Trapping of a Putative Intermediate in the Cytochrome c Nitrite

Aug 5, 2019 - LEARN ABOUT THESE METRICS ... EPR analysis subsequently revealed that the {FeNO}7 moiety ... The ever greater use of the Haber process t...
0 downloads 0 Views 2MB Size
Subscriber access provided by RUTGERS UNIVERSITY

Article

Trapping of a putative intermediate in the cytochrome c nitrite reductase (ccNiR)-catalyzed reduction of nitrite: implications for the ccNiR reaction mechanism Mahbbat Ali, Natalia Stein, Yingxi Mao, Shahid Shahid, Marius Schmidt, Brian Bennett, and A. Andrew Pacheco J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.9b03036 • Publication Date (Web): 05 Aug 2019 Downloaded from pubs.acs.org on August 5, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 18 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Trapping of a putative intermediate in the cytochrome c nitrite reductase (ccNiR)-catalyzed reduction of nitrite: implications for the ccNiR reaction mechanism Mahbbat Ali,† Natalia Stein,† Yingxi Mao,† Shahid Shahid,† Marius Schmidt,‡ Brian Bennett,§,* A. Andrew Pacheco†,* From the †Department of Chemistry and Biochemistry, University of Wisconsin-Milwaukee, Milwaukee, WI 53211, United States; ‡Department of Physics, University of Wisconsin-Milwaukee, Milwaukee, WI 53211, United States; §Department of Physics, Marquette University, Milwaukee, WI 53233, United States.

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 18

Abstract

Cytochrome c Nitrite Reductase (ccNiR) is a periplasmic, decaheme homodimeric enzyme that catalyzes the six-electron reduction of nitrite to ammonia. Under standard assay conditions catalysis proceeds without detected intermediates, and it’s been assumed that this is also true in vivo. However, this report demonstrates that it’s possible to trap a putative intermediate by controlling the electrochemical potential at which reduction takes place. UV/Vis spectropotentiometry showed that nitrite-loaded Shewanella oneidensis ccNiR is reduced in a concerted 2-electron step to generate an {FeNO}7 moiety at the active site, with an associated midpoint potential of +246 mV vs SHE at pH 7. By contrast, cyanide-bound active site reduction is a one-electron process with a midpoint potential of 20 mV, and without a strongfield ligand the active site midpoint potential shifts 70 mV lower still. EPR analysis subsequently revealed that the {FeNO}7 moiety possesses an unusual spectral signature, different from those normally observed for {FeNO}7 hemes, that may indicate magnetic interaction of the active site with nearby hemes. Protein film voltammetry experiments previously showed that catalytic nitrite reduction to ammonia by S. oneidensis ccNiR requires an applied potential of at least 120 mV, well below the midpoint potential for {FeNO}7 formation. Thus, it appears that an {FeNO}7 active site is a catalytic intermediate in the ccNiR-mediated reduction of nitrite to ammonia, whose degree of accumulation depends exclusively on the applied potential. At low potentials the species is rapidly reduced and doesn’t accumulate, while at higher potentials it is trapped, thus preventing catalytic ammonia formation.

INTRODUCTION

Cytochrome c nitrite reductase (ccNiR, also often called NrfA) is a soluble periplasmic bacterial enzyme that catalyzes the 6-electron reduction of nitrite to ammonia. This forms part of an anaerobic respiratory process used by ammonifying bacteria, in which nitrite is the terminal electron acceptor.1-3 In vitro ccNiR can also catalyze the 5-electron reduction of nitric oxide to ammonia, and the 2-electron reduction of hydroxylamine to ammonia,4-5 two processes that have been proposed to have in vivo applications as well. In support of this hypothesis, ccNiR and a related nitrite reducing multi-heme protein have been isolated from bacteria that are incapable of dissimilatory nitrite reduction.6-7 In these organisms ccNiR is believed to help mitigate nitrosative stress arising from intracellular nitrite, nitric oxide and hydroxylamine accumulations.4, 6-8 Active ccNiR is a homodimer, with protomeric molecular mass ranging from 52 to 65 kDa depending on the organism from which it is obtained.6, 9-14 In all homologues studied so far each protomer contains five c-type hemes that are ferric in the resting enzyme (Fig. 1). The active site (heme 1, Fig. 1) is high-spin,

Figure 1. Five hemes, and their associated axial ligands, from one of the Shewanella oneidensis ccNiR protomers (the heme arrangement is virtually identical in ccNiR from other organisms). The S. oneidensis heme midpoint potentials (vs. the standard hydrogen electrode, SHE) previously obtained in the presence of the strong-field (SF) ligand cyanide are shown in red, while those obtained in the presence of only weak-field (WF) ligands such as water are shown in blue (ref. 16). The dashed vertical line represents the dimer interface. Electrons are believed to enter the enzyme via heme 2; from there they can be shuttled from heme to heme, to the active site of either protomer.6, 9-14

and unusually, has a lysine residue in the proximal axial position; its distal axial position is occupied by a labile water in the resting enzyme. The remaining four hemes (hemes 2-5, Fig. 1) are six-coordinate, bis-histidine ligated and low-spin, and serve to efficiently shuttle electrons from the heme 2 entry point to the active site during catalysis.6, 9-14 We are particularly interested in the mechanisms by which the multiheme enzymes ccNiR and hydroxylamine dehydrogenase (HDH15), whose heme arrangements share many superficial similarities, are optimized to catalyze common processes in opposite directions.14, 16-19 This report focuses on the mechanism of the early steps of ccNiR-catalyzed nitrite reduction. Nitrite is an important intermediate in the biological interconversion of ammonia and nitrate. Ammonia – nitrate interconversion by ammonifying and ammonia oxidizing bacteria forms an important part of the biological nitrogen cycle, and has attracted considerable attention from microbiologists, geochemists and enzymologists alike.1-3, 20-24 Though nitrogen is incorporated into biomolecules at the oxidation state of ammonia, at any given moment a large fraction of nitrogen bio-available in the environment is actually in the form of nitrate.24 This is because much of the ammonia not promptly incorporated into biomass is instead

ACS Paragon Plus Environment

Page 3 of 18 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

used as a respiratory electron source by ammonia oxidizing bacteria, which convert it to nitrite; other bacteria then use the nitrite as an electron source, generating nitrate.24 The overall oxidation of ammonia to nitrate is called nitrification. Nitrate can be reduced back to ammonia via nitrite, either for the express purpose of incorporating it into biomass (assimilatory ammonification), or incidentally when microbes use it as an electron sink during anaerobic respiration (dissimilatory ammonification). Alternatively, microbes can also reduce nitrate and nitrite to dinitrogen via nitrous oxide in a different anaerobic respiratory path (denitrification).24 For ecological reasons it is becoming increasingly important to understand the chemical processes that allow interconversion of the bioavailable forms of nitrogen (referred to as “reactive nitrogen”). The ever greater use of the Haber process to fix nitrogen for agricultural consumption has resulted in historically high accumulation of reactive nitrogen species in the biosphere, and this is having many unintended negative consequences that demand urgent attention.22-25 The biggest problem is that ammonia fertilizer is so rapidly converted to nitrate by ammonia oxidizing bacteria. Unlike ammonia, the negatively charged nitrate readily passes out of soil and into waterways, where it promotes algal blooms that consume all available oxygen, thereby killing aerobic aquatic organisms. Steady accumulation in the atmosphere of nitrous oxide, generated during nitrification and denitrification, also poses problems. In contrast to ammonia and nitrate, nitrous oxide is not particularly reactive. However, because of its comparative inertness, it can persist in the atmosphere for prolonged periods of time. This is a problem because nitrous oxide is a potent greenhouse gas (about 300 more potent than carbon dioxide),24 and also currently the single most important ozone depleting emission on earth.26-27 Many of the current problems caused by high accumulations of reactive nitrogen species and nitrous oxide could be mitigated by more efficient use of ammonia fertilizer. A better understanding of ammonia – nitrite interconversion, and other chemical processes in the nitrogen cycle, could in turn greatly aid microbial physiologists and ecologists, and ultimately farmers, in achieving this objective. In the standard ccNiR assay, which uses the strong reductant methyl viologen monocation radical as the electron source, reduction of nitrite to ammonia seems to proceed in one six-electron step, without detection of any intermediates.9, 11 Up until now it has been assumed that ccNiR-catalyzed nitrite reduction in vi-

vo likewise proceeds in one step to generate ammonia exclusively. This would appear to pose a challenge for mechanistic studies, which require intermediate detection as a means of tracking the path of the reaction under study. Recently we showed that when the strong-field ligand cyanide binds to the active site of Shewanella oneidensis ccNiR it has a significant effect on the active site midpoint potential, shifting it by nearly 70 mV in the positive direction, to +27 mV vs SHE (Fig. 1).16 CcNiR’s physiological substrate nitrite is also a strong-field ligand when it binds a metal via the nitrogen atom, and might be expected to have a similar effect on heme 1’s midpoint potential. This is important because S. oneidensis ccNiR-catalyzed nitrite reduction to ammonia does not proceed appreciably at potentials higher than about 120 mV vs SHE,28-29 so if nitrite-bound heme 1 were to reduce at a potential well above 120 mV vs SHE it would effectively generate a stalled catalytic intermediate. In this article we confirm that, like cyanide-bound heme 1, nitrite-bound heme 1 also reduces at a potential that is much too high for ammonia generation, thus allowing a reduced nitrite-loaded heme 1 species to be trapped and studied. The electrochemical behavior of the S. oneidensis enzyme in the presence of nitrite is also found to differ in important ways from its behavior in the presence of cyanide. The possible mechanistic implications of these results are considered. EXPERIMENTAL PROCEDURES General Materials. Reagents were obtained from Fish-

er Scientific, Acros Organics or Sigma-Aldrich, unless specified otherwise, and used without additional purification. The pH-jump activated NO precursor 1(N,N-diethylamino)diazene-1-ium-1,2-diolate (DEANO) was prepared as described by Drago and Paulik.30-31 Protein purification and handling. S. oneidensis ccNiR was purified from a high yield expression system using a protocol described previously,14 but with one modification. After ammonium sulfate fractionation of the cell lysate, followed by hydrophobic interaction and anion-exchange chromatographies at pH 7 (ccNiR is un-retained in the anion-exchange step), the sample was equilibrated in 20 mM bis-tris pH 6.5 and loaded onto a SP-Sepharose cation exchange column (GE), to which ccNiR bound. The ccNiR was then eluted with a 20 mM HEPES pH 7 buffer, and shown to be of high purity by SDS PAGE gel electrophoresis (Fig. S1, Supporting Information). The pure protein was stored in aliquots at 80 C. Crystals of the protein

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

purified by the new method were obtained using the same protocol as described previously.14 Protein structure determination. Data were collected with a home source consisting of a GeniX microfocus X-ray source (Xenocs, France), equipped with a Mar165 detector (Rayonix, USA) and a one-axis goniometer (Huber, Germany). CcNIR crystals were flash-frozen without additional cryoprotectant using a cryogenic nitrogen stream (Oxford CryoJet, Agilent, USA) available at the home source. Data were collected at 110 K overnight. The X-ray data were indexed and integrated with the program iMosflm7.2.2,32 and scaled and merged with the program ‘scala’.33 The initial structure was obtained from pdb-entry 3UBR14 from which the water molecules were removed. Refinement against the 1.66 Å data was achieved with ‘phenix 1.10.1’,34 first with a rigid body refinement, and further by several cycles of restrained reciprocal space refinement with intermittent manual corrections using ‘coot 0.8.9.2’.35 In addition, the structural validation tools implemented in ‘coot’ were extensively used to identify and further correct structural outliers, as detailed in section S2 of Supporting Information. The refined coordinates were deposited in the Protein Data Bank with accession number 6P73. General instrumentation. Routine and time-resolved UV/vis spectra were obtained using one of three Cary 50 (Varian) spectrophotometers. Two of these spectrophotometers are housed in nitrogen-filled anaerobic gloveboxes for obtaining spectra of air-sensitive samples. UV/Vis and EPR spectropotentiometric experiments were carried out in the gloveboxes using BASi Epsilon EC potentiostats to set the appropriate potential. An Ag/AgCl electrode was used as a reference (BASi model RE-5B), and was periodically standardized by comparison with the methyl viologen couple (mo = 0.449 V vs. the standard hydrogen electrode, SHE).36 UV/Visible spectropotentiometric titrations of ccNiR.

UV/Visible spectropotentiometry experiments were performed using a BASi Epsilon EC potentiostat to set the potential, and one of the glovebox-housed spectrophotometers to obtain the spectra at each applied potential. Controlled potential electrolysis of each solution was performed in an optically transparent thin-layer electrode cell similar to one used in earlier work by this group,14 and elsewhere.37-38 A solution containing ccNiR (50 – 70 M protomer),39 varying concentrations of nitrite, and 100 M of the redox mediator 1,2-naphthoquinone-4-sulfonic acid (mo = 168 mV vs SHE, Table S3, and minimal absorbance

Page 4 of 18

in the visible range, section S3 of Supporting Information), was prepared in an appropriate buffer that also contained 200 mM NaCl, and was adjusted to the desired pH. A 50 mM HEPES buffer was used for pH 7.0, and a 50 mM tris buffer for pH 8.0. UV/Visible spectra in the range 300 – 800 nm were collected at 10 mV intervals between +160 mV and +300 mV vs. SHE, after it was clear that the spectral changes of interest occurred in this potential range (section S3, Supporting Information). The data sets were analyzed using programs written within the commercially available software package Mathcad 15 (PTC Software). Data analysis broadly followed the methodology used previously by this group;14 details specific to this work are provided in the Results, and in Supporting Information. Time-resolved Reduction of nitrite-loaded ccNiR by N,N,N′,N′-tetramethyl-p-phenylenediamine. Initially

ccNiR and N,N,N′,N′-tetramethyl-p-phenylenediamine (TMPD) were mixed anaerobically into a solution of 50 mM HEPES buffer, pH 7, with a total volume of 974 L, such that the ccNiR and TMPD concentrations would come to ~10 M and ~1 mM, respectively, when diluted to 1 mL. The solution was monitored by UV/Vis spectroscopy at 30 s intervals for 5 min using one of the glovebox-housed spectrophotometers, after which a 26 L aliquot of nitrite stock was added to the reaction mixture, giving a solution containing 1 mM nitrite. The full reaction mixture was then monitored by UV/Vis spectroscopy every 30 s for another 20 min. The time-resolved data were analyzed using a previously described global fitting routine written within the commercially available software package Mathcad 15 (PTC Software).40-41 Electron paramagnetic resonance (EPR) analysis of S. oneidensis ccNiR. All experiments were carried out in

50 mM HEPES buffer, pH 7, which also contained 200 mM NaCl for electrochemical experiments. Typical ccNiR concentrations for EPR experiments were 50 – 60 M;39 1,2-naphthoquinone-4-sulfonic acid (100 M) was used as the mediator when enzyme reduction was required. For EPR analysis of the 2electron reduced nitrite-loaded ccNiR the samples were prepared in a glovebox using a bulk electrolysis cell with platinum mesh as the working electrode. In a typical experiment 4 mL of ccNiR solution were added to the bulk electrolysis cell, and a potential of 160 mV vs SHE was applied for approximately 3 min intervals with no stirring.42 After each interval the solution was gently mixed by drawing it in and out of a plastic syringe three or four times through a 0.75 mm id PEEK tube. Reduction progress was assessed

ACS Paragon Plus Environment

Page 5 of 18 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

from the current measured at the end of each interval, and by periodically removing subsamples from the cell and checking their UV/Vis spectrum using a 0.02 cm demountable cuvette. When no further changes were detected in the current and UV/Vis spectrum, a 250 µl subsample was removed from the electrolysis cell and transferred into an EPR tube. The EPR tube was then placed into a test tube, which was stoppered with a butyl rubber stopper, removed from the glove box, and immediately submerged in liquid nitrogen. Samples were subsequently stored at 80 C until needed; they were kept on dry ice during transportation to the EPR facility. EPR spectroscopy was carried out at 9.47 GHz on a Bruker EMX-TDU/L E4002501 spectrometer equipped with E532LX digital acquisition, an ER4112 SHQ resonator, a Cold- Edge/Bruker ER4112HV-S5L Stinger cryogen-free 5–100 K closed cycle helium cryocooler, an HP 5350B microwave counter, and an Oxford Instruments ESR900 cryostat and MercuryITC temperature controller. Spectra were typically recorded at 10 K, 5.2 milliwatts of microwave power, 12 G (1.2 mT) magnetic field modulation amplitude at 100 kHz, and 1.2 G (0.12 mT) digital field resolution. Other acquisition parameters were chosen such that the spectral resolution was limited by the modulation amplitude. Background signals were recorded on a frozen water sample and subtracted (Xepr) following field adjustment for precise microwave frequency. RESULTS Structure of S. oneidensis ccNiR. A high-resolution

(1.66 Å) crystal structure of S. oneidensis ccNiR was obtained using the protein purified with the new protocol described in Materials and Methods. Figure S3 compares the structure of the S. oneidensis active site region as obtained in the current study at 110 K, with the 2.59 Å structure obtained previously at room temperature using the Laue method.14 The two structures are seen to be very comparable, showing that neither the temperature difference of data collection nor the purification method had a significant impact. The structures obtained by the two methods are similarly homologous throughout the remainder of the protein, in both protomers. Effect of nitrite on the EPR spectrum of resting S. oneidensis ccNiR. Previous studies showed that the Esche-

richia coli and Desulfovibrio vulgaris ccNiR ferriheme 1 active sites undergo high-spin to low-spin transitions upon binding nitrite.43-44 This transition had minimal effect on the UV/Vis spectrum of the wild type enzyme, but resulted in substantial changes

Figure 2. (a) EPR spectrum of fully oxidized S. oneidensis ccNiR (50 M) in the absence of nitrite.39 (b) Same solution in the presence of 6 mM nitrite. (c) Difference spectrum obtained by subtracting spectrum (b) from Spectrum (a). All of the indicated g values were previously assigned to specific ccNiR hemes; 16 see text for details.

to its EPR spectrum.43 S. oneidensis ccNiR showed analogous behavior, so nitrite binding to it was monitored by EPR. Figure 2a shows the EPR spectrum of as-isolated fully oxidized ccNiR (50 M),39 Fig. 2b shows the result of adding 6 mM nitrite to the sample, and Fig. 2c shows the difference between the two spectra. The EPR signals observed upon incubation of S. oneidensis ccNiR with excess nitrite (Fig. 2b) are very similar to those elicited by addition of a ~5-fold molar excess of cyanide to the enzyme,16 and comparable to those observed for the E. coli enzyme in the presence of excess nitrite.43 By analogy with the previous reports the features with g1 = 2.87, g2 = 2.24, g3 = 1.48 are assigned as a rhombic signal arising from ferriheme 2 (Fig. 1), while the lone broad feature at g = 3.39 is a “highly axial low-spin” (HALS) signal arising from magnetically coupled ferrihemes 3 and 4, possibly overlapping a small contribution from lowspin ferriheme 1 (Fig. 1).16 The spectral contributions from hemes 2 – 5 are essentially unchanged by addition of nitrite, so the difference spectrum of Fig. 2c shows only the signal arising from the fully oxidized, high-spin, nitrite-free active site.45 This signal is complex, exhibiting multiple features. The broad feature with apparent g' > 10, and absorption extending into zero-field,46 is indicative of an integer-spin signal. This feature, together with the peak at g = 3.33 and trough at g = 2.83, have been attributed to the S =

ACS Paragon Plus Environment

Journal of the American Chemical Society

Page 6 of 18

5

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

/2 heme 1, weakly exchange-coupled with one or more of the nearby low-spin hemes (S = ½).11, 16, 43, 4748 The high-spin signal provides an excellent spectroscopic handle for the active site, and titration of ccNiR by nitrite was followed by monitoring its disappearance as nitrite was added. A spectrum was obtained at each nitrite concentration, and in each case the spectrum obtained in the presence of 6 mM nitrite (Fig. 2b) was subtracted from it, to generate a difference spectrum analogous to that of Fig 2c. Figure 3 shows how the fractional occupancy of the active site heme by nitrite changes with added nitrite concentration. The experimental fractional occupancy at the ith nitrite concentration (blue circles) was calculated using Eq. 1, where [FeH1III(NO2)] is the concentration of the nitrite-loaded active site at a given total nitrite concentration NT, FeT is the total ccNiR concentration, (S3.33  S2.83)i is the g = 3.33 peak to g = 2.83 trough separation obtained from the difference spectrum at that nitrite concentration, and (S3.33  S2.83)0 is the corresponding amplitude difference for the nitrite-free sample (Fig 2c). The experimental

[ FeH 1III ( NO2  )]i ( S  S 2.83 )i  1  3.33 FeT ( S3.33  S2.83 )0

Eq. 1

data were modelled reasonably well using Eq. 2 (red trace, Fig. 3), where K is the equilibrium constant for nitrite association with the ferriheme 1 active site. The derivation of Eq. 2 is given in section S4 of Supporting Information. In a least-squares fitting procedure the value of K was systematically varied, and [ FeH 1III ( NO2  )]  ( FeT  NT )  K  1  FeT 2  K 2  NT  2 FeT  K  K 2  NT 2  2 K  NT  1 2K

Eq. 2 III



values of [FeH1 (NO2 )]/FeT at each of the nitrite concentrations shown in Fig 3 were calculated using Eq. 2. The best fit to the experimental data was obtained with a nitrite binding constant K = (5.00.1)103 M1, which translates to a nitrite dissociation constant of K1 = (2.010.04)104 M. Effect of nitrite on the electrochemical properties of S. oneidensis ccNiR. Figure 4 shows the UV/Visible spec-

tral changes observed upon decreasing the potential applied to a solution initially containing 76 M of fully oxidized ccNiR, 100 M 1,2-naphthoquinone-4sulfonic acid and 5 mM nitrite.39 The absorbance increase at 424 nm and decrease at 402 nm are typical of low-spin ferric c-heme reduction;49-50 however, the increases ca. 525 nm and 550 nm that are also charac-

Figure 3. Blue circles: fraction [FeH1III(NO2)]/FeT as a function of nitrite concentration, calculated using Eq. 1. Red line: leastsquares best fit to the experimental data using Eq. 2. The best fit was obtained with a nitrite binding constant K = (5.00.1)103 M1, which translates to a nitrite dissociation constant of K1 = (2.010.04)104 M. The uncertainties reported for K and K1 represent the changes that can be made to the parameters without changing the sum of squares by more than 2%.

teristic of low-spin ferric c-heme reduction are absent, and a single broad increase centered at 555 nm is seen instead. Slices taken in the applied potential domain (Fig. 4b) clearly show a reduction event centered near 240 mV vs SHE. This is much higher than the midpoint potential obtained for the first reduction step of S. oneidensis ccNiR in the absence of strong-field ligands, and much higher even than that obtained in the presence of the strong-field ligand cyanide (20 mV vs NHE).16 Singular value decomposition (SVD) analysis,16, 51-52 and the presence of tight isosbestic points at 341 nm and 415.5 nm in Fig. 4a, reveal only a single spectral component attributable to c-heme reduction. The Fig. 4 data were well fit with a single Nernstian equation (Scheme 1, Eq. 3), in which A,app is the absorbance change observed at wavelength  upon application of potential app, mo is the midpoint potential associated with the reduction event, A(max) is the limiting absorbance change at wavelength  observed at sufficiently low applied potentials, and n is the number of electrons transferred in the Nernstian step. In Eq. 3 app and A,app are the independent and dependent variables, respectively, while mo and A(max)are parameters obtained by non-linear leastsquares fitting of the Fig. 4 data set. The fitting process was performed twice, once with n fixed at 1, and the second time with n fixed at 2.

ACS Paragon Plus Environment

Page 7 of 18 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Figure 4. (a) UV/Vis spectral changes obtained upon exposing a solution initially containing 76 M of fully oxidized ccNiR, 100 M 1,2-naphthoquinone-4-sulfonic acid and 5 mM nitrite to progressively lower potentials.39 Solid blue lines show the experimentally obtained data, whereas the dashed red lines were calculated from least-squares fitting with a Nernstian 2-electron reduction and one spectral component (Scheme 1 and Supporting Information). The fit yields a midpoint potential of mo = 0.246 V. (b) Blue circles: an absorbance difference vs. applied potential slice taken at 424 nm from the spectra of part (a); solid red line: least-squares best fit obtained from global analysis of the data with n fixed at 2 electrons transferred in the Nernstian equation; dotted green line: same as the red line, but with n fixed at 1 electron transferred in the Nernstian equation.

Scheme 1. Nernst equation used to fit the Fig. 4 data. Here Cox refers to fully oxidized ccNiR, and Cred to the enzyme after a single reduction by n electrons. The remaining parameters are as defined in the main text.  Cred Cox +ne  

A , app

 nF 0  Amax(  )  exp  ( m   app )  RT     nF 0  1  exp  ( m   app )   RT 

Eq.3

From the least-squares fit the midpoint potential was calculated to be mo =  V, and the number of electrons transferred was n = 2. The latter is unusual given that in all previous studies ccNiR heme reductions invariably proceeded via one-electron steps (n = 1 in Eq. 3);16 however, the least-squares fit with n fixed at 2 (red trace in Fig. 4b) is clearly superior to that obtained with n fixed at 1 (dotted green trace in Fig. 4b). Scheme 2 provides a chemical interpretation for the two-electron reduction seen in Fig. 4. In the first step of Scheme 2 nitrite binds to the ccNiR heme 1 active site. Using the binding constant K = (5.00.1)103 M1 obtained from EPR (Figs. 2 and 3 above) one can estimate that, in the presence of 5 mM nitrite, the heme 1 active site will be ~96% occupied in the fully oxidized enzyme. Two-electron reduction of the nitrite-loaded active site will then give a species formally referred to as {FeNO}7 using the Enemark-Feltham notation, where {FeNO}7 represents resonance forms of a species that formally has NO bound to FeII, or NO bound to FeIII (Scheme 2).53 The parameter o associated with step 2 of Scheme 2 represents the true standard reduction potential for the process. Thus it appears that, when nitrite is bound at the ccNiR active site, the first enzyme reduction event proceeds via a concerted 2-electron process that to some extent distributes the electrons between the iron and nitrogen moieties (Scheme 2). Scheme 2 predicts that the apparent midpoint potential mo measured by using the empirical Scheme 1 Nernst equation should be pH-dependent, shifting 59 mV per pH unit at 25 C (section S5, Supporting Information). In two spectropotentiometric titrations carried out at pH 8 instead of pH 7 the observed midpoint potential for ccNiR 2-electron reduction was 0.2070.002 V, a negative shift of 39 mV from the value obtained at pH 7. This is less than the predicted shift, but a significant shift nonetheless. Scheme 2 also predicts that the apparent midpoint potential obtained by fitting spectropotentiometric data with the empirical Nernstian of Scheme 1 should decrease if the nitrite concentration is lowered to the point where a significant fraction of the ccNiR active sites are vacant instead of nitrite-loaded. The relationship between mo and nitrite concentration is complex, but it can be readily modeled theoretically, as described in section S5 of Supporting Information, by combining the equilibrium expression of equation 2 with the Nernst equation 3. The relationship depends on the two parameters shown in Scheme 2: the standard re-

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 18

Scheme 2. Chemical interpretation of the nitrite-loaded ccNiR 2-electron reduction. FeH1 represents the active site heme center in various states of oxidation and ligation; see text for details.

duction potential o (adjusted for pH), and the binding constant K. Figure 5 shows how the apparent midpoint potential varies with nitrite concentration. The blue circles show the experimental data obtained with the empirical Nernst equation (Scheme 1 Eq. 3), while the red squares show the theoretical values calculated as described in Supporting Information. The theoretical calculations assumed a standard reduction potential o of  V, which is the mo value obtained at pH 7 using Eq. 3, when most of the heme 1 sites were nitrite-occupied (5 – 20 mM nitrite). However, to obtain the red square data, the value of K was treated as an adjustable parameter in a least-squares fitting routine. The fitting routine yielded the best fit with K = 2600 M1. This value is about half that obtained from the EPR titration described earlier (Fig. 3), but close to the value obtained by Lockwood et al. for E. coli ccNiR (2500300 M1).43 Moreover, given the scatter in the Fig. 5 data, and the small range within which mo changes with nitrite concentration, even the agreement with our EPR titration is reasonable. Indeed, modelling the Fig. 5 data with K fixed at 5000 M1, the value obtained by EPR (Fig. 3), shifts the calculated mo values at each nitrite concentration by at most 7 mV, and generates a theoretical curve that retains the shape of the one obtained by least-squares fit (Fig. 5, green triangles). For completeness Fig. 5 also shows a curve obtained by fixing K at 1400 M1 (purple diamonds), which gave the same sum of squares as obtained with K = 5000 M1. Again the shape of the curve is similar to that of the leastsquares fit. Together, the green and purple curves provide a measure of the sensitivity of the fits to the K value. Most importantly, the midpoint potential calculated using the empirical Nernst equation 3 is seen to vary as predicted theoretically when the nitrite concentration decreases to the point where a significant fraction of the ccNiR active sites are vacant. Finally we note that, in the absence of a strong field ligand such as nitrite, none of the ccNiR hemes reduce at applied potentials greater than 0 V vs SHE,14, 16 so

Figure 5. Blue circles: dependence on nitrite concentration of the apparent midpoint potential obtained by fitting UV/Vis spectropotentiometric data to the empirical Nernst equation (Scheme 1, Eq. 3). Red squares, green triangles and purple diamonds: predicted values of the apparent midpoint potential at a given nitrite concentration, calculated as described in Supporting Information, using o = 0.246 V; the red square values were obtained by varying K to yield the least-squares best fit (K = 2600 M1), while the green triangle values were obtained with K fixed at 5000 M1, and the purple diamond values with K fixed at 1400 M1 (see text for details).

only nitrite-loaded ccNiR can reduce at the high potentials monitored in these experiments. Reduction of nitrite-loaded S. oneidensis ccNiR by N,N,N′,N′-tetramethyl-p-phenylenediamine. The blue

traces in Fig. 6a show the UV/Vis spectral changes observed after mixing S. oneidensis ccNiR with large excesses of nitrite, and the weak reductant N,N,N′,N′tetramethyl-p-phenylenediamine (TMPD, mo = 260 mV vs SHE), at pH 7. Initially ccNiR and 1 mM TMPD were mixed into a solution with a total volume of 974 L. The green trace in Fig. 6a shows the UV/Vis spectrum of this solution, scaled for future dilution to 1 mL. The purple dotted trace that overlays the green one was obtained by least-squares fitting of the green trace with the independently obtained extinction coefficient spectrum of oxidized ccNiR. The fit shows that the solution, once diluted to 1 mL, would contain 11 M of fully oxidized ccNiR. The ccNiR-TMPD solution was monitored by UV/Vis at 30 s intervals for 5 min, and remained unchanged. After 5 min a 26 L aliquot of nitrite stock was added to the reaction mixture, giving a solution containing 1 mM nitrite. The first blue trace above the green one was obtained within 30 s of adding the nitrite, and was set as the t = 0 spectrum. Subsequently spectra were collected every 30 s, but for clarity the blue traces in Fig 6a are at 120 s intervals; Figure 6b

ACS Paragon Plus Environment

Page 9 of 18 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Figure 6. (a) Blue traces: spectral changes seen at 120 s intervals after mixing 11 M oxidized ccNiR with 1 mM each of nitrite and TMPD. Red dashed traces: least-squares best fits using Eq. 4. Green trace: spectrum of the solution containing 11 M oxidized ccNiR and 1 mM TMPD, but prior to addition of nitrite; purple dashed trace: least-squares best fit obtained with the independently known oxidized ccNiR extinction coefficient spectrum. (b) Absorbance vs. time slice from the Fig. 6a spectrum, obtained at 611 nm.

shows an absorbance vs time slice taken at 611 nm. SVD analysis51-52 showed that only two components were needed to faithfully reconstruct a noise-reduced absorbance matrix. The SVD-treated data were then fit to Eq. 4 using a global fitting routine (red traces, Fig. 6a).40-41 In Eq. 4 A,t is the absorbance obtained

A ,t   0(  )  1(  ) (t  q t 2 )

Eq. 4 at wavelength  and time t, the spectral component Λ0 is present immediately after adding nitrite (taken as t = 0), and component Λ1 grows in almost linearly with time (the manually-adjusted parameter q in Eq. 4 al-

Figure 7. Spectral components  0 (a) and 1 (b) generated by fitting the SVD-processed Fig 6 data to Eq. 4 (blue traces). 0 was in turn fit with the extinction coefficient spectra of reduced nitrite-loaded ccNiR (obtained as shown in Supporting Information), and of the blue TMPD+ radical,54 while  1 was fit using only the TMPD+ extinction coefficient spectrum. The red dashed traces show the calculated spectra in each case.

lows small admixture, 0.016%, of a quadratic term that improves the empirical fit). Figure 7 shows the spectral components obtained from the fitting process. The component Λ0 (Fig. 7a) was in its turn fit with the extinction coefficient spectrum of reduced nitrite-loaded ccNiR (obtained from spectropotentiometric titrations, as shown in section S4 of Supporting Information), and that of the blue TMPD+ radical.54 The fit shows that immediately after mixing the solution contained 11.0 M reduced ccNiR and 22.1 M TMPD+, a ratio of 2.03 TMPD+:1 ccNiR. This experiment thus provides independent confirmation that, under weakly reducing conditions, nitrite-loaded ccNiR is reduced in a concerted 2electron process as proposed in Scheme 2; in the pre-

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

sent case the two electrons are supplied by two equivalents of TMPD. The spectral component Λ1 that grows in approximately linearly (Fig. 7b) was fit with the extinction coefficient spectrum of the TMPD+ radical,54 and shows that TMPD continued to oxidize at a rate of ~4.3108 Ms1 after the initial oxidation burst. In a separate study we have shown that ccNiR catalyzes the 1-electron reduction of nitrite to nitric oxide by weak reductants such as ferrocyanide, TMPD and Ru(NH3)62+. This study is being prepared for publication as a separate paper; for the moment we note only that TMPD oxidation during the linear phase is quite slow compared to that associated with the initial reduction of FeH1III(NO2) to {FeH1NO}7, allowing the two processes to be readily deconvoluted by conventional time-resolved spectroscopy. Thus, Figs. 6 and 7 show that ccNiR-catalyzed reduction of free nitrite by TMPD yields ~1.3 M TMPD+ in the first 30 s after reaction initiation, during which time reduction of FeH1III(NO2) to {FeH1NO}7, which generates ~22 M TMPD+, is virtually complete. Preliminary stopped-flow data show that the half-life for {FeH1NO}7 formation is about 6 s (data not shown). Investigation of the S. oneidensis ccNiR {FeH1NO}7 moiety by EPR spectroscopy. To generate ccNiR containing

the {FeH1NO}7 active site for characterization by EPR, a sample was prepared by poising the enzyme at 160 mV vs SHE in the presence of a large excess of nitrite. EPR spectroscopy subsequently revealed two hitherto unobserved sets of features (Fig. 8). The first set exhibited turning points at g = 2.18, 2.12 and 2.05, with a resolved inflection on the crossover feature and a "truncated" trough feature at g = 2.05. The overall line shape was evocative of a low-spin S = ½ {FeNO}7 species, but the g-values and spectral envelope width were not quite consistent with literature values.55-60 Interpretation was further complicated by the presence of an equally broad overlapping signal, clearly of different origin, that was observed without poising the potential (Fig. 8, top), and by a sharp feature at g = 2.0. Attempts to detect 14NI = 1 superhyperfine splittings using lower modulation amplitude (5 G, 0.5 mT) and higher derivatives were unconvincing. Samples were generated using 15NO2 in an attempt to at least detect differences in the g ~ 2 region due to the differences in N (15N > 14N) and the line multiplicity [2 (15NI = 1/2) vs. 3 (14NI = 1)], but the 14NO and 15 NO spectra were indistinguishable within the signalto-noise obtainable with 64 M ccNiR (Fig. 8, bottom). Spin integration showed that the g = 2.2 - 2.0 region contributes a substantial fraction to the total

Page 10 of 18

Figure 8. Green trace: X-band EPR spectrum obtained for fully oxidized ccNiR in the presence of 5 mM nitrite. Blue trace: spectrum obtained for a solution initially containing 64 M of fully oxidized ccNiR and 500 M 14NO2, after exposing it to an applied potential of 160 mV vs SHE.39 Red trace: same as blue trace, but using 500 M 15NO2.

spin envelope, at least twice that contributed by the rhombic heme 2 signal (Fig. S10, Supporting Information). We conjecture that the g = 2.2 - 2.0 features arise from the low-spin {FeNO}7 active site interacting magnetically with nearby hemes, which would explain why they are atypical. To our knowledge there has been only one prior report of heme-heme magnetic interaction that involved an {FeNO}7 species;61 however, such an interaction in ccNiR is plausible given that the active site is known to interact with heme 3, and possibly also heme 4, in the resting enzyme.11, 16, 43, 47-48 The second set of features that appears with 2electron reduction of nitrite-loaded ccNiR is far easier to assign. The co-appearance of a sharp peak with g' = 4.1,46 a sharp crossover at g' = 3.9, and a sharp inverted peak at g' = 2.0 (Fig. 8 middle and bottom spectra) are diagnostic for the MS = 1/2 doublet of an S = 3/2 spin system with |D| >> gBB0S, and a very small rhombic zero-field splitting (E/D  0.015). This signal was observed with multiple batches of ccNiR, using concentrations of nitrite from 500 M – 5 mM and, in each case, the intensity of the S = 3/2 signal measured by EPR correlated with the extent of heme c (i.e. heme 1) reduction determined by electronic absorption spectrophotometry. Under conditions of maximum expression, the S = 3/2 signal’s contribution to the spin envelope is approximately 20% that of the rhombic heme 2 signal (Fig. S10, Supporting Information). The only plausible origin for a signal of this

ACS Paragon Plus Environment

Page 11 of 18 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

type in the ccNiR preparations is a high-spin {FeNO}7 system, likely arising predominantly from antiferromagnetic coupling of high-spin ferric iron to a bound triplet-state NO moiety,55, 62 though the rhombicity allows for some (S = 2, FeII)(S = 1/2, NO) character.63 Such species are relatively common in both protein complexes and model compounds, but only in non-heme iron systems. As mentioned above, all known hemoprotein {FeNO}7 complexes are lowspin with S = 1/2, and are generally described as predominantly low-spin ferroheme (S = 0) with S = 1/2 NO bound.55-60, 64-67 Upadhyay et al. reported an S = 3/2 {FeNO}7 signal in the EPR spectrum of the tetraheme protein cytochrome c554 when this was nitrosylated and 2-electron reduced,61 but attributed it to a minority non-heme impurity in their preparations. The S = 3/2 signals in Fig. 8 may also be due to a non-heme impurity, but two considerations argue against such an assignment. First, SDS-PAGE analysis showed that the ccNiR used in the EPR experiments was of very high purity (Fig. S1, Supporting Information), making it unlikely that the S = 3/2 signal arises from a non-heme protein. The S = 3/2 {FeNO}7 signal could still arise from adventitious iron in the preparation. To test for this possibility a ccNiR sample identical to the ones used to generate the Fig. 8 spectra, but containing 25 M Cu2+ as internal standard, was subjected to ultrafiltration to generate a protein-free filtrate. The filtrate was then treated with enough of the NO precursor DEANO to add 60 M NO to the solution, after which it was immediately reduced electrochemically at an applied potential of 160 mV vs. SHE. EPR analysis of the oxidized and reduced filtrates showed that the total adventitious iron in the preparations was minimal compared to the Cu2+ internal standard (~0.7%, Figs. S11, S12, Supporting Information), and that the S = 3/2 {FeNO}7 signal arising from the filtrate was substantially smaller than that seen in the ccNiR-containing precursor samples (sections S6 and S7, Supporting Information). It is still possible that the S = 3/2 {FeNO}7 signal associated with the ccNiRcontaining sample arises from adventitious iron strongly adsorbed to the protein, though there is no evidence of such iron in the high-resolution ccNiR crystal structure. DISCUSSION Behavior of nitrite-loaded S. oneidensis ccNiR under weakly reducing conditions. The UV/Vis spectropotentiom-

etric results of this study (Figs. 4 and 5) show that, when occupied with nitrite, initial reduction of the

ccNiR active site occurs by a concerted 2-electron process, which thus generates an {FeNO}7 moiety (Scheme 2). The stoichiometry of the reduction was independently verified in a time-resolved UV/Vis spectroscopy experiment that tracked reduction of the nitrite-loaded heme 1 active site by the weak reductant TMPD, which generates blue TMPD+ upon oxidation (Figs. 6 and 7). This experiment clearly showed that two equivalents of TMPD were oxidized for every equivalent of nitrite-loaded ccNiR that was reduced. The TMPD experiment, together with other experiments that used ferrocyanide and Ru(NH3)62+ electron donors (to be published separately), also showed that ccNiR catalyzes nitrite reduction to nitric oxide when weak reductants are used. However, NO formation is very slow when compared to the initial 2-electron reduction of FeH1III(NO2) to {FeH1NO}7, so the two processes are readily deconvoluted (Figs. 6 and 7). EPR characterization of the 2-electron reduced nitriteloaded active site subsequently showed it to possess a novel spectral signature that is different from those normally observed for {FeNO}7 hemes (Fig. 8). Such hemes are typically low-spin S = ½, with spin density mainly on the nitrogen. They have very characteristic features in the g ~2.0 – g ~2.1 range, which usually display at least some degree of hyperfine interaction with the 14NO or 15NO nitrogen.55-60 The spectra for reduced ccNiR in Fig 8 show prominent features in the g ~2.0 – g ~2.2 range, but these look different from the usual low-spin {FeNO}7 heme features, and are not affected by isotopic exchange from 14N to 15N . The features have the right intensities to arise from the low-spin {FeH1NO}7 active site interacting magnetically with nearby hemes (Fig S10, Supporting Information), and this makes a plausible working hypothesis that has a precedent.61 Further testing of this hypothesis will require extensive EPR studies of nitrite-loaded reduced ccNiR at varying temperatures and applied frequencies, possibly buttressed by computational analysis and application of other spectroscopic techniques such as Mössbauer. Such studies are beyond the scope of this report, but are planned for the future. The EPR spectrum of 2-electron reduced nitriteloaded ccNiR also revealed signals characteristic of a high-spin S = 3/2 {FeNO}7 species. The signal has an intensity approximately 20% that of ccNiR heme 2, and might be dismissed as arising from non-heme impurities, except that such an explanation seems inconsistent with the high purity of the samples used in the EPR experiments. SDS-PAGE showed that non-

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ccNiR protein contaminants were less than 5% (Fig. S1), and spin quantitation of the protein-free reaction medium showed that the adventitious iron concentration was similarly low (Figs. S11, S12). Based on these considerations we don’t yet discount the possibility that the high-spin S = 3/2 {FeNO}7 signal actually arises from a small fraction of the ccNiR population in which the active sites have a slightly different heme environment from that of the majority. An S = 3 /2 {FeNO}7 heme would be unprecedented in heme chemistry, and more studies will be needed to explore this possibility. However, there is some independent evidence that 2-electron reduced nitrite-loaded ccNiR can exist in a variety of forms, depending on conditions, as will now be explained. The earliest spectropotentiometric investigation of nitrite-loaded ccNiR performed in our laboratory gave results quite different from those obtained ever since. In this initial single study reduction of nitrite-loaded ccNiR still took place in a concerted 2-electron step, but the observed midpoint potential was 20 mV vs SHE instead of 246 mV (Fig. S13, Supporting Information). Furthermore, the EPR spectrum of the reduced protein also looked different, as shown in Fig. S14 of supporting Information. This spectrum showed a peak feature at g = 2.08 and a broad trough with g ~ 1.9 – 2.04, which together are much more like a typical S = ½ {FeNO}7 signal than the features seen ever since for the S. oneidensis ccNiR. There was only one immediately identifiable difference between the preliminary experiment and all subsequent ones. The ccNiR used in the early experiment was purified by the method of Youngblut et al, in which the final chromatographic step was hydrophobic interaction chromatography with a high-performance column at pH 7.14 In every experiment since then the ccNiR was purified as described in the Experimental Procedures section, whereby the final step involved cation exchange chromatography at pH 6.5. The ccNiR purified by the two methods was otherwise indistinguishable in its properties. In particular both methods yielded ccNiR whose electrochemical properties were comparable in the presence of the strongfield ligand cyanide, and that had comparable kcat and Km values in the standard assay (ccNiR-catalyzed reduction of nitrite by methyl viologen monocation radical). Furthermore, the crystal structure of the enzyme purified by the method described herein shows no obvious differences from the structure obtained with the earlier purification protocol.14 It thus seems that ccNiR can possibly exist in two very closely related forms that nevertheless exhibit significantly dif-

Page 12 of 18

ferent heme 1 electrochemical properties in the presence of nitrite. To further probe this conjecture we purified a batch of ccNiR using the earlier protocol,14 in which the protein was never exposed to a pH below 7, to see if the putative alternate form could be reisolated. However, in this case the ccNiR purified using the earlier protocol exhibited the same electrochemical properties in the presence of excess nitrite that we have observed ever since, namely a midpoint potential of ~240 mV vs SHE. Thus, for now, we have no obvious explanation for our one anomalous early result. Implications of the results for the ccNiR reaction mechanism. Physically, a concerted 2-electron reduction

indicates that addition of a second electron to a redoxactive center is much more favorable than addition of the first. Under these circumstances, applying a potential low enough to prompt the first reduction step simultaneously initiates the second step as well. In the case of ccNiR, 1-electron reduction of the FeH1III(NO2) active site would yield the species referred to as {FeH1NO}6 (dashed arrow, Scheme 3), which represents resonance forms of a moiety that formally has NO bound to FeIII, or NO bound to FeII.53 Ostensibly 1-electron reduction of the FeH1III(NO2) active site to {FeH1NO}6 is more thermodynamically difficult than the subsequent reduction of {FeH1NO}6 to {FeH1NO}7, so the intermediate {FeH1NO}6 doesn’t accumulate, and FeH1III(NO2) appears to reduce directly to {FeH1NO}7 as shown by the solid arrow in Scheme 3. Evolution of ccNiR so that it reduces bound nitrite in a 2-electron process makes sense from a physiological standpoint. Extensive studies with both model and hemoprotein systems have shown that NO typically binds more tightly within {FeNO}7 species when compared to{FeNO}6 species.67-72 Thus, by disfavoring the {FeNO}6 intermediate, the ccNiR design ensures that release of toxic NO is kept to a minimum during catalytic turnover. The experiment with TMPD (Figs. 6 and 7), and similar ones that used the weak reductants Fe(CN)64 or Ru(NH3)62+, highlight this point. With any of these reductants the rate of NO generation was found to be ~10 – 15 M/h per 1M ccNiR, which is very slow. For comparison, kcat for ccNiR-catalyzed reduction of nitrite to ammonia by methyl viologen monocation radical is ~ 800 s1 in the standard ccNiR assay, which uses picomolar quantities of ccNiR.14 The conditions employed in this study, whereby the oxidized active site was first saturated with millimolar

ACS Paragon Plus Environment

Page 13 of 18 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Scheme 3. Here FeH1 represents the active site heme center in various states of oxidation and ligation. The solid arrows show the net reduction observed experimentally, while the dashed ones show the steps that likely underlie the observed process; see text for details.

concentrations of nitrite and then reduced, are unlikely to prevail under the physiological conditions most favorable for ammonia production. Instead, active site reduction likely precedes nitrite binding in vivo.5, 73 As a strong -acceptor, nitrite will bind much more strongly to the electron-rich ferrous active site, thus allowing the site to be fully occupied at much lower concentrations. The relative order of nitrite and electron addition, however, will not affect the thermodynamic conclusions made above, as is shown in Scheme 4. In either case NO release from subsitutionally labile {FeH1NO}6 will be minimized if the standard reduction potential for reduction of {FeH1NO}6 to {FeH1NO}7 is much higher than that for reduction of FeH1III(NO2) to {FeH1NO}6, so that the labile species can’t accumulate. Computational studies described in previous reports predicted that a {FeH1NO}7 active site will constitute a thermodynamic minimum in ccNiR.5,74-75 This in turn led to the suggestion that {FeH1NO}7 constitutes a kinetic trap that sits off of the reaction pathway from nitrite to ammonia, and to suggestions for how the species could be bypassed during catalysis.5,74 The present results confirm that {FeH1NO}7 at the active site constitutes a thermodynamic minimum, but in no way rule out its possible role as a catalytic intermediate, based on the following analysis that i s

Scheme 4. In the experiments described herein active site reduction in the presence of nitrite appears to proceed via Path 1. Under physiological conditions it will likely proceed via Path 2. In either case NO release from subsitutionally labile {FeH1NO}6 will be minimized if the standard reduction potential for reduction of {FeH1NO}6 to {FeH1NO}7 is much higher than that for reduction of FeH1III(NO2) to {FeH1NO}6.

represented schematically in Fig. 9. We calculate the standard free energy G for reduction of free nitrite to ammonia at pH 7 by a standard hydrogen electrode to be a highly favorable ~ 200 kJ/mol (Fig. 9).18, 76 Thus, at an applied potential of 0 mV vs SHE, thermodynamics predict that nitrite reduction to ammonia should be almost quantitative. The results of the present study show that initial nitrite reduction at the active site can occur at applied potentials as high as  mV vs SHE, but that at these high potentials only {FeH1NO}7 is generated. In other preliminary spectropotentiometric studies we have seen no further reduction of {FeH1NO}7 until app reaches ~ 80 mV vs SHE (Fig. S13, Supporting Information).77 Furthermore, protein film voltammetry (PFV) experiments used to probe the catalytic reduction of nitrite to ammonia by electrode-adsorbed S. oneidensis ccNiR have shown that no significant catalytic ammonia formation occurs at potentials above 120 mV vs SHE.28-29 If an applied potential of 0 mV vs SHE is theoretically low enough to reduce free nitrite to ammonia, then the additional 120 mV overpotential required for turnover in the PFV experiments could well be needed to extricate {FeH1NO}7 from its thermodynamically stable hole, as shown schematically in Fig. 9. The stability of {FeH1NO}7 may not be the only barrier to ultimate formation of ammonia. Earlier rapid-

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 9. Cartoon representation of the ccNiR-catalyzed nitrite reduction reaction path, proposed on the basis of the results reported herein. The standard free energies are per mole of product formed, and where applicable, are calculated relative to the standard hydrogen electrode. The question mark denotes an as-yet uncharacterized putative 4-electron reduced intermediate, for which there is some preliminary evidence.17 We propose that weak reductants can’t provide the overpotential necessary to overcome the depicted high activation barrier, thus allowing intermediates to accumulate. Only when a strong reductant such as methyl viologen monocation radical is used does nitrite get reduced directly to ammonia, without accumulation of detectable intermediates.

mixing experiments also revealed a kinetic barrier to the final 2-electron reduction of hydroxylamine to ammonia, by demonstrating that ccNiR cannot catalyze the disproportionation of hydroxylamine to ammonia and nitrite.17 In this disproportionation the reductive half reaction is highly favorable thermodynamically, and should drive the much less favorable oxidative one. Yet in the rapid mixing studies no ammonia was ever produced, while the ccNiR hemes could readily accept electrons from hydroxylamine to generate small amounts of nitrite and heme-bound nitrogenous species.17 It appears that a kinetic barrier to hydroxylamine reduction can only be overcome by applying an overpotential greater than can be supplied by hydroxylamine oxidation, so disproportionation does not occur.17 Based on the combined results of the present study, and the earlier rapid-mixing and PFV investigations,17, 28-29 we propose that the rate-limiting step for ccNiR-catalyzed reduction of nitrite to ammonia is the final 2-electron reduction of hydroxylamine (or a heme-bound hydroxylamine equivalent, Fig. 9),17 and that this step requires a strong reducing agent to overcome it. The hypothesis is consistent with the fact that the kcat value, as measured in electrons transferred per second, is the same for nitrite

Page 14 of 18

and hydroxylamine reduction to ammonia, when methyl viologen monocation radical is the electron source.14 The hypothesis that the final 2-electron reduction is rate-limiting suggests new considerations for ccNiR design parameters. To begin with, avoiding {FeH1NO}7 as a thermodynamic trap becomes less important; a reductant strong enough to reduce ccNiR-bound hydroxylamine (or its equivalent) will also be strong enough to reduce the {FeH1NO}7 intermediate. On the other hand, the hypothesis upends the long-held assumption that ccNiR always catalyzes reduction of nitrite directly to ammonia, without generation of intermediates. While this may be true in the presence of strong reductants such as methyl viologen monocation radical, it is clearly not true under less strongly reducing conditions, whereby the barrier to hydroxylamine reduction can’t be overcome. This is why the {FeH1NO}7 intermediate could be so readily characterized herein. CcNiR’s ability to capture nitrite as {FeH1NO}7 under weakly reducing conditions may be physiologically important for facultative anaerobes such as S. oneidensis. In such bacteria the periplasm can go from being a reducing environment to being an oxidizing one as the bacteria move from anaerobic to aerobic surroundings.78 Unlike in E. coli, in S. oneidensis multi-heme proteins such as ccNiR are expressed under both aerobic and anaerobic conditions, and it’s possible that ccNiR has evolved to capture and hold partially reduced nitrite products rather than to release them, when full reduction to ammonia is not possible. For example, as discussed above, the {FeH1NO}7 moiety would be an effective intermediate for capturing nitrite at high potentials, while minimizing release of NO. Under more reducing conditions that nevertheless fail to provide the overpotential needed for hydroxylamine reduction, other adaptations may help to minimize escape of hydroxylamine or nitrous oxide, two other nitrogenous intermediates that could be generated by ccNiR (Fig. 9);17, 79 future studies to explore this possibility are planned. Summary. Spectropotentiometric and time-resolved UV/Vis experiments have shown that nitrite-loaded S. oneidensis ccNiR can be reduced at high applied potential in a concerted 2-electron step to generate an {FeNO}7 moiety at the heme 1 active site. EPR spectropotentiometry subsequently revealed that the {FeH1NO}7 moiety possesses an unusual spectral signature that is different from those normally observed for {FeNO}7 hemes, and may indicate magnetic interaction of the active site with nearby hemes. The sim-

ACS Paragon Plus Environment

Page 15 of 18 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

plest hypothesis consistent with the results presented herein is that an {FeH1NO}7 moiety is a catalytic intermediate in the ccNiR-mediated reduction of nitrite to ammonia (Fig. 9), whose degree of accumulation depends exclusively on the applied potential. At low potentials the species is rapidly reduced and doesn’t accumulate, while at higher potentials it is trapped, thus stalling catalytic ammonia formation. The {FeH1NO}7 species may be the first of a number of ccNiR catalytic intermediates that can be isolated simply by tuning the applied potential. Earlier rapidmixing studies of ccNiR-hydroxylamine mixtures also hinted at various species that are formally fourelectron reduced relative to the resting enzyme (represented by a question mark in Fig. 9), and that may accumulate at applied potentials between 80 mV and 120 mV.17, 77 Experiments are under way to further characterize these four-electron reduced species. Future studies will also need to explore the structural basis for the very high midpoint potential of the nitrite-loaded heme 1 active site, as well as for its unusual spectroscopic properties. Finally it should be noted that, while the present study shows the first step in ccNiR-catalyzed nitrite reduction to be a concerted 2-electron process, earlier PFV experiments that probed electrocatalytic nitrite reduction by electrode-absorbed S. oneidensis ccNiR consistently showed the rate-limiting step to be a oneelectron event.28-29 This apparent discrepancy illustrates the complementary nature of the PFV and spectropotentiometric approaches. While in the former case steady-state electrocatalysis is monitored, here the initial redox events associated with heme spectral changes are examined. Thus, in pre-steady-state conditions, two-electron redox chemistry appears essential to poise the active site, while subsequent oneelectron transfer steps may ultimately control the rate of catalytic ammonia formation. ASSOCIATED CONTENT Supporting Information Verification of ccNiR purity by SDS-PAGE; S. oneidensis ccNiR structure refinement; Mediators used for spectropotentiometric experiments; Derivation of the equilibrium expression for nitrite binding to the ccNiR active site; Dependence of nitrite-loaded ccNiR’s electrochemical behavior on pH and nitrite concentration; Spin quantitation of the nitrite-loaded ccNiR EPR spectrum after reduction; Quantitation of adventitious iron in the EPR spectra; Outlier result from an early experiment. The Supporting Information is available free of charge on the ACS Publications website.

AUTHOR INFORMATION Corresponding Authors A. Andrew Pacheco, Department of Chemistry and Biochemistry, University of Wisconsin-Milwaukee, Milwaukee, Wisconsin 53211, United States; Telephone: (414) 229-4413; Fax: (414) 229-5530; E-mail: [email protected]. Brian Bennett, Department of Physics, Marquette University, Milwaukee, WI 53233, United States; Telephone: 414-288-6705; E-mail: [email protected].

ACKNOWLEDGMENT This work was supported by National Science Foundation grants MCB-1330809 and MCB-1616824. EPR studies were supported by a National Science Foundation Major Research Instrumentation award (CHE-1532168), Marquette University, and Bruker Biospin. We thank Dr. Jason Tenboer, Zeiss USA, for collecting and processing the ccNiR cryo data with the X-ray home source. We are grateful to Sean J. Elliott and Evan T. Judd from Boston University for their valuable input in the early stages of this project, and to George B. Richter-Addo and Erwin B. Abucayon from the University of Oklahoma, who read the manuscript prior to its submission, and provided many helpful suggestions.

REFERENCES 1. Simon, J., Enzymology and bioenergetics of respiratory nitrite ammonification. FEMS Microbiol. Rev. 2002, 26, 285-309. 2. Ehrlich, H. L., Geomicrobiology. 4th ed.; Marcel Dekker, Inc: New York, NY, 2002. 3. Fenchel, T.; King, G. M.; Blackburn, T. H., Bacterial Biogeochemistry. 2nd ed.; Academic Press: London, 1998. 4. Poock, S.; Leach, E.; Moir, J.; Cole, J.; Richardson, D., Respiratory Detoxification of Nitric Oxide by the Cytochrome c Nitrite Reductase of Escherichia coli. J. Biol. Chem. 2002, 277 (26), 2366423669. 5. Einsle, O.; Messerschmidt, A.; Huber, R.; Kroneck, P. M. H.; Neese, F., Mechanism of the six-electron reduction of nitrite to ammonia by cytochrome c nitrite reductase. J. Am. Chem. Soc. 2002, 124 (39), 11737-11745. 6. Pereira, I.; LeGall, J.; Xavier, A.; Teixeira, M., Characterization of a heme c nitrite reductase from a non-ammonifying microorganism, Desulfovibrio vulgaris Hildenborough. Biochim. Biophys. Acta 2000, 1481, 119-130. 7. Tikhonova, T.; Tikhonov, A.; Trofimov, A.; Polyakov, K.; Boyko, K.; Cherkashin, E.; Rakitina, T.; Sorokin, D.; Popov, V., Comparative structural and functional analysis of two octaheme nitrite reductases from closely related Thioalkalovibrio species. FEBS J. 2012, 279, 4052-4061. 8. Kern, M.; Volz, J.; Simon, J., The oxidative and nitrosative stress defense network of Wollinella succinogenes: cytochrome c nitrite reductase mediates the stress response to nitrite, nitric oxide, hydroxylamine and hydrogen peroxide. Environ. Microbiol. 2011, 13, 2478-2494. 9. Einsle, O.; Messerschmidt, A.; Stach, P.; Bourenkov, G. P.; Bartunik, H. D.; Huber, R.; Kroneck, P. M. H., Structure of cytochrome c nitrite reductase. Nature 1999, 400 (6743), 476-480. 10. Einsle, O.; Stach, P.; Messerschmidt, A.; Simon, J.; Kroger, A.; Huber, R.; Kroneck, P. M. H., Cytochrome c nitrite reductase from Wolinella succinogenes - Structure at 1.6 angstrom resolution, inhibitor binding, and heme-packing motifs. J. Biol. Chem. 2000, 275 (50), 3960839616. 11. Bamford, V. A.; Angove, H. C.; Seward, H. E.; Thomson, A. J.; Cole, J. A.; Butt, J. N.; Hemmings, A. M.; Richardson, D. J., Structure and spectroscopy of the periplasmic cytochrome c nitrite reductase from Escherichia coli. Biochemistry 2002, 41 (9), 2921-2931. 12. Cunha, C. A.; Macieira, S.; Dias, J. M.; Almeida, G.; Goncalves, L. L.; Costa, C.; Lampreia, J.; Huber, R.; Moura, J. J. G.;

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Moura, I.; Romao, M. J., Cytochrome c nitrite reductase from Desulfovibrio desulfuricans ATCC 27774 - The relevance of the two calcium sites in the structure of the catalytic subunit (NrfA). J. Biol.Chem. 2003, 278 (19), 17455-17465. 13. Rodrigues, M.; Oliveira, T.; Pereira, I.; Archer, M., X-ray structure of the membrane-bound cytochrome c quinol dehydrogenase NrfH reveals novel haem coordination. EMBO Journal 2006, 25, 59515960. 14. Youngblut, M.; Judd, E. T.; Srajer, V.; Sayyed, B.; Goelzer, T.; Elliott, S. J.; Schmidt, M.; Pacheco, A. A., Laue crystal structure of Shewanella oneidensis cytochrome c nitrite reductase from a high-yield expression system. J. Biol. Inorg. Chem. 2012, 17, 647-662. 15. The currently accepted name for this enzyme (EC 1.7.2.6) is hydroxylamine dehydrogenase, while its correct systematic name is hydroxylamine:ferricytochrome-c oxidoreductase. In the past the enzyme was commonly referred to as hydroxylamine oxidase or oxidoreductase (HAO), and that term is still commonly used. 16. Stein, N.; Love, D.; Judd, E. T.; Elliott, S. J.; Bennett, B.; Pacheco, A. A., Correlations between the electronic properties of Shewanella oneidensis cytochrome c nitrite reductase (ccNiR) and its structure: effects of heme oxidation state and active site ligation. Biochemistry 2015, 54, 3749-3758. 17. Youngblut, M.; Pauly, D. J.; Stein, N.; Walters, D.; Conrad, J. A.; Moran, G. R.; Bennett, B.; Pacheco, A. A., Shewanella oneidensis cytochrome c nitrite reductase (ccNiR) does not disproportionate hydroxylamine to ammonia and nitrite, despite a strongly favorable driving force. Biochemistry 2014, 53, 2136-2144. 18. Kostera, J.; McGarry, J. M.; Pacheco, A. A., Enzymatic Interconversion of Ammonia and Nitrite: the Right Tool for the Job. Biochemistry 2010, 49, 8546-8553. 19. Kostera, J.; Youngblut, M. D.; Slosarczyk, J. M.; Pacheco, A. A., Kinetic and product distribution analysis of NO reductase activity in Nitrosomonas europaea hydroxylamine oxidoreductase. J. Biol. Inorg. Chem. 2008, 13 (7), 1073-1083. 20. Mowat, C. G.; Chapman, S. J., Multi-heme cytochromes - new structures, new chemistry. J. Chem. Soc. Dalton Trans. 2005, 21, 33813389. 21. Bewley, K. D.; Ellis, K. E.; Firer-Sherwood, M. A.; Elliott, S. J., Multi-heme proteins: nature's electronic multi-purpose tool. Biochim. Biophys Acta 2013, 1827, 938-948. 22. Galloway, J. N.; Townsend, A. R.; Erisman, J. W.; Bekunda, M.; Cai, Z.; Freney, J. R.; Martinelli, L. A.; Seitzinger, S. P.; Sutton, M. A., Transformation of the Nitrogen Cycle: Recent Trends, Questions and Potential Solutions. Science 2008, 320, 889-892. 23. Duce, R. A.; LaRoche, J.; Altieri, K.; Arrigo, K. R.; Baker, A. R.; Capone, D. G.; Cornell, S.; Dentener, F.; Galloway, J.; Ganeshram, R. S.; Geider, R. J.; Jickells, T.; Kuypers, M. M.; Langlois, R.; Liss, P. S.; Liu, S. M.; Middelburg, J. J.; Moore, C. M.; Nickovic, S.; Oschlies, A.; Pedersen, T.; Prospero, J.; Schlitzer, R.; Seitzinger, S.; Sorensen, L. L.; Uematsu, M.; Ulloa, O.; Voss, M.; Ward, B.; Zamora, L., Impacts of Atmospheric Anthropogenic Nitrogen on the Open Ocean. Science 2008, 320, 893-897. 24. Canfield, D. E.; Glazer, A. N.; Falkowski, P. G., The Evolution and Future of Earth’s Nitrogen Cycle. Science 2010, 330, 192-196. 25. Galloway, J. N.; Leach, A. M.; Bleeker, A.; Erisman, J. W., A chronology of human understanding of the nitrogen cycle. Phil. Trans. R. Soc. B. 2013, 368, 20130120. 26. Ravishankara, A. R.; Daniel, J. S.; Portmann, R. W., Nitrous oxide (N2O): the dominant ozone-depleting substance emitted in the 21st century. Science 2009, 326, 123-125. 27. Wuebbles, D. J., Nitrous oxide: no laughing matter. Science 2009, 326, 56-57. 28. Judd, E. T.; Youngblut, M.; Pacheco, A. A.; Elliott, S. J., Direct Electrochemistry of Shewanella oneidensis Cytochrome c Nitrite Reductase: Evidence of Interactions across the Dimeric Interface. Biochemistry 2012, 51 (51), 10175-10185. 29. Judd, E. T.; Stein, N.; Pacheco, A. A.; Elliott, S. J., Hydrogen bonding networks tune proton-coupled redox steps during the enzymatic six-electron conversion of nitrite to ammonia. Biochemistry 2014, 53, 5638-5646. 30. Drago, R. S.; Paulik, F. E., The Reaction of Nitrogen (II) Oxide with Diethylamine. J. Am. Chem. Soc. 1960, 82, 96-98.

Page 16 of 18

31. Maragos, C. M.; Morley, D.; Wink, D. A.; Dunams, T. M.; Saavedra, J. E.; Hoffman, A.; Bove, A. A.; Isaac, L.; Hrabie, J. A.; Keefer, L. K., Complexes of NO with Nucleophiles as Agents for the Controlled Biological Release of Nitric-Oxide - Vasorelaxant Effects. J. Med. Chem. 1991, 34 (11), 3242-3247. 32. Powell, H. R.; Battye, T. G. G.; Kontogiannis, L.; Johnson, O.; Leslie, A. G. W., Integrating macromolecular X-ray diffraction data with the graphical user interface iMosflm. Nat. Protoc. 2017, 12, 1310-1325. 33. Evans, P., Scaling and assessment of data quality. Acta Cryst. D., Biol. Cryst. 2006, 62(Pt 1), 72-82. 34. Adams, P. D.; Afonine, P. V.; Bunkoczi, G.; Chen, V. B.; Davis, I. W.; Echols, N.; Headd, J. J.; Hung, L. W.; Kapral, G. J.; GrosseKuntsleve, R. W., PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta Cryst. D., Biol. Cryst. 2010, 66(Pt 2), 213-221. 35. Emsley, P.; Lohkamp, B.; Scott, W. G.; Cowtan, K., Features and development of Coot. Acta Cryst. D., Biol. Cryst. 2010, 66(Pt 4), 486501. 36. Watanabe, T.; Honda, K., Measurement of the extinction coefficient of the methyl viologen cation radical and the efficiency of its formation by semiconductor photocatalysis. J. Phys. Chem. 1982, 86, 2617-2619. 37. Heineman, W. R.; Norris, B. J.; Goelz, J. F., Measurement of enzyme E0' values by optically transparent thin layer electrochemical cells. Anal. Chem. 1975, 47, 79-84. 38. Pilkington, M. B. G.; Coles, B. A.; Compton, R. G., Construction of an optically transparent thin layer electrode cell for use with oxygen-sensitive species in aqueous and non-aqueous solvents. Anal. Chem. 1989, 61, 1787-1789. 39. To clarify: the concentration of ccNiR protomers is reported herein; the concentration of the full enzyme dimer will be half of the reported protomer concentration. 40. Koebke, K. J.; Pauly, D. J.; Lerner, L.; Liu, X.; Pacheco, A. A., Does the oxidation of nitric oxide by oxymyoglobin share an intermediate with the metmyoglobin-catalyzed isomerization of peroxynitrite? Inorg. Chem. 2013, 52, 7623-7632. 41. Koebke, K. J.; Waletzko, M. T.; Pacheco, A. A., Direct monitoring of the reaction between photochemically generated nitric oxide and Mycobacterium tuberculosis tuncated hemoglobin N wild type and variant forms: an assessment of computational mechanistic predictions. Biochemistry 2016, 55, 686-696. 42. Stirring the solution while applying a potential promoted release of nitric oxide from the active site, as evidenced by the persistence of a high background current (> 100 A), and low accumulation of reduced ccNiR as monitored by UV/Vis spectroscopy. CcNiR's ability to catalyze slow reduction of nitrite to nitric oxide under weakly reducing conditions is briefly discussed herein, and will be more extensively covered in a separate article. 43. Lockwood, C. W. J.; Burlat, B.; Cheesman, M. R.; Kern, M.; Simon, J.; Clarke, T.; Richardson, D. J.; Butt, J. N., Resolution of key roles for the distal pocket histidine in cytochrome c nitrite reductases. J. Am. Chem. Soc. 2015, 137, 3059-3068. 44. Martins, G.; Rodrigues, L.; Cunha, F. M.; Matos, D.; Hildebrandt, P.; Murgida, D. H.; Pereira, I. A. C.; Todorovic, S., Substrate binding to a nitrite reductase induces a spin transition. J. Phys. Chem. B. 2010, 114, 5563-5566. 45. The subtraction isn't perfect; for example, a small residual of the g = 2.24 signal is seen in Fig. 2c. However, the subtraction does isolate the heme 1 signal sufficiently to allow its dependence on nitrite concentration to be easily monitored. 46. g' is the experimentally-observed resonance position where g' = h/BB0, and is distinct from g, the spin-Hamiltonian spectroscopic parameter. 47. Clarke, T. A.; Kemp, G. L.; Van Wonderen, J. H.; Doyle, R. A. S.; Cole, J. A.; Tovell, N.; Cheesman, M. R.; Butt, J. N.; Richardson, D. J.; Hemmings, A. M., Role of a conserved glutamine residue in tuning the catalytic activity of Escherichia coli Cytochrome c nitrite reductase. Biochemistry 2008, 47, 3789-3799. 48. Kemp, G. L.; Clarke, T. A.; Marritt, S. J.; Lockwood, C.; Poock, S. R.; Hemmings, A. M.; Richardson, D. J.; Cheesman, M. R.; Butt, J. N., Kinetic and thermodynamic resolution of the interactions

ACS Paragon Plus Environment

Page 17 of 18 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

between sulfite and the pentahaem cytochrome NrfA from Escherichia coli. Biochem. J. 2010, 431, 73-80. 49. Arciero, D. M.; Collins, M. J.; Haladjian, J.; Bianco, P.; Hooper, A. B., Resolution of the 4 Hemes of Cytochrome-C554 from Nitrosomonas-Europaea by Redox Potentiometry and Optical Spectroscopy. Biochemistry 1991, 30 (48), 11459-11465. 50. Marritt, S. J.; Kemp, G. L.; Xiaoe, L.; Durrant, J. R.; Cheesman, M. R.; Butt, J. N., Spectroelectrochemical characterization of a pentaheme cytochrome in solution and as electrocatalytically active films on nanocrystalline metal-oxide electrodes. J. Am. Chem. Soc. 2008, 130, 8588-8589. 51. Press, W. H.; Teukolsky, S. A.; Vetterling, W. T.; Flannery, B. P., Numerical Recipes the art of scientific computing. 3rd ed.; Cambridge University Press: New York, NY, 2007; pp 65-75. 52. Henry, E. R.; Hofrichter, J., Singular Value Decomposition: Application to Analysis of Experimental Data. In Meth. Enzymol., Brand, L.; Johnson, M. L., Eds. Academic Press: San Diego, 1992; Vol. 210, pp 129-192. 53. Enemark, J. H.; Feltham, R. D., Principles of Structure, Bonding and Reactivity for Metal Nitrosyl Complexes. Coord. Chem. Rev. 1974, 13, 339-406. 54. Albrecht, A. C.; Simpson, W. T., Spectroscopic study of Wurster's Blue and tetramethyl-p-phenylenediamine with assignments of electronic transitions. J. Am. Chem. Soc. 1955, 77, 4454-4461. 55. Westcott, B. L.; Enemark, J. H., Transition Metal Nitrosyls. In Inorganic Electronic Structure and Spectroscopy, Solomon, E. I.; Lever, A. B. P., Eds. John Wiley and Sons, Inc.: New York, NY, 1999; Vol. 2, pp 435-439. 56. Nasri, H.; Ellison, M. K.; Chen, S.; Huynh, B. H.; Scheidt, W. R., Sharing the -bonding. An iron porphyrin derivative with trans accepting axial ligands. Synthesis, EPR and Mossbauer spectra, and molecular structure of two forms of the complex nitronitrosyl(tetrakis(o-pivalamidophenyl)-porphinato)ferrate(II). J. Am. Chem. Soc. 1997, 119, 6274-6283. 57. Wayland, B. B.; Olson, L. W., Spectroscopic studies and bonding model for nitric oxide complexes of iron porphyrins. J. Am. Chem. Soc. 1974, 96, 6037-6041. 58. Stach, P.; Einsle, O.; Schumacher, W.; Kurun, E.; Kroneck, P. M. H., Bacterial cytochrome c nitrite reductase: new structural and functional aspects. J. Inorg. Biochem. 2000, 79, 381-385. 59. Hendrich, M. P.; Upadhyay, A. K.; Riga, J.; Arciero, D. M.; Hooper, A. B., Spectroscopic characterization of the NO adduct of hydroxylamine oxidoreductase. Biochemistry 2002, 41 (14), 4603-4611. 60. Amanullah, S.; Saha, P.; Saha, R.; Dey, A., Synthetic iron porphyrins for probing the differences in the electronic structures of heme a3, heme d, and heme d1 Inorg. Chem. 2019, 58, 152-164. 61. Upadhyay, A. K.; Hooper, A. B.; Hendrich, M. P., NO reductase activity of the tetraheme cytochrome c(554) of Nitrosomonas europaea. J. Am. Chem. Soc. 2006, 128 (13), 4330-4337. 62. Brown, G. C., Reversible Binding and Inhibition of Catalase by Nitric-Oxide. Eur. J. Biochem. 1995, 232 (1), 188-191. 63. Jackson, T. A.; Yikilmaz, E.; Miller, A.-F.; Brunold, T. C., Spectroscopic and computational study of a non-heme iron {FeNO}7 system: exploring geometric and electronic structures of the nitrosyl adduct of iron superoxide dismutase. J. Am. Chem. Soc. 2003, 125, 83488363. 64. Berto, T. C.; Praneeth, V. K. K.; Goodrich, L. E.; Lehnert, N., Iron-porphyrin NO complexes with covalently attached N-donor ligands: formation of a stable six-coordinate species in solution. J. Am. Chem. Soc. 2009, 131, 17116-17126. 65. Berto, T. C.; Xu, N.; Lee, S. R.; McNeil, A. J.; Alp, E. E.; Zhao, J.; Richter-Addo, G. B.; Lehnert, N., Characterization of the bridged hyponitrite complex {[Fe(OEP)]2(-N2O2)}: reactivity of hyponitrite complexes and biological relevance. Inorg. Chem. 2014, 53, 6398-6414. 66. Smith, A. T.; Pazicni, S.; Marvin, K. A.; Stevens, D. J.; Paulsen, K. M.; Burstyn, J. N., Functional divergence of heme thiolate proteins: a classification based on spectroscopic attributes. Chem. Rev. 2015, 115, 2532-2558. 67. Hunt, A. P.; Lehnert, N., Heme nitrosyls: electronic structure implications for function in biology. Acc. Chem. Res. 2015, 48, 21172125.

68. Hoshino, M.; Ozawa, K.; Seki, H.; Ford, P. C., Photochemistry of Nitric-Oxide Adducts of Water-Soluble Iron(III) Porphyrin and Ferrihemoproteins Studied by Nanosecond Laser Photolysis. J. Am. Chem. Soc. 1993, 115 (21), 9568-9575. 69. Bohle, S. D.; Hung, C.-H., Ligand-promoted rapid nitric oxide dissociation from ferrous porphyrin nitrosyls. J. Am. Chem. Soc. 1995, 117, 9584-9585. 70. Ellison, M. K.; Scheidt, W. R., Synthesis, molecular structures, and properties of six-coordinate [Fe(OEP)(L)(NO)] + derivatives: Elusive nitrosyl ferric porphyrins. J. Am. Chem. Soc. 1999, 121 (22), 5210-5219. 71. Ford, P. C.; Lorkovic, I. M., Mechanistic aspects of the reactions of nitric oxide with transition-metal complexes. Chem. Rev. 2002, 102 (4), 993-1017. 72. Praneeth, V. K. K.; Paulat, F.; Berto, T. C.; De Beer George, S.; Nather, C.; Sulok, C. D.; Lehnert, N., Electronic structure of sixcoordinate iron(III) - porphyrin NO adducts: the elusive iron (III) NO(radical) state and its influence on the properties of these complexes. J. Am. Chem. Soc. 2008, 130, 15288-15303. 73. Bykov, D.; Neese, F., Substrate binding and activation in the active site of cytochrome c nitrite reductase: a density functional study. J. Biol. Inorg. Chem 2011, 16, 417-430. 74. Bykov, D.; Neese, F., Reductive activation of the heme ironnitrosyl intermediate in the reaction mechanism of cytochrome c nitrite reductase: a theoretical study. J. Biol. Inorg. Chem. 2012, 17 (5), 741-760. 75. Bykov, D.; Neese, F., Six-electron reduction of nitrite to ammonia by cytochrome c nitrite reductase: insights from density functional theory studies. Inorg. Chem. 2015, 54, 9303-9316. 76. Atkins, P.; Overton, T.; Rourke, J.; Weller, M.; Armstrong, F.; Hagerman, M., Shriver & Atkins' Inorganic Chemistry. 5th ed.; W. H. Freeman and Co.: New York, NY, 2010; pp 147-178. 77. Stein, N. Spectroscopic and electrochemical studies of Shewanella oneidensis cytochrome c nitrite reductase, and improving cheme expression systems. Ph.D., University of Wisconsin-Milwaukee, Dept. of Chemistry and Biochemistry, Milwaukee, WI, 2014. 78. Nicholls, D. G.; Ferguson, S. J., Bioenergetics 3. In Bioenergetics 3, Academic Press: San Diego, CA, 2002; pp 134-154. 79. Caranto, J. D.; Vilbert, A. C.; Lancaster, K. M., Nitrosomonas europaea cytochrome P460 is a direct link between nitrification and nitrous oxide emission. PNAS 2016, 113, 14704-14709.

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 18

Table of Contents Artwork

ACS Paragon Plus Environment

18