Next Article in Journal
Glycosylation Tunes Neuroserpin Physiological and Pathological Properties
Next Article in Special Issue
Insights into the FMNAT Active Site of FAD Synthase: Aromaticity Is Essential for Flavin Binding and Catalysis
Previous Article in Journal
Mechanisms of Multidrug Resistance in Cancer Chemotherapy
Previous Article in Special Issue
Riboflavin: The Health Benefits of a Forgotten Natural Vitamin
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Reactions of Plasmodium falciparum Ferredoxin:NADP+ Oxidoreductase with Redox Cycling Xenobiotics: A Mechanistic Study

by
Mindaugas Lesanavičius
1,
Alessandro Aliverti
2,
Jonas Šarlauskas
1 and
Narimantas Čėnas
1,*
1
Department of Xenobiotics Biochemistry, Institute of Biochemistry of Vilnius University, Saulėtekio 7, LT-10257 Vilnius, Lithuania
2
Department of Biosciences, Università degli Studi di Milano, via Celoria 26, I-20133 Milano, Italy
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2020, 21(9), 3234; https://doi.org/10.3390/ijms21093234
Submission received: 6 April 2020 / Revised: 29 April 2020 / Accepted: 30 April 2020 / Published: 2 May 2020
(This article belongs to the Special Issue Flavin Adenine Dinucleotide (FAD): Biosynthesis and Function)

Abstract

:
Ferredoxin:NADP+ oxidoreductase from Plasmodium falciparum (PfFNR) catalyzes the NADPH-dependent reduction of ferredoxin (PfFd), which provides redox equivalents for the biosynthesis of isoprenoids and fatty acids in the apicoplast. Like other flavin-dependent electrontransferases, PfFNR is a potential source of free radicals of quinones and other redox cycling compounds. We report here a kinetic study of the reduction of quinones, nitroaromatic compounds and aromatic N-oxides by PfFNR. We show that all these groups of compounds are reduced in a single-electron pathway, their reactivity increasing with the increase in their single-electron reduction midpoint potential (E17). The reactivity of nitroaromatics is lower than that of quinones and aromatic N-oxides, which is in line with the differences in their electron self-exchange rate constants. Quinone reduction proceeds via a ping-pong mechanism. During the reoxidation of reduced FAD by quinones, the oxidation of FADH. to FAD is the possible rate-limiting step. The calculated electron transfer distances in the reaction of PfFNR with various electron acceptors are similar to those of Anabaena FNR, thus demonstrating their similar “intrinsic” reactivity. Ferredoxin stimulated quinone- and nitro-reductase reactions of PfFNR, evidently providing an additional reduction pathway via reduced PfFd. Based on the available data, PfFNR and possibly PfFd may play a central role in the reductive activation of quinones, nitroaromatics and aromatic N-oxides in P. falciparum, contributing to their antiplasmodial action.

1. Introduction

The emergence of a malarial parasite Plasmodium falciparum resistance to available drugs, e.g., chloroquine or artemisinin ([1] and references therein), results in both a demand for new antimalarial agents and a better understanding of their mechanisms of action. P. falciparum is particularly vulnerable to oxidative stress, which might be caused by its lack of the antioxidant enzymes catalase and glutathione peroxidase [2]. For this reason, redox cycling compounds such as quinones, aromatic nitrocompounds and aromatic N-oxides, which frequently exhibit antiplasmodial in vitro activity at micromolar or lower concentrations, were a subject of great interest for a number of years ([3,4,5,6,7,8] and references therein). However, only fragmental data are available on their reactions with P. falciparum redox enzymes [6,9,10,11].
It is commonly accepted that the single-electron reduction of quinones and other classes of prooxidant compounds is performed by flavin-dependent dehydrogenases-electrontransferases such as NADPH:cytochrome P-450 reductase (P-450R), ferredoxin:NADP+ oxidoreductase (FNR) or NO-synthase (NOS) ([12,13,14] and references therein). These enzymes, working in conjunction with physiological single-electron acceptors, transform a two-electron (hydride) transfer into a single-electron one by stabilizing the neutral (blue) semiquinone form of the flavin nucleotide as the reaction intermediates [15,16,17].
In P. falciparum, an FAD-containing ferredoxin:NADP+ oxidoreductase (PfFNR, EC 1.18.1.2) is localized in a nonphotosynthetic plastid organelle called apicoplast [18,19], which performs the biosynthesis of isoprenoids and fatty acids and is essential for the parasite’s survival. PfFNR supplies redox equivalents to the apicoplast redox system via a Fe2S2-protein ferredoxin (PfFd) [18]. PfFd is characterized by a standard redox potential (E07.5) of -0.26 V and possesses about 50% amino acid sequence homology with plant ferredoxins [18]. PfFNR is characterized by E07 = −0.28 V [19]; it possesses low homology (20–30%) with plant FNRs, displaying unique large insertions and deletions [20]. The protein complex formation is attributed to the electrostatic interaction between the basic residues of PfFNR and acidic residues of PfFd and is sensitive to ionic strength [18,19,21].
PfFNR reduces quinones and nitroaromatic compounds in a single-electron way and, based on currently available data, may be considered as an important source of their radicals in P. falciparum [6,11]. In this work, we extended the studies of PfFNR using a large number of nonphysiological electron acceptors with different structures, reduction potentials and electrostatic charges. Our results provide a general insight into their reduction mechanisms and highlight the specific features of PfFNR relevant to these processes.

2. Results

2.1. Steady-State Kinetics and Substrate Specificity Studies of PfFNR

In a previous study, juglone (5-hydroxy-1,4-naphthoquinone) was identified as one of the most active nonphysiological electron acceptors of PfFNR [6]. In this work a series of parallel lines was obtained in double-reciprocal plots at varied concentrations of juglone and fixed concentrations of NADPH (Figure 1). This indicates that the quinone-reductase reaction catalyzed by PfFNR follows a “ping-pong” mechanism. As deduced from Equation (A1) (Appendix A), the kcat value for the juglone reduction at an infinite NADPH concentration is equal to 63.2 ± 4.1 s−1, and the values of the bimolecular rate constants (or catalytic efficiency constants, kcat/Km) for NADPH and juglone are equal to 6.0 ± 0.4 × 105 M−1s−1 and 1.1 ± 0.1 × 106 M−1s−1, respectively.
In order to assess the substrate specificity of PfFNR, we examined the reduction of three series of electron acceptors, namely quinones (Q), nitroaromatic compounds (ArNO2) and aromatic N-oxides (ArN→O), whose single-electron reduction midpoint potentials (E17) vary from 0.01 V to −0.575 V. In addition, several single-electron acceptors, ferricyanide, Fe(EDTA) and benzylviologen have been studied. The apparent reduction maximal rate constants, kcat(app), of electron acceptors at 100 µM NADPH, and their respective kcat/Km, are given in Table 1. The kcat values for a number of less-active oxidants were not determined because of a nearly linear dependence of the reaction rate on their concentrations.
The log kcat/Km of ArNO2 exhibits a linear dependence on their E17 (Table 1 and Figure 2). In general, the log kcat/Km values of quinones and aromatic N-oxides are higher than those of nitroaromatics and are characterized by a parabolic dependence on their E17 values (Figure 3). It is important to note that the reactivity of the single-electron acceptor benzylviologen matches the reactivity of quinones (Figure 2).
Previously, we found that PfFNR reduces quinones and nitroaromatic compounds in a single-electron way [6,11]. Here, we found that the PfFNR-catalyzed oxidation of NADPH by 100–200 µM tirapazamine was accompanied by O2 consumption at a rate close to that of NADPH oxidation. The addition of 50-µM cytochrome c to the reaction mixture resulted in its reduction at a rate representing 180–190% of that of NADPH oxidation. Superoxide dismutase (100 U/mL) inhibited the reduction of cytochrome c by 40–65%. This shows that the single-electron flux in the PfFNR-catalyzed reduction of ArN→O is equal to 90–95% and that cytochrome c is reduced by their radicals, which are under a steady state with the O2/O2−. couple.
Pyridine nucleotide analogues of NAD(P)+ are frequently used in the analysis of mechanisms of NAD(P)H-oxidizing flavoenzymes. The kcat of the transhydrogenase reaction of PfFNR, the formation of reduced 3-acetylpyridineadenine dinucleotide phosphate (AcPyPH) from AcPyP+ at the expense of NADPH, did not depend on the NADPH concentration, in the 25–200 µM range, and was equal to 3.2 ± 0.4 s−1 (Figure 3). However, in this case, the kcat/Km for the oxidant decreased with the increase in the NADPH concentration (Figure 3). This shows that NADPH acts as a competitive inhibitor to AcPyP+, occupying the pyridine nucleotide binding site of the reduced enzyme form. As deduced from Equation (A2) (Appendix A), the Kis of NADPH, describing the effect of NADPH on the slopes in the Lineweaver–Burk plots, is equal to 140 ± 20 µM, and the kcat/Km for AcPyP+ at (NADPH) = 0 is equal to 9.3 ± 0.8 × 103 M−1s−1.
The affinity of PfFNR for its physiological oxidant, PfFd, decreases with the ionic strength of the medium, due to the electrostatic character of their interaction [18,21]. In order to assess the role of electrostatic interactions in the reactions of PfFNR with nonphysiological oxidants, we examined the effects of ionic strength on their reduction rates. It was reported that the NADPH-ferricyanide reductase reaction of PfFNR was inhibited by high concentrations of ferricyanide, which acted as a competitive inhibitor with respect to NADPH (Ki = 230 µM) [20]. However, when the phosphate buffer was used instead of 0.1-M Tris-HCl [20], the substrate inhibition by ferricyanide was absent. This enabled us to perform a more thorough analysis of its reduction kinetics.
The data of Figure 4 show a bell-shape dependence of log kcat/Km for ferricyanide, Fe(EDTA) and benzylviologen on the ionic strength of the solution, irrespective of the opposite electrostatic charge of the latter oxidant. In contrast, the kcat/Km for the uncharged electron-acceptor tetramethyl-1,4-benzoquinone did not depend on the ionic strength (Figure 4).

2.2. Kinetics of PfFNR Oxidation under Multiple Turnover Conditions

In order to get insight into the enzyme reoxidation mechanism, we investigated the spectral changes of PfFNR-bound FAD during its multiple turnover under aerobic conditions in the presence of NADPH and tetramethyl-1,4-benzoquinone. This electron acceptor does not absorb light at ≥460 nm, and its semiquinone form is rapidly reoxidized by oxygen [26]. In control experiments performed in the absence of quinone, the initial fast phase of FAD reduction by NADPH monitored at 460 nm is followed by its slower reoxidation by oxygen (Figure 5A). Importantly, this is accompanied by a transient increase in absorbance at 600 nm at the same time scale (Figure 5A). The addition of quinone accelerates the reoxidation of FADH and the decay of the 600-nm absorbing species by about two orders of magnitude (Figure 5B). This shows that, under our experimental conditions, O2 plays a negligible role in the kinetics of enzyme reoxidation. This process is also accompanied by a transient increase in absorbance at 600 nm. The kinetics of reoxidation were analyzed by the method of chance [27] using Equation (1), where kox is the apparent first-order rate constant of enzyme reoxidation, (NADPH)0 is the initial NADPH concentration, (Ered)max is the maximal concentration of the reduced enzyme formed during the turnover and t1/2(off) is the time interval between the formation of the half-maximal amount of Ered and its decay to the half-maximal value:
k ox = [ NADPH ] 0 [ E red ] max · t 1 / 2 ( off )  
It was assumed that complete FAD reduction corresponds to the maximal ∆A460 after the enzyme mixing with NADPH in the absence of quinone (Figure 5A). This 460-nm absorbance change was close to that expected using the value of ∆ε460 = 7.8 mM−1cm−1 for the absorbance difference between the oxidized and two-electron reduced PfFNR [19,28,29]. For the reoxidation of PfFNR with oxygen (Figure 5A), we obtained a kox = 0.18 ± 0.02 s−1, which was close to the enzyme NADPH oxidase activity under a steady state. The dependence of kox on the tetramethyl-1,4-benzoquinone concentration (Figure 5C) gives an apparent bimolecular rate constant of 1.34 ± 0.37 × 105 M−1s−1, which is comparable with the steady-state kcat/Km for this oxidant (Table 1) and kox(max) = 155 ± 32 s−1. However, the later value may lack sufficient accuracy, because it was impossible to obtain a saturating concentration (sufficiently high kox values) due to the limited solubility of the oxidant.

2.3. NADP+ Inhibition Studies

At a fixed juglone concentration, the reaction product NADP+ acted as a competitive inhibitor towards NADH (Figure 6A) with Kis = 1.42 ± 0.13 mM, as deduced from Equation (A2) (Appendix A). In turn, at a fixed concentration of 50 µM NADH, NADP+ acted as an uncompetitive inhibitor towards juglone (Figure 6B) with Kii = 1.83 ± 0.19 mM, as deduced from Equation (A3) (Appendix A), describing the effects of NADP+ on the intercepts in the Lineweaver–Burk plots. As compared with the previous data obtained in 0.05-M Hepes, the use of 0.1-M phosphate decreased the kcat/Km for NADPH and increased the Kis of NADP+ almost by one order of magnitude [28,29].

2.4. Stimulation of Quinone- and Nitroreductase Activity of PfFNR by Ferredoxin

The complex of PfFNR with PfFd is characterized by micromolar Kd values [19,21]. Therefore, it is important to characterize the reduction of nonphysiological electron acceptors in the presence of both redox proteins. PfFd stimulated the reduction of quinones and nitroaromatics by PfFNR, concomitantly causing a biphasicity of the corresponding Lineweaver-Burk plots (Figure 7A). The maximal rates of their “slower” phase (low concentrations and higher kcat/Km of the oxidant) were close to the rates of cytochrome c reduction at corresponding PfFd concentrations (Figure 7B), which, in turn, are equal to the rate of PfFd reduction by PfFNR. In 0.1-M K-phosphate, pH 7.0, the maximal rate of this reaction at saturating the PfFd concentration is 16 ± 2.5 s−1, on the one-electron base, which is close to the previously determined values, 13–15 s−1 [19]. The Km(app) for PfFd is 5.2 ± 1.3 µM, a value significantly higher than that previously reported [19], which may be attributed to the higher ionic strength of the medium.

3. Discussion

These results provide a general insight into the mechanism of reduction of several groups of nonphysiological oxidants by PfFNR and complement the data on the mechanism of its interaction with the physiological electron-acceptor PfFd [18,19]. First, let us consider the possible rate-limiting step of the reaction.
Since PfFNR follows a ping-pong mechanism (Figure 1), its reductive and oxidative half-reactions can proceed independently, with the kcat expressed as 1/kcat = 1/kred(max) + 1/kox(max), where kred(max) and kox(max) are the maximal rates of the reductive and oxidative half-reactions, respectively. Our data show that, in the reduction of tetramethyl-1,4-benzoquinone, the overall catalytic process should be partly limited by both oxidative and reductive half-reactions, because the maximal rate of PfFNR reduction by NADPH at pH 7.0 is 125–148 s−1 [19,29], and a comparable value, kox(max) = 155 ± 32 s−1, was obtained for tetramethyl-1,4-benzoquinone under multiple turnover conditions (Figure 5C). On the other hand, the rate of the enzyme reduction with a fixed concentration of NADPH should be the same, irrespectively of the oxidant used, and the observed kcat(app) differences (Table 1) should be attributed to different maximal rates of reoxidation. Thus, for oxidants 1–4, 7, 10 and 13, whose kcat(app) values are close to that of tetramethyl-1,4-benzoquinone (Table 1), the overall catalytic process should be partly limited by both oxidative and reductive reactions, as well. The lower values of kcat(app) for other compounds (Table 1) imply that, in these cases, the process is limited by the oxidative half-reaction. The nature of these differences will be the object of our future studies.
Since reduced PfFNR is reoxidized in a single-electron fashion, the oxidative half-reaction must proceed through the two steps of FADH → FADH. and FADH. →FAD. There are several arguments supporting FAD semiquinone oxidation as a possible rate-limiting step of the overall process: (a) The 600-nm absorbing species is formed during the reoxidation of reduced PfFNR by quinones, which may point to a transient FADH. accumulation [13,30] (Figure 5B); (b) NADPH acts as a competitive inhibitor with respect to the oxidant in the reduction of AcPyP+ by PfFNR (Figure 3). In contrast, NADPH does not inhibit quinone reduction (Figure 1). This shows that quinones and AcPyP+ oxidize different redox forms of PfFNR, which possess different affinities for NADPH. Since AcPyP+ is an obligatory two-electron (hydride) acceptor, it can be reduced only by a two-electron reduced FAD. This argues against its involvement as a rate-limiting step in quinone reduction, and (c) NADP+ acts as a competitive inhibitor with respect to NADPH (Figure 6A) and as an uncompetitive inhibitor with respect to quinones (Figure 6B). This may be attributed to a specific case of a ping-pong mechanism, where NADP+ binds relatively tightly to the oxidized enzyme form but binds weakly or not at all to its reduced state [31]. Since NADP+ binds tightly to the two-electron reduced form of PfFNR [28,29], this also argues against its involvement as a rate-limiting step in quinone reduction. Thus, PfFNR may have properties in common with FNRs from Anabaena PCC7118. and from spinach, where the oxidation of FADH. is the rate-limiting step of reactions with various nonphysiological electron acceptors [13,30,32,33,34].
Moreover, our study discloses some specific properties of PfFNR relevant to the reduction of nonphysiological redox agents. The experiments at varied ionic strengths may characterize the surface region of PfFNR that interacts with charged oxidants. In this regard, the observed bell-shape dependences of the reactivity of oppositely charged oxidants on the ionic strength of the solution (Figure 4) were unexpected. Similar, although less pronounced, dependences were observed in the reactions of FNR from Anabaena PCC7118 [13]. A possible explanation of this phenomenon is that the oxidants may interact with both the negatively charged Glu-314 (Glu-301 in Anabaena PCC7118 FNR [18]) and positively charged Lys-287 (Arg-274 in Anabaena PCC7118 FNR), the latter participating in the binding of PfFd, which are located close to the dimethylbenzene part of the isoalloxazine ring [20,21]. This is in line with the finding that ferricyanide competes with the nicotinamide mononucleotide part of NADP+ for binding close to the FAD isoalloxazine ring [20]. Other positively charged residues participating in the binding of PfFd, such as Arg-98, Arg-290 and Lys-308, are too distant from the isoalloxazine ring [21] and may not be likely to interact with low molecular weight oxidants.
Our data point to an absence of a strict substrate specificity in the reduction of quinones, ArNO2, and ArN→O by PfFNR, with the exception of an increase in their log kcat/Km with an increase in E17 (Figure 3). The latter feature points to an “outer-sphere” electron transfer mechanism of their reduction, which is established for FNR from Anabaena PCC7119, P-450R and NOS [12,13,14]. According to this mechanism, the bimolecular rate constant of the electron transfer between the reactants (k12) is expressed as:
k12 = (k11 × k22 × K × f)1/2,
where k11 and k22 are the electron self-exchange rate constants of the reactants, K is the equilibrium constant of the reaction (log K = ΔE1/0.059 V) and f is expressed as:
log f = (log K)2/4log (k11 × k22/Z2),
where Z is the frequency factor, 1011 M−1s−1 [35]. According to Equations (2 and 3), in the reaction of the electron donor with a series of homologous electron acceptors (which display the same k22), log k12 will exhibit a parabolic (quadratic) dependence on ∆E1 with a slope 8.45 V−1 at ∆E1 = ±0.15 V. In particular, the lower reactivity of ArNO2 as compared with quinones and ArN→O, which possess similar E17 values (Figure 3), is explained by their k22 = ~106 M−1s−1 [36], which are much lower than those of quinones and ArN→O, ~108 M−1s−1 [36,37].
According to the model of Mauk et al. [38], at an infinite ionic strength, where electrostatic interactions are absent, the k11 value of metalloproteins for the reactions with the inorganic complexes is related to the distance of the electron transfer, Rp:
Rp (Å) = 6.3 − 0.35 ln k11
We have applied this approach for the analysis of the single-electron oxidation of P-450R, FNR from Anabaena PCC7118 and NOS by quinones, nitroaromatics and inorganic complexes [12,13,14]. Their reactions with Q and ArNO2 are characterized by Rp ranging from 3.4 to 5.0 Å, whereas, for the reactions with hydrophilic ferricyanide and Fe(EDTA), which are incapable of entering the protein globule, the Rp values are much higher (Table 2). However, it is possible that this procedure gives slightly overestimated distances in the case of flavoproteins, since the dimethylbenzene part of the flavin isoalloxazine ring in P-450R, FNR and NOS is partly exposed to the solvent [39,40,41]. Thus, these values may be useful only for an approximate assessment of the “intrinsic” flavoenzyme reactivity. The estimation of k11 and Rp for the reactions of PfFNR is complicated by the unknown E17 value for the FAD/FADH. couple. Based on the available data [19], the FAD semiquinone state in PfFNR is not stabilized. Thus, the values of k11 were calculated tentatively assuming 15% and 5% FAD. formations at the equilibrium (Appendix B). The resulting values of Rp for PfFNR are given in Table 2.
Typically, they are larger than those of P-450R but close to those of FNR from Anabaena PCC7118. This shows that low molecular weight oxidants may access the FAD isoalloxazine ring of both representatives of FNR with similar ease, in spite of some differences in their surroundings [20,21,41].
Finally, the stimulation of the nonphysiological acceptor reductase reactions of PfFNR by PfFd (Figure 7A,B) shows that PfFd provides an alternative pathway for their reduction via reduced PfFd. Since the redox potentials of both proteins are similar [18,19], this may be most easily explained by a better accessibility of the active center of PfFd. This phenomenon has been previously observed in reactions of bovine adrenodoxin reductase and adrenodoxin and Anabaena PCC7118 FNR and Fd [13,42]; thus, it may be a general feature of this group of redox proteins.
In conclusion, our comprehensive study characterized the mechanism of reactions of PfFNR with redox-cycling xenobiotics, which may be instrumental in the further development of redox-active antiplasmodial agents [3,4,5,6,7,8]. In terms of kcat/Km for the reduction of quinones and nitroaromatics (Table 1 and Figure 2), the reactivity of PfFNR is considerably higher as compared to other P. falciparum flavoenzymes, glutathione reductase, thioredoxin reductase and type 2 NADH dehydrogenase [9,10]. To the best of our knowledge, the reactions of the above groups of xenobiotics with other electrontransferases of the P. falciparum mitochondrial respiratory chain, namely dihydroorotate dehydrogenase, succinate dehydrogenase and malate:quinone oxidoreductase, have not been studied so far. Thus, based on the available data, PfFNR and possibly PfFd may play a central role in the reductive activation of pro-oxidant xenobiotics relevant for malaria chemotherapy. This also points to a versatility of the properties of PfFNR, which may be relevant for the design of new antiplasmodial agents, because another intriguing approach to this task is the development of compounds that may bind at the PfFNR-PfFd interface and inhibit the physiological reduction of PfFd [43,44]. Thus, an attractive possibility could be the derivatization of these ligands, chalcones or alkaloids [43,44] by redox-active quinone, ArNO2 or ArN→O moieties. These hybrid molecules may combine two mechanisms of antiplasmodial action: the inhibition of the electron supply to the apicoplast redox system and redox cycling.

4. Materials and Methods

4.1. Enzymes and Reagents

Recombinant P. falciparum ferredoxin:NADP+ oxidoreductase and ferredoxin were prepared as previously described [19], and their concentrations were determined spectrophotometrically according to ε454 = 10.0 mM−1 cm−1 and ε424 = 9.68 mM−1 cm−1, respectively. Compounds 2,4,6-trinitrotoluene (TNT) and 2,4,6-trinitrophenyl-N-methylnitramine (tetryl) were synthesized as described [45]. The 5-Nitrothiophene-2-carbonic acid morpholide was synthesized as described [25]. The 7-substituted tirapazamines, quinoxaline 1,4-dioxide and 1-oxide of tirapazamine (Figure 8) were synthesized as described [23,46,47].
All compounds were characterized by determining their melting point, as well as their 1H-NMR, UV and IR spectra. The purity of the compounds, determined using a high-performance liquid chromatography system equipped with a mass spectrometer (LCMS-2020, Shimadzu, Kyoto, Japan), was >98%. Cytochrome c, NADPH, superoxide dismutase and other compounds were obtained from Sigma-Aldrich (St. Louis, MO, USA) and used as received.

4.2. Steady-State Kinetic Studies

All kinetic experiments were carried out spectrophotometrically using a PerkinElmer Lambda 25 UV–VIS spectrophotometer (PerkinElmer, Waltham, MA, USA) in 0.1-M K-phosphate buffer (pH 7.0) containing 1-mM EDTA at 25 °C. The steady-state parameters of the reactions, the catalytic constants (kcat(app.)) and the bimolecular rate constants (or catalytic efficiency constants, kcat/Km) of the oxidants at fixed concentrations of NADPH were obtained by fitting the kinetic data to the parabolic expression using SigmaPlot 2000 (v. 11.0, SPSS Inc., Chicago, IL, USA). They correspond to the reciprocal intercepts and slopes of Lineweaver-Burk plots, (E)/V vs. 1/(oxidant) respectively, where V is the reaction rate, and (E) is the enzyme concentration. kcat represents the number of molecules of NADPH oxidized by a single active center of the enzyme per second. In the case of two-substrate reactions or inhibitions, the data were fitted to Equations (A1–3). The rates of PfFNR-catalyzed NADPH oxidation in the presence of quinones, nitroaromatic compounds or tirapazamine derivatives were determined using the value Δε340 = 6.2 mM−1 cm−1. The rates were corrected for the intrinsic NADPH-oxidase activity of the enzyme, determined as 0.12 s−1. In separate experiments, in which 50-µM cytochrome c was included in the reaction mixture, its tirapazamine-mediated reduction was measured using the value Δε550 = 20 mM−1 cm−1. Ferricyanide reduction rate was determined using the value ∆ε420 = 1.03 mM−1cm−1. The reduction rate of AcPyP+ was determined using the value ∆ε363 = 5.6 mM−1cm−1 [48]. The rates of oxygen consumption during the reactions were monitored under identical conditions using a Clark electrode (Rank Brothers Ltd., Bottisham, UK).

4.3. Presteady-State Kinetic Studies

Enzyme rapid kinetic studies were performed using a SX20 stopped-flow spectrophotometer (Applied Photophysics, Leatherhead, UK) under aerobic conditions. The enzyme reduction by NADPH and its reoxidation was monitored at 460 and 600 nm, respectively. During turnover studies, the enzyme in the first syringe (6.0–7.0 µM after mixing) was mixed with the contents of the second syringe (50-µM NADPH and 100–500-µM tetramethyl-1,4-benzoquinone after mixing).

Author Contributions

M.L. performed kinetic experiments, A.A. purified enzymes, J.Š. synthesized compounds and N.Č. designed and supervised the experiments and wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the European Social Fund (Measure No. 09.33-LMT-K-712, grant No. DOTSUT-34/09.3.3.-LMT-K712-01-0058/LSS-600000-58) (M.L., J.Š and N.Č.) and by the Fund Linea-2-2016/2017 grant by the Universita degli Studi di Milano (A.A.).

Acknowledgments

We thank Saulius Klimašauskas for access to the stopped-flow facility.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses or interpretation of data; in the writing of the manuscript or in the decision to publish the results.

Abbreviations

AcPyP+ 3-Acetylpyridineadenine dinucleotide phosphate
ArNO2aromatic nitrocompound
ArN→Oaromatic N-oxide
E17single-electron reduction midpoint potential at pH 7.0
E07two-electron (standard) reduction midpoint potential at pH 7.0
Fdferredoxin
FNRferredoxin:NADP+ oxidoreductase
kcatcatalytic constant
kcat/Kmbimolecular rate constant
Qquinone

Appendix A

Kinetic parameters of steady-state reactions according to a ping-pong mechanism were calculated according to Equation (A1):
V [ E ] = k c a t   [ S ] [ Q ] K m ( S ) [ Q ] + K m   ( Q ) [ S ] + [ S ] [ Q ]
where S stands for NADPH, and Q stands for the electron acceptor. The competitive inhibition constant (Kis) of NADP+ (I) vs. NADPH (S) was calculated according to Equation (A2):
V [ E ] = k cat   [ S ] K m ( S ) ( 1 + [ I ] K is ) + [ S ]  
For the transhydrogenase reaction, S stands for AcPyP+, and I stands for NADPH. The noncompetitive inhibition constant (Kii) of NADP+ vs. the electron acceptor was calculated according to Equation (A3):
V [ E ] = k cat   [ Q ] K m ( Q ) + [ Q ] ( 1 + [ I ] K ii )

Appendix B

The k22 values of ferricyanide and Fe(EDTA)at infinite ionic strengths are equal to 4.6 × 105 M−1s−1 and 6.9 × 104 M−1s−1, respectively [38]. At the ionic strength of the solution, 1.10 M, the k12 values of ferricyanide and Fe(EDTA) are equal to 1.0 × 106 M−1s−1 and 1.0 × 104 M−1s−1, respectively (Figure 4). Using E7(FAD/FADH) = −0.28 V for PfFNR [19], the calculations according to the Nernst equation give E7(FAD/FADH.) = −0.308 V (15% FADH. stabilization) and E7(FAD/FADH.) = −0.337 V (5% FADH. stabilization). The electron self-exchange rate constants of PfFNR, k11, are calculated after the rearrangement of Equation (3) [49]:
log k11 = log k12 − log k22 − 0.5 log K + log Z − [(log Z − log k12)2 + log K (log Z − log k12)]1/2
For the reactions of PfFNR with ferricyanide, Equation (A4) gives k11 = 10−4 M−1s−1 (15% FADH. stabilization) and k11 = 4.1 × 10−5 M−1s−1 (5% FADH. stabilization). Similarly, for the reactions with Fe(EDTA), it yields k11 = 3.5 × 10−4 M−1s−1 (15% FADH. stabilization) and k11 = 1.3 × 10−4 M−1s−1 (5% FADH. stabilization). For the reactions of PfFNR with Q and ArNO2, the approximate k11 values may be estimated from the data of Figure 3 at ∆E17 = 0, where k12 = (k11 × k22)1/2. At a 15% FADH. stabilization, i.e., at the E17 of the oxidant being equal to −0.308 V, the k11 values are equal to 79 M−1s−1 (quinones) and to 63 M−1s−1 (nitroaromatics). At a 5% FADH. stabilization, i.e., at the E17 of the oxidant being equal to −0.337 V, the k11 values are equal to 40 M−1s−1 (quinones) and to 6.3 M−1s−1 (nitroaromatics).

References

  1. Bhatt, S.; Weiss, D.J.; Cameron, E.; Bisancia, D.; Mappin, B.; Dalrymple, U.; Battle, K.; Moyes, C.L.; Henry, A.; Eckhoff, P.A.; et al. The effect of malaria control on Plasmodium falciparum in Africa between 2000 and 2015. Nature 2015, 526, 207–211. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Müller, S. Redox and antioxidant systems of the malaria parasite Plasmodium falciparum. Mol. Microbiol. 2004, 53, 1291–1305. [Google Scholar] [CrossRef] [PubMed]
  3. Vennerstrom, L.J.; Eaton, J.W. Oxidants, oxidant drugs, and malaria. J. Med. Chem. 1988, 31, 1269–1277. [Google Scholar] [CrossRef] [PubMed]
  4. Grellier, P.; Šarlauskas, J.; Anusevičius, Ž.; Marozienė, A.; Houeee-Levin, C.; Schrevel, J.; Čėnas, N. Antiplasmodial activity of nitroaromatic and quinoidal compounds: Redox potential vs inhibition of erythrocyte glutathione reductase. Arch. Biochem. Biophys. 2001, 393, 199–206. [Google Scholar] [CrossRef]
  5. Vicente, E.; Lima, L.M.; Bongard, E.; Charnaud, S.; Villar, R.; Solano, B.; Burguete, A.; Perez-Silanes, S.; Aldana, I.; Vivas, L.; et al. Synthesis and structure-activity relationship of 3-phenylquinoxaline 1,4-di-N-oxide derivatives as antimalarial agents. Eur. J. Med. Chem. 2008, 43, 1903–1910. [Google Scholar] [CrossRef] [PubMed]
  6. Grellier, P.; Marozienė, A.; Nivinskas, H.; Šarlauskas, J.; Aliverti, A.; Čėnas, N. Antiplasmodial activity of quinones: Roles of aziridinyl substituents and the inhibition of Plasmodium falciparum glutathione reductase. Arch. Biochem. Biophys. 2010, 494, 32–39. [Google Scholar] [CrossRef]
  7. Pal, C.; Bandyopadhyay, H. Redox active antiparasitic drugs. Antiox. Redox. Signal. 2012, 17, 555–582. [Google Scholar] [CrossRef]
  8. Quiliano, M.; Pabón, A.; Ramirez-Calderon, G.; Barea, C.; Deharo, E.; Galiano, S.; Aldana, I. New hydrazine and hydrazide quinoxaline 1,4-di-N-oxide derivatives: In silico ADMET, antiplasmodial and antileishmanial activity. Bioorg. Med. Chem. Lett. 2017, 27, 1820–1825. [Google Scholar] [CrossRef]
  9. Morin, C.; Besset, T.; Moutet, J.C.; Fayolle, M.; Brückner, M.; Limosin, D.; Becker, K.; Davioud-Charvet, E. The aza-analogues of 1,4-naphthoquinones are potent substrates and inhibitors of plasmodial thioredoxin and glutathione reductases and of human erythrocyte glutathione reductase. Org. Biomol. Chem. 2008, 6, 2731–2742. [Google Scholar] [CrossRef]
  10. Dong, C.K.; Patel, V.; Yang, J.C.; Dvorin, J.D.; Duraisingh, M.T.; Clardy, J.; Wirth, D.F. Type II NADH dehydrogenase of the respiratory chain of Plasmodium falciparum and its inhibitors. Bioorg. Med. Chem. Lett. 2009, 19, 972–975. [Google Scholar] [CrossRef] [Green Version]
  11. Marozienė, A.; Lesanavičius, M.; Davioud-Charvet, E.; Aliverti, A.; Grellier, P.; Šarlauskas, J.; Čėnas, N. Antiplasmodial activity of nitroaromatic compounds: Correlation with their reduction potential and inhibitory action on Plasmodium falciparum glutathione reductase. Molecules 2019, 24, 4509. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Čėnas, N.; Anusevičius, Ž.; Bironaitė, D.; Bachmanova, G.I.; Archakov, A.I.; Öllinger, K. The electron transfer reactions of NADPH:cytochrome P450 reductase with nonphysiological oxidants. Arch. Biochem. Biophys. 1994, 315, 400–406. [Google Scholar] [CrossRef] [PubMed]
  13. Anusevičius, Ž.; Martínez-Júlvez, M.; Genzor, C.G.; Nivinskas, H.; Gómez-Moreno, C.; Čėnas, N. Electron transfer reactions of Anabaena PCC 7119 ferredoxin:NADP+ reductase with nonphysiological oxidants. Biochim. Biophys. Acta. 1997, 1320, 247–255. [Google Scholar] [CrossRef] [Green Version]
  14. Anusevičius, Ž.; Nivinskas, H.; Šarlauskas, J.; Sari, M.A.; Boucher, J.L.; Čėnas, N. Single-electron reduction of quinone and nitroaromatic xenobiotics by recombinant rat neuronal nitric oxide synthase. Acta Biochim. Pol. 2013, 60, 217–222. [Google Scholar] [CrossRef]
  15. Vermilion, J.L.; Ballou, D.P.; Massey, V.; Coon, M. Separate roles for FMN and FAD in catalysis by liver microsomal NADPH-cytochrome P-450 reductase. J. Biol. Chem. 1981, 256, 266–277. [Google Scholar]
  16. Batie, C.J.; Kamin, H. Electron transfer by ferredoxin:NADP+ reductase. Rapid-reaction evidence for participation of a ternary complex. J. Biol. Chem. 1984, 259, 11976–11985. [Google Scholar]
  17. Matsuda, H.; Kimura, S.; Iyanagi, T. One-electron reduction of quinones by the neuronal nitric oxide synthase reductase domain. Biochim. Biophys. Acta 2000, 1459, 106–116. [Google Scholar] [CrossRef] [Green Version]
  18. Kimata-Ariga, Y.; Kurisu, G.; Kusunoki, M.; Aoki, S.; Sato, D.; Kobayashi, T.; Kita, K.; Horii, T.; Hase, T. Cloning and characterization of ferredoxin and ferredoxin-NADP+ reductase from human malaria parasite. J. Biochem. 2007, 141, 421–428. [Google Scholar] [CrossRef]
  19. Balconi, E.; Pennati, A.; Crobu, D.; Pandini, V.; Cerutti, R.; Zanetti, G.; Aliverti, A. The ferredoxin-NADP+ reductase/ferredoxin electron transfer system of Plasmodium falciparum. FEBS J. 2009, 276, 4249–4260. [Google Scholar] [CrossRef]
  20. Milani, M.; Balconi, E.; Aliverti, A.; Mastrangelo, E.; Seeber, F.; Bolognesi, M.; Zanetti, G. Ferredoxin-NADP+ reductase from Plasmodium falciparum undergoes NADP+-dependent dimerization and inactivation: Functional and crystallographic analysis. J. Mol. Biol. 2007, 367, 501–513. [Google Scholar] [CrossRef]
  21. Kimata-Ariga, Y.; Yuasa, S.; Saitoh, T.; Fukuyama, H.; Hase, T. Plasmodium-specific basic amino acid residues important for the interaction with ferredoxin on the surface of ferredoxin-NADP+ reductase. J. Biochem. 2018, 164, 231–237. [Google Scholar] [CrossRef] [PubMed]
  22. Wardman, P. Reduction potentials of one-electron couples involving free radicals in aqueous solution. J. Phys. Chem. Ref. Data 1989, 18, 1637–1755. [Google Scholar] [CrossRef] [Green Version]
  23. Hay, M.P.; Gamage, S.A.; Kovacs, M.S.; Pruijn, F.B.; Anderson, R.F.; Patterson, A.V.; Wilson, W.R.; Brown, J.M.; Denny, W.A. Structure-activity relationships of 1,2,4-benzotriazine 1,4-dioxides as hypoxia-selective analogues of tirapazamine. J. Med. Chem. 2003, 46, 169–182. [Google Scholar] [CrossRef] [PubMed]
  24. Anderson, R.F.; Shinde, S.S.; Hay, M.P.; Denny, W.A. Potentiation of the cytotoxicity of the anticancer agent tirapazamine by benzotriazine N-oxides: The role of redox equilibria. J. Am. Chem. Soc. 2006, 128, 245–249. [Google Scholar] [CrossRef] [PubMed]
  25. Breccia, A.; Busi, F.; Gattavechia, E.; Tamba, M. Reactivity of nitro-thiophene derivatives with electron and oxygen radicals studied by pulse radiolysis and polarographic techniques. Radiat. Environ. Biophys. 1990, 29, 153–160. [Google Scholar] [CrossRef] [PubMed]
  26. O’Brien, P.J. Molecular mechanisms of quinone cytotoxicity. Chem. Biol. Interact. 1991, 80, 1–41. [Google Scholar] [CrossRef]
  27. Chance, B. A simple relationship for a calculation of the “on” velocity constant in enzyme reactions. Arch. Biochem. Biophys. 1957, 71, 130–136. [Google Scholar] [CrossRef]
  28. Crobu, D.; Canevari, G.; Milani, M.; Pandini, V.; Vanoni, M.A.; Bolognesi, M.; Zanetti, G.; Aliverti, A. Plasmodium falciparum ferredoxin-NADP+ reductase His286 plays a dual role in NADP(H) binding and catalysis. Biochemistry 2009, 48, 9525–9533. [Google Scholar] [CrossRef]
  29. Baroni, S.; Pandini, V.; Vanoni, M.A.; Aliverti, A. A single tyrosine hydroxyl group almost entirely controls the NADPH specificity of Plasmodium falciparum ferredoxin-NADP+ reductase. Biochemistry 2011, 51, 3819–3826. [Google Scholar] [CrossRef]
  30. Anusevičius, Ž.; Misevičienė, L.; Medina, M.; Martinez-Julvez, M.; Gomez-Moreno, C.; Čėnas, N. FAD semiquinone stability regulates single- and two-electron reduction of quinones by Anabaena PCC7119 ferredoxin:NADP+ reductase and its Glu301Ala mutant. Arch. Biochem. Biophys. 2005, 437, 144–150. [Google Scholar] [CrossRef]
  31. Valiauga, B.; Williams, E.M.; Ackerley, D.F.; Čėnas, N. Reduction of quinones and nitroaromatic compounds by Escherichia coli nitroreductase A (NfsA): Characterization of kinetics and substrate specificity. Arch. Biochem. Biophys. 2017, 614, 14–22. [Google Scholar] [CrossRef] [PubMed]
  32. Massey, V.; Matthews, R.G.; Foust, G.P.; Howell, L.G.; Williams, C.H., Jr.; Zanetti, G.; Ronchi, S. A new intermediate in TPNH-linked flavoproteins. In Pyridine Nucleotide-Dependent Dehydrogenases; Sund, H., Ed.; Springer: Berlin, Germany, 1970; pp. 393–411. [Google Scholar]
  33. Corrado, M.E.; Aliverti, A.; Zanetti, G.; Mayhew, S.G. Analysis of the oxidation-reduction potentials of recombinant ferredoxin-NADP+ reductase from spinach chloroplasts. Eur. J. Biochem. 1996, 239, 662–667. [Google Scholar] [CrossRef] [PubMed]
  34. Medina, M.; Martínez-Júlvez, M.; Hurley, J.K.; Tollin, G.; Gómez-Moreno, C. Involvement of glutamic acid 301 in the catalytic mechanism of ferredoxin-NADP+ reductase from Anabaena PCC 7119. Biochemistry 1998, 37, 2715–2728. [Google Scholar] [CrossRef]
  35. Marcus, R.; Sutin, N. Electron transfers in chemistry and biology. Biochim. Biophys. Acta. 1985, 811, 265–322. [Google Scholar] [CrossRef]
  36. Wardman, P.; Dennis, M.F.; Everett, S.A.; Patel, K.B.; Stratford, M.R.; Tracy, M. Radicals from one-electron reduction of nitro compounds, aromatic N-oxides and quinones: The kinetic basis for hypoxia-selectve, bioreductive drugs. Biochem. Soc. Symp. 1995, 61, 171–194. [Google Scholar]
  37. Nemeikaitė-Čėnienė, A.; Šarlauskas, J.; Jonušienė, V.; Marozienė, A.; Misevičienė, L.; Yantsevich, A.V.; Čėnas, N. Kinetics of flavoenzyme-catalyzed reduction of tirapazamine derivatives: Implications for their prooxidant cytotoxicity. Int. J. Mol. Sci. 2019, 20, 4602. [Google Scholar] [CrossRef] [Green Version]
  38. Mauk, A.G.; Scott, R.A.; Gray, H.B. Distances of electron transfer to and from metalloprotein redox sites in reactions with inorganic complexes. J. Am. Chem. Soc. 1980, 102, 4360–4363. [Google Scholar] [CrossRef]
  39. Wang, M.; Roberts, D.L.; Paschke, R.; Shea, T.M.; Masters, B.S.S.; Kim, J.J.P. Three-dimensional structure of NADPH-cytochrome P450 reductase: Prototype for FMN- and FAD-containing enzymes. Proc. Natl. Acad. Sci. USA 1997, 94, 8411–8416. [Google Scholar] [CrossRef] [Green Version]
  40. Xia, C.; Misra, I.; Iyanagi, T.; Kim, J.J. Regulation of interdomain interactions by calmodulin in inducible nitric-oxide synthase. J. Biol. Chem. 2009, 284, 30708–30717. [Google Scholar] [CrossRef] [Green Version]
  41. Hurley, J.K.; Morales, R.; Martinez-Julvez, M.; Brodie, T.B.; Medina, M.; Gomez-Moreno, C.; Tollin, G. Structure-function relationships in Anabaena ferredoxin/ferredoxin:NADP+ reductase electron transfer: Insight from site-directed mutagenesis, transient absorption spectroscopy and X-ray crystallography. Biochim. Biophys. Acta 2002, 1554, 5–21. [Google Scholar] [CrossRef] [Green Version]
  42. Čėnas, N.K.; Marcinkevičienė, J.A.; Kulys, J.K.; Usanov, S.A. A negative cooperativity between NADPH and adrenodoxin on binding to NADPH:adrenodoxin reductase. FEBS Lett. 1990, 259, 338–340. [Google Scholar] [CrossRef] [Green Version]
  43. Suwito, H.; Jumina; Mustofa; Pudjiastuti, P.; Fanani, M.Z.; Kimata-Ariga, Y.; Katahira, R.; Kawakami, T.; Fujiwara, T.; Hase, T.; et al. Design and synthesis of chalcone derivatives as inhibitors of the ferredoxin – ferredoxin-NADP+ reductase interaction of Plasmodium falciparum: Pursuing new antimalarial agents. Molecules 2014, 19, 21473–21488. [Google Scholar] [PubMed] [Green Version]
  44. Pudjiastuti, P.; Puspaningsih, N.N.T.; Siswanto, I.; Fanani, M.Z.; Kimata-Ariga, Y.; Hase, T.; Sarker, S.D.; Nahar, L. Inhibitory activity and docking analysis of antimalarial agents from Stemona sp. toward ferredoxin-NADP+ reductase from malaria parasites. J. Parasitol. Res. 2018, 2018, 3469132. [Google Scholar] [CrossRef] [PubMed]
  45. Čėnas, N.; Nameikaitė-Čėnienė, A.; Sergedienė, E.; Nivinskas, H.; Anusevičius, Ž.; Šarlauskas, J. Quantitative structure-activity relationships in enzymatic single-electron reduction of nitroaromatic explosives: Implications for cytotoxicity. Biochim. Biophys. Acta 2001, 1528, 31–38. [Google Scholar] [CrossRef]
  46. Monge, A.; Martinez-Crespo, F.J.; Lopez de Cerain, A.; Palop, J.A.; Narro, S.; Senador, V.; Marin, A.; Sainz, Y.; Gonzalez, M. Hypoxia-selective agents derived from 2-quinoxalinecarbonitrile 1,4-di-N-oxides. 2. J. Med. Chem. 1995, 38, 4488–4494. [Google Scholar] [CrossRef] [PubMed]
  47. Boyd, M.; Hay, M.P.; Boyd, P.D.W. Complete 1H, 13C and 15N NMR assignment of tirapazamine and related 1,2,4-benzotriazine N-oxides. Magn. Reson. Chem. 2006, 44, 948–954. [Google Scholar] [CrossRef]
  48. Kaplan, N.O.; Ciotti, M.M. Chemistry and properties of the 3-acetylpyridine analogue of diphosphopyridine nucleotide. Arch. Biochem. Biophys. 1956, 221, 823–832. [Google Scholar]
  49. Pladziewicz, J.R.; Carney, M.H. Reduction of ferricenium ion by horse heart ferricytochrome c. J. Am. Chem. Soc. 1982, 104, 3544–3545. [Google Scholar] [CrossRef]
Figure 1. Steady-state kinetics of a reduction of juglone by NADPH catalyzed by PfFNR. NADPH concentrations: 200 µM (1), 150 µM (2), 100 µM (3), 75 µM (4), 50 µM (5) and 25 µM (6).
Figure 1. Steady-state kinetics of a reduction of juglone by NADPH catalyzed by PfFNR. NADPH concentrations: 200 µM (1), 150 µM (2), 100 µM (3), 75 µM (4), 50 µM (5) and 25 µM (6).
Ijms 21 03234 g001
Figure 2. Dependence of the reactivity of quinones, nitroaromatic compounds and aromatic N-oxides on their single-electron reduction midpoint potentials. Relationship between the log kcat/Km of quinones (solid circles), nitroaromatics (blank triangles) and N-oxides (blank circles) and their single-electron reduction midpoint potentials at pH 7.0 (E17). Numbers and reduction potentials of compounds are given in Table 1.
Figure 2. Dependence of the reactivity of quinones, nitroaromatic compounds and aromatic N-oxides on their single-electron reduction midpoint potentials. Relationship between the log kcat/Km of quinones (solid circles), nitroaromatics (blank triangles) and N-oxides (blank circles) and their single-electron reduction midpoint potentials at pH 7.0 (E17). Numbers and reduction potentials of compounds are given in Table 1.
Ijms 21 03234 g002
Figure 3. Steady-state kinetics of the reduction of 3-acetylpyridineadenine dinucleotide phosphate by PfFNR. NADPH concentrations: 200 µM (1), 150 µM (2), 100 µM (3), 50 µM (4) and 25 µM (5).
Figure 3. Steady-state kinetics of the reduction of 3-acetylpyridineadenine dinucleotide phosphate by PfFNR. NADPH concentrations: 200 µM (1), 150 µM (2), 100 µM (3), 50 µM (4) and 25 µM (5).
Ijms 21 03234 g003
Figure 4. Effects of the ionic strength on the reactivity of PfFNR towards the electron acceptors. The dependence of log kcat/Km for ferricyanide (1), Fe(EDTA) (2), benzylviologen (3) and tetramethyl-1,4-benzoquinone (4) on the ionic strength of the phosphate buffer at pH 7.0 is shown.
Figure 4. Effects of the ionic strength on the reactivity of PfFNR towards the electron acceptors. The dependence of log kcat/Km for ferricyanide (1), Fe(EDTA) (2), benzylviologen (3) and tetramethyl-1,4-benzoquinone (4) on the ionic strength of the phosphate buffer at pH 7.0 is shown.
Ijms 21 03234 g004
Figure 5. Turnover of PfFNR in the presence of NADPH and oxidants. (A,B) The kinetics of the absorbance changes at 600 nm (1) and 460 nm (2) during the reduction of Pf FNR (6.0 µM) by 50-µM NADPH and its subsequent reoxidation by oxygen (A) or 250-µM tetramethyl-1,4-benzoquinone (B) (concentrations after mixing). (C) The dependence of an apparent first-order reoxidation rate constant on the concentration of tetramethyl-1,4-benzoquinone.
Figure 5. Turnover of PfFNR in the presence of NADPH and oxidants. (A,B) The kinetics of the absorbance changes at 600 nm (1) and 460 nm (2) during the reduction of Pf FNR (6.0 µM) by 50-µM NADPH and its subsequent reoxidation by oxygen (A) or 250-µM tetramethyl-1,4-benzoquinone (B) (concentrations after mixing). (C) The dependence of an apparent first-order reoxidation rate constant on the concentration of tetramethyl-1,4-benzoquinone.
Ijms 21 03234 g005
Figure 6. Inhibition of the juglone reductase reaction of PfFNR by NADP+. (A) Inhibition at varied NADPH concentrations in the presence of 100-µM juglone, and (B) inhibition at varied juglone concentrations in the presence of 100-µM NADPH, NADP+ concentrations: 0 mM (1), 1.0 mM (2), 2.0 mM (3), 3.0 mM (4) and 5.0 mM (5).
Figure 6. Inhibition of the juglone reductase reaction of PfFNR by NADP+. (A) Inhibition at varied NADPH concentrations in the presence of 100-µM juglone, and (B) inhibition at varied juglone concentrations in the presence of 100-µM NADPH, NADP+ concentrations: 0 mM (1), 1.0 mM (2), 2.0 mM (3), 3.0 mM (4) and 5.0 mM (5).
Ijms 21 03234 g006
Figure 7. Stimulation of quinone- and nitroreductase reactions of PfFNR by ferredoxin. (A) The dependence of the PfFNR-catalyzed NADPH oxidation rate on the concentration of p-nitroacetophenone (1,4) or 2-hydroxy-1,4-naphthoquinone (2,3) in the absence of PfFd (1,2), and in the presence of 1.7-µM (3) or 4.0-µM (4) PfFd; concentration of NADPH is 100 µM. (B) The dependence of the cytochrome c reduction rate by PfFNR on the concentration of PfFd (solid circles), concentration of NADPH, 100 µM and concentration of cytochrome c, 50 µM. Blank circles show the doubled maximal rates of the ”slower” phase of NADPH oxidation in the presence of tetramethyl-1,4-benzoquinone (1), p-nitroacetophenone (2,4) and 2-hydroxy-1,4-naphthoquinone (3) at corresponding concentrations of PfFd. The maximal rates were obtained by the fitting of kinetic data of the “slower” phase (5–6 lower concentrations of oxidant) to the parabolic expression.
Figure 7. Stimulation of quinone- and nitroreductase reactions of PfFNR by ferredoxin. (A) The dependence of the PfFNR-catalyzed NADPH oxidation rate on the concentration of p-nitroacetophenone (1,4) or 2-hydroxy-1,4-naphthoquinone (2,3) in the absence of PfFd (1,2), and in the presence of 1.7-µM (3) or 4.0-µM (4) PfFd; concentration of NADPH is 100 µM. (B) The dependence of the cytochrome c reduction rate by PfFNR on the concentration of PfFd (solid circles), concentration of NADPH, 100 µM and concentration of cytochrome c, 50 µM. Blank circles show the doubled maximal rates of the ”slower” phase of NADPH oxidation in the presence of tetramethyl-1,4-benzoquinone (1), p-nitroacetophenone (2,4) and 2-hydroxy-1,4-naphthoquinone (3) at corresponding concentrations of PfFd. The maximal rates were obtained by the fitting of kinetic data of the “slower” phase (5–6 lower concentrations of oxidant) to the parabolic expression.
Ijms 21 03234 g007
Figure 8. Structural formulae of aromatic N-oxides used in this work. Derivatives of 3-amino-1,2,4-benzotriazine-1,4-dioxide (tirapazamine) (1), 3-amino-1,2,4-benzotriazine 1-oxide (2) and quinoxaline-1,4-dioxide. Formulae were drawn using ISIS/Draw (v. 2002, MDL Information Systems, San Leandro, CA, USA).
Figure 8. Structural formulae of aromatic N-oxides used in this work. Derivatives of 3-amino-1,2,4-benzotriazine-1,4-dioxide (tirapazamine) (1), 3-amino-1,2,4-benzotriazine 1-oxide (2) and quinoxaline-1,4-dioxide. Formulae were drawn using ISIS/Draw (v. 2002, MDL Information Systems, San Leandro, CA, USA).
Ijms 21 03234 g008
Table 1. Steady-state rate constants of the reduction of nonphysiological electron acceptors by NADPH catalyzed by PfFNR. (NADPH) = 100 µM, 0.1-M K-phosphate + 1.0-mM EDTA, pH 7.0, 25 °C.
Table 1. Steady-state rate constants of the reduction of nonphysiological electron acceptors by NADPH catalyzed by PfFNR. (NADPH) = 100 µM, 0.1-M K-phosphate + 1.0-mM EDTA, pH 7.0, 25 °C.
No.CompoundE17 (V) [22,23,24,25]kcat(app) (s−1)kcat/Km (M−1s−1)
Quinones
12-Methyl-1,4-benzoquinone0.0132.4 ± 4.12.8 ± 0.2 × 105
22,5-Dimethyl-1,4-benzoquinone−0.0726.0 ± 2.31.8 ± 0.2 × 105
35-Hydroxy-1,4-naphthoquinone−0.0935.7 ± 5.11.1 ± 0.1 × 106
45,8-Dihydroxy-1,4-naphthoquinone−0.1125.0 ± 3.84.1 ± 0.5 × 105
59,10-Phenanthrene quinone−0.1220.3 ± 3.32.0 ± 0.3 × 105
61,4-Naphthoquinone−0.1516.9 ± 2.13.1 ± 0.3 × 105
72-Methyl-1,4-naphthoquinone−0.2026.5 ± 3.22.6 ± 0.3 × 105
8Tetramethyl-1,4-benzoquinone−0.2633.1 ± 4.37.2 ± 0.8 × 104
9Benzylviologen−0.3544.8 ± 0.62.1 ± 0.1 × 104
109,10-Anthraquinone-2-sulphonate−0.3827.0 ± 3.21.6 ± 0.2 × 105
112-Hydroxy-1,4-naphthoquinone−0.412.6 ± 0.39.8 ± 0.2 × 103
122-Methyl-3-hydroxy-1,4-naphthoquinone−0.4614.0 ± 1.21.1 ± 0.1 × 104
Nitroaromatic Compounds
13Tetryl−0.19140.0 ± 5.12.1 ± 0.4 × 105
142,4,6-Trinitrotoluene a−0.253 1.3 ± 0.3 × 104
15Nifuroxime a−0.255 3.3 ± 0.2 × 104
16Nitrofurantoin a−0.255 6.8 ± 0.5 × 104
171,4-Dinitrobenzene−0.257 9.1 ± 0.8 × 104
181,2-Dinitrobenzene a−0.287 1.1 ± 0.2 × 104
195-Nitrothiophene-2-carbonic acid morpholide−0.305 2.0 ± 0.2 × 104
204-Nitrobenzaldehyde a−0.325 4.0 ± 0.3 × 103
213,5-Dinitrobenzoic acid a−0.344 3.9 ± 0.5 × 103
221,3-Dinitrobenzene a−0.348 2.7 ± 0.2 × 103
234-Nitroacetophenone a−0.355 3.2 ± 0.4 × 103
242-Nitrothiophene−0.390 2.2 ± 0.2 × 103
254-Nitrobenzoic acid a−0.425 4.5 ± 0.4 × 102
264-Nitrobenzyl alcohol−0.475 3.9 ± 0.2 × 102
27Nitrobenzene a−0.485 5.5 ± 0.6 × 101
Aromatic N-Oxides
287-CF3-tirapazamine−0.34511.5 ± 2.05.2 ± 0.4 × 104
297-Cl-tirapazamine−0.40014.8 ± 1.33.7 ± 0.4 × 104
307-F-tirapazamine−0.400 2.7 ± 0.2 × 104
313-Amino-1,2,4-benzotriazine-1,4-dioxide (tirapazamine)−0.456 4.4 ± 0.5 × 103
327-CH3-tirapazamine−0.474 5.0 ± 0.6 × 103
337-C2H5O-tirapazamine−0.494 4.5 ± 0.5 × 103
343-Amino-1,2,4-benzotriazine-1-oxide−0.568 3.2 ± 0.2 × 103
35Quinoxaline-1,4-dioxide−0.575 8.2 ± 0.9 × 102
Inorganic Complexes
36Ferricyanide b0.4147.9 ± 4.03.0 ± 0.4 × 106
37Fe (EDTA)0.12 4.3 ± 0.2 × 104
a Taken from Reference [11]. b Calculated on a single-electron base. Catalytic efficiency constants, kcat/Km; reduction maximal rate constants, kcat(app) and single-electron reduction midpoint potentials, E17.
Table 2. Distances of the electron transfer (Rp) in reactions of flavin-dependent electrontransferases with nonphysiological electron acceptors, calculated according to Equation (4).
Table 2. Distances of the electron transfer (Rp) in reactions of flavin-dependent electrontransferases with nonphysiological electron acceptors, calculated according to Equation (4).
FlavoproteinReactionRp (Å)
QArNO2Fe (CN)3−6Fe (EDTA)
P-450R, rat [12]FMNH − e → FMNH.,
E17 = −0.270 V
3.44.28.17.3
n-NOS, rat [14]FMNH − e → FMNH.,
E17 = −0.274 V
4.73.9--
FNR,
Anabaena PCC7118 a
FADH. − e − H+ → FAD,
E17 = −0.280 V
5.04.49.210.4–11.4
PfFNR, this workFADH. − e − H+ → FAD,
E17 = −0.308 V b
4.84.99.59.1
FADH. − e − H+ → FAD,
E17 = −0.337 V c
5.05.69.89.4
b Calculated according to the data of Reference [13], using the value of E7 (FAD/FADH.) = −0.280 V [30]. b,c The values of E17 are calculated assuming 15% and 5% FADH. stabilizations at the equilibrium, respectively.

Share and Cite

MDPI and ACS Style

Lesanavičius, M.; Aliverti, A.; Šarlauskas, J.; Čėnas, N. Reactions of Plasmodium falciparum Ferredoxin:NADP+ Oxidoreductase with Redox Cycling Xenobiotics: A Mechanistic Study. Int. J. Mol. Sci. 2020, 21, 3234. https://doi.org/10.3390/ijms21093234

AMA Style

Lesanavičius M, Aliverti A, Šarlauskas J, Čėnas N. Reactions of Plasmodium falciparum Ferredoxin:NADP+ Oxidoreductase with Redox Cycling Xenobiotics: A Mechanistic Study. International Journal of Molecular Sciences. 2020; 21(9):3234. https://doi.org/10.3390/ijms21093234

Chicago/Turabian Style

Lesanavičius, Mindaugas, Alessandro Aliverti, Jonas Šarlauskas, and Narimantas Čėnas. 2020. "Reactions of Plasmodium falciparum Ferredoxin:NADP+ Oxidoreductase with Redox Cycling Xenobiotics: A Mechanistic Study" International Journal of Molecular Sciences 21, no. 9: 3234. https://doi.org/10.3390/ijms21093234

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

Article Metrics

Back to TopTop