the one-electron-reduced species was confirmed by redox titration of the reconstituted HsIYDs using xanthine and xanthine oxidase in the presence of 3-fluorotyrosine (F-Tyr, Figure 2). HsIYD containing 7-Cl-FMN produced a long-wavelength absorbance between 550 and 650 nm that is reminiscent of the absorbance of the neutral FMNsq of native HsIYD.12 Equivalent reduction of HsIYD containing 8-MeO-FMNsq generated a related absorbance at a shorter wavelength (500–600 nm) with respect to those of 7-Cl-FMNsq and FMNsq. Stabilization of these semiquinone states is consistent with the ability of the reconstituted HsIYD to support the dehalogenation described above. Otherwise, significant destabilization of the semiquinone intermediate would have severely limited dehalogenation as observed previously with 5-deazaFMN.20
Figure 2.
Formation of FMNsq during reduction of IYD. HsIYD reconstituted alternatively with (A) 7-Cl-FMN and (B) 8-MeO-FMN was then reduced under anaerobic conditions by xanthine and xanthine oxidase in the presence of 0.5 mM F-Tyr over 2 h.12,25,26 Titrations began with the oxidized cofactors indicated by the FMNox designation.
Further analysis of the reconstituted IYDs provided additional mechanistic information beyond the redox steps of catalysis. Previously, the acid/base chemistry of FMN had been shown to control the efficiency of type II isopentenyl diphosphate:dimethylallyl diphosphate (IPP:DMAPP) isomerase through a correlation between the estimated pKa of N5 and catalytic efficiency.19 The acidic form of the N5 position within a reduced flavin has the potential to donate a proton to the halogenated carbon of I-Tyr since these atoms are separated by only ~4 Å in the crystal of I-Tyr and HsIYD containing oxidized FMN (FMNox) (Figure 3).12 No other residues within a 5 Å radius are better positioned for proton transfer. However, the rate of dehalogenation is not enhanced by increasing the estimated acidity of the N5 by substituting FMN with 7-Cl-FMN (Table 2). Hence, the source of a proton for I-Tyr activation remains to be determined and does not likely involve FMN. Flavoproteins within the same structural superfamily appear to rely on solvent as a proton source during reduction of nitroaromatic substrates.14 Currently, this is the most reasonable source for IYD as well.
Figure 3.
Amino acid residues within 5 Å of the site of I-Tyr protonation in HsIYD based on PDB 4TTC.12 Carbon atoms of I-Tyr, FMNox, and amino acid side chains are rendered in light magenta, yellow, and light cyan, respectively.
The aromatic ring of halotyrosine stacks intimately with the isoalloxazine ring of FMNox in all crystal structures characterized to date (Figure 3). Similar stacking is expected in the active complex with FMNhq and this is inferred by the tight and competitive binding of an N-methylpyridone derivative (N-MePyr) that mimics the proposed protonated substrate intermediate (Scheme 1 and Table 3).46 The FMN analogues provided an opportunity to correlate this stacking between the electron-rich FMNhq species and the electron-poor intermediate analogue. Competitive inhibition of I-Tyr dehalogenation was compared for HsIYD reconstituted with 7-Cl-FMN, 8-MeO-FMN, and native FMN (Table 3 and Figure S6). The pyridone derivative strongly inhibited all of the reconstituted HsIYDs as expected from a prior examination of IYD that had been extracted from porcine thyroids.46 Additionally, the binding affinity as determined by Ki was enhanced by ~60% when the electron-poor 7-Cl-FMN was replaced by the electron-rich 8-MeO-FMN in the HsIYD active site. Thus, FMN likely participates in substrate binding and orientation as well as serving as a donor of electrons. The significance of the π stacking is also evident from the ability of 2-iodophenol to adopt a similar orientation as the corresponding atoms of I-Tyr above FMN in a crystal structure of HhIYD.7 Still, the overall contribution of this noncovalent association is not easily deciphered since the trend of Ki does not match the trend of the Km values for I-Tyr dehalogenation (Tables 2 and 3). Because neither proton nor electron transfer strongly controls turnover of IYD, attention was directed back to the most obvious high-energy species proposed in catalysis, the dearomatized substrate intermediate (Scheme 1).
Table 3.
Competitive Binding of N-MePyr to Reconstituted HsIYDsa
![]() |
Cofactor | KI (nM) |
8-MeO-FMN | 121 ± 17 | |
FMN | 163 ±4 | |
7-Cl-FMN | 198 ± 2 |
KI values represent an average of two independent trials and the error represents their range (see Figure S6 for details).
Correlation between Dehalogenation Efficiency and Substrate Electronics.
As a complement to the use of flavin analogues above, a series of substrate analogues were designed to perturb the electronic properties of I-Tyr and consequently those of its intermediates formed during catalysis. Electron-donating substituents ortho to the phenol of I-Tyr have the potential to stabilize the dearomatized, electron-deficient intermediate while concurrently diminishing its propensity for accepting an electron from FMNhq (Scheme 1). Electron-withdrawing groups induce the reciprocal response to dearomatization and electron transfer. Accordingly, substrates alternatively containing −OMe, −Me, −F, and −CN were prepared by iodination of their tyrosine precursors that were generated from a Negishi cross-coupling between an iodoalanine derivative and the appropriate para-iodophenol (see the Supporting Information). The consequences of these substituents on binding to IYD containing FMNox were conveniently surveyed by the ability of bound substrate to quench flavin fluorescence.24,47 The native substrate I-Tyr bound most tightly to HsIYD, but all of the analogues bound with Kd values in the low μM range (Table 4). No trend based on electronics was apparent and the greatest influence could be the size of the substituent because substrates containing the −OMe and −CN bound somewhat weaker than those containing −Me, −F, and −H. In vivo, HsIYD is thought to process both I-Tyr and I2-Tyr and hence tolerance to the 3,5-substituted tyrosines had been anticipated.11 Previously, the Kd of I2-Tyr with HsIYD was determined to be similar to that of Me-I-Tyr and one order of magnitude larger than that for I-Tyr.12
Table 4.
Substituent Effects on Binding and Deiodination with HsIYDa
Steady-state kinetic characterization of these tyrosine derivatives reveals that the native substrate I-Tyr exhibits the lowest Km value, but otherwise there is no obvious trend between ortho substitution and Km (Table 4 and Figure 4A). In contrast, both kcat and kcat/Km reveal a clear dependence on substrate properties. Electron-deficient substrates were processed much less efficiently than the electron-rich substrates by a range of 38-fold as measured by kcat and 29-fold by kcat/Km. This trend suggests the importance of stabilizing the C-protonated intermediate proposed in the mechanism over a dependence on electron transfer (Scheme 1). Additionally, the substituent effects fit well to a linear free-energy correlation using σm (Figure 4B).48 This parameter was selected based on the placement of the substituents meta to the halogen-bearing carbon. The sensitivity factor (ρ) derived from the kcat/Km values is −2.2 ± 0.3 with an R2 value of 0.934. Similar treatment of the kcat values provides a ρ of −2.4 ± 0.5 with an R2 of 0.851 (Figure S11). A negative value is consistent with an electron-deficient species in a rate-controlling process. The magnitude of this ρ is similar to that used to confirm the generation of an electron-deficient species during key steps in turnover of two flavin-dependent enzymes, UDP-galactose mutase and IPP:DMAPP isomerase.19,38
Figure 4.
Substituent effects on catalytic deiodination by HsIYD. (A) Concentration dependence of deiodination for the indicated substrates in 100 mM potassium phosphate pH 7.4 and 0.5% sodium dithionite. Each data point represents an average of two individual observations and error bars represent their range. Kinetic parameters were determined from the best fit of data (solid lines) to Michaelis–Menten kinetics using Origin 9. (B) Linear free-energy correlation between turnover and substituent constant σm.48 The sensitivity to substituents ρ is defined as the slope of the linear best fit to the data.
The 29- to 38-fold effect on catalysis created by perturbing the substrate electronics is significantly more pronounced than the effects of the FMN reducing potential and solvent properties (viscosity and deuteration). To test the generality of this major determinant, binding and turnover of the same set of substrates were repeated with the bacterial HhIYD used in the initial studies. This enzyme exhibited even less variation than HsIYD in its affinity (Kd) for the substrates, and binding of I-Tyr to HhIYD was among the weakest variant examined (Table 5). These values should be considered most useful for comparing relative affinities since they reflect binding to the oxidized (FMNox) state of IYD while catalysis involves the reduced state containing FMNhq. Association of I-Tyr and other substrates to the reduced enzyme is not easily detected due to the immediate reoxidation of FMNhq as part of the standard turnover. However, the affinity of an inert analogue, fluorotyrosine, to the reduced enzyme was estimated from its competitive inhibition of I-Tyr dehalogenation by HhIYD. Its Ki (10.7 ± 0.1 μM, Figure S8) is within the same order of magnitude as its Kd (46 ± 2 μM) measured previously with HhIYD containing FMNox.7
Table 5.
Substituent Effects on Binding and Deiodination of I-Tyr with HhIYDa
For HhIYD, I-Tyr may not be a native substrate since the physiological role of IYD has not yet been well documented in organisms other than Chordata.11,49 However, at least one of the halotyrosines is likely to be physiologically important because tyrosine derivatives in general are bound and processed by HhIYD much more efficiently than their simple halophenol counterparts.7 Dehalogenation of the substrates by HhIYD established the same trend as that observed for HsIYD (Figure S9). The electron-rich I-Me-Tyr and I-MeO-Tyr were processed most efficiently, and the electron-poor I-F-Tyr and I-CN-Tyr were processed least efficiently. Sensitivity to these substituents (ρ = −2.8 ± 0.9) was experimentally equivalent to that of HsIYD, although the linear correlation was weaker with an R2 of 0.777 (Figure S10). The kcat values for HhIYD also responded similarly with a ρ of −2.1 ± 0.2 and an R2 of 0.867 (Figure S11). Thus, substrate electronics dominate the efficiency of turnover for both the human and bacterial enzymes.
The substituents on I-Tyr also control the acidity of the phenolic proton and thus affect the balance of the protonated and deprotonated substrate under the assay conditions of pH 7.4. The phenolate form is thought to bind with significantly more affinity to HsIYD than the phenol form and this preference has been invoked to explain Kd values as low as 0.2 μM for I-Tyr but greater than 1000 μM for Tyr.12 Association of the phenolate anion to the active site of IYD is likely stabilized by hydrogen bonding offered by the protein backbone and the 2′-hydroxy group of FMN.11,12 If the phenolate form is then the true substrate (Scheme 1), Kd values calculated for the total substrate may not reflect the true affinity of the phenolate forms of the ligands. More importantly for this study, the steady-state rate constants and their response to the substituents may need a re-evaluation based on the approximate concentrations of phenolate rather than the total substrate. At the two extremes, 98% of I-CN-Tyr should exist in its phenolate form based on an estimated pKa of 5.7, whereas only 3.4% of I-Me-Tyr should exist in its phenolate form based on an estimated pKa of 8.9 (Table S1). Under this scenario, the differences between their kcat/Km values would increase to almost 1000-fold. The actual state of protonation within the active site is not yet known, so this analysis may exaggerate the power of the substituents but at least the trend remains consistent whether the total substrate or only its phenolate form is considered when calculating the steady-state constants.
CONCLUSIONS
The dominant effect of substituents on I-Tyr dehalogenation by human and bacterial IYD most likely arises from the varying stability and steady-state concentrations of the dearomatized intermediate that subsequently accepts an electron from FMNhq before halide elimination. This sensitivity to substrate electronics would not have been observed if enzyme reduction by dithionite or halide release had controlled turnover. Contrary to expectations, neither electron nor proton transfer strongly control the efficiency of dehalogenation under neutral conditions. The weak dependence on reduction potential indicates electron transfer only moderately contributes to the overall rate of catalysis and the subsequent lysis of the carbon–halogen bond was already shown to be relatively fast based on the nearly equivalent rates of deiodination and debromination.16,17
Proton transfer associated with dearomatization of the substrate is similarly not rate limiting since solvent deuterium isotope effects are minimal under neutral conditions. However, this transfer within an internal partitioning of intermediates may begin to limit catalysis under acidic conditions (pL ≤ 6.0) that supports turnover efficiency greater than that at neutral conditions. The proton source remains to be determined and does not likely involve the reduced FMNhq since an inverse correlation was observed between the acidity of its N5 proton and the kcat of I-Tyr dehalogenation. This contrasts the behavior of another flavin-dependent enzyme, IPP:DMAPP isomerase, that does utilize the N5 proton.19 An inverse correlation was evident for the acidity of the substrate phenol and turnover of HsIYD and HhIYD. Thus, the phenol is also not a likely candidate for proton transfer particularly since the substrate may bind in the phenolate form without its proton. In addition, the pH rate profile does not reflect the pKa of the phenol and the pH response of HhIYD with I-Tyr is equivalent to that of HsIYD with I2-Tyr despite the difference in the phenol acidities of 2 pKa units.12
The largest variable in catalysis is the steady-state concentration of the electrophilic intermediate as evident from the strong correlation between rate and substrate substituents. A very similar conclusion was reached after extensive analysis of tyrosine phenol lyase, a PLP-dependent enzyme that also requires proton transfer to dearomatize its tyrosine substrate before bond lysis (Scheme 2).15,50 Proton transfer did not control this later enzyme as much as stabilization of the dearomatized intermediate. Future efforts to expand the scope of substrates for IYD may now focus on electron-rich halophenols and active site mutations that stabilize the electron-deficient intermediate of substrates lacking the zwitterion of halotyrosine.
Supplementary Material
ACKNOWLEDGMENTS
This research was supported in part by the National Institute of General Medicine (RO1 GM130937 to S.E.R. and T32 GM080189 to J.D.)
Footnotes
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.biochem.3c00041.
The Supporting Information is available free of charge on the ACS Publication website.
Experimental methods, synthetic protocols for FMN and substrate analogues, and Figures S1-S12 detailing binding and catalytic characterization of HhIYD and HsIYD (PDF)
The authors declare no competing financial interest.
Contributor Information
Anton Kozyryev, Department of Chemistry, Johns Hopkins University, Baltimore, Maryland 21218, United States.
Daniel Lemen, Department of Chemistry, Johns Hopkins University, Baltimore, Maryland 21218, United States.
Jessica Dunn, Chemistry Biology Interface Graduate Program, Johns Hopkins University, Baltimore, Maryland 21218, United States.
Steven E. Rokita, Department of Chemistry, Johns Hopkins University, Baltimore, Maryland 21218, United States; Chemistry Biology Interface Graduate Program, Johns Hopkins University, Baltimore, Maryland 21218, United States
REFERENCES
- (1).Zhai H; Zhang X; Zhu X; Liu J; Ji M Formation of Brominated Disinfection Byproducts During Chloramination of Drinking Water: New Polar Species and Overall Kinetics. Environ. Sci. Technol 2014, 48, 2579–2588. [DOI] [PubMed] [Google Scholar]
- (2).Sutton NB; Atashgahi S; Saccenti E; Grotenhuis T; Smidt H; Rijnaarts HHM Microbial Community Response of an Organohalide Respiring Enrichment Culture to Permanganate Oxidation. PLoS One 2015, 10, No. e0134615. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (3).Huang C-H; Ren F-R; Shan G-Q; Qin H; Mao L; Zhu B-Z Molecular Mechanism of Metal-Independent Decomposition of Organic Hydroperoxides by Halogenated Quinoid Carcinogens and the Potential Biological Implications. Chem. Res. Toxicol 2015, 28, 831–837. [DOI] [PubMed] [Google Scholar]
- (4).Ang T-F; Maiangwa J; Salleh AB; Normi YM; Leow TC Dehalogenases: From Improved Performance to Potential Microbial Dehalogenation Applications. Molecules 2018, 23, 1100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (5).Pimviriyakul P; Chaiyen PA Complete Bioconversion Cascade for Dehalogenation and Denitration by Bacterial Flavin-Dependent Systems. J. Biol. Chem 2018, 293, 18525–18539. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (6).Mikkonen A; Yläranta K; Tiirola M; Dutra LAL; Salmi P; Romantschuk M; Copley S; Ikäheimo J; Sinkkonen A Successful Aerobic Bioremediation of Groundwater Contaminated with Higher Chlorinated Phenols by Indigenous Degrader Bacteria. Water Res. 2018, 138, 118–128. [DOI] [PubMed] [Google Scholar]
- (7).Ingavat N; Kavran JM; Sun Z; Rokita SE Active Site Binding Is Not Sufficient for Reductive Deiodination by Iodotyrosine Deiodinase. Biochemistry 2017, 56, 1130–1139. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (8).Hu J; Su Q; Schlessman JL; Rokita SE Redox Control of Iodotyrosine Deiodinase. Protein Sci. 2019, 28, 68–78. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (9).Sun Z; Xu B; Spisak S; Kavran JM; Rokita SE The Minimal Structure for Iodotyrosine Deiodinase Function Is Defined by an Outlier Protein from Thermophilic Bacteria Thermotoga Neapolitana. J. Biol. Chem 2021, 297, No. 101385. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (10).Sun Z; Rokita SE Towards a Halophenol Dehalogenase from Iodotyrosine Deiodinase Via Computational Design. ACS Catal. 2018, 8, 11783–11793. [Google Scholar]
- (11).Sun Z; Su Q; Rokita SE The Distribution and Mechanism of Iodotyrosine Deiodinase Defied Expectations. Arch. Biochem. Biophys 2017, 632, 77–87. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (12).Hu J; Chuenchor W; Rokita SE A Switch between One- and Two-Electron Chemistry of the Human Flavoprotein Iodotyrosine Deiodinase Is Controlled by Substrate. J. Biol. Chem 2015, 290, 590–600. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (13).Akiva E; Copp JN; Tokuriki N; Babbit PC Evolutionary and Molecular Foundations of Multiple Contemporary Functions of the Nitroreductase Superfamily. Proc. Natl. Acad. Sci. U.S.A 2017, 114, E9549–E9558. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (14).Pitsawong W; Haynes CA; Koder RL; Rodgers DW; Miller A-F Mechanism-Informed Refinement Reveals Altered Substrate-Binding Mode for Catalytically Competent Nitroreductase. Structure 2017, 25, 978–987. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (15).Faleev NG; Spirina SN; Demidkina TV; Phillips RS Mechanism of Catalysis by Tyrosine Phenol Lyase from Erwinia Herbicol. Multiple Kinetic Isotope Effets for the Reaction with Adequate Substrates. JCS Perk. Trans II 1996, 2001–2004. [Google Scholar]
- (16).Bobyk KD; Ballou DP; Rokita SE Rapid Kinetics of Dehalogenation Promoted by Iodotyrosine Deiodinase from Human Thyroid. Biochemistry 2015, 54, 4487–4494. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (17).Phatarphekar A; Rokita SE Functional Analysis of Iodotyrosine Deiodinase from Drosophila Melanogaster. Protein. Sci 2016, 25, 2187–2195. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (18).McMillen DF; Golden DM Hydrocarbon Bond Dissociation Energies. Annu. Rev. Phys. Chem 1982, 33, 493–532. [Google Scholar]
- (19).Thibodeaux CJ; Chang W.-c.; Liu H.-w. Linear Free Energy Relationships Demonstrate a Catalytic Role for the Flavin Mononucleotide Coenzyme of the Type II Isopentenyl Diphosphate: Dimethylally Diphosphate Isomerase. J. Am. Chem. Soc 2010, 132, 9994–9996. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (20).Su Q; Boucher PA; Rokita SE Conversion of a Dehalogenase to a Nitroreductase by Swapping Its Flavin Cofactor with a 5-Deazaflavin Analog. Angew. Chem., Int. Ed 2017, 56, 10862–10866. [DOI] [PubMed] [Google Scholar]
- (21).Ross AJ; Lang HL; Jackson RFW Much Improved Conditions for the Negishi Cross-Coupling of Iodoalanine Derived Zinc Reagents with Aryl Halides. J. Org. Chem 2010, 75, 245–248. [DOI] [PubMed] [Google Scholar]
- (22).Cochrane JR; White JM; Wille U; Hutton CA Total Synthesis of Mycocyclosin. Org. Lett 2012, 14, 2402–2405. [DOI] [PubMed] [Google Scholar]
- (23).Bergström M; Suresh G; Naidu VR; Unelius CR N-Iodosuccinimide (NIS) in Direct Aromatic Iodination. Eur. J. Org. Chem 2017, 2017, 3234–3239. [Google Scholar]
- (24).Warner JR; Copley SD Pre-Steady-State Kinetic Studies of the Reductive Dehalogenation Catalyzed by Tetrachlorohydroquinone Dehalogenase. Biochemistry 2007, 46, 13211–13222. [DOI] [PubMed] [Google Scholar]
- (25).Massey V Flavins Flavoproteins Proceedings of the International Symposium; Curti B; Ronchi S; Zanetti G, Eds.; Gruyter & Co.: Berlin, 1991; pp 59–66. [Google Scholar]
- (26).van den Heuvel RHH; Fraaije MW; Van Berkel WJH Redox Properties of Vanillyl-Alcohol Oxidase. Methods Enzymol. 2002, 353, 177–186. [DOI] [PubMed] [Google Scholar]
- (27).Thomas SR; McTamney PM; Adler JM; LaRonde-LeBlanc N; Rokita SE Crystal Structure of Iodotyrosine Deiodinase, a Novel Flavoprotein Responsible for Iodide Salvage in Thyroid Glands. J. Biol. Chem 2009, 284, 19659–19667. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (28).Robinson RM; Klancher CA; Rodriquez PJ; Sobrado P Flavin Oxidation in Flavin-Dependent N-Monooxygenases. Protein Sci. 2019, 28, 90–99. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (29).Hardy RC; Cottington RL Viscosity of Deuterium Oxide and Water in the Range 5° to 125 °C. J. Res. Natl. Bur. Stand 1949, 42, 573–578. [Google Scholar]
- (30).Rahman MQ; Chuah K-S; Macdonald ECA; Trusler JPM; Ramaesh K The Effect of pH, Dilution, and Temperature on the Viscosity of Ocular Lubricants–Shift in Rheological Parameters and Potential Clinical Significance. Eye 2012, 26, 1579–1584. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (31).Gadda G; Sobrado P Kinetic Solvent Viscosity Effects as Probes for Studying the Mechanisms of Enzyme Action. Biochemistry 2018, 57, 3445–3453. [DOI] [PubMed] [Google Scholar]
- (32).Schowen BK; Schowen RL Solvent Isotope Effects on Enzyme Systems. In Methods in Enzymology 1982; Vol. 87, pp 551–606. [PubMed] [Google Scholar]
- (33).Jorgensen EC Thyroid Hormones; Li CH, Ed.; Academic Press: New York, 1978; pp 57–107. [Google Scholar]
- (34).Tegoni M; Gervais M; Desbois A Resonance Raman Study on the Oxidized and Anionic Semiquinone Forms of Flavocytochrome B2 and L-Lactate Monooxygenase. Influence of the Structure and Environment of the Isoalloxazine Ring on the Flavin Function. Biochemistry 1997, 36, 8932–8946. [DOI] [PubMed] [Google Scholar]
- (35).Xu D; Kohli RM; Massey V The Role of Threonine 37 in Flavin Reactivity of the Old Yellow Enzyme. Proc. Natl. Acad. Sci 1999, 96, 3556–3561. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (36).Light DR; Walsh CT Flavin Analogs as Mechanistic Probes of Adrenodoxin Reductase-Dependent Electron Transfer to the Cholesterol Side Chain Cleavage Cytochrome P-450 of the Adrenal Cortex. J. Biol. Chem 1980, 255, 4264–4277. [PubMed] [Google Scholar]
- (37).Hasford JJ; Rizzo CJ Linear Free Energy Substituent Effects on Flavin Redox Chemistry. J. Am. Chem. Soc 1998, 120, 2251–2255. [Google Scholar]
- (38).Sun HG; Ruszczycky MW; Chang W.-c.; Thibodeaux CJ; Liu H.-w. Nucleophilic Participation of Reduced Flavin Coenzyme in Mechanism of UDP-Galactopyranose Mutase. J. Biol. Chem 2012, 287, 4602–4608. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (39).Edmonson D; Ghisla S Flavoenzyme Structure and Function. Approaches Using Flavin Analogues. In Flavoprotein Protocols. Methods Molecular Biology; Humana Press, 1999; Vol. 131, pp 157–179. [DOI] [PubMed] [Google Scholar]
- (40).Thibodeaux CJ; Chang W.-c.; Liu H.-w. Unraveling Flavoenzyme Reaction Mechanisms Using Flavin Analogues and Linear Free Energy Relationships. In Methods in Enzymology; Elsevier, 2019; Vol. 620, pp 167–188. [DOI] [PubMed] [Google Scholar]
- (41).Mansurova M; Koay MS; Gärtner W Synthesis and Electrochemical Properties of Structurally Modified Flavin Compounds. Eur. J. Org. Chem 2008, 2008, 5401–5406. [Google Scholar]
- (42).Macheroux P; Bornemann S; Ghisla S; Thorneley RNF Studies with Flavin Analogs Provide Evidence That a Protonated Reduced FMN Is the Substrate-Induced Transient Intermediate in the Reaction of Escherichia Coli Chorismate Synthase. J. Biol. Chem 1996, 271, 25850–25858. [DOI] [PubMed] [Google Scholar]
- (43).Walsh CT; Fisher J; Spencer R; Graham DW; Ashton WT; Brown JE; Brown R; Rogers EF Chemical and Enzymatic Properties of Riboflavin Analogues. Biochemistry 1978, 17, 1942–1951. [DOI] [PubMed] [Google Scholar]
- (44).Eckstein JW; Hastings JW; Ghisla S Mechanism of Bacterial Bioluminescence: 4a,5-Dehydroflavin Analogs as Models for Luciferase Hydroperoxide Intermediates and Effect of Substituents at the 8-Position of Flavin on Luciferase Kinetics. Biochemistry 1993, 32, 404–411. [DOI] [PubMed] [Google Scholar]
- (45).Mukherjee A; Rokita SE Single Amino Acid Switch between a Flavin-Dependent Dehalogenase and Nitroreductase. J. Am. Chem. Soc 2015, 137, 15342–15345. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (46).Kunishima M; Friedman JE; Rokita SE Transition-State Stabilization by a Mammalian Reductive Dehalogenase. J. Am. Chem. Soc 1999, 121, 4722–4723. [Google Scholar]
- (47).McTamney PM; Rokita SE A Mammalian Reductive Deiodinase Has Broad Power to Dehalogenate Chlorinated and Brominated Substrates. J. Am. Chem. Soc 2009, 131, 14212–14213. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (48).Hansch C; Leo A; Taft RW A Survey of Hammett Substituent Constants and Resonance and Field Parameters. Chem. Rev 1991, 91, 165–195. [Google Scholar]
- (49).Phatarphekar A; Su Q; Eun SH; Chen X; Rokita SE The Importance of a Halotyrosine Dehalogenase for Drosophila Fertility. J. Biol. Chem 2018, 293, 10314–10321. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (50).Phillips RS; Demidkina TV; Faleev NG The Role of Substrate Strain in the Mechanism of the Carbon-Carbon Lyases. Bioorg. Chem 2014, 57, 198–205. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.