<<

Utah State University DigitalCommons@USU

All Graduate Theses and Dissertations Graduate Studies

8-2018

Electron Flow and Management in Living Systems: Advancing Understanding of Electron Transfer to

Rhesa N. Ledbetter Utah State University

Follow this and additional works at: https://digitalcommons.usu.edu/etd

Part of the Biochemistry Commons

Recommended Citation Ledbetter, Rhesa N., "Electron Flow and Management in Living Systems: Advancing Understanding of Electron Transfer to Nitrogenase" (2018). All Graduate Theses and Dissertations. 7197. https://digitalcommons.usu.edu/etd/7197

This Dissertation is brought to you for free and open access by the Graduate Studies at DigitalCommons@USU. It has been accepted for inclusion in All Graduate Theses and Dissertations by an authorized administrator of DigitalCommons@USU. For more information, please contact [email protected]. ELECTRON FLOW AND MANAGEMENT IN LIVING SYSTEMS: ADVANCING UNDERSTANDING OF ELECTRON TRANSFER TO NITROGENASE

by

Rhesa N. Ledbetter

A dissertation submitted in partial fulfillment of the requirements for the degree

of

DOCTOR OF PHILOSOPHY

in

Biochemistry

Approved:

______Lance C. Seefeldt, Ph.D. Scott A. Ensign, Ph.D. Biochemistry Biochemistry Major Professor Committee Member

______Bruce Bugbee, Ph.D. Sean J. Johnson, Ph.D. Plant Physiology Biochemistry Committee Member Committee Member

______Nicholas E. Dickenson, Ph.D. Mark R. McLellan, Ph.D. Biochemistry Vice President for Research and Committee Member Dean of the School of Graduate Studies

UTAH STATE UNIVERSITY Logan, Utah

2018 ii

Copyright © Rhesa N. Ledbetter 2018

All Rights Reserved

iii ABSTRACT

Electron Flow and Management in Living Systems:

Advancing Understanding of Electron Transfer to Nitrogenase

by

Rhesa N. Ledbetter, Doctor of Philosophy

Utah State University, 2018

Major Professor: Dr. Lance C. Seefeldt Department: Biochemistry

Diazotrophic microorganisms use the nitrogenase to reduce dinitrogen

(N2) into a biologically available ammonia (NH3) and play a critical role in the global . Although there is a wealth of knowledge on many aspects of this enzyme, we have a limited understanding of how low potential reductants, in the form of reduced flavodoxin and , are generated and their interaction with nitrogenase. This research provides new insight into the mechanism of electron flow to nitrogenase and introduces a bioelectrochemical system for the production of fixed nitrogen.

A new mechanism known as electron bifurcation was demonstrated to generate reduced flavodoxin for in vinelandii. The reaction, facilitated by a protein complex known as FixABCX, used NADH as the electron donor to couple the exergonic reduction of coenzyme Q to the endergonic reduction of flavodoxin without the use of ATP or an .

Further, the rate limiting step of nitrogenase catalysis was revealed to be events associated with Pi release rather than protein-protein dissociation as previously iv suggested. Because nitrogenase is readily reduced by artificial chemical reductants such as dithionite, few in vitro studies investigating the nitrogenase catalytic cycle had been conducted using physiological electron donors, such as flavodoxin. By comparing the rate constants of key steps in the nitrogenase catalytic cycle using dithionite and flavodoxin, novel insights into the mechanism of biological nitrogen fixation, most relevant to how it occurs in vivo, were discovered.

Biological nitrogen fixation is an attractive option for the sustainable generation of nitrogen fertilizer. Here, it was demonstrated that phototrophic, nitrogen-fixing can grow in association with an electrode, in which electrons are supplied from water oxidation and the energy provided by infra-red light. This biohybrid approach, while in its early stages, has potential for the generation of bioavailable nitrogen for food production.

Nitrogen fixation is one of the most biologically challenging reactions. Thus, understanding the fundamental mechanisms of electron flow and management can offer fresh insight into the elegant mechanisms nature has evolved. The insight may also contribute to advances in other bio-inspired applications.

(236 pages)

v PUBLIC ABSTRACT

Electron Flow and Management in Living Systems:

Advancing Understanding of Electron Transfer to Nitrogenase

Rhesa N. Ledbetter

Nitrogen is a critical nutrient for growth and reproduction in living .

Although the Earth’s atmosphere is composed of ~80% nitrogen gas (N2), it is inaccessible to most living organisms in that form. Biological nitrogen fixation, however, can be performed by microbes that harbor the enzyme nitrogenase. This enzyme converts

N2 into bioavailable ammonia (NH3) and accounts for at least half of the “fixed” nitrogen on the planet. The other major contributor to ammonia production is the industrial Haber-

Bosch process. While the Haber-Bosch process has made significant advances in sustaining the global food supply through the generation of fertilizer, it requires high temperature and pressure and fossil fuels. This makes nitrogenase an ideal system for study, as it is capable of performing this challenging chemistry under ambient conditions and without fossil fuels.

Nitrogenase requires energy and electrons to convert N2 into NH3. The work presented here examined how the enzyme receives electrons to perform the reaction. It was discovered that some microbes employ a novel mechanism that adjusts the energy state of the electrons so that nitrogenase can accept them. Further, the slowest step that takes place in nitrogenase once the electrons are taken up was identified. Finally, by capitalizing on fundamental knowledge, a biohybrid system was designed to grow vi nitrogen-fixing bacteria in association with electrodes for light-driven production of fixed nitrogen that has potential to be used as a fertilizer for plant growth.

Gaining an in-depth understanding of nitrogenase provides insight into one of the most challenging biological reactions, and the newfound knowledge may be a catalyst in developing more efficient systems for sustainable ammonia production.

vii ACKNOWLEDGMENTS

Thank you to the numerous people who have been part of my Ph.D. journey. And while science isn’t a sport, it has become clear to me that it’s most definitely made up of a team.

My head coach, Lance Seefeldt: What a pleasure it has been working for you!

There are so many qualities I admire in you—your ability to get things done without procrastinating, your gentle and kind personality, and your excellent communication skills. Above all though, I most admire your integrity. You are simply a good person that

I look up to. Thank you for believing in me, even on the many days I had less than stellar data to show. Also, I will always appreciate you letting me work part-time for Utah

Public Radio. Not every advisor would allow that, and I am thankful.

My assistant coaches, Ph.D. committee: Thank you for always having an open door. The time you spent (without batting an eye) helping me troubleshoot and answering questions means so much. You are a fine group of people, and I am lucky to have worked with each one of you.

My teammates, lab comrades: We’ll always carry special memories of spending years battling equipment (I’m glad we don’t have to curse at the GC printer anymore), coming up with the most ingenious of ways to do an experiment, and celebrating the successes and sharing in the disappointments of science. I would not be where I am today without all of you. Thank you for investing in me.

My buddies on other teams, collaborators across the country: Much of the work in this dissertation would not have been possible without you. You are exceptional viii scientists, who have contributed to my growth and success more than you know. But even more important to me is the friendships that have blossomed from working with you over the years.

My boosters, funding agencies: Much gratitude to the U.S. Department of

Energy, Office of Science, Basic Energy Sciences Program and NASA for funding the research projects in this dissertation.

My biggest cheerleader, Mom: You endlessly cheered me on from the sidelines, especially on the days I wanted to quit. I’ll never forget you saying, “Well, let’s think about this.” And somehow I always came back around. You also listened to every presentation I needed to practice with a smile…even when you had no idea what I was talking about.

My biggest fan, Tim: It’s hard to write this, because you deserve more than my words could ever express. You have supported me in everything I have wanted to do without complaint…literally. Thank you for making me a better person, always making me laugh, and like my mom, listening to every scientific presentation I ever had to give…yes, even the one associated with this dissertation…two or three times. Now, I can’t wait to move home!

Rhesa N. Ledbetter

ix CONTENTS

Page

ABSTRACT ...... iii

PUBLIC ABSTRACT ...... v

ACKNOWLEDGMENTS ...... vii

LIST OF TABLES ...... xi

LIST OF FIGURES ...... xii

CHAPTER

I. INTRODUCTION ...... 1 Biological nitrogen fixation ...... 1 Nitrogenase ...... 2 Physiological electron donors to nitrogenase ...... 4 Generation of reductant for nitrogen fixation ...... 6 The protein cycle ...... 12 Capitalizing on biological nitrogen fixation ...... 13 Conclusion ...... 15 References ...... 16

II. THE ELECTRON BIFURCATING FIXABCX COMPLEX FROM : GENERATION OF LOW-POTENTIAL REDUCING EQUIVALENTS FOR NITROGENASE CATALYSIS ...... 26 Abstract ...... 26 Introduction ...... 27 Materials and methods ...... 30 Results and discussion ...... 42 Acknowledgments ...... 62 Author contributions ...... 62 References ...... 63 Appendix A: Supporting information ...... 73

III. EVIDENCE THAT THE Pi RELEASE EVENT IS THE RATE LIMITING STEP IN THE NITROGENASE CATALYTIC CYCLE ...... 96 Abstract ...... 96 Introduction ...... 97 Materials and methods ...... 100 Results and discussion ...... 105 Funding ...... 115 Acknowledgments ...... 115 x Author contributions ...... 115 References ...... 116

IV. CO2 and N2 FIXATION BY MICROBIAL ELECTROCATALYSIS ...... 124 Abstract ...... 124 Introduction ...... 124 Materials and methods ...... 127 Results and discussion ...... 129 Acknowledgments ...... 136 Author contributions ...... 136 References ...... 136

V. SUMMARY AND FUTURE DIRECTIONS ...... 142 Introduction ...... 142 Electron bifurcation by FixABCX ...... 142 Electron transfer to nitrogenase ...... 149 + Microbial electrosynthesis for NH4 production ...... 152 Conclusion ...... 155 Acknowledgments ...... 156 References ...... 156

APPENDICES ...... 162 Appendix B: Evidence that the Pi release event is the rate limiting step in the nitrogenase catalytic cycle (Full manuscript) ...... 162 Appendix C: Evidence that the Pi release event is the rate limiting step in the nitrogenase catalytic cycle (Supporting information) ...... 203

CURRICULUM VITAE ...... 214 xi LIST OF TABLES

Table Page

1-1 Flavodoxins and in A. vinelandii ...... 6

A1 Primers for Δfix mutant generation ...... 74

A2 Primers for overexpression of FixABCX ...... 75

A3 The inter-cross-links and intra-cross-links within FixABCX complex during cross-linking reaction ...... 76

A4 Overview of the EPR signals observed for FixABCX ...... 82

A5 Calculated Gibbs free energies of three possible reactions catalyzed by the Fix complex ...... 83

A6 Lifetimes of the ASQ species observed in Rh. rubrum FixAB and A. vinelandii FixABCX ...... 84

3-1 Steady-state kcat values for Fe protein and MoFe protein with different reductants...... 107

B1 Steady-state kcat values for Fe protein and MoFe protein with different reductants...... 178

B2 Number of ATP hydrolyzed per electron transferred for reduction of different substrates by nitrogenase under different electron flux conditions ...... 189

C1 Electron distribution for H2 and NH3 production under different electron flux conditions with dithionite or flavodoxin ...... 204

xii LIST OF FIGURES

Figure Page

1-1 nitrogenase ...... 3

1-2 Structural representations of flavodoxin and ferredoxin ...... 4

1-3 Pyruvate ferredoxin/flavodoxin oxidoreductase ...... 7

1-4 Rhodobacter nitrogen fixation complex ...... 9

1-5 Electron bifurcation by the FixABCX complex ...... 11

1-6 The iron protein cycle ...... 13

1-7 Essential steps in a phototrophic, nitrogen-fixing microbe ...... 15

2-1 Phenotype of A. vinelandii DJ wt, Δfix, Δrnf1, ΔfixΔrnf1, Δrnf2, and ΔfixΔrnf2 strains...... 42

2-2 Proposed mechanism and electron transfer pathway in FixABCX electron bifurcation ...... 45

2-3 SDS-PAGE gel of purified FixABCX complex from A. vinelandii ...... 47

2-4 FixABCX complex composition...... 49

2-5 UV-visible spectrum of the FixABCX complex ...... 51

2-6 EPR spectra of FixABCX ...... 53

2-7 Electron bifurcation by the FixABCX complex ...... 56

2-8 Transient absorption spectra of FixAB and FixABCX ...... 59

A1 Scheme for ∆fix mutant generation ...... 85

A2 Scheme for overexpression of fix genes under the nifH promoter in A. vinelandii ...... 86

A3 Protein identification within purified FixABCX complex ...... 87

A4 Size distribution of FixABCX using dynamic light scattering ...... 88

A5 Evaluation of subunit homology models ...... 89

xiii A6 Predicted structural features in Fix subunits ...... 90

A7 Protein interface of FixB ...... 91

A8 Structural model of FixABCX ...... 92

A9 EPR temperature profiles of FixABCX ...... 93

A10 Comparison of EPR spectra of FixABCX from A. vinelandii and the FixX subunit from Ro. castenholzii ...... 94

3-1 Overview of nitrogenase catalysis ...... 98

3-2 Fe protein cycle with rate constants for each step ...... 99

3-3 Reduction of Fe protein by dithionite, methyl viologen, and flavodoxin in the presence of MoFe protein...... 108

3-4 Primary electron transfer from Fe protein to MoFe protein in presence of dithionite and flavodoxin...... 110

3-5 Kinetics of the reduction of Fe protein by dithionite and flavodoxin in the presence and absence of nucleotides ...... 111

3-6 Pi release during nitrogenase catalysis with dithionite and flavodoxin ...... 113

3-7 Updated Fe protein cycle with flavodoxin as the reductant ...... 115

4-1 A microbial electrosynthesis cell for light-driven CO2 and N2 reduction by R. palustris ...... 129

4-2 Spectra of incandescent and IR lights used in the study...... 131

4-3 Growth curves of R. palustris grown under phototrophic, nitrogen fixing conditions with incandescent and IR light ...... 131

4-4 Photoautotrophic growth of R. palustris in a microbial electrosynthesis cell with IR light ...... 133

4-5 Development of a biohybrid system for fertilizer production, in which microbial biomass is used a source of fixed N for plant growth...... 135

5-1 Possible electron transfer pathways in FixABCX ...... 144

5-2 UV-visible spectrum of FixABCX without non-covalently bound cofactors ...... 145

xiv 5-3 FixABCX structural model illustrating a conserved Cys residue that may form a covalent bond with the bifurcating flavin ...... 146

5-4 Scheme for coupling the FixABCX electron bifurcation assay to nitrogenase activity...... 149

5-5 Mediated vs. direct electron transfer...... 154

B1 Overview of nitrogenase catalysis ...... 165

B2 Fe protein cycle with rate constants for each step ...... 166

B3 Reduction of Fe protein by dithionite, methyl viologen, and flavodoxin in the presence of MoFe protein...... 179

B4 Primary electron transfer from Fe protein to MoFe protein in presence of dithionite or flavodoxin...... 181

B5 Kinetics of the reduction of Fe protein by dithionite and flavodoxin in the presence and absence of nucleotides ...... 183

B6 Interaction of Fe protein with flavodoxin ...... 184

B7 Mapping of the Fe protein-flavodoxin interaction site ...... 186

B8 Pre-steady state and steady state ATP hydrolysis during nitrogenase catalysis with dithionite or flavodoxin ...... 190

B9 Pi release during nitrogenase catalysis using dithionite or flavodoxin ...... 192

B10 Updated Fe protein cycle with flavodoxin as the reductant ...... 193

C1 Acetylene reduction by wt Mo-nitrogenase using dithionite or flavodoxin as reductants under different electron flux conditions ...... 205

C2 Electron flux dependence on proton reduction activity by wild-type Mo- nitrogenase using dithionite or flavodoxin as reductants ...... 206

C3 Electron flux dependence on N2 reduction by wild-type Mo-nitrogenase using dithionite or flavodoxin as reductants ...... 207

C4 Electron flux dependence on the total activity of proton and N2 reduction by wild-type Mo-nitrogenase using dithionite or flavodoxin as reductants under N2...... 208

C5 Electron flux dependence on MoFe protein activity for N2 reduction using dithionite or flavodoxin as reductants ...... 209 xv C6 The isosbestic point of flavodoxin ...... 210

C7 Electrostatic interactions between Fe protein and flavodoxin ...... 211

C8 Inhibition of nitrogenase activity by high concentration of flavodoxin under different electron flux conditions ...... 212

C9 ATP hydrolysis and electron utilization for proton reduction under various electron flux conditions using dithionite or flavodoxin as reductant ...... 213

CHAPTER 1

INTRODUCTION

BIOLOGICAL NITROGEN FIXATION

Nitrogen is an essential element for all living organisms as it is a constituent of proteins, nucleic acids, and other biomolecules.1,2 While the majority of Earth’s atmosphere is composed N2 gas, the triple bond between nitrogen atoms makes it chemically inert and unavailable to most living organisms. Therefore, atmospheric N2 must be transformed or “fixed” into a form cells can utilize during their growth and reproduction.1

Prokaryotes with the enzyme nitrogenase have the ability to transform N2 into

2,3 NH3 through a process know as biological nitrogen fixation. This process is carried out by both free-living and symbiotic microorganisms known as .3,4 Although a relatively small number of microorganisms have the ability to reduce N2, those that do are phylogenetically and metabolically diverse, ranging from aquatic phototrophs (e.g.

Rhodopseudomonas) to soil-dwelling bacteria (e.g. Azotobacter and Klebsiella).4,5 Others form close symbiotic associations with plants. For example, rhizobia colonize the root system of legumes (e.g. soybeans, peas, alfalfa, and peanuts), giving plants access to fixed N in exchange for carbon from the plants.6 Biological nitrogen fixation provides over 50% of the fixed nitrogen on Earth2,7 with 80% of that coming from symbiotic associations and the remaining from free-living nitrogen-fixing organisms.8

2

Nitrogenases are a family of found in a limited number of microorganisms that catalyze the reduction of N2 to NH3. The stoichiometry of the reaction under “optimal” conditions is as follows (Eq. 1-1):

+ - N2 + 8 H + 16 MgATP + 8 e → 2 NH3 + H2 + 16 MgADP + 16 Pi (Eq. 1-1)

Within the nitrogenase family, there are three genetically distinct homologues, including the molybdenum (Mo), vanadium (V), and iron (Fe-only) nitrogenases.9–11

While the basic architecture and mechanism appear to be similar among the three, the presence of Mo, V, or Fe in the metal cluster is one differentiating factor.2,9,10,12,13 Of the three nitrogenases, Mo nitrogenase is the best characterized and described below.

Mo nitrogenase is a two-component metalloprotein complex composed of the iron

(Fe) protein and molybdenum-iron (MoFe) protein (Figure 1-1).2 The Fe protein (NifH) is an α2 homodimer whose role is to deliver electrons to the MoFe protein. The two subunits of the Fe protein homodimer are bridged by an electron-carrying [4Fe-4S] cluster, and each subunit contains an ATP-binding site.14 The MoFe (NifDK) protein is

15 an α2β2 heterotetramer with each αβ heterodimer constituting a catalytic half. Each half contains two metal clusters. The P-cluster [8Fe-7S] accepts electrons from the Fe protein and shuttles them to the FeMo- [7Fe-9S-Mo-C-homocitrate] where substrate reduction occurs.16 During catalysis, the Fe and MoFe protein transiently associate as the

+ - N2 + 8 H + 16 MgATP + 8 e → 2 NH3 + H2 + 16 MgADP + 16 Pi 3

Fe protein

Fe protein MoFe protein Fldred/Fdred + 2 ATP N2 + 8 H 1 e- 1 e-

Fldox/Fdox 2 ADP + 2 P [4Fe-4S] [8Fe-7S] [7Fe-9S- 2 NH + H i P-cluster C-Mo-homocitrate] 3 2 M-cluster Catalytic cycle (8x) Figure 1-1. Molybdenum nitrogenase. Top: Structural representation of Fe (green) and MoFe (α-subunit= blue; β-subunit= red) proteins of nitrogenase, including both catalytic halves (PDB ID 4WZB). Bottom: Electron flow through one catalytic half of nitrogenase.

Fe protein delivers electrons one-at-a-time to the MoFe protein. At least two ATP are

17 hydrolyzed for each electron transferred. Following eight catalytic cycles, one N2 is fully reduced to two NH3 (Figure 1-1).

Although there is a wealth of knowledge on many aspects of nitrogenase, a fundamental understanding of how the low-potential reducing equivalents needed to drive nitrogen fixation are generated and their impact on the nitrogenase catalytic cycle is limited. This gap in knowledge inspired two questions addressed in this dissertation: 1)

How is low-potential reductant generated for nitrogen fixation in Azotobacter vinelandii?

2) How is the rate limiting step of nitrogenase catalysis affected in the presence of a non- physiological (sodium dithionite) and physiological (Fld) reductant? The final question applied a basic understanding of biological nitrogen fixation to offer fresh insight into the

4 following question: 3) How can diazotrophic bacteria be harnessed to produce sustainable fixed nitrogen for food production?

PHYSIOLOGICAL ELECTRON DONORS TO NITROGENASE

As shown in Eq. 1-1, reducing power in the form of low electrons is one essential element for nitrogen fixation. Small proteins, flavodoxin

(Fld) and ferredoxin (Fd), serve as the direct electron donor to the Fe protein of nitrogenase in vivo.18 The redox properties of Fld and Fd are derived from bound cofactors, flavin mononucleotide (FMN) and FeS cluster(s), respectively (Figure 1-2).18

Flds are 1- or 2-electron carriers and exhibit three oxidation states: oxidized (Ox), semiquinone (Sq) and hydroquinone (Hq) forms with redox potential ranging from -50 to

-250 mV for the Sq/Ox couple and -370 to -450 mV for the Hq/Sq couple.18 Because the

Fe protein of nitrogenase requires electrons <-400 mV, only the Hq/Sq couple has sufficient driving force.

Figure 1-2. Structural representations of NifF flavodoxin (red; FMN=yellow) and ferredoxin I (green; FeS clusters= yellow/orange) from the A. vinelandii (PDB IDs 5K9B and 1FER).

5 Fds are distinguished by the type and number of FeS clusters they contain and have a wide range of redox potentials (0 to -645 mV) tuned by the polypeptide structure.18 The Fds relevant to nitrogen fixation have primarily been ascribed to those containing the [2Fe-2S] or 2[4Fe-4S] clusters.18,19

In many cases, diazotrophs possess several Flds and Fds, and it has been demonstrated that more than one can serve as an electron donor to nitrogenase.20,21 This has presented challenges in assigning specific Flds and Fds to nitrogenase activity in the cell. An additional complexity is that these electron carriers can act in diverse metabolic processes (e.g. I, ) and have overlapping functions.18,22 The only diazotroph in which the direct electron donor has been firmly established is

Klebsiella pneumoniae. NifF is a flavodoxin encoded by the nifF gene, which is part of the nitrogenase (nif) gene cluster. Studies in which nifF was mutated rendered K. pneumoniae cells unable to grow under nitrogen-fixing conditions, implicating NifF as the sole electron donor for diazotrophic growth.23,24

Like K. pneumoniae, A. vinelandii harbors a homologous nifF gene. However, in addition to nifF, there is genetic evidence of at least one additional Fld and several Fds or

Fd-like proteins (Table 1-1).18,25,26 NifF (Fld) and FdI (Fd) have both been demonstrated to donate electrons to Mo nitrogenase in vitro;27–29 however, their role in vivo is still being elucidated. Cells with gene disruptions in nifF and fdxA alone as well as a nifF/fdxA double mutant grew equally as well as wild-type cells under nitrogen-fixing conditions.30,31 These growth studies confirmed the presence of alternative electron carriers that play either a compensatory or co-functioning role.18 In addition to the cell growth studies, a whole-cell nitrogenase acetylene reduction assay was performed on

6 Table 1-1. Flavodoxins and ferredoxins in A. vinelandii. Adapted from Reyntjens et al.25

Gene Protein Cofactor Subunit Mr Function nifF NifF/ FMN 19,700 Under nif promoter; role in electron FldII transfer to Mo-nitrogenase20,30 fldA FldI FMN 19,600 Role in electron transfer to nitrate reductase26 fdxN FdxN 2 [4Fe-4S] 9,600 nif-related; functions in M-cluster biosynthesis;32 role in electron transfer is not established fixFd FixFd 2 [4Fe-4S] 7,200 Unknown; in the fixABCX operon25 vnfFd VnfFd 2 [4Fe-4S] 6,975 Likely role in electron transfer to V- nitrogenase33 fdxA FdI [3Fe-4S][4Fe-4S] 13,000 Unknown; not integral for nitrogen- fixation20 fixX FixX 2 [4Fe-4S] 10,700 Part of FixABCX complex for

generation of reductant for N2- fixation (Chapter 2)

cells with the nifF disruption.30 This mutant displayed a 30% reduction in nitrogenase activity and suggested that NifF serves as a source of electrons to nitrogenase, but is not the only electron carrier in A. vinelandii.18,30

Although, identifying the specific primary electron donor(s) to nitrogenase in many diazotrophs is difficult, it is clear that Flds and/or Fds play an essential role in the electron transfer process. The pathways to generate the reduced Fld or Fd, however, are just beginning to be uncovered in many diazotrophs.

GENERATION OF REDUCTANT FOR NITROGEN FIXATION

Four major systems capable of reducing Fld and Fd have been identified in diazotrophs thus far, including pyruvate-ferredoxin/flavodoxin oxidoreductase (PFOR),

7 ferredoxin-NADP+ reductase, the Rhodobacter Nitrogen Fixation (RNF) system, and the

Fix complex.

Pyruvate ferredoxin/flavodoxin oxidoreductase (PFOR)

PFOR catalyzes the reversible oxidation of pyruvate to acetyl CoA and CO2 with

Fld or Fd as the electron acceptor. In K. pneumonia, the nifJ gene encodes for a PFOR system specific for generating the reduced Fld needed for nitrogen fixation (Figure 1-

3).34–36 Like nifF (encodes for Fld), disruption of the nifJ gene in K. pneumonia results in non-nitrogen-fixing phenotype (Nif-), implicating NifJ and NifF as the only path by which to supply the reducing equivalents for nitrogenase catalysis.

Glucose

Pyruvate + NifF Fld (oxidized) CoA

NifJ Nitrogenase PFOR

CO2 + NifF Fld (reduced) Acetyl-CoA Figure 1-3. Schematic illustrating the oxidation of pyruvate to generate reduced Fld for nitrogenase in K. pneumoniae.

8 Ferredoxin-NADP+ reductase

Ferredoxin-NADP+ reductase is best known for its role in , in which uses Fd to catalyze the reduction of NAD(P) (Eq. 1-2).37–39

+ + + (Eq. 1-2) 2 Fdred + NADP + 2 H 2 Fdox + NADPH + H

However, some diazotrophic microbes, including A. vinelandii, Rhodobacter capsulatus, and nitrogen-fixing cyanobacteria, have been shown to harbor ferredoxin-NADP+ reductases as well. For nitrogen fixation, the reaction must run in reverse to generate reduced Fd. While this is possible, the forward reaction is generally favored, since the reduction potential of Fd is often lower than that of NADPH (Em= -320 mV). The ferredoxin-NADP+ reductases have been purified from A. vinelandii and Rb. capsulatus and demonstrated to reduce Fd.40–44 It was shown that the ferredoxin-NADP+ reductase from A. vinelandii bound specifically to FdI and that electron transfer between the two was facilitated.43 To determine if this would provide a direct line for electron transfer to nitrogenase, in vitro acetylene reduction assays were conducted. No substrate reduction occurred, suggesting that ferredoxin-NADP+ reductase and FdI do not provide a route for electron transfer to nitrogenase. Interestingly, attempts to delete the ferredoxin-NADP+ reductase gene were lethal, indicating this protein is playing some other crucial role in the cell. Studies in cyanobacteria, on the other hand, suggest that ferredoxin-NADP+ reductase likely generates reductant for nitrogenase, as heterocysts (specialized nitrogen- fixing compartments) reduce acetylene when NADPH is generated.40,45

9 Rhodobacter Nitrogen Fixation (RNF)

The RNF complex is a putative reversible membrane-bound complex that couples

Fld/Fd to the pyridine nucleotide pool.46 First identified in the diazotroph Rb. capsulatus,

RNF complexes are distributed among diverse nitrogen-fixing and non-nitrogen-fixing microbes.46 Ten rnf-related genes have been identified in Rb. capsulatus, and while gene composition and organization can vary among organisms, a general model for generating reduced Fld/Fd for nitrogen fixation is shown in Figure 1-4.18,22,46 Initial observations of protein sequences revealed the certain RNF proteins share similar primary structure with sodium ion pumps18,47,48 and although there are limited biochemical studies, some have indeed demonstrated ion transport.46,49 Thus, a proton or sodium ion motive force generated from electron transport processes appears to be used to drive the reduction of

Fld/Fd (Em= <-400 mV) using a pyridine nucleotide like NADH (Em= -320 mV) as the electron donor.46

Na+ or H+

RNF complex

NADH NAD+ Fldox/Fdox Fldred/Fdred Nitrogenase Figure 1-4. Model for the generation of reduced Fld/Fd for nitrogen fixation using the RNF complex.

10 Among diazotrophs, RNF appears to be the primary pathway for generating reductant for nitrogen fixation in Rb. capsulatus. Disruption of genes in rnf operon of Rb. capsulatus resulted in a significant decrease in nitrogenase activity and abolished the ability of cells to grow under nitrogen-fixing conditions.50,51 Additionally, nitrogenase activity increased 50-100% over wild-type upon overexpression of rnf genes,52 supporting its nitrogenase-related function in this .

While most organisms only have a single RNF complex, A. vinelandii harbors two: one that is nif-regulated (rnf1) and another constitutively expressed (rnf2). Gene disruptions in rnf1 and rnf2 delayed diazotrophic growth, however, Δrnf1, Δrnf2, and

Δrnf1rnf2 mutants could still fix nitrogen, albeit at decreased activities.53 It was suggested that the RNF complexes did not play role in electron transfer to nitrogenase, but rather served a regulatory role in expression and maturation of the enzyme in A. vinelandii.53 However, more recent research presented in Chapter 2 has also demonstrated a role in electron transfer to nitrogenase.

Fix complex

Rhizobia form associations with plants and in the early 1980s, a suite of genes essential to their ability to fix nitrogen were discovered.54–56 These genes were termed fix, and one gene cluster, known as fixABCX, is conserved among rhizobia as well as some free-living diazotrophs, such as A. vinelandii, Rhodopseudomonas palustris, and

Rhodospirillum rubrum.

A study in Azorhizobium caulinodans indicated that FixABCX may promote the formation of functional nitrogenase complexes,57 and others proposed roles in electron transfer. While FixABCX could function in multiple capacities, in was not until 2004 that

11 a function in electron transport became apparent. Mutants of fix genes in Rh. rubrum had 25% of nitrogenase activity compared to wild-type, but in vitro nitrogenase assays demonstrated full functionality of the enzyme, suggesting that Fix provides an electron transfer pathway in vivo.22,58

These genetic studies prompted further investigation into how FixABCX might generate the low-potential electrons needed for nitrogen fixation. The protein sequences encoded by fixABCX are homologous to electron-transferring flavoproteins (Etfs).59–62

Some Etfs are electron bifurcating, meaning they couple an endergonic and exergonic reaction to maximize energy conservation. By harnessing the energy from the exergonic reaction, cells can drive thermodynamically unfavorable reactions.63–65 All known electron bifurcation processes reduce or oxidize Fld/Fd; thus, it was hypothesized that

FixABCX takes two electrons from NADH and sends one electron to the a high-potential donor proposed to be CoQ (based on FixCX homology to an Etf-ubiquinone oxidoreductase) and the other electron to reduce Fld/Fd (Figure 1-5).58,65

Fld/Fd Nitrogenase FixABCX (-460 mV) (-440 mV)

e- - NADH 2 e (-320 mV)

e- CoQ (+10 mV) Figure 1-5. Pathway of electron bifurcation by the FixABCX complex to generate reduced Fd/Fld for nitrogen fixation.

12 While there has been significant speculation on how FixABCX might provide a route for electron transfer to nitrogenase, Chapter 2 provides the first biochemical evidence of an electron bifurcation pathway to generate reductant.

Multiple pathways have been identified to generate the low potential reductant needed for nitrogenase catalysis. This reductant, in the form of Fld/Fd, then provides electrons to nitrogenase for N2 reduction.

THE IRON PROTEIN CYCLE

Reduced Fld/Fd delivers an electron (<-400 mV) to the [4Fe-4S] cluster of the Fe protein. The reduced Fe protein with two bound ATP then transiently associates with

MoFe protein to deliver the electron.66 A series of events known as “the Fe protein cycle” take place in the Fe protein upon its interaction with MoFe protein.67 First,

-1 conformationally-gated electron transfer occurs (kET= 140 s ), in which the MoFe protein

P-cluster transfers its electron to FeMo-cofactor. The oxidized P-cluster is quickly back- filled with the electron from the reduced Fe protein.68 This is followed by several steps,

-1 -1 which occur in the following order: ATP hydrolysis (kATP= 70 s ), Pi release (kpi= 16 s ),

-1 dissociation of Fe protein from MoFe protein (kdiss= 6 s ), and finally re-oxidation of the

Fe protein and nucleotide exchange (Figure 1-6).66,67 This completes the Fe protein cycle, and another round of electron transfer to MoFe protein can take place.

In addition to the physiological reductants, Fld/Fd, the Fe protein also readily accepts electrons from the small molecule sodium dithionite (Na-dithionite). Because of this, the majority of in vitro studies on nitrogenase have been conducted using this chemical reductant due to ease of use. The rate-limiting step of catalysis with Na-

13

Association Eletron transfer/ ATP hydrolysis Pi release Dissociation

M M Red Red Red M M

M

-

e

P

P

P P P

-

e

+

+

Ox

Ox

ATP Red ATP

ADP

ADP

ADP/Pi

ADP/Pi

Ox

ATP ATP

Red

ADP ADP

Re-reduction and nucleotide exchange

Figure 1-6. The Fe protein cycle. Ovals represent ½ of the MoFe protein (P= P-cluster; M= FeMo-cofactor) and squares represent the Fe protein (Red= Reduced; Ox= Oxidized). Adapted from Patil et al.2

- dithionite appeared to be dissociation of the Fe protein from MoFe protein (kdiss= 6 s

1),17,67 but a recent study of the Fe protein cycle with the physiological reductant Fld revealed the rate-limiting step is Pi release (Chapter 3).

CAPITALIZING ON BIOLOGICAL NITROGEN FIXATION

Fundamental knowledge gained from studying the enzyme nitrogenase is essential as new technologies for sustainable fixed nitrogen production are considered. The availability of fixed nitrogen on Earth is often the major limiter of primary productivity.69

In fact, biological nitrogen fixation on its own no longer meets the world’s needs due to population growth and increased food demands over the past century.2 This has prompted the development of additional processes, including Haber-Bosch. Developed in the early

1900s, Haber Bosch manufactures nearly all of the chemically-fixed nitrogen (primarily

14 7,70,71 used in fertilizer) using an iron catalyst to make NH3 from N2 and H2. While this process has revolutionized the global food industry, it requires high temperature and pressure, consumes about 2% of the world’s annual energy production, burns fossil fuels, and emits greenhouse gasses.72 Due to the environmental impacts and unsustainable nature of Haber Bosch, there is significant interest in greener approaches to generate fixed nitrogen on a large scale.73,74 Further, with plans for human space travel to Mars, an alternative to Haber Bosch for fertilizer production will be a necessity for sustainable food production.

Nitrogenase functions optimally under mild conditions and processes developed to capitalize on it would help eliminate or minimize fossil fuel use.74 The challenge of biological nitrogen fixation is not necessarily generating the fixed nitrogen, it is accomplishing it at a scale and price that is not only practical, but competitive with industry.75 While many scientists might argue that “alternative” approaches have a long way to go, they agree that continued efforts in this area are needed.74,76,77

Recent advances in biotechnological applications of biological nitrogen fixation have primarily focused on two approaches: 1) engineering plants so they can associate with nitrogen-fixing microbes and 2) engineering plants to harbor a functional nitrogenase.78 However, another approach is emerging, in which the potential of photosynthetic diazotrophs attached to electrodes is being explored.79 An electrode offers economical source of electrons and bypasses the need for organic acids or thiosulfate typically required by the microbes, and recent work using R. palustris established its

79 ability to accept electrons from an electrode using CO2 as the sole carbon source.

15

N 2 CO2

ATP Nitrogenase RuBisCo

e- Biomaterial

NH3 (N, C)

Organic acids, thiosulfate, H2, electrode Figure 1-7. Essential steps in a phototrophic, nitrogen-fixing microbe, illustrating how these diazotrophs can use the energy from light and electrons from various sources to power nitrogenase for the production of fixed nitrogen and organic carbon. An approach like this could provide insights into applications for sustainable fixed nitrogen production.

Further, the phototrophic diazotrophs offer several key features ideal for biological nitrogen fixation on a larger-scale. They can achieve light-driven nitrogen fixation, in addition to fixing CO2 (Figure 1-7). Overall, this approach capitalizes on the light-driven synthesis of ATP with electrons provided electrochemically to produce fixed carbon and nitrogen. Initial studies as well as continued directions aimed at demonstrating CO2 and

N2 reduction via microbial electrosynthesis are addressed in Chapter 4.

CONCLUSION

Nitrogen fixation is arguably one of the most important processes on the planet, and it is hoped that the new findings presented in the following chapters provide a catalyst for further discoveries into one of the most intricate, yet, powerful biological enzymes on

Earth.

16 REFERENCES

(1) United Nations Environment Programme, and Woods Hole Research Center. (2007)

Reactive nitrogen in the environment: Too much or too little of a good thing. United

Nations Environment Programme.

(2) Patil, B. S., Hessel, V., Seefeldt, L. C., Dean, D. R., Hoffman, B. M., Cook, B. J., and

Murray, L. J. (2000) Nitrogen fixation, in Ullmann’s Encyclopedia of Industrial

Chemistry, pp 1–21. Wiley-VCH Verlag GmbH & Co. KGaA.

(3) Bruijn, F. J. de. (2015) Biological nitrogen fixation, in Principles of Plant-Microbe

Interactions, pp 215–224. Springer.

(4) Vitousek, P. M., Menge, D. N. L., Reed, S. C., and Cleveland, C. C. (2013) Biological nitrogen fixation: Rates, patterns and ecological controls in terrestrial ecosystems. Philos.

Trans. R. Soc. B Biol. Sci. 368.

(5) Dos Santos, P. C., Fang, Z., Mason, S. W., Setubal, J. C., and Dixon, R. (2012)

Distribution of nitrogen fixation and nitrogenase-like sequences amongst microbial genomes. BMC Genomics 13, 162.

(6) Maróti, G., and Kondorosi, É. (2014) Nitrogen-fixing Rhizobium-legume symbiosis: are polyploidy and host peptide-governed symbiont differentiation general principles of endosymbiosis? Front. Microbiol. 5, 326.

(7) Smil, V. (2001) Enriching the earth: Fritz Haber, Carl Bosch, and the transformation of world food production. MIT Press, Cambridge, Mass.

(8) Bishnoi, U. (2015) PGPR Interaction: An Ecofriendly Approach Promoting the

Sustainable Agriculture System, in Advances in Botanical Research, pp 81–113.

Academic Press.

17 (9) Burgess, B. K., and Lowe, D. J. (1996) Mechanism of Molybdenum Nitrogenase.

Chem. Rev. 96, 2983–3012.

(10) Hu, Y., and Ribbe, M. W. (2016) Biosynthesis of the metalloclusters of nitrogenases.

Annu. Rev. Biochem. 85, 455–483.

(11) Eady, R. R. (1996) Structure−function relationships of alternative nitrogenases.

Chem. Rev. 96, 3013–3030.

(12) Seefeldt, L. C., Hoffman, B. M., and Dean, D. R. (2009) Mechanism of Mo- dependent nitrogenase. Annu. Rev. Biochem. 78, 701–722.

(13) Harris, D. F., Lukoyanov, D. A., Shaw, S., Compton, P., Tokmina-Lukaszewska, M.,

Bothner, B., Kelleher, N., Dean, D. R., Hoffman, B. M., and Seefeldt, L. C. (2018)

Mechanism of N2 reduction catalyzed by Fe-nitrogenase involves reductive elimination of H2. Biochemistry 57, 701–710.

(14) Georgiadis, M. M., Komiya, H., Chakrabarti, P., Woo, D., Kornuc, J. J., and Rees,

D. C. (1992) Crystallographic structure of the nitrogenase iron protein from Azotobacter vinelandii. Science 257, 1653–1659.

(15) Kim, J., and Rees, D. C. (1992) Crystallographic structure and functional implications of the nitrogenase molybdenum–iron protein from Azotobacter vinelandii.

Nature 360, 553–560.

(16) Kim, J., and Rees, D. C. (1992) Structural models for the metal centers in the nitrogenase molybdenum-iron protein. Science 257, 1677–1682.

(17) Thorneley, R. N. F., and Lowe, D. J. Kinetics and mechanism of the nitrogenase enzyme, pp 221–284. Wiley-Interscience Publications, New York.

18 (18) Saeki, K. (2004) Electron transport to nitrogenase: Diverse routes for a common destination, in Genetics and regulation of nitrogen fixation in free-living bacteria, pp

257–290. Kluwer Academic Publishers.

(19) Fukuyama, K. (2011) Ferredoxins containing one [4Fe–4S] center, in Encyclopedia of Inorganic and . John Wiley & Sons, Ltd.

(20) Martin, A. E., Burgess, B. K., Iismaa, S. E., Smartt, C. T., Jacobson, M. R., and

Dean, D. R. (1989) Construction and characterization of an Azotobacter vinelandii strain with mutations in the genes encoding flavodoxin and ferredoxin I. J. Bacteriol. 171,

3162–3167.

(21) Yakunin, A. F., Gennaro, G., and Hallenbeck, P. C. (1993) Purification and properties of a nif-specific flavodoxin from the photosynthetic bacterium Rhodobacter capsulatus. J. Bacteriol. 175, 6775–6780.

(22) Edgren, T. (2006) Electron transport to nitrogenase. Dissertation, Stockholm

University, Stockholm, Sweden.

(23) Roberts, G. P., MacNeil, T., MacNeil, D., and Brill, W. J. (1978) Regulation and characterization of protein products coded by the nif (nitrogen fixation) genes of

Klebsiella pneumoniae. J. Bacteriol. 136, 267–279.

(24) John, R. T. S., Johnston, H. M., Seidman, C., Garfinkel, D., Gordon, J. K., Shah, V.

K., and Brill, W. J. (1975) Biochemistry and genetics of Klebsiella pneumoniae mutant strains unable to fix N2. J. Bacteriol. 121, 759–765.

(25) Reyntjens, B., Jollie, D. R., Stephens, P. J., Gao-Sheridan, H. S., and Burgess, B. K.

(1997) Purification and characterization of a fixABCX-linked 2[4Fe-4S] ferredoxin from

Azotobacter vinelandii. J. Biol. Inorg. Chem. 2, 595–602.

19 (26) Gangeswaran, R., and Eady, R. R. (1996) Flavodoxin 1 of Azotobacter vinelandii:

Characterization and role in electron donation to purified assimilatory nitrate reductase.

Biochem. J. 317, 103–108.

(27) Benemann, J. R., Yoch, D. C., Valentine, R. C., and Arnon, D. I. (1971) The electron transport system in nitrogen fixation by Azotobacter, III: Requirements for

NADPH-supported nitrogenase activity. Biochim. Biophys. Acta, Bioenerg. 226, 205–

212.

(28) Hageman, R. V., and Burris, R. H. (1978) Nitrogenase and nitrogenase reductase associate and dissociate with each catalytic cycle. Proc. Natl. Acad. Sci. U. S. A. 75,

2699–2702.

(29) Yoch, D. C., and Arnon, D. I. (1972) Two biologically active ferredoxins from the aerobic nitrogen-fixing bacterium, Azotobacter vinelandii. J. Biol. Chem. 247, 4514–

4520.

(30) Bennett, L. T., Jacobson, M. R., and Dean, D. R. (1988) Isolation, sequencing, and mutagenesis of the nifF gene encoding flavodoxin from Azotobacter vinelandii. J. Biol.

Chem. 263, 1364–1369.

(31) Morgan, T. V., Lundell, D. J., and Burgess, B. K. (1988) Azotobacter vinelandii ferredoxin I: cloning, sequencing, and mutant analysis. J. Biol. Chem. 263, 1370–1375.

(32) Jiménez-Vicente, E., Navarro-Rodríguez, M., Poza-Carrión, C., and Rubio, L. M.

(2014) Role of Azotobacter vinelandii FdxN in FeMo-co biosynthesis. FEBS Lett. 588,

512–516.

(33) Raina, R., Bageshwar, U. K., and Das, H. K. (1993) The ORF encoding a putative ferredoxin-like protein downstream of the vnfH gene in Azotobacter vinelandii is

20 involved in the vanadium-dependent alternative pathway of nitrogen fixation. Mol.

Gen. Genet. MGG 236, 459–462.

(34) Hill, S., and Kavanagh, E. P. (1980) Roles of nifF and nifJ gene products in electron transport to nitrogenase in Klebsiella pneumoniae. J. Bacteriol. 141, 470–475.

(35) Shah, V. K., Stacey, G., and Brill, W. J. (1983) Electron transport to nitrogenase.

Purification and characterization of pyruvate:flavodoxin oxidoreductase. The nifJ gene product. J. Biol. Chem. 258, 12064–12068.

(36) Nieva-Gómez, D., Roberts, G. P., Klevickis, S., and Brill, W. J. (1980) Electron transport to nitrogenase in Klebsiella pneumoniae. Proc. Natl. Acad. Sci. 77, 2555–2558.

(37) Carrillo, N., and Ceccarelli, E. A. (2003) Open questions in ferredoxin-NADP+ reductase catalytic mechanism. Eur. J. Biochem. 270, 1900–1915.

(38) Karplus, P. A., and Faber, H. R. (2004) Structural aspects of plant ferredoxin:

NADP+ oxidoreductases. Photosynth. Res. 81, 303–315.

(39) Medina, M., and Gómez-Moreno, C. (2004) Interaction of ferredoxin–NADP+ reductase with its substrates: Optimal interaction for efficient electron transfer.

Photosynth. Res. 79, 113–131.

(40) Bittel, C., Tabares, L. C., Armesto, M., Carrillo, N., and Cortez, N. (2003) The oxidant-responsive diaphorase of Rhodobacter capsulatus is a ferredoxin (flavodoxin)-

NADP(H) reductase. FEBS Lett. 553, 408–412.

(41) Nogués, I., Pérez-Dorado, I., Frago, S., Bittel, C., Mayhew, S. G., Gómez-Moreno,

C., Hermoso, J. A., Medina, M., Cortez, N., and Carrillo, N. (2005) The ferredoxin-

NADP(H) reductase from Rhodobacter capsulatus: Molecular structure and catalytic mechanism. Biochemistry 44, 11730–11740.

21 (42) Isas, J. M., and Burgess, B. K. (1994) Purification and characterization of a

NADP+/NADPH-specific flavoprotein that is overexpressed in FdI-strains of Azotobacter vinelandii. J. Biol. Chem. 269, 19404–19409.

(43) Jung, Y.-S., Roberts, V. A., Stout, C. D., and Burgess, B. K. (1999) Complex formation between Azotobacter vinelandii Ferredoxin I and its physiological electron donor NADPH-ferredoxin reductase. J. Biol. Chem. 274, 2978–2987.

(44) Isas, J. M., Yannone, S. M., and Burgess, B. K. (1995) Azotobacter vinelandii

NADPH:ferredoxin reductase cloning, sequencing, and overexpression. J. Biol. Chem.

270, 21258–21263.

(45) Razquin, P., Fillat, M. F., Schmitz, S., Stricker, O., Böhme, H., Gómez-Moreno, C., and Peleato, M. L. (1996) Expression of ferredoxin-NADP+ reductase in heterocysts from

Anabaena sp. Biochem. J. 316, 157–160.

(46) Biegel, E., Schmidt, S., González, J. M., and Müller, V. (2011) Biochemistry, evolution and physiological function of the Rnf complex, a novel ion-motive electron transport complex in prokaryotes. Cell. Mol. Life Sci. 68, 613–634.

(47) Hayashi, M., Nakayama, Y., and Unemoto, T. (2001) Recent progress in the Na(+)- translocating NADH-quinone reductase from the marine Vibrio alginolyticus. Biochim.

Biophys. Acta, Gen. Subj. 1505, 37–44.

(48) Steuber, J. (2001) Na+ translocation by bacterial NADH:quinone oxidoreductases: an extension to the complex-I family of primary redox pumps. Biochim. Biophys. Acta,

Bioenerg. 1505, 45–56.

22 (49) Tremblay, P.-L., Zhang, T., Dar, S. A., Leang, C., and Lovley, D. R. (2013) The

Rnf complex of Clostridium ljungdahlii Is a proton-translocating ferredoxin:NAD+ oxidoreductase essential for autotrophic growth. mBio 4, e00406-12.

(50) Schmehl, M., Jahn, A., Vilsendorf, A. M. zu, Hennecke, S., Masepohl, B.,

Schuppler, M., Marxer, M., Oelze, J., and Klipp, W. (1993) Identification of a new class of nitrogen fixation genes in Rhodobacter capsalatus: A putative membrane complex involved in electron transport to nitrogenase. Mol. Gen. Genet. MGG 241, 602–615.

(51) Saeki, K., and Kumagai, H. (1998) The rnf gene products in Rhodobacter capsulatus play an essential role in nitrogen fixation during anaerobic DMSO-dependent growth in the dark. Arch. Microbiol. 169, 464–467.

(52) Jeong, H. S., and Jouanneau, Y. (2000) Enhanced nitrogenase activity in strains of

Rhodobacter capsulatus that overexpress the rnf genes. J. Bacteriol. 182, 1208–1214.

(53) Curatti, L., Brown, C. S., Ludden, P. W., and Rubio, L. M. (2005) Genes required for rapid expression of nitrogenase activity in Azotobacter vinelandii. Proc. Natl. Acad.

Sci. 102, 6291–6296.

(54) Renalier, M. H., Batut, J., Ghai, J., Terzaghi, B., Gherardi, M., David, M.,

Garnerone, A. M., Vasse, J., Truchet, G., and Huguet, T. (1987) A new symbiotic cluster on the pSym megaplasmid of Rhizobium meliloti 2011 carries a functional fix gene repeat and a nod locus. J. Bacteriol. 169, 2231–2238.

(55) Corbin, D., Barran, L., and Ditta, G. (1983) Organization and expression of

Rhizobium meliloti nitrogen fixation genes. Proc. Natl. Acad. Sci. U. S. A. 80, 3005–

3009.

23 (56) Batut, J., Terzaghi, B., Ghérardi, M., Huguet, M., Terzaghi, E., Garnerone, A. M.,

Boistard, P., and Huguet, T. (1985) Localization of a symbiotic fix region on Rhizobium meliloti pSym megaplasmid more than 200 kilobases from the nod-nif region. Mol. Gen.

Genet. 199, 232–239.

(57) Kaminski, P. A., Norel, F., Desnoues, N., Kush, A., Salzano, G., and Elmerich, C.

(1988) Characterization of the fixABC region of Azorhizobium caulinodans ORS571 and identification of a new nitrogen fixation gene. Mol. Gen. Genet. 214, 496–502.

(58) Edgren, T., and Nordlund, S. (2004) The fixABCX genes in Rhodospirillum rubrum encode a putative membrane complex participating in electron transfer to nitrogenase. J.

Bacteriol. 186, 2052–2060.

(59) Tsai, M. H., and Saier, M. H. (1995) Phylogenetic characterization of the ubiquitous electron transfer flavoprotein families ETF-alpha and ETF-beta. Res. Microbiol. 146,

397–404.

(60) Garcia Costas, A., Poudel, S., Miller, A. F., Schut, G., Ledbetter, R., Fixen, K.,

Seefeldt, L. C., Adams, M., Harwood, C., Boyd, E., and Peters, J. (2017) Defining electron bifurcation in the electron-transferring flavoprotein family. J. Bacteriol. 199, e00440-17.

(61) Weidenhaupt, M., Rossi, P., Beck, C., Fischer, H. M., and Hennecke, H. (1996)

Bradyrhizobium japonicum possesses two discrete sets of electron transfer flavoprotein genes: fixA, fixB and etfS, etfL. Arch. Microbiol. 165, 169–178.

(62) Arigoni, F., Kaminski, P. A., Hennecke, H., and Elmerich, C. (1991) Nucleotide sequence of the fixABC region of Azorhizobium caulinodans ORS571: similarity of the

24 fixB product with eukaryotic flavoproteins, characterization of fixX, and identification of nifW. Mol. Gen. Genet. 225, 514–520.

(63) Peters, J. W., Miller, A. F., Jones, A. K., King, P. W., and Adams, M. W. (2016)

Electron bifurcation. Curr. Opin. Chem. Biol. 31, 146–152.

(64) Buckel, W., and Thauer, R. K. (2013) Energy conservation via electron bifurcating ferredoxin reduction and proton/Na+ translocating ferredoxin oxidation. Biochim.

Biophys. Acta, Bioenerg. 1827, 94–113.

(65) Herrmann, G., Jayamani, E., Mai, G., and Buckel, W. (2008) Energy conservation via electron-transferring flavoprotein in anaerobic bacteria. J. Bacteriol. 190, 784–791.

(66) Pence, N., Tokmina-Lukaszewska, M., Yang, Z.-Y., Ledbetter, R. N., Seefeldt, L.

C., Bothner, B., and Peters, J. W. (2017) Unraveling the interactions of the physiological reductant flavodoxin with the different conformations of the Fe protein in the nitrogenase cycle. J. Biol. Chem. 292, 15661–15669.

(67) Duval, S., Danyal, K., Shaw, S., Lytle, A. K., Dean, D. R., Hoffman, B. M., Antony,

E., and Seefeldt, L. C. (2013) Electron transfer precedes ATP hydrolysis during nitrogenase catalysis. Proc. Natl. Acad. Sci. 110, 16414–16419.

(68) Danyal, K., Dean, D. R., Hoffman, B. M., and Seefeldt, L. C. (2011) Electron transfer within nitrogenase: Evidence for a deficit-spending mechanism. Biochemistry 50,

9255–9263.

(69) Bernhard, A. (2010) The nitrogen cycle: Processes, players, and human impact. Nat.

Educ. 3, 25.

(70) Haber, F. Über die darstellung des ammoniaks aus stickstoff und wasserstoff.

Naturwissenschaften 10, 1041–1049.

25 (71) Haber, F. Neue arbeitsweisen- Wissenschaft und wirtschaft nach dem driege die naturwissenschaften. Naturwissenschaften 11, 753–756.

(72) Ritter, S. K. (2008) The Haber-Bosch reaction: An early chemical impact on sustainability. Chem. Eng. News.

(73) Jewess, M., and Crabtree, R. H. (2016) Electrocatalytic nitrogen fixation for distributed fertilizer production? ACS Sustain. Chem. Eng. 4, 5855–5858.

(74) Chen, J. G., Crooks, R. M., Seefeldt, L. C., Bren, K. L., Bullock, R. M.,

Darensbourg, M. Y., Holland, P. L., Hoffman, B., Janik, M. J., Jones, A. K., Kanatzidis,

M. G., King, P., Lancaster, K. M., Lymar, S. V., Pfromm, P., Schneider, W. F., and

Schrock, R. R. (2018) Beyond fossil fuel–driven nitrogen transformations. Science 360, eaar6611.

(75) Zhang, S. (2016) A chemical reaction revolutionized farming 100 years ago. Now it needs to go. Wired.

(76) Ritter, S. K. (2017) New ways to nab nitrogen. ACS Cent. Sci. 3, 512–514.

(77) Vicente, E. J., and Dean, D. R. (2017) Keeping the nitrogen-fixation dream alive.

Proc. Natl. Acad. Sci. 114, 3009–3011.

(78) Oldroyd, G. E., and Dixon, R. (2014) Biotechnological solutions to the nitrogen problem. Curr. Opin. Biotechnol. 26, 19–24.

(79) Bose, A., Gardel, E. J., Vidoudez, C., Parra, E. A., and Girguis, P. R. (2014)

Electron uptake by iron-oxidizing phototrophic bacteria. Nat. Commun. 5, 3391.

26 CHAPTER 2

THE ELECTRON BIFURCATING FIXABCX PROTEIN COMPLEX FROM

AZOTOBACTER VINELANDII: GENERATION OF LOW-POTENTIAL REDUCING

EQUIVALENTS FOR NITROGENASE CATALYSIS

ABSTRACT

The biological reduction of dinitrogen (N2) to ammonia (NH3) by nitrogenase is

an energetically demanding reaction that requires low-potential electrons and ATP;

however, pathways used to deliver the electrons from central metabolism to the

reductants of nitrogenase, ferredoxin or flavodoxin, remain unknown for many

diazotrophic microbes. The FixABCX protein complex has been proposed to reduce

flavodoxin or ferredoxin using NADH as the electron donor in a process known as

electron bifurcation. Herein, the FixABCX complex from Azotobacter vinelandii was

purified and demonstrated to catalyze an electron bifurcation reaction: oxidation of

NADH (Em= -320 mV) coupled to reduction of flavodoxin semiquinone (Em= -460 mV)

and reduction of coenzyme Q (Em= +10 mV). Knocking out fix genes rendered Δrnf A.

vinelandii cells unable to fix dinitrogen, confirming that the FixABCX system provides

another electron delivery route to nitrogenase. Characterization of the purified FixABCX

complex revealed the presence of flavin and iron- cofactors confirmed by native

mass spectrometry, electron paramagnetic resonance spectroscopy, and transient

______*Coauthored by Rhesa N. Ledbetter, Amaya M. Garcia Costas, Carolyn E. Lubner, David W. Mulder, Monika Tokmina-Lukaszewska, Jacob H. Artz, Angela Patterson, Timothy S. Magnuson, Zackary J. Jay, H. Diessel Duan, Jacquelyn Miller, Mary H. Plunkett, John P. Hoben, Brett M. Barney, Ross P. Carlson, Anne-Frances Miller, Brian Bothner, Paul W. King, John W. Peters, and Lance C. Seefeldt (2017) Biochemistry, 56:4177-4190. Copyright 2017, American Chemical Society. Reprinted with permission.

27 absorption spectroscopy. Transient absorption spectroscopy further established the presence of a short-lived flavin semiquinone radical, suggesting that a thermodynamically unstable flavin semiquinone may participate as an intermediate in electron transfer to flavodoxin. A structural model of FixABCX, generated using chemical cross-linking in conjunction with homology modeling, revealed plausible electron transfer pathways to both high- and low-potential acceptors. Overall, this study informs on a mechanism for electron bifurcation, offering insight into a unique method for delivery of low-potential electrons required for energy-intensive biochemical conversions.

INTRODUCTION

Biological dinitrogen (N2) fixation is performed by diazotrophic microbes, which harbor the enzyme nitrogenase. This enzyme converts N2 into bioavailable ammonia

(NH3) (Eq. 2-1) and accounts for at least half of the production of fixed nitrogen on

Earth.1-3

+ - N2 + 8 H + 16 MgATP + 8 e → 2 NH3 + H2 + 16 MgADP + 16 Pi (Eq. 2-1)

As summarized by Eq. 2-1, the biological reduction of N2 is an energy-demanding reaction, requiring both ATP and low reduction-potential electrons. These electrons are provided by small redox proteins, ferredoxin (Fd) and flavodoxin (Fld), which serve as direct electron donors to the iron (Fe) protein of nitrogenase.4–10 The redox active iron- sulfur cluster(s) of Fd typically access one redox couple (FdOx/Red) with midpoint

4,11 reduction potentials (Em) ranging from 0 to -645 mV. Two redox couples of the flavin

28 in Fld are accessible, including the oxidized quinone/semiquinone (FldOx/Sq) and

Sq/Hq 5,9,12 Ox/Sq semiquinone/hydroquinone (Fld ) couples. In general, the Em of the Fld couple ranges from –50 to –250 mV and the FldSq/Hq couple from -370 to -500 mV.4,7

Only the FldSq/Hq couple of Fld has enough driving force to donate electrons to nitrogenase. While much is known about other aspects of biological nitrogen fixation, pathways for delivery of the low-potential reducing equivalents for Fd and Fld reduction are not well understood for many diazotrophs.4 It has been shown that the nitrogen-fixing organism Klebsiella pneumoniae uses the anaerobic oxidation of pyruvate to reduce

Fld,13,14 and it was proposed that other microbes likely use energy associated with the proton motive force to drive reduction of a low-potential electron donor.15,16

Recently, a new mechanism for generating reductant for nitrogen fixation was put forward. Flavin-based electron bifurcation (FBEB), considered a third fundamental form of energy conservation, couples exergonic and endergonic electron transfer reactions to limit free energy loss in biological systems.17–19 FBEB exploits a favorable electron transfer event in order to drive a thermodynamically unfavorable reaction without the use of ATP or an electrochemical gradient.17–19 Several bifurcating complexes, all of which contain a flavin as the proposed site of bifurcation, have been identified in anaerobic bacteria and archaea.17,18 While these enzymes catalyze a variety of reactions, all characterized thus far use Fld/Fd as an electron donor or acceptor.20–24 For example, the electron transferring flavoprotein/butyryl-coenzyme A (Etf-Bcd) bifurcating system uses

Sq/Hq the electron donor NADH (Em= -320 mV) to reduce Fld (Em = -430 mV) or Fd (Em= -

405 mV). This thermodynamically unfavorable reaction is achieved by coupling it to an

17,22,23 exergonic one, in this case, the reduction of crotonyl-CoA (Em= -10 mV). The

29 bifurcation of electrons to both a high- and a low-potential acceptor results in an overall thermodynamically favorable reaction.17,22,23,25

Homologs of bifurcating electron transfer flavoproteins (Etfs), known as FixAB, have been found in physiologically and phylogenetically distinct nitrogen fixing organisms such as Azotobacter vinelandii,26,27 Rhodopseudomonas palustris,28

Rhodospirillum rubrum,29 and Sinorhizobium meliloti.30,31 Previous studies demonstrated that disrupting the Fix system in R. palustris, Rh. rubrum, and S. meliloti completely abolishes or significantly impairs their ability to grow under nitrogen fixing conditions, suggesting that Fix may serve as a source of electrons to nitrogenase.28–30 Given the homology between FixAB and known bifurcating Etfs, it was hypothesized that the Fix system uses electron bifurcation to generate low-potential reducing equivalents for nitrogenase.20

While the FixABCX complex is clearly linked to nitrogen fixation in many diazotrophs, its specific role has not been firmly established, nor has its ability to generate reductant for nitrogenase via electron bifurcation been demonstrated. Here, we report electron bifurcation by the FixABCX complex from the obligate aerobe A. vinelandii and characterize the Fix proteins using advanced biochemical and spectroscopic tools. Further, we provide evidence that FixABCX provides electrons for nitrogenase in A. vinelandii cells. Overall, this work establishes a new pathway for the generation of low-potential reductant required by nitrogenase and elucidates a mechanism by which biology can overcome thermodynamic barriers to accomplish a difficult reductive reaction.

30 MATERIALS AND METHODS

Construction of A. vinelandii ∆fix mutant

A. vinelandii strain DJ was the wild type strain used for physiological studies and mutant construction. Strains UW195 (∆rnf1) and UW207 (∆rnf2) were kindly donated by

Luis Rubio.32 The ∆fix mutant was generated by gene disruption with an antibiotic resistance cassette as shown in Figure A1 using primers in Table A1. Briefly, two 1.2 kb

DNA fragments that included 0.3 kb of fixA (Avin_10520) and 0.1 kb of fixC

(Avin_10540) respectively were obtained by PCR and cloned sequentially in pT7-7 ampicillin resistant vector using NdeI/BamHI and BamHI/HindIII as restriction cloning sites. The kanamycin resistance (KmR) gene (aph), was isolated as a 1.5-kb BamHI fragment from mini-Tn533 and inserted between the 1.2 kb upstream and downstream regions previously cloned in pT7-7 vector using BamHI restriction cloning site. The final construct was transformed into A. vinelandii strain DJ as described previously.34,35 KmR transformants were screened for sensitivity to ampicillin (AmpS); AmpS derivatives were assumed to have arisen from a double crossover recombination event in which the wild- type fixABC genes were replaced by the aph-containing cassette (Figure A1). This replacement was confirmed by PCR and sequencing (Figure A1, Table A1). Burk’s media36 with no nitrogen source or supplemented with ammonium acetate (1 g/L) was used for physiological analyses of A. vinelandii DJ and ∆rnf and ∆fix mutants.

Homologous overexpression of FixABCX

The fix genes of A. vinelandii were homologously overexpressed by being placed under control of the nifH promoter, which is used for the transcription of genes associated with the molybdenum-dependent nitrogenase (Figure A2). The 4.3 kb fix operon, which

31 includes six genes (fixfd, Avin_10510; fixA, Avin_10520; fixB, Avin_10530; fixC,

Avin_10540; fixX, Avin_10550; ORF6, Avin_10560) (Figure A2), was PCR amplified using primers specified in Table A2. The PCR product was digested with XbaI and

BamHI and inserted into a slightly modified pUC19 vector for blue-white screening.

Using NdeI and BamHI restriction sites, the fix genes were then cloned into a vector built specifically to support the insertion of genes behind the nifH promoter.37 The fix operon was inserted between two segments of DNA from the nif-region of A. vinelandii. The flanking regions served as sites for homologous recombination within the chromosome, allowing replacement of nifHD with fix genes (Figure A2). The presence of a streptomycin resistance gene (SmR) between the flanking regions provided a selectable marker upon transformation of the plasmid into A. vinelandii.34 The proper insertion of the fix genes into the final plasmid and the chromosome of A. vinelandii were confirmed by PCR and sequencing (Table A2).

Growth of recombinant A. vinelandii and purification of the FixABCX complex

Recombinant A. vinelandii was grown in a 100 L fermenter (New Brunswick

BioFlo 610, Hauppauge, NY) at Utah State University’ Synthetic Biomanufacturing

Institute. Cells were grown in Burk’s medium supplemented with 10 mM urea as a nitrogen source.36 In the presence of urea, the nifH promoter is repressed. When the cells reached an OD600nm of ~1.8, they were harvested in a stacked disk centrifuge (TSE 10,

GEA Westfalia, Northvale, NJ) at 14,200 x g and resuspended in Burk’s medium with no source of fixed nitrogen. Upon removal of fixed nitrogen source, the nifH promoter is derepressed and gene expression occurs. A. vinelandii was derepressed for 5 h to achieve optimal expression of Fix proteins. Because nif genes were replaced with fix genes, Fix

32 rather than nitrogenase was expressed. Following the derepression, cells were harvested and stored at -80°C until further use. Fermenter conditions were maintained as follows throughout the growth and derepression: 30°C, pH 7, 20% dissolved oxygen, and agitation at ~200 rpm.

All purification steps were performed anaerobically under an argon atmosphere.

100 g of wet cell paste was resuspended in 50 mM HEPES, pH 8, 150 mM NaCl, 1 mM dithiothreitol (DTT) and 2 mg DNase at a cell to buffer ratio of 1:5. Flavin adenine dinucleotide (FAD) (0.25 mM) was also added to the lysis buffer as it increased the flavin cofactor occupancy of the Fix complex. Lysis was achieved with a French pressure cell

(SLM Aminco FA-078, Aminco, Rochester, NY) at 200 MPa. The cell extract was centrifuged at 8,000 x g for 15 min to remove cell debris. A second centrifugation was then conducted at 50,000 x g for 2 h to obtain the cell membrane fraction. The membranes were solubilized for 1 h at 4°C in 50 mM HEPES, pH 8, 150 mM NaCl, 1%

(w/v) n-dodecyl β-D-maltoside (DDM), and 1 mM DTT at a pellet to buffer ratio of approximately 1:6. The solubilized membrane was obtained by centrifuging at 50,000 x g for 1 h and then diluted 3-fold in Buffer A (50 mM HEPES, pH 8, 0.02% (w/v) DDM, and 1mM DTT) to reduce the salt and detergent concentration before loading onto a 100 ml Q-Sepharose column. The Q-Sepharose column was prewashed with 2 column volumes of Buffer B (50 mM HEPES, pH 8, 1 M NaCl, 0.02% (w/v) DDM, and 1 mM

DTT) and then equilibrated with 2 column volumes of Buffer A. Once the protein was loaded, unbound proteins were removed with 2 column volumes of Buffer A, followed by elution of bound proteins with a salt gradient of 15-60% (5 column volumes). FixABCX eluted between 32-36% NaCl. Fractions were pooled and diluted to less than 100 mM

33 NaCl for concentration on a 15 ml Q-Sepharose column pre-washed and equilibrated as described above. After loading, the column was washed with 2 column volumes of

Buffer A and bound protein eluted with 500 mM NaCl. The resultant concentrated fraction was loaded onto a Sephacryl S-200 column equilibrated with 50 mM HEPES, pH

8, 150 mM NaCl, 0.02% (w/v) DDM, and 1 mM DTT. Fractions containing FixABCX were pooled and concentrated using an Amicon (EMD Millipore, Billerica, MA) concentrator with a 100 kDa cutoff membrane and stored in liquid nitrogen. The purity of

FixABCX was determined using SDS-PAGE and protein concentration was measured using the DC Protein Assay (Bio-Rad, Hercules, CA).

Heterologous overexpression and purification of R. palustris FixAB

R. palustris FixAB was co-expressed in E. coli strain NiCo21(DE3) (New

England Biolabs, Ipswich, MA) transformed with plasmids based on pMCSG28 and pMCSG21 (DNASU Plasmid Repository, Arizona State University) modified to carry the fixA gene with a C-terminal His-tag and the fixB gene with an N-terminal His-tag, respectively. Cells were grown in terrific broth (TB) supplemented with riboflavin (20 mg/L) and MgSO4 (2 mM) along with carbenicillin (100 μg/mL) and spectinomycin (100

μg/ml) at 37°C with shaking at 200 RPM to an OD600 of ~2. After fully cooling the culture to 18-20°C, fix gene expression was induced with β-D-1-thiogalactopyranoside

(IPTG) (0.1 mM) and cells were grown for an additional 12 hours at this lower temperature. Cells were harvested by centrifugation at 11,900 x g at 4°C for 6 min, and the pellet was stored at -80°C.

The cell pellet was suspended in BugBuster (80 ml) (EMD Millipore, Billerica,

MA) containing 4-(2-aminoethyl)benzenesulfonyl fluoride hydrochloride (AEBSF) (1

34 mM), Tris(2-carboxyethyl)phosphine hydrochloride (TCEP) (1 mM), FAD (1 mM), benzonase nuclease (20 µL) and rLysozyme (2 µL) (EMD Millipore, Billerica, MA), and further incubated at 4°C for 2 hours with stirring. After centrifugation at 20,000 x g for

30 min at 4°C, the supernatant was filtered through a 0.22 μm syringe filter. The resulting protein solution was mixed with 3 mL pre-equilibrated Ni-NTA resin and incubated overnight at 4°C with stirring. The next day, the mixture was transferred to a column at

4°C. After collecting the flow-through, the column was washed with TPGT buffer (20 mM Tris, pH 7.8, 500 mM KCl, 10% (w/v) glycerol and 1 mM TCEP) containing 20 mM, 40 mM and then 50 mM imidazole in sequence using 20, 2 and 2 column volumes for each, respectively. Finally, the column was developed with 2 column volumes of

TPGT buffer containing 100 mM imidazole and the eluate was collected in different fractions. After SDS-PAGE analysis, imidazole was removed from the pooled pure fractions by passage over a 10DG column (Bio-Rad Laboratories, Hercules, CA) equilibrated with BPGT buffer (20 mM Bis-Tris propane, pH 9.0, 200 mM KCl, 10%

(w/v) glycerol and 1 mM TCEP). Any apo-flavin sites were then reconstituted by overnight incubation of the protein in 1 mM FAD at 4°C. Excess flavin was removed by gel filtration on a 10DG column (see above) prior to prompt use or flash freezing in liquid nitrogen and storage at -80°C.

Heterologous overexpression and purification of Ro. castenholzii FixX

Ro. castenholzii FixX was overexpressed in E. coli using a pCDFDuet-1 vector modified to include a C-terminal strep-tag. Transformed cells were grown in Luria-

Bertani (LB) broth containing streptomycin (50 μg/ml) at 37°C and 250 RPM to an OD600 of 0.4-0.5. To induce expression of the fixX gene, IPTG (1.5 mM) was added to the cell

35 culture. Ammonium Fe(III) citrate (4 mM), L- (2 mM), and sodium fumarate

(10 mM) were also added. Ammonium Fe(III) citrate and cysteine increased iron-sulfur cluster occupancy and sodium fumarate served as an electron acceptor during anaerobic metabolism. The flasks were sealed with rubber stoppers containing cannulas sparging argon into the cell suspension and suited with exhaust tubes flowing into a water trap.

Cells were incubated for 4-6 h at room temperature and then anaerobically transferred to centrifugation bottles in a Coy chamber (Coy Laboratories, Grass Lake, MI) under a nitrogen atmosphere. Cells were harvested at 5,000 x g for 10 min.

All purification steps were performed anaerobically under a nitrogen or argon atmosphere. Wet cell pellets were suspended in lysis buffer (50 mM Tris-HCl, pH 8, 150 mM NaCl, 1 mM sodium dithionite (Na-dithionite), 5% glycerol, 1 μL/mL supernatant of a super saturated phenylmethylsulfonyl fluoride (PMSF) solution, 1 mg/10 mL DNAse, and 5 mg/10 mL lysozyme) at a cell to buffer ratio of 1:5. Cells were lysed using a cell bomb (Parr Bomb Instrument Co., Moline, IL) by slowly increasing the pressure to 1500

PSI and equilibrating for 30 min before collecting the lysate. This process was repeated and the supernatant collected by centrifugation at 18,000 x g for 20 min. FixX was purified using a single step affinity column purification using Strep-Tactin Superflow

Plus (Qiagen, Hilden, Germany) resin pre-equilibrated with wash buffer containing 50 mM Tris-HCl, pH 8, 150 mM NaCl, and 1 mM Na-dithionite. Once the protein was loaded, the column was washed until the baseline absorbance returned. The protein was then eluted with 50 mM Tris-HCl, pH 8, 150 mM NaCl, 1 mM Na-dithionite, and 2.5 mM D-desthiobiotin as a brown band and anaerobically concentrated using an Amicon

(EMD Millipore, Billerica, MA) centrifugal filter with a membrane molecular mass

36 cutoff of 10 kDa. The concentrated protein was flash frozen with liquid nitrogen for further analysis.

Heterologous overexpression and purification of Rh. rubrum FixAB

FixAB from Rh. rubrum was overexpressed and purified as described previously.26 Briefly, E. coli BL21 cells containing the plasmid pET101/d-

TOPO::fixAB26 were cultured anaerobically in LB supplemented with 2.8 mM glucose,

17 mM KH2PO4,72 mM K2HPO4 and 15 mg/L riboflavin at 20°C until early exponential growth phase. At this point gene expression was induced with the addition of 0.5 mM

IPTG, and cells were cultured under the same conditions for an additional 12 h.

Harvested cells were resuspended in degassed lysis buffer (20 mM Tris-Cl, pH 7.6, 10 mM imidazole, 500 mM NaCl, 5 mg/ml lysozyme, and 1 mg/ml Dnase I) and submitted to three freeze/thaw cycles. The lysate was added to a nickel-NTA column (Qiagen,

Hilden, Germany) that had been equilibrated with degassed lysis buffer. The column was washed with degassed lysis buffer, and FixAB was eluted into anaerobic vials using a 10-

300 mM imidazole gradient in degassed 20 mM Tris-HCl, pH 7.6, and 500 mM NaCl.

Yellow fractions indicating the presence of FixAB were subjected to buffer exchange using a PD-10 column (GE-Healthcare, Little Chalfont, UK) and stored in desalting buffer (20 mM Tris-HCl, pH7.6, and 100 mM NaCl). Purified FixAB was incubated overnight at 4°C with 1 mM FAD in desalting buffer to reconstitute FAD cofactors lost during purification.38,39 Unbound FAD was rinsed by concentrating the samples with

Amicon membrane (cutoff 30 kDa) followed by dilution with desalting buffer containing

0.5 µM FAD. Reconstituted samples were stored at -20°C. Purity and identity of the fractions were assessed with SDS-PAGE and mass spectrometry, respectively.

37 Dynamic light scattering

Dynamic light scattering (DLS) was performed on DynaPro NanoStar (Wyatt

Technology, Santa Barbara, CA) to determine the size distribution and hydrodynamic radii of species in solution. Samples were filtered through a 0.1 μm syringe filter prior to the experiments. DLS was measured aerobically using 10 μL of buffer as the control and also 1 mg/mL enzyme using a disposable cyclic olefin copolymer cuvette at 25°C.

Protein identification, chemical cross-linking, and model construction

Protein identification from gel bands and solution digestion was performed according to standard protocols recommended by the manufacturers using a trypsin

(Promega, Madison, WI) protease:complex ratio of 1:50-1:100 overnight and pepsin

(Sigma, St. Louis, MO) protease:complex ratio of 1:10 for 60 s. Proteins were identified as described5 using a maXis Impact UHR-QTOF instrument (Bruker Daltonics, Billerica,

MA) interfaced with a Dionex 3000 nano-uHPLC (Thermo-Fisher, Waltham, MA) followed by data analysis in Peptide Shaker.40 Chemical cross-linking was performed using 20 μg of the FixABCX complex and 1 mM bis(sulfosuccinimidyl)suberate (BS3)

(Thermo-Fisher, Waltham, MA) in 50mM HEPES, pH 7.2, 150mM NaCl buffer at room temperature for 1 h. The reaction was quenched by addition of 120 mM Tris (final concentration). The resulting mixture was separated by SDS-PAGE (4-20% linear gradient gel, Bio-Rad, Hercules, CA) and stained with Coomassie Brilliant Blue

(Thermo-Fisher, Waltham, MA). Protein bands of interest were excised from the gel and digested with trypsin as described above. For cross-link mapping, a Spectrum

Identification Machine, SIM, was used.41 Precursor and fragment ion tolerances were set to ±20ppm. Intact protein analysis was performed as described previously using a Bruker

38 Micro-TOF mass spectrometer (Bruker Daltonics, Billerica, MA).42 Non-covalent mass spectrometry experiments were conducted on SYNAPT G2-Si instrument (Waters,

Milford, MA) in a similar fashion as described.43 Briefly, the FixABCX complex sample was buffer exchanged to 200 mM ammonium acetate, pH 7 (Sigma, St. Louis, MO) using

3 kDa molecular mass cutoff spin filters (Pall Corporation, Port Washington, NY) and infused from in-house prepared gold-coated borosilicate glass capillaries to electrospray source at protein concentration of 2–3 μM and the rate around 90 nl/min. The instrument was tuned to enhance performance in the high mass-to-charge range. Settings were as follows: source temperature 30°C, capillary voltage 1.7 kV, trap bias voltage 16 V and argon flow in collision cell (trap) 7 ml/min. Transfer collision energy was held at 10 V while trap energy varied between 10–200 V. Data analysis was performed in MassLynx software version 4.1 (Waters, Milford, MA).

Protein homology models were generated by Phyre2,44 and energy minimized models were docked using ClusPro2 with restrictions derived from chemical cross- linking experiments.45–48 The flavin and iron sulfur cluster co-factors were added using

SwissDock49,50 for individual subunits and eventually added as rigid bodies to the final

FixABCX complex model. Molecular graphics were created using the UCSF Chimera package.51

Electron paramagnetic resonance spectroscopy

Electron paramagnetic resonance (EPR) spectra were recorded with a Bruker E-

500 spectrometer (X-band, 9.38 Ghz) equipped with a SHQ resonator (Bruker, Billerica,

MA), an in-cavity cryogen free VT system (ColdEdge Technologies, Allentown, PA) and

MercuryiTC temperature controller (Oxford Instruments, Abingdon, UK). Spin

39 quantifications were determined by double integration of the spectra after manual baseline subtraction in the OriginPro software package and referenced to copper- triethylamine standards (75-125 μM) measured at the same conditions. To assist with spectral deconvolution and assignment of g-factors (±0.003), computer simulations of the experimental spectra were carried out in MatLab using the EasySpin package and ‘esfit’ fitting function incorporating g-strain to replicate line broadening.

Electron bifurcation assays

FixABCX electron bifurcation activity was measured anaerobically in a UV-Vis spectrophotometer (Varian Cary 50 Bio, Agilent Technologies, Santa Clara, CA) using quartz cuvettes (d= 1 cm). All assays were carried out in 50 mM HEPES, pH 7.5, 10% glycerol and 0.02% DDM and contained the following: 0.8 μM FixABCX (1.7 nmol flavin/nmol FixABCX and 9.1 nmol Fe/nmol FixABCX), 85 μM FldSq, 200 μM NADH,

Sq Ox and 300 μM Coenzyme Q1 (CoQ1). NADH, Fld , and Fld were monitored at 340 nm

(ε= 6.2 mM-1 cm-1), 580 nm (ε= 5.7 mM-1 cm-1), and 450 nm (ε= 11.3 mM-1 cm-1), respectively.12,52 The Fld used in the assays as the low-potential electron acceptor was purified in the hydroquinone state as previously described.5 FldHq was exposed to oxygen for a short period of time, upon which the majority (>80%) of the FldHq converted to the semiquinone form. The protein was then degassed with argon, and the absorbance of the semiquinone species was measured at 580 nm. The concentration was then calculated using the extinction coefficient (ε= 5.7 mM-1 cm-1).

40 Thermodynamic calculations

+ Standard reduction midpoint potentials of the NAD /NADH (Em= -320 mV),

Ox/Sq Sq/Hq CoQ/CoQH2 (Em= +10 mV), Fld (Em= -180 mV), Fld (Em= -460 mV) half reactions were converted to standard Gibbs free energies (∆G°’) with the equation,

∆G°'= -nFE°' (Eq. 2-2) where n= electrons (mol), F= Faraday constant (96,485.34 J/V·mol) (Table A2-5). The

°' standard Gibbs free energy of reaction (∆Grxn; J/mol) can be calculated for the reaction of interest using the equation,

°' °' °' ∆Grxn= ∑|vi|∆G(products) - ∑|vi|∆G(reactants) (Eq. 2-3)

°' where |푣푖| = stoichiometric coefficient, 푖 = product or reactant, ∆G(products or reactants)= standard Gibbs free energy of the products and reactants (J/mol) calculated from Eq. 2-2.

Transient absorption spectroscopy

The ultrafast (100 fs to 5.1 ns) transient absorption spectroscopy (TAS) spectrometer employed in this study uses an amplified 4W Ti:sapphire laser (Libra,

Coherent, 800 nm, 1 kHz, 100 fs pulse width), and the Helios spectrometer (Ultrafast

Systems LLC, Sarasota, FL). A fraction of the 800 nm Libra output was frequency- doubled in beta barium borate (BBO) to produce the desired pump wavelength (480 nm for the data described here) for sample excitation, which was then directed into the

Helios. The pump pulses were passed through a depolarizer and chopped by a synchronized chopper to 500 Hz before reaching the sample. The pump pulse energy was

1.1 µJ per pulse at the sample. Another fraction of the 800 nm Libra output was guided

41 directly into the Helios for generation of the probe. Within the spectrometer, a white light continuum of wavelengths including 340–800 nm was generated using a 2 mm thick

CaF2 crystal. This beam was split into a probe and a reference beam. The probe beam was focused into the sample where it was overlapped with the pump beam. The transmitted probe and reference beams were then focused into optical fibers coupled to multichannel spectrometers with CMOS sensors with 1 kHz detection rates. The reference signal is used to correct the probe signal for pulse-to-pulse fluctuations in the white-light continuum. The time delay between the pump and probe pulses was controlled by a motorized delay stage. For all transient absorption measurements, the sample was made in an Mbraun glove-box (N2 atmosphere), sealed in a 2 mm quartz cuvette and constantly stirred to prevent photodegradation. Rh. rubrum FixAB and A. vinelandii FixABCX concentrations were approximately 150 µM and 43 µM, respectively, and they were measured in their as-isolated state (mostly oxidized). The Rh. rubrum dimer contained

>0.7 nmol flavin/nmol FixAB, and the A. vinelandii FixABCX complex contained 1.7 nmol flavin/nmol FixABCX. For the purpose of this study, light initiated the formation of the semiquinone intermediates for each flavin through generation of FAD excited state and electron donation from nearby protein residues or the other flavins. Qualitatively, this experiment shows which type of semiquinone is formed for a particular FAD site and suggests how thermodynamically stable that intermediate is, based on its lifetime. All experiments were conducted at room temperature. The change in absorbance signal (∆A) was calculated from the intensities of signals detected after sequential probe pulses with and without the pump pulse excitation. The data collection (350 pump shots per time point) was carried out three consecutive times and then averaged. The experiment was

42 repeated three times for A. vinelandii FixABCX and once for Rh. rubrum FixAB. Data were corrected for spectral chirp using SurfaceXplorer (Ultrafast Systems LLC, Sarasota,

FL). ASQ signals were fit in Igor Pro with a double exponential function. The 550 nm emissive feature in Rh. rubrum FixAB is due to stimulated emission.53

RESULTS AND DISCUSSION

Effect of fix and rnf deletions on nitrogen fixation in A. vinelandii

To determine whether the Fix system is associated with diazotrophic growth in A. vinelandii, as in R. palustris, Rh. rubrum, and S. meliloti,28,29,31 a deletion mutant lacking fixABC was generated (Figure A1). This mutant, unlike those of R. palustris, Rh. rubrum, and S. meliloti exhibited equally robust growth on solid media with and without added fixed nitrogen (Figure 2-1). This nitrogen fixing (Nif+) phenotype was also observed when the A. vinelandii fix mutant expressed the vanadium and iron-only alternative nitrogenases or when cultured under low aeration conditions (data not shown).

Figure 2-1. Phenotype of A. vinelandii DJ wt (A), Δfix (B), Δrnf1 (C), ΔfixΔrnf1 (D), Δrnf2 (E), and ΔfixΔrnf2 (F) strains. Wild type and mutant strains were cultivated in Burk’s medium supplemented with ammonium acetate (+N) or with no added fixed nitrogen (-N). All samples were grown aerobically.

43 Although the results appeared to indicate that Fix is not associated with nitrogen fixation in A. vinelandii and were consistent with a previous study by Wientjens,27 we also tested the possibility of redundancy,54 thinking that an alternative complex could participate in balancing the chemical energy and reductant pools in the absence of Fix, masking the effect of the ΔfixABC mutation.

Rnf (Rhodobacter nitrogen fixation) is a membrane-bound complex found in some diazotrophs that, like Fix, has been hypothesized to generate reductant in the form of Fd or Fld for nitrogen fixation.16,55,56 Nitrogen-fixing organisms that have genes coding for Rnf typically do not have genes coding for Fix and vice versa. Interestingly, A. vinelandii is one of the few known diazotrophs with genes coding for both complexes.57

The physiological implications of having both of Rnf and Fix are not altogether clear, but given that A. vinelandii fixes nitrogen under highly oxic conditions, the combination of

Rnf and Fix could confer the ability to fine-tune the redox status of the cell to a high degree under a wide range of oxygen tensions.

It has been suggested that Rnf uses the energy of the proton motive force to generate reduced Fd/Fld15,16 while Fix uses electron bifurcation.18,20 Neither the Rnf nor

Fix complex, however, has previously been directly implicated in electron delivery to nitrogenase. Curatti et al. showed that A. vinelandii possesses two Rnf complexes, Rnf1 associated with nitrogen fixation and Rnf2, which is expressed constitutively.32 Δrnf1 mutants, although still able to grow diazotrophically (Figure 2-1), consistently exhibit a long lag in nif gene expression and diazotrophic growth during derepression studies.32 To test whether Rnf and Fix might be compensating for one another, we transformed Δrnf mutants of A. vinelandii with the Δfix construct (Figure A1)32 to generate double mutants

44 of A. vinelandii lacking either Rnf1 and Fix, or Rnf2 and Fix. The Δfix-Δrnf1 A. vinelandii strain was able to grow when fixed nitrogen was supplied, but no growth was observed under nitrogen fixing conditions (Figure 2-1). Therefore, cells lacking both

Rnf1 and Fix displayed a non-nitrogen fixing (Nif-) phenotype. Our data provide support for a role of both Fix and Rnf1 in providing reducing equivalents to nitrogenase and indicate that the two complexes have compensatory activities, such that either one can replace the other to ensure electron flow. Furthermore, this is the first study to link the

Fix complex to nitrogen fixation in A. vinelandii. The Δfix-Δrnf2 A. vinelandii mutant control strain was able to grow with or without added fixed nitrogen (Figure 2-1), confirming that Rnf1, but not Rnf2, can support nitrogen fixation.

Characterization of the FixABCX complex

To understand how Fix is able to support nitrogen fixation, we characterized the

FixABCX complex and its activity in vitro. The fix operon of A. vinelandii consists of six genes, fixFd (annotated as fixP in some literature), fixA, fixB, fixC, fixX, and Avin_10560

(designated ORF6) (Figure A2). FixAB is homologous to the previously characterized

EtfAB, which is part of the Etf-Bcd bifurcating complex. FixCX is similar to Etf ubiquinone oxidoreductase (Etf-QO), involved in transfer of electrons to the quinone pool in the membrane.29 Based on the similarity to other Etf systems, it is hypothesized that the Fix complex oxidizes NADH and bifurcates one electron to the high-potential quinone pool (exergonic branch) and drives the other electron to a low-potential acceptor in the form of Fd or Fld (endergonic branch) (Figure 2-2).20

45

Figure 2-2. (A) Proposed mechanism of FixABCX electron bifurcation in A. vinelandii, operating within the framework of the known oxidation-reduction potentials of NADH, Fld/Fd, and CoQ. It is hypothesized that crossed potentials of the bifurcating flavin (a- FAD) promote electron bifurcation based on the appearance of a short-lived ASQ. Once the first electron transfers into the exergonic branch, the second electron, sitting at a very low reduction potential, is thermodynamically unstable and immediately transfers into the endergonic branch composed of the low potential iron-sulfur clusters of FixX. The exergonic branch is set up to ensure electron delivery to c-FAD where electrons accumulate before being transferred to CoQ. The specific oxidation-reduction potential levels are qualitative unless otherwise noted. (B) Electron transfer pathways illustrated in FixABCX structural model generated by docking (ClusPro2) four homology models (Phyre2) (Figure A2-8).

46 Most microbes with fix genes only have fixABCX; homologs fixFd and ORF6 have not been investigated in fix gene clusters of any other species thus far.27,58 Although

FixFd and ORF6 were not specifically addressed in this study, it is thought that FixFd, a small iron-sulfur protein,27,59 may serve as a low-potential acceptor in Fix electron bifurcation. ORF6 is described as being ferritin-like, but studies on its function have yet to be conducted.

The fix genes from A. vinelandii were homologously overexpressed under control of the nifH promoter (Figure A2), which regulates transcription of genes encoding molybdenum nitrogenase. This approach allows gene(s) of interest to be overexpressed upon removal of a fixed nitrogen source. Based on the hypothesis that the FixABCX complex delivers one electron per bifurcation reaction to the quinone pool (coenzyme Q,

CoQ), it was anticipated that the enzyme could be associated with the membrane fraction, as observed for Etf-QO.60,61 Indeed, a previous study found that heterologously overexpressed Rh. rubrum FixC was localized to the membrane fraction in Escherichia coli.29 Furthermore, an N-terminal lipophilic tail and transmembrane helix were predicted on the FixC subunit based on the amino acid sequence (SACS MEMSAT2 and Phyre 2 prediction software).44,62 Purification attempts for the FixABCX complex confirmed membrane association. After trying a range of NaCl concentrations (0-1 M) and a variety of non-ionic detergents to solubilize the complex, the detergent n-dodecyl β-D-maltoside

(DDM) used for many membrane proteins,61,63 was found to be most effective for the isolation of intact FixABCX in a soluble and active form.

Upon extraction of the FixABCX complex from the membrane, the DDM- solubilized fraction was subjected to anion-exchange and size-exclusion chromatography.

47 SDS-PAGE analysis of the purified FixABCX complex revealed four bands corresponding in size to FixA, FixB, FixC, and FixX (Figure 2-3) and was >80% homogenous (defined by percent band (Image Lab, BioRad)). Mass spectrometry confirmed the identities of the four Fix subunits (Figure A3), and densitometry of the stained gel indicated that the band intensities were consistent with a subunit stoichiometry of 1:1:1:1 for A:B:C:X.

Dynamic light scattering DLS was used to determine the oligomeric state of the FixABCX complex on the basis of particle size. The complex in buffer containing 0.02% DDM displayed an average

Figure 2-3. SDS-PAGE gel of purified (~80%) FixABCX complex from A. vinelandii (extra lanes were removed). All Fix protein bands were verified by mass spectrometry (Figure A2-3) and according to densitometry, the subunit stoichiometry of FixABCX is 1:1:1:1.

48 hydrodynamic radius of 5.1 ± 0.3 nm, corresponding to an estimated molecular mass of

150 kDa versus the expected 129 kDa for the FixABCX heterotetramer (Figure A4). This slightly larger-than-expected complex is consistent with attachment of DDM to the protein tetramers. While there was also some contribution from larger aggregates, which are very prominent in the DLS output due to their much larger scattering efficiencies

(≈12-16 nm), the aggregates comprised <1% of the population (Figure A4). Buffer containing DDM was used as a negative control and produced micelles with a radius of

≈4.3 nm (Figure A4). Overall, the DLS data suggest FixABCX exists as individual heterotetramers in solution.

FixABCX ligand composition and structural model

Multiple electron transfer cofactors, including flavins and iron-sulfur clusters, have been identified in electron bifurcating enzyme complexes, with a single flavin proposed to serve as the site of bifurcation.17,18,20,25 Based on common sequence motifs for ligand binding, FixABCX was predicted to have 3 FAD moieties, one each in FixA,

FixB, and FixC, and two [4Fe-4S] clusters in FixX. Native mass spectrometry (MS) was employed to reveal the subunit stoichiometry and ligand composition of the complex.

All complex members were first identified based on fragments produced by two different proteases, trypsin and pepsin. Additionally, proteolysis reactions revealed that

FixABCX and a small amount of Fld (NifF) purify together. The next step was to determine the molecular masses of the intact proteins. Reverse phase LC-MS confirmed full-length FixA and that FixB and FixC subunits did not contain N-terminal Met residues. The purified FixABCX complex, under anaerobic conditions, was analyzed using native MS conditions by direct infusion on a Synapt G2-Si. Under standard

49 conditions (60V collision energy), the complex dissociated into subcomplexes of

FixAB and FixCX. A consistent ion signal for the tetrameric FixABCX complex could not be maintained despite the presence of all four proteins and cofactors in the sample after size-based purification. The FixAB component contained two FAD molecules and closely matched the expected size (71,414.8 Da vs predicted MW = 71,413.7 Da) (Figure

2-4A and B); whereas, the FixCX subcomplex containing FAD and two [4Fe-4S]

Figure 2-4. A. vinelandii FixABCX complex composition. (A) Native mass spectrum of FixAB containing two FAD cofactors (71,414.8 Da; calculated MW = 71,413.7 Da) generated during FixABCX complex activation in the gas phase. Red diamonds signify charge state envelope centered around charge 18+. The unmarked masses are the charge state envelop of the molecular chaperone DnaK. (B) Sub-complexes obtained during FixABCX complex activation. (C) SDS-PAGE gel of FixABCX complex (right) cross- linked with BS3 reagent (left). The most distinguished bands in cross-linking reaction, migrating around the 60 kDa, 70 kDa, and 88 kDa molecular weight marker, were identified as FixCX, FixAB and FixBC dimers respectively. (D) Protein-protein interaction map based on highest scoring cross-links (score 7 and higher; red and blue lines indicate intra and inter cross-links, respectively). A complete list of generated cross- links can be found in Table A2-3.

50 clusters was slightly heavier than expected (60,310.1 Da vs predicted MW = 59,932.9

Da: FixCX) (Figure 2-4B). The additional mass of 377.2 Da matches very closely to riboflavin (Rf) (MW = 376.4 Da), raising the intriguing possibility that it may also be associated with the complex (Figure 2-4B).

The UV-visible spectrum of FixABCX revealed a distinct peak at 428 nm, as well as oxidized flavin signatures with shoulders near 460 nm and 365 nm (Figure 2-5 (Black trace)). After denaturation of FixABCX and removal of the protein by centrifugation, the soluble fraction demonstrated clear signatures of flavins with broad maxima at 370 and

450 nm (Figure 2-5 (Red trace)).

The band at 428 nm is not a usual signature of flavins, and such a band has not been observed in FixAB from other diazotrophs that have been expressed in E. coli.

Thus, it is likely that this species is associated with FixC or FixX and may be covalently attached to the protein, as it does not remain in the supernatant when denatured protein is removed. The signatures of the [4Fe-4S] clusters are not evident in as-purified FixABCX, but are not expected to be prominent because of scattering and their broad and relatively weak signals (400-420 nm, ε≈ 2-4 mM-1 cm-1).64

Once we had established the subunit stoichiometry and ligand composition of the

FixABCX complex, we investigated the quaternary interactions that define the active enzyme. Models for each of the Fix subunits were generated using Phyre244, and the figures of merit are provided in Figures A5-A8. To validate these subunit models and to elucidate protein-protein interactions within the FixABCX complex, chemical cross- linking was used. The bis(sulfosuccinimidyl)suberate (BS3) cross-linking reagent is homobifunctional, reacting with primary amines and hydroxyl groups to form covalent

51

Figure 2-5. UV-visible spectrum of the FixABCX complex from A. vinelandii demonstrating the presence of flavin. Black: as-purified FixABCX, red: cofactors released upon denaturation of FixABCX, and blue: flavins as observed in the spectrum of FixAB from R. palustris, vertically offset by -0.05 AU for clarity. The two spectra representing FixABCX were corrected for Raleigh scattering by fitting the baseline to the equation for scattering and then subtracting the fit from the measured spectrum to obtain a corrected spectrum. A. vinelandii FixABCX was prepared in 50 mM Tris-HCl, pH 8, 150 mM NaCl, and 0.02% (w/v) DDM. R. palustris FixAB was prepared in 20 mM Bis-Tris propane, pH 9, 200 mM KCl, 10% (w/v) glycerol and 1 mM tris(2 carboxyethyl)phosphine (TCEP).

bonds. Cross-linked samples were separated using 1D SDS-PAGE to confirm connections between subunits and to enrich cross-linked species in samples subsequently analyzed by liquid chromatography tandem mass spectrometry (LC-MS/MS) (Figure 2-

4C and Table A3). Overall, the reaction of the FixABCX complex with BS3 reagent produced over 200 connections. For initial model building of the FixABCX complex and to confirm the subunit homology models, a reduced set of cross-links of very high confidence were selected (probability score >7, significance is >3) (Figure 2-4D).

52 The resulting model of the quaternary structure provides insightful hypotheses to be tested in next generation of experiments (Figures 2-2B, A8E). The MS-based model places the cofactors in locations that suggest a pair of possible paths for electron transfer, both emanating from the flavin in FixA. One provides a plausible route to the

[4Fe-4S] clusters in FixX and the other can provide a path to the site at which quinone is found to be bound in the functional homolog of FixC, the Etf-QO (Figures 2-2B, A8E).65

While some of the distances between cofactors are longer than ideal for direct electron transfer, conformational changes could be invoked to bring them sufficiently close,22,39,66 but we also note that the model places conserved Trp and Tyr in locations that could provide electron transfer paths between cofactors (Figure A8E).67–69 Thus, our model provides testable hypotheses for which amino acids are expected to alter complex stability or internal electron transfer upon mutagenesis. These amino acids will be targets of experiments to elucidate fundamental elements of bifurcating activity and test attribution of bifurcating activity to FixABCX.

EPR spectroscopy

FixABCX was analyzed by low-temperature continuous wave X-band EPR spectroscopy to assess its flavin radical content and identity of iron-sulfur clusters. The

EPR spectrum of the as-purified enzyme showed a very weak fast relaxing, broad signal that was strongest near 5 K along with a weak radical signal near g = 2 (Figure 2-6,

Table A4). Upon reduction with either NADH or Na-dithionite, the broad signal almost disappeared and was replaced with multiple signals in the S = ½ region indicative of FeS clusters and a flavin radical. Both treatments yielded overlapping rhombic g = 2.07 (g =

2.072, 1.940, 1.895) and axial g = 2.04 (g = 2.041, 1.944, 1.944) signals consistent with

53

Figure 2-6. EPR spectra of FixABCX from A. vinelandii prepared in 50 mM HEPES pH 7.5, 150 mM NaCl, and 5% glycerol. Black: As-prepared FixABCX (100 μM); Blue: FixABCX (100 μM) reduced with NADH (1 mM); Red: FixABCX (100 μM) reduced with sodium dithionite (10 mM). Simulations of spectra of NADH- and Na-dithionite- reduced FixABCX are shown in lighter colored traces. Microwave frequency, 9.38 GHz; microwave power, 1 mW; modulation frequency, 100 kHz; modulation amplitude, 10.0 G; sample temperature, 10 K.

the presence of two iron-sulfur clusters, in addition to an isotropic g = 2.0 signal consistent with a flavin radical that was dominant above 50 K (Figure A9). The temperature dependencies of the rhombic g = 2.07 and axial g = 2.04 signals were characteristic of [4Fe-4S] clusters,70 in agreement with the sequence prediction for a pair of [4Fe-4S] clusters. The optimal temperature (Topt) for the rhombic g = 2.07 signal was

54 near 15 K with broadening above 30 K (Figure A9). The axial g = 2.04 signal was faster relaxing with a lower Topt between 5 and 10 K and broadening above 15 K. Also present for the NADH-reduced sample was a highly temperature dependent axial g = 2.03 signal (g = 2.030, 2.00, 2.00) observed at 10 K. The signal appeared to correlate with a slight loss in intensity from the axial g = 2.04 signal compared to the Na-dithionite treatment, suggesting that this feature could arise from the spin interaction between the faster relaxing [4Fe-4S] cluster and a flavin radical.

Further EPR analysis of the FixX subunit alone (from Ro. castenholzii) identified very similar rhombic g = 2.07 and axial g = 2.04 signals indicating that these signals in

FixABCX originate from the FixX subunit (Figure A10). Line broadening at the high- and low-field edges of the reduced spectra was observed for both FixABCX and FixX and is indicative of spin coupling between the two [4Fe-4S] clusters.71 For FixABCX, the different relative intensities of the rhombic g = 2.07 and axial g = 2.04 signals suggest that the corresponding [4Fe-4S] clusters have different Em values. Accordingly, for

FixABCX both clusters are thought to have Em values below that of Na-dithionite (≤-440 mV vs. SHE, pH 7.5), since the sample appeared to be only partially reduced by Na- dithionite. We hypothesize that the [4Fe-4S] cluster giving rise to the axial g = 2.04 signal has a lower Em, due to its weaker contribution to the spectrum, based on simulations (Table A4).

Altogether, the two [4Fe-4S] cluster EPR signals show striking resemblance to the

EPR signals assigned to the two [4Fe-4S] clusters in the NADH-dependent ferredoxin:NADP+ oxidoreductase I (Nfn) bifurcating enzyme, as the signals share similar g-values, temperature dependence, and oxidation-reduction properties.25 For the

55 Nfn enzyme, structural and biophysical analyses showed that two [4Fe-4S] clusters in the large subunit form an electron-transfer chain from the flavin site of electron- bifurcation to an external ferredoxin redox partner. Spin-coupling between the two clusters is thought to facilitate electron-transfer between the redox cofactors. It was also found for the Nfn enzyme that one of the [4Fe-4S] clusters had an unusually low Em that creates a thermodynamically favorable electron transfer pathway from a highly unstable semiquinone intermediate formed during electron bifurcation. An analogous model is suggested here for FixABCX where the two low-potential [4Fe-4S] clusters in the FixX subunit would form a thermodynamically favorable pathway for facile electron transfer between the site of bifurcation and Fd/Fld.

Electron bifurcation by the FixABCX complex

Electron bifurcation by the FixABCX complex is proposed to drive the endergonic reduction of FldSq by coupling it to the exergonic reduction of CoQ using

NADH as an electron donor (Figures 2-2, 2-7A). Electron bifurcation by the FixABCX complex was demonstrated by incubating FixABCX with the electron donor NADH, high

Sq potential acceptor coenzyme Q1 (CoQ1), and low-potential acceptor Fld . CoQ1 served as an amphipathic analog of the physiological acceptor coenzyme Q8 (CoQ8) found in the electron transport chain of A. vinelandii.72,73 FldSq was used as the low-potential acceptor in these assays since FldHq has been shown to be a direct electron donor to nitrogenase.5,6

Similarly, the previously characterized electron bifurcating Etf-Bcd complex could direct electrons to either Fd or Fld,22,23 suggesting that Fld could likely serve as an electron acceptor from FixABCX as well.

56

Figure 2-7. Electron bifurcation by the FixABCX complex of A. vinelandii under anaerobic conditions. (A) Overview of electron bifurcation by the FixABCX complex showing the NADH bifurcating an electron to the high-potential acceptor, CoQ1 and the other to the low-potential acceptor, FldSq. Evidence of electron bifurcation was obtained using UV-visible spectrophotometry. NADH oxidation as well as FldSq reduction and oxidation were monitored over time as signatures of the bifurcation reaction. (B) NADH oxidation at 340 nm, C) FldSq reduction/oxidation at 580 nm, and D) formation of FldOx at 450 nm. Final concentrations of components added to the reactions were as follows: Sq 0.8 μM FixABCX, 85 μM Fld , 200 μM NADH, and 300 μM CoQ1. Data were normalized and demonstrate the overall change in absorbance.

For the electron bifurcation assay, the oxidation of NADH (340 nm), disappearance of FldSq (580 nm), and formation of FldOx (450 nm) were monitored over time. CoQ1 reduction (λmax= 274 nm) was not recorded, due to significant interference at that wavelength. While the simultaneous presence of several redox active components in

57 the assays posed a challenge, monitoring activities at multiple wavelengths provided evidence of electron bifurcation. Control reactions for NADH oxidation (diaphorase)

Sq activity linked to CoQ1 reduction in the absence of Fld yielded a specific activity of 396

± 44 nmol/min/mg with a turnover frequency of 51 ± 6 min-1 (Figure 2-7B (No FldSq)).

Bifurcation reactions containing the additional component needed for bifurcation (FldSq, in addition to the control reactants FixABCX, NADH, and CoQ1) exhibited 25% greater

NADH oxidation activity with a specific activity of 524 ± 21 nmol/min/mg and turnover frequency of 68 ± 3 min-1 (Figure 2-7B (All components)). It is important to note that the specific activity values are not adjusted for flavin occupancy, since it could not be confirmed how much of the protein was fully occupied (how flavins were distributed among the flavin binding sites).

The Fld-dependent increase in NADH consumption is consistent with bifurcation, and previous studies on the Etf-Bcd bifurcating complex also demonstrated increased

22,23 NADH consumption in a bifurcating reaction. Omission of FixABCX or CoQ1 resulted in baseline NADH oxidation activity (Figure 2-7B). To decipher whether the enhanced consumption of NADH in the presence of FldSq could be attributed to the endergonic reduction of FldSq to FldHq (the bifurcation reaction), disappearance of FldSq was monitored at 580 nm where optical absorption by the hydroquinone species is negligible but absorbance by the semiquinone is strong (ε= 5.7 mM-1 cm-1). Reactions monitoring the disappearance of FldSq were difficult to interpret on their own, due to background activity in the absence of FixABCX (Figure 2-7C). In vivo, CoQ is in the membrane, whereas, Fld is in the cytoplasm so direct contact between them is diminished and spontaneous electron transfer between the two species is suppressed (Table A5). In

58 vitro experiments however lack this separation, and our controls revealed that FldSq is able to donate electrons to CoQ1 in the absence of FixABCX (Figure 2-7C). Thus, the

FldSq concentration decreased while an increase at 450 nm, characteristic of FldOx, was

Sq observed (Figure 2-7D). This non-enzymatic reduction of CoQ1 by Fld is consistent with the favorable ΔG° of -36.7 kJ/mol (Table A5). However, the increase in absorbance at 450 nm produced by spontaneous FldSq oxidation enabled us to account for this reaction that detracts from the apparent yield of the bifurcating reaction (Figure 2-7C

(All components)). The FldOx species was not detected in the bifurcating reaction assays indicating that FldHq, the expected product of the bifurcation reaction, was formed instead

Hq (Figure 2-7D). However, we do not expect Fld to accumulate in the presence of CoQ1 due to the favorability of electron transfer between these two that will return Fld to its semiquinone state and thus conceal the true extent of FldHq production (ΔG°= -154 kJ/mol) (Eq. 2-4, Table A5).

Hq + Sq + NADH + 2 Fld + 2 CoQ + 3 H = 2 CoQH2 + 2 Fld + NAD (Eq. 2-4)

Overall, the absence of FldOx in conjunction with the increased activity of the electron bifurcation reaction in comparison to diaphorase activity, suggest that the

FixABCX complex can in fact perform electron bifurcation. To further investigate the flavin cofactors in FixABCX, transient absorption spectroscopy (TAS) was conducted.

Transient absorption spectroscopy

A comparative TAS study between FixAB from Rh. rubrum and FixABCX from

A. vinelandii was performed in order to gain qualitative information on spectral

59 contributions from individual flavins. TAS of the incomplete Fix complex, FixAB, revealed an anionic semiquinone (ASQ) absorption at 365 nm with corresponding oxidized (Ox) flavin bleach at 447 nm. The ASQ absorption decays with two components: a short-lived component with a lifetime of tens of picoseconds and a longer short-lived component with a lifetime of 1000 picoseconds (Figure 2-8A, (Red trace),

Table A6). A similar ASQ signal was also observed in the A. vinelandii FixABCX complex (Figure 2-8A (Black trace)). This suggests that two ASQ species can be formed in the FixAB unit, consistent with the presence of two flavins. Bifurcating flavins have

Ox/SQ been shown to exhibit crossed potentials, whereby the Em is at lower potential than

SQ/HQ the Em with the result that the high-energy SQ intermediate does not appreciably accumulate relative to Ox and HQ.

Figure 2-8. Transient absorption spectra of as-prepared Rh. rubrum FixAB (red traces) and A. vinelandii FixABCX (black traces). (A) Kinetic traces of ASQ signal (dots) at 365 nm. ASQ decay shows half-lives of ~15 ps and ~1000 ps when fit with a double exponential function (solid lines). (B) An NSQ absorption peak (~565-650 nm) is only observed in the A. vinelandii FixABCX complex.

60 A short-lived ASQ has been observed in the bifurcating Nfn from Pyrococcus furiosus and is proposed to play a central role in bifurcation.19,25 The current observation of relatively short-lived ASQ signals in Rh. rubrum FixAB and A. vinelandii FixABCX is consistent with the work on Nfn, and suggests that neither flavin significantly stabilizes a one-electron reduced species. Based on comparison with Etf, the flavin in FixA (a-FAD)

(analogous to the β-FAD in EtfB) is the proposed site of bifurcation and likely corresponds to the tens of picosecond component in our TAS experiments.22 This assignment then implies that the FixB b-FAD acts as an electron transferring flavin, consistent with a slightly longer-lived ASQ, since accumulation of electrons at this site may congest electron flow and impede bifurcation. This may additionally play a role in gating electron transfer, ensuring that only one electron follows the exergonic path, and thus restricting the lower potential electron to travel down the endergonic branch, similar to the mechanism of electron bifurcation in P. furiosus Nfn.25 Due to the crossed potentials believed to characterize bifurcating flavins, the more negative potential electron of the bifurcating flavin (ASQ to Ox transition) would provide enough driving force to reduce the low-potential iron-sulfur clusters (predicted ≤-440 mV vs. SHE, pH

7.5) in FixX directly. In the complete FixABCX complex from A. vinelandii, an additional peak is observed in the TAS corresponding to a neutral semiquinone (NSQ) and is assigned to the c-FAD in FixC (Figure 2-8B). We propose that the c-FAD serves to accumulate two electrons from successive bifurcations in order to permit 2-electron reduction of CoQ.

61 Proposed mechanism for FixABCX electron bifurcation

The data presented herein establish that the FixABCX complex from A. vinelandii

Sq can bifurcate electrons from NADH to CoQ1 and Fld . The biochemical and biophysical evidence is consistent with a proposed mechanism of Fix electron bifurcation initiated at the flavin in FixA, which would accept a pair of electrons from NADH and pass one electron to the quinone pool via the flavins in FixB and FixC, and the other to Fld/Fd via the low potential [4Fe-4S] clusters in FixX, with the energetic cost being paid by favorable transfer of the former electron to the quinone pool (Figure 2-2). TAS on the

FixABCX complex provides evidence by demonstrating the presence of a short-lived anionic flavin semiquinone consistent with the ability to support rapid and efficient electron transfer. The findings here provides experimental support for the mechanism proposed previously for a homologous bifurcating Etf that assigns the site of bifurcation to the flavin bound by FixA (EtfB).22 This assignment is further supported by the demonstration by Sato et al. that the flavin reduced by NADH is the one bound in FixA

(EtfB), but that electron transfer to the flavin in FixB (EtfA) is favorable.39 Thus, upon reduction of the FixA flavin, transfer of an electron through the exergonic branch to CoQ via Fix B and FixC can leave the second electron in a very unstable, highly energetic

Ox/Sq flavin semiquinone state with a low Em able to drive reduction of Fld/Fd, via FixX

(Figure 2-2).

The mechanism presented here offers a means by which Fd/Fld can be reduced by

NADH in a reaction that is thermodynamically favorable overall and identifies a new pathway by which low-potential reducing equivalents can be generated to drive nitrogen fixation.

62 ACKNOWLEDGMENTS

The authors would like to thank Stefan Nordlund at Stockholm University and

Tomas Edgren at Umeå University for the plasmid containing Rh. rubrum FixAB and their guidance in Fix protein expression and purification and Luis Rubio for the A. vinelandii rnf mutants. The authors would also like to acknowledge Utah State

University’s Synthetic Biomanufacturing Institute for conducting large-scale and Montana State University’s Microfabrication Facility for help in preparation of gold-coated borosilica capillaries for non- covalent mass spectrometry.

This work is supported as part of the Biological and Electron Transfer and Catalysis

(BETCy) EFRC, an Energy Frontier Research Center funded by the U.S. Department of

Energy, Office of Science (DE-SC0012518). The Mass Spectrometry Facility at MSU is supported in part by the Murdock Charitable Trust and an NIH IDEA program grant

P20GM103474.

AUTHOR CONTRIBUTIONS

RNL, MHP, BMB, and LCS, FixABCX genetics; RNL, TSM and LCS, FixABCX purification and characterization; AMGC and JWP, fix and rnf mutational analyses and

FixAB purification; MTL, AP, and BB, Mass spectrometry and chemical cross-linking;

MTL, AP, BB, and AFM, Molecular modeling; CEL and PWK, Transient absorption spectroscopy; DWM, JHA, and PWK, Electron paramagnetic resonance; HDD, and

AFM, Dynamic light scattering; JPH, RNL, and AFM, Flavin content; ZJJ, RNL, and

RPC, Thermodynamics; JM, FixX purification; All authors involved in data interpretation, writing, and revisions.

63 REFERENCES

(1) Kim, J., and Rees, D. C. (1994) Nitrogenase and biological nitrogen fixation.

Biochemistry 33, 389–397.

(2) Burgess, B. K., and Lowe, D. J. (1996) Mechanism of molybdenum nitrogenase.

Chem. Rev. 96, 2983–3012.

(3) Seefeldt, L. C., Hoffman, B. M., and Dean, D. R. (2009) Mechanism of Mo- dependent nitrogenase. Annu. Rev. Biochem. 78, 701–722.

(4) Saeki, K. (2004) Electron transport to nitrogenase: Diverse routes for a common destination, in Genetics and Regulation of Nitrogen Fixation in Free-living Bacteria

(Klipp, W., Masepohl, B., Gallon, J. R., and Newton, W. E., Eds.), pp 257–290. Kluwer

Academic Publishers, Dordrecht, The Netherlands.

(5) Yang, Z. Y., Ledbetter, R., Shaw, S., Pence, N., Tokmina-Lukaszewska, M., Eilers,

B., Guo, Q., Pokhrel, N., Cash, V. L., Dean, D. R., Antony, E., Bothner, B., Peters, J. W., and Seefeldt, L. C. (2016) Evidence that the Pi release event is the rate-limiting step in the nitrogenase catalytic cycle. Biochemistry 55, 3625–3635.

(6) Martin, A. E., Burgess, B. K., Iismaa, S. E., Smartt, C. T., Jacobson, M. R., and Dean,

D. R. (1989) Construction and characterization of an Azotobacter vinelandii strain with mutations in the genes encoding flavodoxin and ferredoxin I. J. Bacteriol. 171, 3162–

3167.

(7) Deistung, J., and Thorneley, R. N. (1986) Electron transfer to nitrogenase:

Characterization of flavodoxin from Azotobacter chroococcum and comparison of its redox potentials with those of flavodoxins from Azotobacter vinelandii and Klebsiella pneumoniae (nifF-gene product). Biochem. J. 239, 69–75.

64 (8) Mortenson, L. E. (1964) Ferredoxin and ATP, requirements for nitrogen fixation in cell-free extracts of Clostridium pasteurianum. Proc. Natl. Acad. Sci. U.S.A. 52, 272–

279.

(9) Yates, M. G. (1972) Electron transport to nitrogenase in Azotobacter chroococcum:

Azotobacter flavodoxin hydroquinone as an electron donor. FEBS Lett. 27, 63–67.

(10) Thorneley, R. N., and Deistung, J. (1988) Electron-transfer studies involving flavodoxin and a natural redox partner, the iron protein of nitrogenase: Conformational constraints on protein-protein interactions and the kinetics of electron transfer within the protein complex. Biochem. J. 253, 587–595.

(11) Cammack, R. (1992) Iron-sulfur clusters in enzymes: Themes and variations, in

Advances in Inorganic Chemistry (Cammack, R., Ed.), pp 281–322. Academic Press, San

Diego.

(12) Klugkist, J., Voorberg, J., Haaker, H., and Veeger, C. (1986) Characterization of three different flavodoxins from Azotobacter vinelandii. Eur. J. Biochem. 155, 33–40.

(13) Ludden, P. W. (1991) Energetics of and sources of energy for biological nitrogen fixation, in Current Topics in Bioenergetics (Lee, C. P., Ed.), pp 369–390. Academic

Press.

(14) Shah, V. K., Stacey, G., and Brill, W. J. (1983) Electron transport to nitrogenase.

Purification and characterization of pyruvate:flavodoxin oxidoreductase. The nifJ gene product. J. Biol. Chem. 258, 12064–12068.

(15) Saeki, K., and Kumagai, H. (1998) The rnf gene products in Rhodobacter capsulatus play an essential role in nitrogen fixation during anaerobic DMSO-dependent growth in the dark. Arch. Microbiol. 169, 464–467.

65 (16) Schmehl, M., Jahn, A., Vilsendorf, A. M. zu, Hennecke, S., Masepohl, B.,

Schuppler, M., Marxer, M., Oelze, J., and Klipp, W. (1993) Identification of a new class of nitrogen fixation genes in Rhodobacter capsalatus: A putative membrane complex involved in electron transport to nitrogenase. Molec. Gen. Genet. 241, 602–615.

(17) Buckel, W., and Thauer, R. K. (2013) Energy conservation via electron bifurcating ferredoxin reduction and proton/Na+ translocating ferredoxin oxidation. Biochim.

Biophys. Acta, Bioenerg. 1827, 94–113.

(18) Peters, J. W., Miller, A. F., Jones, A. K., King, P. W., and Adams, M. W. (2016)

Electron bifurcation. Curr. Opin. Chem. Biol. 31, 146–152.

(19) Nitschke, W., and Russell, M. J. (2012) Redox bifurcations: Mechanisms and importance to life now, and at its origin. Bioessays 34, 106–109.

(20) Herrmann, G., Jayamani, E., Mai, G., and Buckel, W. (2008) Energy conservation via electron-transferring flavoprotein in anaerobic bacteria. J. Bacteriol. 190, 784–791.

(21) Li, F., Hinderberger, J., Seedorf, H., Zhang, J., Buckel, W., and Thauer, R. K. (2008)

Coupled rerredoxin and crotonyl coenzyme A (CoA) reduction with NADH catalyzed by the butyryl-CoA dehydrogenase/Etf complex from Clostridium kluyveri. J. Bacteriol.

190, 843–850.

(22) Chowdhury, N. P., Mowafy, A. M., Demmer, J. K., Upadhyay, V., Koelzer, S.,

Jayamani, E., Kahnt, J., Hornung, M., Demmer, U., Ermler, U., and Buckel, W. (2014)

Studies on the mechanism of electron bifurcation catalyzed by electron transferring flavoprotein (Etf) and butyryl-CoA dehydrogenase (Bcd) of Acidaminococcus fermentans. J. Biol. Chem. 289, 5145–5157.

66 (23) Chowdhury, N. P., Klomann, K., Seubert, A., and Buckel, W. (2016) Reduction of flavodoxin by electron bifurcation and sodium ion-dependent re-oxidation by NAD+ catalysed by ferredoxin:NAD+reductase (Rnf). J. Biol. Chem. 291, 11993–12002.

(24) Bertsch, J., Parthasarathy, A., Buckel, W., and Müller, V. (2013) An electron- bifurcating caffeyl-CoA reductase. J. Biol. Chem. 288, 11304–11311.

(25) Lubner, C. E., Jennings, D. P., Mulder, D. W., Schut, G. J., Zadvornyy, O. A.,

Hoben, J., Tokmina-Lukaszewska, M., Berry, L., Nguyen, D., Lipscomb, G. L., Bothner,

B., Jones, A. K., Miller, A. F., King, P. W., Adams, M. W. W., and Peters, J. W. (2017)

Mechanistic insights into energy conservation by flavin-based electron bifurcation. Nat.

Chem. Biol. 13, 655–659.

(26) Edgren, T. (2006) Electron transport to nitrogenase. Stockholm University,

Stockholm, Sweden.

(27) Wientjens, R. (1993) The involvement of the fixABCX genes and the respiratory chain in the electron transport to nitrogenase in Azotobacter vinelandii. The Department of Chemistry and Biochemistry, Agricultural University, Wageningen, The Netherlands.

(28) Huang, J. J., Heiniger, E. K., McKinlay, J. B., and Harwood, C. S. (2010) Production of hydrogen gas from light and the inorganic electron donor thiosulfate by

Rhodopseudomonas palustris. Appl. Environ. Microbiol. 76, 7717–7722.

(29) Edgren, T., and Nordlund, S. (2004) The fixABCX genes in Rhodospirillum rubrum encode a putative membrane complex participating in electron transfer to nitrogenase. J.

Bacteriol. 186, 2052–2060.

67 (30) Ruvkun, G. B., Sundaresan, V., and Ausubel, F. M. (1982) Directed transposon

Tn5 mutagenesis and complementation analysis of Rhizobium meliloti symbiotic nitrogen fixation genes. Cell 29, 551–559.

(31) Earl, C. D., Ronson, C. W., and Ausubel, F. M. (1987) Genetic and structural analysis of the Rhizobium meliloti fixA, fixB, fixC, and fixX genes. J. Bacteriol. 169,

1127–1136.

(32) Curatti, L., Brown, C. S., Ludden, P. W., and Rubio, L. M. (2005) Genes required for rapid expression of nitrogenase activity in Azotobacter vinelandii. Proc. Natl. Acad.

Sci. U.S.A. 102, 6291–6296.

(33) de Lorenzo, V., Herrero, M., Jakubzik, U., and Timmis, K. N. (1990) Mini-Tn5 transposon derivatives for insertion mutagenesis, promoter probing, and chromosomal insertion of cloned DNA in gram-negative eubacteria. J. Bacteriol. 172, 6568–6572.

(34) Page, W. J., and von Tigerstrom, M. (1979) Optimal conditions for transformation of

Azotobacter vinelandii. J. Bacteriol. 139, 1058–1061.

(35) Jacobson, M. R., Cash, V. L., Weiss, M. C., Laird, N. F., Newton, W. E., and Dean,

D. R. (1989) Biochemical and genetic analysis of the nifUSVWZM cluster from

Azotobacter vinelandii. Molec. Gen. Genet. 219, 49–57.

(36) Toukdarian, A., and Kennedy, C. (1986) Regulation of nitrogen metabolism in

Azotobacter vinelandii: isolation of ntr and glnA genes and construction of ntr mutants.

EMBO J. 5, 399–407.

(37) Sarma, R., Barney, B. M., Hamilton, T. L., Jones, A., Seefeldt, L. C., and Peters, J.

W. (2008) Crystal structure of the L protein of Rhodobacter sphaeroides light-

68 independent protochlorophyllide reductase with MgADP bound: a homologue of the nitrogenase Fe protein. Biochemistry 47, 13004–13015.

(38) Lewis, J. A., and Escalante-Semerena, J. C. (2006) The FAD-dependent tricarballylate dehydrogenase (TcuA) enzyme of Salmonella enterica converts tricarballylate into cis-aconitate. J Bacteriol. 188, 5479–5486.

(39) Sato, K., Nishina, Y., and Shiga, K. (2013) Interaction between NADH and electron- transferring flavoprotein from Megasphaera elsdenii. J. Biochem. 153, 565–572.

(40) Vaudel, M., Burkhart, J. M., Zahedi, R. P., Oveland, E., Berven, F. S., Sickmann, A.,

Martens, L., and Barsnes, H. (2015) PeptideShaker enables reanalysis of MS-derived proteomics data sets. Nat. Biotechnol. 33, 22–24.

(41) Lima, D. B., de Lima, T. B., Balbuena, T. S., Neves-Ferreira, A. G. C., Barbosa, V.

C., Gozzo, F. C., and Carvalho, P. C. (2015) SIM-XL: A powerful and user-friendly tool for peptide cross-linking analysis. J. Proteomics 129, 51–55.

(42) Poudel, S., Tokmina-Lukaszewska, M., Colman, D. R., Refai, M., Schut, G. J., King,

P. W., Maness, P. C., Adams, M. W. W., Peters, J. W., Bothner, B., and Boyd, E. S.

(2016) Unification of [FeFe]- into three structural and functional groups.

Biochim. Biophys. Acta, Gen. Subj. 1860, 1910–1921.

(43) Luo, M. L., Jackson, R. N., Denny, S. R., Tokmina-Lukaszewska, M., Maksimchuk,

K. R., Lin, W., Bothner, B., Wiedenheft, B., and Beisel, C. L. (2016) The CRISPR RNA- guided surveillance complex in Escherichia coli accommodates extended RNA spacers.

Nucleic Acids Res. 44, 7385–7394.

69 (44) Kelley, L. A., Mezulis, S., Yates, C. M., Wass, M. N., and Sternberg, M. J. E.

(2015) The Phyre2 web portal for protein modeling, prediction and analysis. Nat.

Protocols 10, 845–858.

(45) Kozakov, D., Beglov, D., Bohnuud, T., Mottarella, S. E., Xia, B., Hall, D. R., and

Vajda, S. (2013) How good is automated protein docking? Proteins 81, 2159–2166.

(46) Kozakov, D., Brenke, R., Comeau, S. R., and Vajda, S. (2006) PIPER: an FFT-based protein docking program with pairwise potentials. Proteins 65, 392–406.

(47) Comeau, S. R., Gatchell, D. W., Vajda, S., and Camacho, C. J. (2004) ClusPro: an automated docking and discrimination method for the prediction of protein complexes.

Bioinformatics 20, 45–50.

(48) Comeau, S. R., Gatchell, D. W., Vajda, S., and Camacho, C. J. (2004) ClusPro: a fully automated algorithm for protein-protein docking. Nucleic Acids Res. 32, W96-99.

(49) Grosdidier, A., Zoete, V., and Michielin, O. (2011) SwissDock, a protein-small molecule docking web service based on EADock DSS. Nucleic Acids Res. 39, W270-277.

(50) Grosdidier, A., Zoete, V., and Michielin, O. (2011) Fast docking using the

CHARMM force field with EADock DSS. J. Comput. Chem. 32, 2149–2159.

(51) Pettersen, E. F., Goddard, T. D., Huang, C. C., Couch, G. S., Greenblatt, D. M.,

Meng, E. C., and Ferrin, T. E. (2004) UCSF Chimera--a visualization system for exploratory research and analysis. J. Comput. Chem. 25, 1605–1612.

(52) McComb, R. B., Bond, L. W., Burnett, R. W., Keech, R. C., and Bowers, G. N.

(1976) Determination of the molar absorptivity of NADH. Clin. Chem. 22, 141–150.

70 (53) Enescu, M., Lindqvist, L., and Soep, B. (1998) Excited-state dynamics of fully reduced flavins and flavoenzymes studied at subpicosecond time resolution. Photochem.

Photobiol. 68, 150–156.

(54) Wang, Z., and Zhang, J. (2009) Abundant indispensable redundancies in cellular metabolic networks. Genome Biol. Evol. 1, 23–33.

(55) Jeong, H. S., and Jouanneau, Y. (2000) Enhanced nitrogenase activity in strains of

Rhodobacter capsulatus that overexpress the rnf genes. J. Bacteriol. 182, 1208–1214.

(56) Sarkar, A., Köhler, J., Hurek, T., and Reinhold-Hurek, B. (2012) A novel regulatory role of the Rnf complex of Azoarcus sp. strain BH72. Mol. Microbiol. 83, 408–422.

(57) Boyd, E. S., Costas, A. M. G., Hamilton, T. L., Mus, F., and Peters, J. W. (2015)

Evolution of molybdenum nitrogenase during the transition from anaerobic to aerobic metabolism. J. Bacteriol. 197, 1690–1699.

(58) Markowitz, V. M., Chen, I. M. A., Palaniappan, K., Chu, K., Szeto, E., Grechkin, Y.,

Ratner, A., Jacob, B., Huang, J., Williams, P., Huntemann, M., Anderson, I., Mavromatis,

K., Ivanova, N. N., and Kyrpides, N. C. (2012) IMG: the Integrated Microbial Genomes database and comparative analysis system. Nucleic Acids Res. 40, D115-122.

(59) Reyntjens, B., Jollie, D. R., Stephens, P. J., Gao-Sheridan, H. S., and Burgess, B. K.

(1997) Purification and characterization of a fixABCX-linked 2[4Fe-4S] ferredoxin from

Azotobacter vinelandii. J. Biol. Inorg. Chem. 2, 595–602.

(60) Watmough, N. J., and Frerman, F. E. (2010) The electron transfer flavoprotein:

Ubiquinone oxidoreductases. Biochim. Biophys. Acta, Bioenerg. 1797, 1910–1916.

(61) Usselman, R. J., Fielding, A. J., Frerman, F. E., Watmough, N. J., Eaton, G. R., and

Eaton, S. S. (2008) Impact of mutations on the midpoint potential of the [4Fe-4S]+1,+2

71 cluster and on catalytic activity in electron transfer flavoprotein-ubiquinone oxidoreductase (ETF-QO). Biochemistry 47, 92–100.

(62) Jones, D. T., Taylor, W. R., and Thornton, J. M. (1994) A model recognition approach to the prediction of all-helical membrane protein structure and topology.

Biochemistry 33, 3038–3049.

(63) le Maire, M., Champeil, P., and Møller, J. V. (2000) Interaction of membrane proteins and lipids with solubilizing detergents. Biochim. Biophys. Acta, Biomembr.

1508, 86–111.

(64) Sweeney, W. V., and Rabinowitz, J. C. (1980) Proteins Containing 4Fe-4S Clusters:

An Overview. Ann. Rev. Biochem. 49, 139–161.

(65) Zhang, J., Frerman, F. E., and Kim, J.-J. P. (2006) Structure of electron transfer flavoprotein-ubiquinone oxidoreductase and electron transfer to the mitochondrial ubiquinone pool. Proc. Natl. Acad. Sci. U.S.A. 103, 16212–16217.

(66) Demmer, J. K., Rupprecht, F. A., Eisinger, M. L., Ermler, U., and Langer, J. D.

(2016) Ligand binding and conformational dynamics in a flavin-based electron- bifurcating enzyme complex revealed by Hydrogen-Deuterium Exchange Mass

Spectrometry. FEBS Letters 590, 4472–4479.

(67) Dryhurst, G., and Elving, P. J. (1969) Electrochemical oxidation-reduction paths for pyrimidine, cytosine, purine and adenine: Correlation and application. Talanta 16, 855–

874.

(68) Bonetti, C., Stierl, M., Mathes, T., van Stokkum, I. H. M., Mullen, K. M., Cohen-

Stuart, T. A., van Grondelle, R., Hegemann, P., and Kennis, J. T. M. (2009) The role of

72 key amino acids in the photoactivation pathway of the Synechocystis Slr1694 BLUF domain. Biochemistry 48, 11458–11469.

(69) Gray, H. B., and Winkler, J. R. (2015) Hole hopping through tyrosine/tryptophan chains protects proteins from oxidative damage. Proc. Natl. Acad. Sci. U.S.A. 112,

10920–10925.

(70) Rupp, H., Rao, K. K., Hall, D. O., and Cammack, R. (1978) Electron spin relaxation of iron-sulphur proteins studied by microwave power saturation. Biochim. Biophys. Acta,

Protein Struct. 537, 255–269.

(71) Mathews, R., Charlton, S., Sands, R. H., and Palmer, G. (1974) On the nature of the spin coupling between the iron-sulfur clusters in the eight-iron ferredoxins. J. Biol. Chem.

249, 4326–4328.

(72) Swank, R. T., and Burris, R. H. (1969) Restoration by ubiquinone of Azotobacter vinelandii reduced nicotinamide adenine dinucleotide oxidase activity. J. Bacteriol. 98,

311–313.

(73) Knowles, C. J., and Redfearn, E. R. (1969) and ubiquinone patterns during growth of Azotobacter vinelandii. J. Bacteriol. 97, 756–760.

73

APPENDIX A

SUPPORTING INFORMATION

THE ELECTRON BIFURCATING FIXABCX PROTEIN COMPLEX FROM

AZOTOBACTER VINELANDII: GENERATION OF LOW-POTENTIAL

REDUCING EQUIVALENTS FOR NITROGENASE CATALYSIS

74 Table A1. Primers for Δfix mutant generation.

Primer name Sequence Purpose fix101 5’-GGTGGGAATTCCGATCTGCATGGCGCCCG-3’ Construct fix102 5’-CGGCGGATCCGGCATAGGAAGTGGCCAGGGT-3’ Construct fix103 5’-TCAAGGATCCCTTCATCGCCTCACGTTCG-3’ Construct fix104 5’-CGTGAAGCTTCTGCGGGCGCTGTACTTC-3’ Construct fix107 5’-ACATTCGGTGCAATGGTCGG-3’ Verification fix108 5’-CCATCATGCGCCAGGGTG-3’ Verification fix105 5’-GTCACGTCTGGGTCTGCAT -3’ Verification fix106 5’-CGGTCAGTGCCGGCAGG-3’ Verification fix109 5’-GCCACCGACAGCGCCAG-3’ Verification

75 Table A2. Primers for overexpression of FixABCX.

Primer name Sequence Purpose

BB2290 5’- NNNTCTAGACATATGCCCTACAAGATCAACGGTTCC-3’ Construct

BB2291 5’-NNNGGATCCTCAGTTCTTGGCCGAGCGG-3’ Construct

BBP2286 5’-CTGGAAGGCACCAATGAAATGCG-3’ Verification

BBP2287 5’-GCTGTGCACCATCCTCACCCAG-3’ Verification

BBP2288 5’-CGCCATCGAGAAGATCATTCCCGAC-3’ Verification

BBP2289 5’-GTCAACGGCATCCATCGCGAAG-3’ Verification

BB2283 5’-TCAGAGGGCTTGGTAGCGCGGTAC-3’ Verification

76 Table A3 (starts on next page). Protein-protein interactions captured, (A) the inter- cross-links and (B) intra-cross-links within FixABCX complex during cross-linking reaction with BS3. Mapping of intra-subunit cross-links onto the homology models was consistent with the expected spanning distance of 20-25 Å between C-alpha chains. It should be noted that a subset of cross-links mapped to regions spanning 26-30 Å (for Lys-Lys) and higher. This is to be expected and has been explained by protein flexibility using known systems.1

77 A

25.4 30.8 33.9 34.8 35.8 37.2 37.4 39.1 42.1 42.2 42.9 46.5 13.5 20.9 21.5 24.4 25.9 26.1 26.1 26.3 27.1 27.4 27.8 28.0 30.0 30.2 30.9 31.0 31.2 31.2 31.4 31.8 31.9 32.3 33.1 33.6 33.6 33.7 34.0 34.5 34.8 35.4 37.0 38.2 38.5 38.6 39.5 40.1 40.4 40.7 40.9 42.4 42.6 42.6 46.2 46.3 47.8 50.1 49.7 50.5 51.8 51.9 52.0 52.9 53.2 53.7 53.8 54.3 55.7 Cα to Cα Distance to Cα Cα 8 3 1 51 49 31 49 49 75 41 42 13 42 49 49 13 13 85 41 158 227 157 225 225 227 157 225 107 158 246 272 246 158 351 351 158 137 165 249 243 249 246 241 165 158 246 246 238 246 249 246 349 158 115 351 351 157 157 375 375 349 359 349 349 338 341 137 227 341 Protein 2 Protein Cross-linked residue Cross-linked 1 84 31 84 31 71 31 31 71 72 49 72 51 75 72 49 75 72 98 72 75 72 41 42 92 25 51 72 51 85 42 41 98 92 51 51 42 103 166 246 243 251 249 246 349 207 260 139 242 249 102 349 207 139 350 351 227 165 115 260 232 227 227 107 115 225 115 350 233 Protein 1 Protein Cross-linked residue Cross-linked FIXB FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA Protein 2 Protein FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXB FIXB FIXB FIXB FIXB FIXB FIXB FIXB FIXB FIXA FIXB FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXB FIXB Protein 1 Protein TDRQGGEVRADAVILADGVNSR - KEIQVER - TDRQGGEVRADAVILADGVNSR ESITLGVGCMLSDFKQQKISPYELLDR KEIQVER - ESITLGVGCMLSDFKQQKISPYELLDR KEIQVER - ESITLGVGCMLSDFKQQKISPYELLDR KEIQVER - ESITLGVGCMLSDFKQQKISPYELLDR KEIQVER - MHSVVCIKQVPDSAQIR - ESITLGVGCMLSDFKQQK - MHSVVCIKQVPDSAQIR KLADALGVELYGVVICGPGER - KIMEDMPVDLIFTGK - KLADALGVELYGVVICGPGER DYRNEPYTKALTDLVNSYQPEILMLGATTLGR - KEIQVERR - DYRNEPYTKALTDLVNSYQPEILMLGATTLGR DKFGGTVTVVTMGPPMAEAALR - AADKPFSAANLAEYKK - DKFGGTVTVVTMGPPMAEAALR AADKPFSAANLAEYKK - DKFGGTVTVVTMGPPMAEAALR AADKPFSAANLAEYKK - DKFGGTVTVVTMGPPMAEAALR NLASTRPTFGGSLLC(57.02146)TILTQR - IVDVDTAKKEIQVER - NLASTRPTFGGSLLC(57.02146)TILTQR GETPGS(156.0786)KNVQGAILYADAIEKIIPDFR - IVDVDTAKK - GETPGS(156.0786)KNVQGAILYADAIEKIIPDFR IVDVDTAKK - GETPGSK(156.0786)NVQGAILYADAIEKIIPDFR AADKPFSAANLAEYKK(156.0786)KLDDSFVMK - QVPDSAQIR - AADKPFSAANLAEYKK(156.0786)KLDDSFVMK QVPDSAQIR - AADKPFSAANLAEYKKK(156.0786)LDDSFVMK QVPDSAQIR - AADKPFSAANLAEYKKKLDDS(156.0786)FVMK EAGVLTICETTVTDLLIEGGKVVGVRTDRQGGEVR - VFAPKPRSK - EAGVLTICETTVTDLLIEGGKVVGVRTDRQGGEVR VFAPKPRSK - EAGVLTICETTVTDLLIEGGKVVGVRTDRQGGEVR HVWVC(57.02146)IESERGVVHPVSWELLGEGRK - IVDVDTAKK - HVWVC(57.02146)IESERGVVHPVSWELLGEGRK Peptide Sequence (additional mass modifications indicated) modifications mass Peptide Sequence (additional VLQIERGETPGSK KIMEDMPVDLIFTGKQTIDGDTAQVGPGIAK - AFAGSDTLATSYALSAVIR - ESITLGVGC(57.02146)MLSDFKQQK - AFAGSDTLATSYALSAVIR ESITLGVGC(57.02146)MLSDFKQQK - AFAGSDTLATSYALSAVIR ESITLGVGC(57.02146)MLSDFKQQK - AFAGSDTLATSYALSAVIR SKRAELIESHDS(156.0786)DPKNLAEAALAK - ADAVILADGVNSR - SKRAELIESHDS(156.0786)DPKNLAEAALAK VAAGIDTVEMIGLKGSPTVVSK - AADK(156.0786)PFSAANLAEYKK - VAAGIDTVEMIGLKGSPTVVSK AADKPFSAANLAEYKK - VAAGIDTVEMIGLKGSPTVVS(156.0786)K AADKPFSAANLAEYKK - VAAGIDTVEMIGLKGS(156.0786)PTVVSK DKFGGTVTVVTMGPPMAEAALR - THPS(156.0786)IAPLIAGSEMK - DKFGGTVTVVTMGPPMAEAALR FDVIVVGAGMAGNAAAYTLAKGGLK - RAELIESHDS(156.0786)DPK - FDVIVVGAGMAGNAAAYTLAKGGLK AFAGSDTLATSYALS(156.0786)AVIR - NKQFFTTYPDLLS(156.0786)R - AFAGSDTLATSYALS(156.0786)AVIR NKQFFTTYPDLLSR - AFAGSDTLATS(156.0786)YALS(156.0786)AVIR HELKVWDRVAAGIDTVEMIGLKGSPTVVS(156.0786)K - LDDSFVMKDLK - HELKVWDRVAAGIDTVEMIGLKGSPTVVS(156.0786)K AADK(156.0786)PFSAANLAEYKK(156.0786)K - QTIDGDTAQVGPGIAKR - AADK(156.0786)PFSAANLAEYKK(156.0786)K QTIDGDTAQVGPGIAKR - AADK(156.0786)PFSAANLAEYK(156.0786)KK LKDK(156.0786)FGGTVTVVTMGPPMAEAALRK - QGGEVRADAVILADGVNSR - LKDK(156.0786)FGGTVTVVTMGPPMAEAALRK VWDRVAAGIDTVEMIGLKGSPTVVSKVFAPK(156.0786)PR - AADKPFSAANLAEYK - VWDRVAAGIDTVEMIGLKGSPTVVSKVFAPK(156.0786)PR ESITLGVGCMLSDFKQQKIS(156.0786)PYELLDR - QGVPAIINPYDLFALEEALRLKDK - ESITLGVGCMLSDFKQQKIS(156.0786)PYELLDR QGVPAIINPYDLFALEEALRLKDK - ESITLGVGCMLSDFKQQKIS(156.0786)PYELLDR GKEICGEAFQHGADKAYLLQHEILRDYR - VAAGIDTVEMIGLKGSPTVVS(156.0786)K - GKEICGEAFQHGADKAYLLQHEILRDYR VAAGIDTVEMIGLKGSPTVVS(156.0786)K - GKEICGEAFQHGADKAYLLQHEILRDYR VAAGIDTVEMIGLKGSPTVVS(156.0786)K - GKEICGEAFQHGADKAYLLQHEILRDYR VAAGIDTVEMIGLKGSPTVVS(156.0786)K - GKEICGEAFQHGADKAYLLQHEILRDYR DKFGGTVTVVTMGPPMAEAALRKCLS(156.0786)FGADDAILVSDR - KLDDS(156.0786)FVMK - DKFGGTVTVVTMGPPMAEAALRKCLS(156.0786)FGADDAILVSDR VHFDQWFNKK(156.0786)AREAGVLTICETTVTDLLIEGGK - MHSVVCIKQVPDS(156.0786)AQIR - VHFDQWFNKK(156.0786)AREAGVLTICETTVTDLLIEGGK MHSVVCIKQVPDS(156.0786)AQIR - VHFDQWFNKK(156.0786)AREAGVLTICETTVTDLLIEGGK MHSVVCIKQVPDS(156.0786)AQIR - VHFDQWFNKK(156.0786)AREAGVLTICETTVTDLLIEGGK QQKISPYELLDRMKTHPS(156.0786)IAPLIAGS(156.0786)EMK - CLSFGADDAILVS(156.0786)DR - QQKISPYELLDRMKTHPS(156.0786)IAPLIAGS(156.0786)EMK CLS(156.0786)FGADDAILVSDR - QQKISPYELLDRMKTHPS(156.0786)IAPLIAGS(156.0786)EMK DKFGGTVTVVTMGPPMAEAALRKC(57.02146)LSFGADDAILVS(156.0786)DR - KLDDSFVMKDLK - DKFGGTVTVVTMGPPMAEAALRKC(57.02146)LSFGADDAILVS(156.0786)DR KLDDSFVMKDLK - DK(156.0786)FGGTVTVVTMGPPMAEAALRKC(57.02146)LSFGADDAILVSDR KLDDS(156.0786)FVMKDLK - DKFGGTVTVVTMGPPMAEAALRKC(57.02146)LSFGADDAILVSDR GETPGSK(156.0786)NVQGAILYADAIEKIIPDFR - LKDK(156.0786)FGGTVTVVTMGPPMAEAALRK - GETPGSK(156.0786)NVQGAILYADAIEKIIPDFR LKDK(156.0786)FGGTVTVVTMGPPMAEAALRK - GETPGSKNVQGAILYADAIEK(156.0786)IIPDFR GETPGSKNVQGAILYADAIEK(156.0786)IIPDFR - VWDRVAAGIDTVEMIGLKGSPTVVS(156.0786)K - GETPGSKNVQGAILYADAIEK(156.0786)IIPDFR DKFGGTVTVVTMGPPMAEAALRKCLS(156.0786)FGADDAILVSDR - VHFDQWFNKK(156.0786)AR - DKFGGTVTVVTMGPPMAEAALRKCLS(156.0786)FGADDAILVSDR HVWVC(57.02146)IESERGVVHPVSWELLGEGRK - LK(156.0786)DKFGGTVTVVTMGPPMAEAALRK - HVWVC(57.02146)IESERGVVHPVSWELLGEGRK LKDK(156.0786)FGGTVTVVTMGPPMAEAALRK - HVWVC(57.02146)IESERGVVHPVSWELLGEGRK HVWVC(57.02146)IESERGVVHPVSWELLGEGRK - VWDRVAAGIDTVEMIGLKGSPTVVS(156.0786)K - HVWVC(57.02146)IESERGVVHPVSWELLGEGRK HELKVWDRVAAGIDTVEMIGLKGSPTVVS(156.0786)K - ES(156.0786)ITLGVGC(57.02146)MLSDFKQQK - HELKVWDRVAAGIDTVEMIGLKGSPTVVS(156.0786)K ES(156.0786)ITLGVGC(57.02146)MLSDFKQQK - HELKVWDRVAAGIDTVEMIGLKGSPTVVS(156.0786)K ES(156.0786)ITLGVGC(57.02146)MLSDFKQQK - HELKVWDRVAAGIDTVEMIGLKGSPTVVS(156.0786)K KC(57.02146)LSFGADDAILVSDRAFAGSDTLATSYALSAVIR - ESITLGVGCMLSDFKQQKIS(156.0786)PYELLDR - KC(57.02146)LSFGADDAILVSDRAFAGSDTLATSYALSAVIR ESITLGVGCMLSDFKQQKIS(156.0786)PYELLDR - KC(57.02146)LSFGADDAILVSDRAFAGSDTLATSYALSAVIR ESITLGVGCMLSDFKQQKIS(156.0786)PYELLDR - KC(57.02146)LSFGADDAILVSDRAFAGSDTLATSYALSAVIR FGGTVTVVTMGPPMAEAALRKC(57.02146)LSFGADDAILVS(156.0786)DR - GETPGSKNVQGAILYADAIEKIIPDFR - FGGTVTVVTMGPPMAEAALRKC(57.02146)LSFGADDAILVS(156.0786)DR GETPGSKNVQGAILYADAIEKIIPDFR - FGGTVTVVTMGPPMAEAALRKC(57.02146)LSFGADDAILVS(156.0786)DR GETPGSKNVQGAILYADAIEKIIPDFR - FGGTVTVVTMGPPMAEAALRKC(57.02146)LSFGADDAILVS(156.0786)DR FGGTVTVVTMGPPMAEAALRKC(57.02146)LSFGADDAILVSDR - HVWVC(57.02146)IESERGVVHPVSWELLGEGRK - FGGTVTVVTMGPPMAEAALRKC(57.02146)LSFGADDAILVSDR 20 31 15 31 41 65 58 65 24 64 63 36 42 11 75 29 46 11 25 31 25 11 31 42 81 31 81 36 19 18 56 25 11 36 55 12 29 55 65 11 65 36 12 22 12 22 30 12 13 24 10 16 16 32 27 34 19 43 24 41 30 27 41 25 23 29 36 38 16 Protein 2 Protein Peaks Matched 86 44 62 44 35 61 62 62 60 61 79 58 52 58 74 47 43 64 74 49 80 59 41 67 66 68 74 35 53 89 58 65 35 87 58 68 89 55 53 34 58 48 59 48 49 57 57 74 62 54 55 41 51 36 53 53 87 57 77 51 58 49 49 51 33 34 67 101 101 Protein 1 Protein Peaks Matched 8.1 4.4 5.5 4.4 5.1 5.1 8.2 6.2 8.9 9.6 5.3 7.3 5.3 0.1 8.9 2.4 0.8 2.2 2.6 2.4 2.2 8.9 2.2 2.4 2.7 6.3 9.8 1.1 4.3 5.3 7.5 0.8 1.9 9.8 9.8 5.4 11.1 10.5 10.5 13.7 14.9 19.0 11.1 14.9 14.9 11.9 19.0 19.0 13.6 14.9 12.7 17.4 13.6 17.4 12.7 12.7 16.2 16.2 11.1 17.0 11.1 10.9 10.9 10.2 17.0 10.9 17.0 12.5 18.5 ppm 7.62 7.82 7.62 7.73 8.51 9.55 9.52 7.03 7.72 9.05 7.64 8.11 7.13 7.34 7.21 7.31 8.41 7.74 7.21 7.13 9.85 7.10 7.14 7.15 7.25 7.11 8.04 7.03 7.11 9.94 7.31 7.11 7.22 7.45 7.91 7.13 7.31 9.74 9.96 7.91 7.53 7.34 8.12 7.03 7.29 8.42 10.77 12.75 12.13 12.83 12.65 14.36 15.05 10.04 10.66 11.05 14.85 10.05 15.66 12.08 14.65 14.35 14.55 12.13 11.38 11.93 11.27 10.86 10.06 Primary Score Primary 4839.566 5727.943 4185.177 5727.943 6080.245 5727.939 5969.066 5969.066 4186.197 5727.939 6949.570 4034.089 6080.255 4112.079 7025.418 4094.064 4034.089 4112.079 5727.934 6080.245 5727.934 4112.079 4029.078 4839.566 7025.418 5727.934 7025.418 4094.064 5471.799 4839.566 6949.570 4190.183 4112.079 4094.064 6949.570 4094.064 5727.934 6949.570 5969.066 4839.566 5969.066 4094.064 4094.064 4185.174 4094.064 5727.949 4112.079 3340.753 5967.053 5727.934 5471.799 4185.174 4185.174 4094.064 6080.245 4094.064 5471.799 4190.194 4839.566 4190.194 4839.566 6080.245 4190.194 4112.079 4112.079 6080.245 4029.076 6080.252 5412.792 Experimental M+H Experimental A with B with A C with A

78 56.8 57.0 57.2 57.2 58.0 59.4 59.4 60.0 60.4 60.9 61.7 62.8 63.6 64.9 65.3 67.5 13.7 30.2 31.2 34.1 37.4 37.7 40.1 45.1 43.5 44.0 44.8 47.3 49.3 50.8 60.9 61.8 63.3 72.2 72.3 79.5 82.1 18.7 19.5 20.9 21.9 22.6 24.7 28.4 29.2 33.4 34.3 34.7 37.8 38.3 41.5 53.9 34.8 36.8 44.8 53.5 56.0 56.9 58.0 34.0 35.2 38.3 40.1 40.2 41.1 42.2 44.8 45.2 1 3 7 7 3 25 25 29 25 25 36 36 17 72 17 17 36 25 25 25 25 17 25 36 36 338 225 338 232 227 227 341 338 232 338 137 137 233 246 292 292 166 166 249 359 249 246 249 351 166 338 338 272 198 158 196 386 350 232 233 225 227 238 249 166 207 258 246 41 25 29 29 25 42 71 71 25 49 25 71 54 54 54 54 72 54 54 72 25 49 92 49 54 71 25 25 29 41 233 232 260 225 227 225 341 338 227 225 225 267 246 249 249 131 267 158 267 267 285 260 375 198 198 225 227 137 233 238 225 258 349 338 341 338 338 338 FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXB FIXB FIXB FIXB FIXC FIXC FIXC FIXC FIXC FIXC FIXB FIXC FIXC FIXC FIXB FIXB FIXB FIXC FIXC FIXC FIXX FIXX FIXX FIXX FIXX FIXX FIXX FIXX FIXX FIXB FIXX FIXB FIXX FIXX FIXB FIXX FIXX FIXC FIXX FIXX FIXX FIXX FIXX FIXX FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXC FIXX FIXX FIXX FIXX FIXX FIXX FIXX FIXX FIXX FIXX FIXC FIXC FIXX FIXC FIXC FIXC FIXC FIXC FIXC FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXB FIXB FIXB FIXB FIXB FIXB FIXB FIXB FIXB FIXB FIXB FIXB FIXB FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXA FIXB FIXB FIXB FIXB SGEIIEVPFSMIETDIITK - AADKPFSAANLAEYKKK - SGEIIEVPFSMIETDIITK VAAGIDTVEMIGLKGSPTVVSK - AADKPFSAANLAEYKK - VAAGIDTVEMIGLKGSPTVVSK NLASTRPTFGGSLLCTILTQR - QQKIS(156.0786)PYELLDR - NLASTRPTFGGSLLCTILTQR MPEIFHKNKQFFTTYPDLLS(156.0786)R - VMAMPEPDASR - MPEIFHKNKQFFTTYPDLLS(156.0786)R SGEIIEVPFSMIETDIITK - GGLKVLQIERGETPGS(156.0786)K - SGEIIEVPFSMIETDIITK NQENFQLVWDLAKVLGAEVGAS(156.0786)RP - IEEKLFQDR - NQENFQLVWDLAKVLGAEVGAS(156.0786)RP VWDRVAAGIDTVEMIGLKGSPTVVSKVFAPKPR - VDSGRPHIR - VWDRVAAGIDTVEMIGLKGSPTVVSKVFAPKPR VDSGRPHIR - VWDRVAAGIDTVEMIGLKGSPTVVSKVFAPKPR KLADALGVELYGVVICGPGERGK - THPS(156.0786)IAPLIAGSEMK KLADALGVELYGVVICGPGERGK - VAAGIDTVEMIGLKGSPTVVS(156.0786)K - AADKPFSAANLAEYKK - VAAGIDTVEMIGLKGSPTVVS(156.0786)K AADKPFSAANLAEYKK - VAAGIDTVEMIGLKGSPTVVS(156.0786)K AADKPFSAANLAEYKK - VAAGIDTVEMIGLKGSPTVVS(156.0786)K VLGAEVGASRPIVQAGWAELDR - AADKPFS(156.0786)AANLAEYK - VLGAEVGASRPIVQAGWAELDR FDVIVVGAGMAGNAAAYTLAKGGLKVLQIERGETPGSK - GSPTVVSK - FDVIVVGAGMAGNAAAYTLAKGGLKVLQIERGETPGSK GSPTVVSKVFAPKPR - FDVIVVGAGMAGNAAAYTLAKGGLKVLQIER GSPTVVSK - FDVIVVGAGMAGNAAAYTLAKGGLKVLQIERGETPGSK AADKPFSAANLAEYKK VLGAEVGASRPIVQAGWAELDRQVGQSGK - MAERFDVIVVGAGMAGNAAAYTLAKGGLK - MHS(156.0786)VVCIK - MAERFDVIVVGAGMAGNAAAYTLAKGGLK AADK(156.0786)PFSAANLAEYKK(156.0786)K - QTIDGDTAQVGPGIAKR - AADK(156.0786)PFSAANLAEYKK(156.0786)K QTIDGDTAQVGPGIAKR - AADKPFSAANLAEYK(156.0786)K(156.0786)K VLGAEVGASRPIVQAGWAELDR - LDDS(156.0786)FVMKDLK(156.0786)K - VLGAEVGASRPIVQAGWAELDR KLDDS(156.0786)FVMK(156.0786)DLK - VLGAEVGASRPIVQAGWAELDR FDVIVVGAGMAGNAAAYTLAKGGLK - NLASTRPTFGGSLLC(57.02146)TILTQR - FDVIVVGAGMAGNAAAYTLAKGGLK NLASTRPTFGGSLLC(57.02146)TILTQR - FDVIVVGAGMAGNAAAYTLAKGGLK KLADALGVELYGVVICGPGERGKEIC(57.02146)GEAFQHGADK - QFFTTYPDLLSR - KLADALGVELYGVVICGPGERGKEIC(57.02146)GEAFQHGADK QQKISPYELLDRMKTHPSIAPLIAGSEMK - NLASTRPTFGGSLLC(57.02146)TILTQR QQKISPYELLDRMKTHPSIAPLIAGSEMK - VWDRVAAGIDTVEMIGLKGSPTVVSKVFAPK(156.0786)PR - AADKPFSAANLAEYK - VWDRVAAGIDTVEMIGLKGSPTVVSKVFAPK(156.0786)PR IS(156.0786)PYELLDRMKTHPSIAPLIAGSEMK - SGEIIEVPFS(156.0786)MIETDIITK - IS(156.0786)PYELLDRMKTHPSIAPLIAGSEMK EGSNLAMTTGMLAAQTLVELRAADKPFS(156.0786)AANLAEYKK - MSKIEEKLFQDR - EGSNLAMTTGMLAAQTLVELRAADKPFS(156.0786)AANLAEYKK MSKIEEKLFQDR - EGSNLAMTTGMLAAQTLVELRAADKPFS(156.0786)AANLAEYKK YRVDSGRPHIRIKDPDHC(57.02146)TECLEK - VFAPKPRSKRAELIESHDS(156.0786)DPK - YRVDSGRPHIRIKDPDHC(57.02146)TECLEK IICQDSGNLEWEWPRGGFGILFKFG - NLAS(156.0786)TRPTFGGSLLC(57.02146)TILTQR - IICQDSGNLEWEWPRGGFGILFKFG VLQIERGETPGSKNVQGAILYADAIEK - NLAS(156.0786)TRPTFGGSLLC(57.02146)TILTQR - VLQIERGETPGSKNVQGAILYADAIEK VWDRVAAGIDTVEMIGLKGSPTVVS(156.0786)K - VDS(156.0786)GRPHIRIKDPDHCTECLEK - VWDRVAAGIDTVEMIGLKGSPTVVS(156.0786)K VDS(156.0786)GRPHIRIKDPDHCTECLEK - VWDRVAAGIDTVEMIGLKGSPTVVS(156.0786)K VDSGRPHIRIK(156.0786)DPDHCTECLEK - VWDRVAAGIDTVEMIGLKGSPTVVS(156.0786)K ES(156.0786)ITLGVGCMLSDFKQQKIS(156.0786)PYELLDR - TVRPKLYIAAGIS(156.0786)GAIQHR - ES(156.0786)ITLGVGCMLSDFKQQKIS(156.0786)PYELLDR TVRPKLYIAAGIS(156.0786)GAIQHR - ESITLGVGCMLSDFK(156.0786)QQKIS(156.0786)PYELLDR GETPGSKNVQGAILYADAIEK(156.0786)IIPDFR - VWDRVAAGIDTVEMIGLKGSPTVVS(156.0786)K - GETPGSKNVQGAILYADAIEK(156.0786)IIPDFR FDVIVVGAGMAGNAAAYTLAKGGLKVLQIERGETPGS(156.0786)K - YRVDSGRPHIRIKDPDHCTECLEK - FDVIVVGAGMAGNAAAYTLAKGGLKVLQIERGETPGS(156.0786)K EGSNLAMTTGMLAAQTLVELRAADKPFSAANLAEYKK - IKDPDHCTECLEKQC(57.02146)TVCCPAGCYTR - EGSNLAMTTGMLAAQTLVELRAADKPFSAANLAEYKK IKDPDHCTECLEKQC(57.02146)TVCCPAGCYTR - EGSNLAMTTGMLAAQTLVELRAADKPFSAANLAEYKK IKDPDHCTECLEKQC(57.02146)TVCCPAGCYTR - EGSNLAMTTGMLAAQTLVELRAADKPFSAANLAEYKK IKDPDHCTECLEKQC(57.02146)TVCCPAGCYTR - EGSNLAMTTGMLAAQTLVELRAADKPFSAANLAEYKK QCTVCCPAGCYTRETNGKVTLVTDGCLECGTC(57.02146)RIICQDSGNLEWEWPR - GSPTVVSKVFAPKPR - QCTVCCPAGCYTRETNGKVTLVTDGCLECGTC(57.02146)RIICQDSGNLEWEWPR GSPTVVSKVFAPKPR - QCTVCCPAGCYTRETNGKVTLVTDGCLECGTC(57.02146)RIICQDSGNLEWEWPR GSPTVVSKVFAPKPR - QCTVCCPAGCYTRETNGKVTLVTDGCLECGTC(57.02146)RIICQDSGNLEWEWPR GSPTVVSKVFAPKPR - QCTVCCPAGCYTRETNGKVTLVTDGCLECGTC(57.02146)RIICQDSGNLEWEWPR K(156.0786)LADALGVELYGVVICGPGERGKEIC(57.02146)GEAFQHGADK - QQKIS(156.0786)PYELLDR - K(156.0786)LADALGVELYGVVICGPGERGKEIC(57.02146)GEAFQHGADK ETNGKVTLVTDGCLECGTCRIIC(57.02146)QDSGNLEWEWPR - VMAMPEPDASRSGEIIEVPFSMIETDIITK - ETNGKVTLVTDGCLECGTCRIIC(57.02146)QDSGNLEWEWPR MKTHPSIAPLIAGSEMKEYAAHLIPEGGYNAIPQVYGDGWMIAGDAGQFVNGIHR - GS(156.0786)PTVVSK - MKTHPSIAPLIAGSEMKEYAAHLIPEGGYNAIPQVYGDGWMIAGDAGQFVNGIHR HELKVWDRVAAGIDTVEMIGLK(156.0786)GSPTVVS(156.0786)K - FDVIVVGAGMAGNAAAYTLAKGGLK - HELKVWDRVAAGIDTVEMIGLK(156.0786)GSPTVVS(156.0786)K FDVIVVGAGMAGNAAAYTLAKGGLK - HELK(156.0786)VWDRVAAGIDTVEMIGLKGSPTVVS(156.0786)K HELK(156.0786)VWDRVAAGIDTVEMIGLKGSPTVVS(156.0786)K - ESITLGVGC(57.02146)MLSDFKQQK - HELK(156.0786)VWDRVAAGIDTVEMIGLKGSPTVVS(156.0786)K MSKIEEK(156.0786)LFQDR - FGGTVTVVTMGPPMAEAALRKC(57.02146)LS(156.0786)FGADDAILVSDR VLGAEVGASRPIVQAGWAELDRQVGQS(156.0786)GK - ES(156.0786)ITLGVGC(57.02146)MLSDFKQQK - VLGAEVGASRPIVQAGWAELDRQVGQS(156.0786)GK VDS(156.0786)GRPHIRIKDPDHCTECLEKQCTVCCPAGCYTR - NQENFQLVWDLAKVLGAEVGAS(156.0786)RP - VDS(156.0786)GRPHIRIKDPDHCTECLEKQCTVCCPAGCYTR LADALGVELYGVVICGPGERGKEICGEAFQHGADK - ESITLGVGC(57.02146)MLSDFKQQKIS(156.0786)PYELLDR - LADALGVELYGVVICGPGERGKEICGEAFQHGADK VLQIERGETPGSKNVQGAILYADAIEK(156.0786)IIPDFR - IK(156.0786)DPDHCTECLEKQCTVCC(57.02146)PAGCYTR - VLQIERGETPGSKNVQGAILYADAIEK(156.0786)IIPDFR IKDPDHCTECLEK(156.0786)QCTVCCPAGC(57.02146)YTRETNGK - ESITLGVGCMLSDFKQQKIS(156.0786)PYELLDR - IKDPDHCTECLEK(156.0786)QCTVCCPAGC(57.02146)YTRETNGK KCLSFGADDAILVSDRAFAGSDTLATSYALSAVIRK - IKDPDHCTECLEKQCTVC(57.02146)CPAGC(57.02146)YTRETNGK - KCLSFGADDAILVSDRAFAGSDTLATSYALSAVIRK DLAGSVATTLGTGLVADC(57.02146)TELVIDTETRNLASTRPTFGGSLLC(57.02146)TILTQR - QQKIS(156.0786)PYELLDR - DLAGSVATTLGTGLVADC(57.02146)TELVIDTETRNLASTRPTFGGSLLC(57.02146)TILTQR KLADALGVELYGVVICGPGERGKEICGEAFQHGADK - IKDPDHCTECLEKQC(57.02146)TVCCPAGC(57.02146)YTRETNGK - KLADALGVELYGVVICGPGERGKEICGEAFQHGADK IKDPDHC(57.02146)TECLEKQCTVCCPAGC(57.02146)YTRETNGK - KLADALGVELYGVVICGPGERGKEICGEAFQHGADK IKDPDHCTECLEKQC(57.02146)TVCCPAGC(57.02146)YTRETNGK - KLADALGVELYGVVICGPGERGKEICGEAFQHGADK ETNGKVTLVTDGC(57.02146)LECGTCRIIC(57.02146)QDSGNLEWEWPR - VWDRVAAGIDTVEMIGLKGS(156.0786)PTVVSK - ETNGKVTLVTDGC(57.02146)LECGTCRIIC(57.02146)QDSGNLEWEWPR QCTVCCPAGCYTRETNGKVTLVTDGC(57.02146)LECGTCRIIC(57.02146)QDSGNLEWEWPR - KCLSFGADDAILVS(156.0786)DR - QCTVCCPAGCYTRETNGKVTLVTDGC(57.02146)LECGTCRIIC(57.02146)QDSGNLEWEWPR K(156.0786)IMEDMPVDLIFTGK(156.0786)QTIDGDTAQVGPGIAKR - IK(156.0786)DPDHCTECLEKQCTVCCPAGC(57.02146)YTR - K(156.0786)IMEDMPVDLIFTGK(156.0786)QTIDGDTAQVGPGIAKR 7 6 5 4 8 8 20 35 37 14 16 37 29 33 33 34 36 36 44 29 31 29 29 37 62 27 22 13 37 30 19 72 30 31 39 24 64 20 16 35 35 40 40 31 87 60 17 44 17 40 21 34 89 37 88 82 87 15 34 34 15 77 34 47 16 34 34 31 67 36 57 55 71 53 55 71 53 35 33 55 84 42 45 55 59 44 78 50 56 40 58 50 57 52 82 66 49 57 49 64 37 36 98 98 58 62 42 98 98 58 63 66 58 63 39 35 55 42 42 69 70 44 93 69 46 39 46 127 103 121 100 142 100 100 106 105 5.4 6.4 6.4 0.4 8.2 0.9 3.5 2.8 8.9 5.0 8.9 3.3 0.3 4.6 9.3 3.3 1.7 1.7 0.2 0.2 1.7 1.7 3.9 6.1 3.9 0.2 8.6 5.4 6.4 1.4 6.4 6.4 9.6 2.0 1.5 2.0 8.9 18.5 17.8 13.2 16.3 17.4 13.4 13.4 13.4 12.8 12.5 17.9 17.9 16.7 16.3 10.2 16.4 10.0 11.3 14.8 11.5 19.7 11.7 12.3 13.2 14.0 13.5 15.5 18.8 15.5 15.5 15.5 8.83 7.19 9.55 7.24 7.94 8.35 8.96 7.83 8.75 7.01 7.02 9.23 7.19 7.72 8.53 8.93 8.23 7.33 7.03 7.81 8.95 7.02 8.35 7.43 7.05 8.12 7.32 7.24 9.94 7.41 8.13 8.43 9.94 8.54 7.41 7.31 7.11 7.71 7.69 8.53 7.90 8.64 8.10 7.41 7.80 14.24 14.20 12.96 14.17 14.97 13.12 11.48 10.65 12.30 10.42 10.32 10.34 10.72 18.85 10.72 11.92 18.75 19.05 18.94 18.96 19.35 14.86 14.16 5412.792 6080.252 4034.089 6949.570 4094.064 4700.564 4839.612 4190.183 4190.183 4700.564 4190.183 4029.075 4029.076 6080.255 5727.939 6080.252 5471.799 5727.944 5727.944 5727.943 5500.014 6949.570 4185.177 5727.949 7024.433 4034.089 4185.177 4839.566 4185.177 4839.566 4190.194 5471.799 4839.566 4112.079 5412.792 4185.186 4112.079 7078.324 7078.324 5727.934 5727.934 7063.375 7078.324 7429.417 7078.324 7024.429 4700.564 7429.417 4700.564 5727.934 5727.944 5727.943 7429.417 5471.799 7429.417 7429.417 7429.417 4112.079 7024.433 7024.433 5727.944 7025.418 7024.433 6949.570 5727.944 7024.433 7063.375 7024.433 B with C B with X B with X with C A with X with A

79 45.8 47.3 47.8 49.1 52.4 54.8 57.5 61.4 66.4 36 25 41 42 56 17 249 249 338

36 42 25 54 54 54 54 116 275

FIXC FIXX FIXC FIXC FIXX FIXC FIXC FIXC FIXX FIXX FIXC FIXX FIXX FIXC FIXX FIXX FIXX FIXC KAREAGVLTICETTVTDLLIEGGK - IKDPDHCTECLEK - KAREAGVLTICETTVTDLLIEGGK ETNGKVTLVTDGC(57.02146)LECGTCRIICQDSGNLEWEWPR - GETPGSKNVQGAILYADAIEKIIPDFR - ETNGKVTLVTDGC(57.02146)LECGTCRIICQDSGNLEWEWPR GETPGSKNVQGAILYADAIEKIIPDFR - ETNGKVTLVTDGC(57.02146)LECGTCRIICQDSGNLEWEWPR GETPGSKNVQGAILYADAIEKIIPDFR - ETNGKVTLVTDGC(57.02146)LECGTCRIICQDSGNLEWEWPR THPS(156.0786)IAPLIAGSEMKEYAAHLIPEGGYNAIPQVYGDGWMIAGDAGQFVNGIHR - VDSGRPHIR - THPS(156.0786)IAPLIAGSEMKEYAAHLIPEGGYNAIPQVYGDGWMIAGDAGQFVNGIHR ETNGKVTLVTDGCLECGTCRIIC(57.02146)QDSGNLEWEWPR - AADKPFSAANLAEYKKKLDDS(156.0786)FVMK - ETNGKVTLVTDGCLECGTCRIIC(57.02146)QDSGNLEWEWPR VLQIERGETPGSKNVQGAILYADAIEK(156.0786)IIPDFR - IK(156.0786)DPDHCTECLEKQCTVCC(57.02146)PAGCYTR - VLQIERGETPGSKNVQGAILYADAIEK(156.0786)IIPDFR IKDPDHCTECLEKQCTVC(57.02146)CPAGCYTRETNGK - ESITLGVGCMLSDFK(156.0786)QQKIS(156.0786)PYELLDR - IKDPDHCTECLEKQCTVC(57.02146)CPAGCYTRETNGK ESITLGVGCMLSDFKQQKIS(156.0786)PYELLDR - IKDPDHCTECLEK(156.0786)QCTVC(57.02146)CPAGCYTRETNGK 55 34 54 24 19 39 40 36 16 43 40 44 47 52 59 59 59 127 8.9 3.2 1.9 1.9 1.9 17.4 17.4 14.5 15.6 9.94 7.51 9.94 7.22 7.20 9.94 9.63 10.04 14.44 7024.405 7063.375 7024.405 7063.375 4185.174 7024.405 7024.405 7024.405 6949.570

80 B

5.4 9.4 5.1 10.5 13.6 14.1 17.1 18.5 21.1 21.1 21.4 23.2 23.9 24.7 25.4 26.9 27.6 29.0 29.0 29.8 30.5 30.5 30.8 31.1 33.4 33.7 36.2 37.1 38.0 40.0 40.0 40.1 40.5 40.8 41.7 43.1 43.6 44.5 45.6 47.5 48.5 49.7 52.4 54.3 11.9 13.0 14.3 14.8 14.9 15.2 15.7 16.0 16.1 16.8 19.0 19.0 19.5 20.7 23.1 24.1 25.8 26.3 26.6 26.6 27.1 α to Cα Distance Cα to α C 3 8 1 41 41 42 83 83 25 29 75 72 85 75 72 72 233 272 233 115 115 115 359 351 351 243 165 272 139 260 359 373 260 373 249 362 362 243 260 359 233 349 351 243 351 350 115 350 386 233 251 386 249 207 107 232 207 227 107 107 242 207 137 107 157 Peptide 2 Peptide Cross-linkedresidue 8 8 8 3 8 1 25 41 29 42 29 42 25 29 25 42 41 41 42 25 25 25 25 29 29 72 72 85 75 72 75 51 51 51 75 231 260 275 165 165 165 116 246 362 233 249 115 165 231 233 218 231 233 260 231 139 338 260 218 264 225 207 227 207 233 Peptide 1 Peptide Cross-linkedresidue VWDRVAAGIDTVEMIGLKGSPTVVSK KAREAGVLTICETTVTDLLIEGGKVVGVR GETPGSKNVQGAILYADAIEKIIPDFR - MPEIFHKNK - GETPGSKNVQGAILYADAIEKIIPDFR MPEIFHKNK - GETPGSKNVQGAILYADAIEKIIPDFR THPSIAPLIAGSEMKEYAAHLIP - ESITLGVGCMLSDFK THPSIAPLIAGSEMKEYAAHLIP- GETPGSKNVQGAILYADAIEK - MKTHPSIAPLIAGSEMK - GETPGSKNVQGAILYADAIEK MKTHPSIAPLIAGSEMK - GETPGSKNVQGAILYADAIEK MKTHPSIAPLIAGSEMK - GETPGSKNVQGAILYADAIEK VVGVRTDRQGGEVRADAVILADGVNSR - KLDDSFVMK - VVGVRTDRQGGEVRADAVILADGVNSR ESITLGVGCMLSDFKQQKIS(156.0786)PYELLDR - DLKKYR - ESITLGVGCMLSDFKQQKIS(156.0786)PYELLDR DKFGGTVTVVTMGPPMAEAALR - QTIDGDTAQVGPGIAKR - DKFGGTVTVVTMGPPMAEAALR NLAEAALAKLFTQHPNLEQEIAK - AELIESHDS(156.0786)DPK - NLAEAALAKLFTQHPNLEQEIAK TDRQGGEVRADAVILADGVNSR - K(156.0786)LDDSFVMKDLK - TDRQGGEVRADAVILADGVNSR Peptide Sequence (additional mass modifications indicated) modifications mass (additional Sequence Peptide QGGEVRADAVILADGVNSR - MKTHPS(156.0786)IAPLIAGSEMK - QGGEVRADAVILADGVNSR VVGVRTDRQGGEVRADAVILADGVNSR - KLDDS(156.0786)FVMK - VVGVRTDRQGGEVRADAVILADGVNSR ESITLGVGC(57.02146)MLSDFKQQK - LDDSFVMKDLK(156.0786)KYR - ESITLGVGC(57.02146)MLSDFKQQK LDDSFVMKDLKK(156.0786)YR - ESITLGVGC(57.02146)MLSDFKQQK FDVIVVGAGMAGNAAAYTLAKGGLK - ESITLGVGC(57.02146)MLSDFK - FDVIVVGAGMAGNAAAYTLAKGGLK ESITLGVGC(57.02146)MLSDFK - FDVIVVGAGMAGNAAAYTLAKGGLK EAGVLTICETTVTDLLIEGGKVVGVRTDR - ESITLGVGC(57.02146)MLSDFK - EAGVLTICETTVTDLLIEGGKVVGVRTDR ESITLGVGC(57.02146)MLSDFKQQK - AADKPFSAANLAEYKK(156.0786)K - ESITLGVGC(57.02146)MLSDFKQQK AADKPFSAANLAEYKKK - ESITLGVGC(57.02146)MLSDFK(156.0786)QQK GETPGSKNVQGAILYADAIEK - ES(156.0786)ITLGVGC(57.02146)MLSDFK - GETPGSKNVQGAILYADAIEK MHSVVC(57.02146)IKQVPDSAQIRVHPVTNTIMR - GSPTVVS(156.0786)K - MHSVVC(57.02146)IKQVPDSAQIRVHPVTNTIMR FDVIVVGAGMAGNAAAYTLAKGGLK(156.0786)VLQIER - VLQIERGETPGSK - FDVIVVGAGMAGNAAAYTLAKGGLK(156.0786)VLQIER VLQIERGETPGSK - FDVIVVGAGMAGNAAAYTLAK(156.0786)GGLKVLQIER FDVIVVGAGMAGNAAAYTLAKGGLKVLQIERGETPGSK - VWVMDDASYVGTHYR - FDVIVVGAGMAGNAAAYTLAKGGLKVLQIERGETPGSK VWVMDDASYVGTHYR - FDVIVVGAGMAGNAAAYTLAKGGLKVLQIERGETPGSK FDVIVVGAGMAGNAAAYTLAKGGLKVLQIER - VVGVRTDRQGGEVRADAVILADGVNSR - FDVIVVGAGMAGNAAAYTLAKGGLKVLQIER KIMEDMPVDLIFTGK - FGGTVTVVTMGPPMAEAALRKCLS(156.0786)FGADDAILVSDR KIMEDMPVDLIFTGK - FGGTVTVVTMGPPMAEAALRKCLSFGADDAILVS(156.0786)DR FDVIVVGAGMAGNAAAYTLAKGGLKVLQIERGETPGS(156.0786)K - YTIIRVHFDQWFNKK - FDVIVVGAGMAGNAAAYTLAKGGLKVLQIERGETPGS(156.0786)K YTIIRVHFDQWFNKK - FDVIVVGAGMAGNAAAYTLAKGGLK(156.0786)VLQIERGETPGSK YTIIRVHFDQWFNKK - FDVIVVGAGMAGNAAAYTLAKGGLKVLQIERGETPGS(156.0786)K DKFGGTVTVVTMGPPMAEAALRKCLS(156.0786)FGADDAILVSDR - KIMEDMPVDLIFTGK - DKFGGTVTVVTMGPPMAEAALRKCLS(156.0786)FGADDAILVSDR KIMEDMPVDLIFTGK - DKFGGTVTVVTMGPPMAEAALRK(156.0786)CLSFGADDAILVSDR QQK(156.0786)ISPYELLDRMKTHPSIAPLIAGSEMK - FDVIVVGAGMAGNAAAYTLAKGGLK - QQK(156.0786)ISPYELLDRMKTHPSIAPLIAGSEMK ISAGMMGTGFLYTNKESITLGVGCMLSDFKQQK - GGLKVLQIERGETPGSKNVQGAILYADAIEK - ISAGMMGTGFLYTNKESITLGVGCMLSDFKQQK DK(156.0786)FGGTVTVVTMGPPMAEAALRKCLS(156.0786)FGADDAILVSDR - HELKVWDR - DK(156.0786)FGGTVTVVTMGPPMAEAALRKCLS(156.0786)FGADDAILVSDR HELKVWDR - DKFGGTVTVVTMGPPMAEAALRK(156.0786)CLSFGADDAILVS(156.0786)DR HELKVWDR - DKFGGTVTVVTMGPPMAEAALRK(156.0786)CLS(156.0786)FGADDAILVSDR HELKVWDRVAAGIDTVEMIGLKGSPTVVS(156.0786)K - MHSVVC(57.02146)IKQVPDSAQIR - HELKVWDRVAAGIDTVEMIGLKGSPTVVS(156.0786)K MHSVVC(57.02146)IKQVPDSAQIR - HELKVWDRVAAGIDTVEMIGLKGSPTVVS(156.0786)K MHSVVC(57.02146)IKQVPDSAQIR - HELKVWDRVAAGIDTVEMIGLKGSPTVVS(156.0786)K EGSNLAMTTGMLAAQTLVELRAADKPFS(156.0786)AANLAEYK(156.0786)K - ISPYELLDRMK - EGSNLAMTTGMLAAQTLVELRAADKPFS(156.0786)AANLAEYK(156.0786)K MAERFDVIVVGAGMAGNAAAYTLAKGGLK - GETPGSKNVQGAILYADAIEK(156.0786)IIPDFR - MAERFDVIVVGAGMAGNAAAYTLAKGGLK MHSVVC(57.02146)IKQVPDS(156.0786)AQIR - KC(57.02146)LSFGADDAILVS(156.0786)DR - MHSVVC(57.02146)IKQVPDS(156.0786)AQIR KC(57.02146)LS(156.0786)FGADDAILVSDR - MHSVVC(57.02146)IKQVPDS(156.0786)AQIR KC(57.02146)LSFGADDAILVS(156.0786)DR - MHSVVC(57.02146)IKQVPDS(156.0786)AQIR KC(57.02146)LSFGADDAILVS(156.0786)DR - MHSVVC(57.02146)IKQVPDS(156.0786)AQIR KC(57.02146)LSFGADDAILVS(156.0786)DR - MHSVVC(57.02146)IKQVPDS(156.0786)AQIR FDVIVVGAGMAGNAAAYTLAKGGLK(156.0786)VLQIERGETPGS(156.0786)K - QFFTTYPDLLSR - FDVIVVGAGMAGNAAAYTLAKGGLK(156.0786)VLQIERGETPGS(156.0786)K QFFTTYPDLLSR - FDVIVVGAGMAGNAAAYTLAK(156.0786)GGLKVLQIERGETPGS(156.0786)K IS(156.0786)PYELLDRMKTHPSIAPLIAGS(156.0786)EMK - YTIIRVHFDQWFNKK(156.0786)AR - IS(156.0786)PYELLDRMKTHPSIAPLIAGS(156.0786)EMK FGGTVTVVTMGPPMAEAALRKCLSFGADDAILVS(156.0786)DR - CLSFGADDAILVS(156.0786)DR - FGGTVTVVTMGPPMAEAALRKCLSFGADDAILVS(156.0786)DR ISAGMMGTGFLYTNKESITLGVGCMLS(156.0786)DFK(156.0786)QQK - KLDDSFVMKDLK(156.0786)K - ISAGMMGTGFLYTNKESITLGVGCMLS(156.0786)DFK(156.0786)QQK KKLDDSFVMK(156.0786)DLK - ISAGMMGTGFLYTNKESITLGVGCMLS(156.0786)DFK(156.0786)QQK KKLDDSFVMK(156.0786)DLK - ISAGMMGTGFLYTNKESITLGVGCMLS(156.0786)DFK(156.0786)QQK IVDVDTAKK(156.0786)EIQVER - C(57.02146)LSFGADDAILVSDRAFAGSDTLATSYALS(156.0786)AVIR VHFDQWFNKKAREAGVLTIC(57.02146)ETTVTDLLIEGGK - ES(156.0786)ITLGVGC(57.02146)MLSDFK - VHFDQWFNKKAREAGVLTIC(57.02146)ETTVTDLLIEGGK ISPYELLDRMKTHPS(156.0786)IAPLIAGS(156.0786)EMK - MK(156.0786)THPS(156.0786)IAPLIAGSEMK - ISPYELLDRMKTHPS(156.0786)IAPLIAGS(156.0786)EMK VFAPK(156.0786)PRSKRAELIESHDS(156.0786)DPK - VAAGIDTVEMIGLKGSPTVVSKVFAPK(156.0786)PR AADKPFSAANLAEYK(156.0786)KKLDDSFVMKDLKK - ESITLGVGCMLSDFK(156.0786)QQKIS(156.0786)PYELLDR - AADKPFSAANLAEYK(156.0786)KKLDDSFVMKDLKK ISAGMMGTGFLYTNKESITLGVGCMLSDFK(156.0786)QQK - ISAGMMGTGFLYTNKESITLGVGC(57.02146)MLSDFK - ISAGMMGTGFLYTNKESITLGVGCMLSDFK(156.0786)QQK ISAGMMGTGFLYTNKES(156.0786)ITLGVGCMLSDFKQQK - AADKPFSAANLAEYKK(156.0786)KLDDS(156.0786)FVMK - ISAGMMGTGFLYTNKES(156.0786)ITLGVGCMLSDFKQQK 7 9 9 9 50 20 32 15 28 15 31 10 17 32 38 32 37 34 52 36 33 16 36 16 71 34 27 47 19 31 45 17 67 34 18 36 31 17 22 17 36 21 22 51 33 23 39 36 23 30 50 37 39 12 38 52 39 26 26 30 31 30 41 Peptide 2 Peptide Peaks Matched Peaks 47 49 51 57 64 57 59 60 58 40 58 38 85 25 40 38 37 67 37 66 69 36 76 57 50 56 68 44 58 73 68 56 46 43 56 56 74 71 49 77 33 33 47 65 95 31 94 39 63 34 57 94 36 74 72 49 66 38 89 84 105 110 104 111 104 Peptide 1 Peptide Peaks Matched Peaks 0.9 1.4 0.9 0.6 5.1 6.1 5.9 4.4 3.9 4.2 0.8 0.8 4.6 4.2 3.6 9.2 4.0 9.5 2.3 9.5 9.5 2.3 3.2 2.4 5.3 4.0 8.1 7.7 4.0 7.9 5.6 5.6 9.6 4.0 9.0 8.1 1.6 14.0 11.1 16.7 11.0 14.9 14.0 11.2 14.9 14.9 10.4 12.3 13.8 10.4 12.3 15.5 13.3 15.4 13.3 16.4 16.6 15.6 15.6 17.7 15.7 15.6 17.7 17.7 15.6 ppm 9.81 7.03 8.42 7.29 9.32 7.29 9.14 7.12 7.63 7.32 9.70 7.13 2.60 9.31 7.52 7.14 8.43 7.42 8.33 7.12 8.44 8.53 9.86 7.61 9.03 7.61 7.51 8.32 7.52 7.41 7.40 8.31 7.35 7.04 7.15 7.51 3.21 7.76 7.31 7.45 7.03 7.65 9.95 8.03 7.62 7.03 14.39 12.08 14.16 10.06 14.16 10.37 13.36 11.18 12.64 13.44 12.36 10.14 12.64 14.58 14.39 14.78 10.16 12.05 12.64 PrimaryScore 7078.324 5471.803 4186.197 6080.252 6080.255 6080.252 4034.089 4029.075 4185.176 4112.079 6080.255 4112.079 4839.612 4839.566 3165.743 6080.255 4112.079 4094.064 4186.197 4112.079 4186.197 6949.570 4094.064 4186.197 4186.197 4034.089 5727.943 4190.194 4186.197 5727.943 6949.570 4190.194 5727.943 4186.197 5727.943 5727.943 5727.939 5727.943 4839.566 5727.939 5727.939 5969.066 6950.565 5727.934 4185.176 4185.176 4185.174 5471.799 5969.066 2866.561 5500.014 4185.177 5471.799 4185.176 4185.174 5500.014 5500.014 4185.176 5727.934 5727.934 6080.245 5471.799 4112.079 5969.066 5727.939 Experimental M+H Experimental FixC FixA

81

28.7 30.9 34.1 36.4 36.5 37.3 39.0 46.0 52.0 53.4 58.2 60.1 21.6 38.8 3 72 75 71 242 107 238 242 249 241 227 225 249 292 232 225 232 233 225 102 227 233 157 264 264 225 166 166 LDYQLLTYVSRIVDVDTAKK - SKRAELIESHDSDPK - LDYQLLTYVSRIVDVDTAKK NLASTRPTFGGSLLC(57.02146)TILTQR - GKEICGEAFQHGADK - NLASTRPTFGGSLLC(57.02146)TILTQR HELKVWDRVAAGIDTVEMIGLKGSPTVVSK - PMAEAALRKCLS(156.0786)FGADDAILVSDR - HELKVWDRVAAGIDTVEMIGLKGSPTVVSK NLAEAALAKLFTQHPNLEQEIAK(156.0786)RAV - VAAGIDTVEMIGLKGSPTVVS(156.0786)K - NLAEAALAKLFTQHPNLEQEIAK(156.0786)RAV VAAGIDTVEMIGLKGSPTVVS(156.0786)K - NLAEAALAKLFTQHPNLEQEIAK(156.0786)RAV HELKVWDRVAAGIDTVEMIGLKGSPTVVS(156.0786)K - MHSVVC(57.02146)IKQVPDSAQIR - HELKVWDRVAAGIDTVEMIGLKGSPTVVS(156.0786)K VWDRVAAGIDTVEMIGLKGSPTVVSKVFAPKPR - KIMEDMPVDLIFTGKQTIDGDTAQVGPGIAK - VWDRVAAGIDTVEMIGLKGSPTVVSKVFAPKPR NLAS(156.0786)TRPTFGGSLLCTILTQRHRPQMATVRPR - TVRPKLYIAAGIS(156.0786)GAIQHR - NLAS(156.0786)TRPTFGGSLLCTILTQRHRPQMATVRPR VAAGIDTVEMIGLKGSPTVVSKVFAPK(156.0786)PRSK - K(156.0786)CLSFGADDAILVS(156.0786)DR - VAAGIDTVEMIGLKGSPTVVSKVFAPK(156.0786)PRSK VAAGIDTVEMIGLKGSPTVVSKVFAPK(156.0786)PR - VFAPK(156.0786)PRSKRAELIESHDS(156.0786)DPK - VAAGIDTVEMIGLKGSPTVVSKVFAPK(156.0786)PR VFAPKPRSKRAELIES(156.0786)HDS(156.0786)DPK - VAAGIDTVEMIGLKGSPTVVSKVFAPK(156.0786)PR VFAPKPRS(156.0786)KRAELIESHDS(156.0786)DPK - VAAGIDTVEMIGLKGSPTVVSKVFAPK(156.0786)PR VFAPKPRSKRAELIESHDS(156.0786)DPK - VAAGIDTVEMIGLKGSPTVVSK(156.0786)VFAPK(156.0786)PR C(57.02146)LSFGADDAILVSDRAFAGS(156.0786)DTLATSYALSAVIR - GSPTVVSK(156.0786)VFAPKPRSK(156.0786)R - C(57.02146)LSFGADDAILVSDRAFAGS(156.0786)DTLATSYALSAVIR 30 50 43 66 47 31 31 30 34 74 74 33 32 27 48 95 48 76 92 59 46 49 49 73 73 43 58 62 9.6 9.6 9.6 5.6 2.1 2.1 9.6 17.7 12.0 17.6 10.9 10.0 13.3 14.3 7.93 9.21 9.14 7.82 8.03 8.44 7.72 9.13 9.04 14.69 14.33 14.07 14.88 14.88 6080.245 5500.014 6080.255 6949.570 5500.014 6080.255 6080.245 6080.245 4190.183 5500.014 5500.014 6080.245 6080.255 4034.089 FixB

82 Table A4. Overview of the EPR signals observed for FixABCX from A. vinelandii.

Signal Name g-values Topt (K) Samples Percent Possible present contribution source

Fast relaxing, * 5 K As-purified * Unknown broad NADH *

Rhombic 2.072, 1.940, 15 K NADH 86 [4Fe-4S]1+ 2.07 1.895 Na-dithionite 87

Axial 2.04 2.041, 1.944, 5 – 10 K NADH 9 [4Fe-4S]1+ 1.944 Na-dithionite 12

Isotropic 2.005 > 50 K As-purified * Flavin 2.005 NADH 1 radical Na-dithionite 1

Axial 2.03 2.030, 2.00, 10 K NADH 4 Interacting 2.00 flavin radical and [4Fe-4S]1+

*Not determined

83 Table A5. Calculated Gibbs free energies of three possible reactions catalyzed by the Fix complex and three possible non-enzymatic reactions.

° Reaction ∆Grxn (kJ/mol) Sq Hq + -36.7 2 NADH + 2 Fld + CoQ = 2 Fld + CoQH2 + 2 NAD Sq Hq + + 118 NADH + CoQH2 + 4 Fld = 4 Fld + CoQ + NAD + 3 H Hq + Sq + -154 NADH + 2 Fld + 2 CoQ + 3 H = 2 CoQH2 + 2 Fld + NAD Sq + Ox -36.7 2 Fld + CoQ + 2 H = 2 Fld + CoQH2 + + -63.7 NADH + CoQ + H = CoQH2 + NAD Ox Sq + + NADH + 2 Fld = 2 Fld + NAD + H -27.0

Sq – Hq Ox – Sq – Fld + e = Fld , E°= -460 mV; Fld + e = Fld , E°= -180 mV; CoQ + 2 e = CoQH2, E°= + 10 mV; NAD+ + 2 e– = NADH, E°= -320 mV

84 Table A6. Fitting of the ASQ kinetic curves for Rh. rubrum FixAB and A. vinelandii FixABCX reveal a biphasic decay, with one component having a lifetime of tens of picoseconds and a slightly longer lifetime of one thousand picoseconds. Fitting was performed in Igor Pro using a double exponential fit function.

Sample ASQ τ1 ASQ τ2

R. rubrum FixAB 14 ps 970 ps

A. vinelandii FixABCX 14 ps 1140 ps

85

Figure A1. Scheme for ∆fix mutant generation.

86 fixfd% fixA% fixB% fixC% fixX%ORF6%

nifHp% nifH% nifD% nifK%

fixfd% fixA% fixB% fixC% fixX%ORF6% SmR% nifHp% nifK%

BB2290% 2286% 2287% 2288% 2289% BB2291% % % % % % %

2283% % Figure A2. Scheme for overexpression of fix genes under the nifH promoter in A. vinelandii.

87

Figure A3. Protein identification within FixABCX complex purified from A. vinelandii. The identity of FixA (panel A), FixB (panel B), FixC (panel C) and FixX (panel D) subunits was confirmed based on 54, 22, 59 and 9 unique tryptic peptides that correspond to 79%, 63%, 85% and 73% sequence coverage (top row). Peptides and proteins are represented by blue and red spheres, respectively. Identified peptides were rendered in blue on primary sequence (middle row) and homology models of corresponding subunits (bottom row). Homology models were generated based on following protein templates: FixA, 4L2I; FixB, 3CLR, FixC, 4K2X; FixX, 2GMH.

88

Figure A4. Size distribution of protein complexes as revealed by dynamic light scattering, comparing buffer containing DDM (negative control) and the purified FixABCX complex. (A) Buffer containing 0.02% DDM (w/v) with a determined hydrodynamic radius of 4.32 nm. (B) 1.0 mg/mL of FixABCX complex with a determined hydrodynamic radius of 4.69 nm accounted for 99.4% of the total population. Triplicate measurements yielded an average hydrodynamic radius of 5.1 ± 0.3 nm.

89 ProQ2&quality&assessment& Clashes& Ramachandran&analysis& Alignment&confidence& Disorder&

FixA&

FixB&

FixC&

FixX&

Figure A5. Evaluation of subunit homology models. ProQ2 quality assessment: a support vector machine is used to predict local and global quality of the model;2 Clashes: indicates which residues lie too close to one another due to incorrect side chain and/or backbone placement; Ramachandran analysis: provides information on residues position (favorable/allowed/disallowed) in the model and potential problems with the backbone phi/psi angles; Alignment confidence: reliability of the pair-wise query-template alignment as reported by HHsearch. The confidence values are obtained from the posterior probabilities calculated in the forward- backward algorithm; Disorder: identifies dynamically flexible regions. This prediction has been made by the knowledge-based Disopred method (PSIPRED server).3 Overall, models for each subunit were obtained with the highest template confidence of 100% (with the exception of 99.8% for FixX). However, it was impossible to attain complete sequence coverage, primarily due to insufficient structural information in the database. For example, sequence coverage for FixA, FixB, FixC and FixX were 92%, 88%, 82% and 96%, respectively.

90

FixA% FixB% FixC% FixX%

FixA% FixB% FixC% FixX%

Figure A6. Predicted structural features in Fix subunits. Conservation: sequence analysis based on Jensen-Shannon divergence;4 Largest pocket: the pockets are detected by the fpocket2 program and highlighted in red.5 This provides a tool for tracking small molecule binding sites, pockets for molecular docking or detection of subpockets of conformational ensembles. The highest degree of conservation (from green to red color) is associated with cofactor binding sites. A helix (Ala338- Leu353) on the C-terminal domain of FixB appears to be critical for A and B subunit interactions. Further, large pockets are detected around FixA, FixB and FixX cofactor binding sites. In addition to flavin binding sites, two large cavities are present between the N-terminal and C-terminal domain of FixB (important for interactions with FixA) and near the predicted transmembrane region (Met299-Leu325) on FixC.

91 FixB%

Figure A7. Protein interface. PI-site interface residues6 and ProtinDB interface residues provide information on prediction of protein-protein interaction sites based on information extracted from protein complexes stored in the . Interface residues are colored in red.

92

Figure A8. Structural model of FixABCX from A. vinelandii. Fix subunit homology models were built using Phyre2 and had the closest match to following templates from the Protein Data Bank (PDB): FixA, 4L2I; FixB, 3CLR, FixC, 4K2X; FixX, 2GMH. Chemical cross-linking data was used to produce distance constraints for generating structural models. (A) ClusPro2 docking model of FixAB subunits (visualized in Chimera). FixA in blue, FixB in tan. (B) ClusPro2 docking model of FixCX subunits (visualized in Chimera). FixC in green, FixX in red. (C) FixABCX complex model generated by docking (ClusPro2) four homology models (Phyre2). Restraints derived from the chemical cross-linking experiment. The final model was evaluated and satisfied two requirements: 1) Most of the cross-links could be explained by the model and 2) cofactor distances of <14 Å that enable electron transfer. (D) Cofactors and protein residues that could function in electron transfer. (E) Distances between cofactors taking into account the presence of conserved Trp and Tyr side chains that may convey electrons between cofactors. FixA-Tyr37 is conserved as Tyr in 44 of the 53 FixAs of which we are aware, FixB-Trp275 is conserved as Trp in 53 out of 53 FixBs of which we are aware, and FixC-Tyr277 is Tyr in 8/8 sequences examined.

93

Figure A9. EPR temperature profiles of FixABCX from A. vinelandii reduced with either Na-dithionite or NADH. (A) FixABCX (100 µM) reduced with Na-dithionite (10 mM). (B) FixABCX (100 µM) reduced with NADH (1 mM). The spectra were recorded at 5, 10, 15, 20, 30, 40, 50, 60 and 80 K. Simulations of the overall spectra at 10 K are shown in the lighter shade of blue. The individual signal components and g-values comprising the overall simulated spectra at 10 K are shown at the top (blue, axial 2.04; red, rhombic 2.07 magenta, isotropic 2.0; green, axial 2.03) and varied slightly for each treatment (Table S4). The overall signal intensities for NADH (0.40 spins/mol protein) and Na- dithionite treatments (0.45 spins/mol protein) were similar at 10 K, indicating partial reduction of the protein complex. Microwave frequency, 9.38 GHz; microwave power, 1 mW; modulation frequency, 100 kHz; modulation amplitude, 10.0 G.

94

Figure A10. Comparison of EPR spectra of reduced FixABCX from A. vinelandii (Av) and the reduced, individual subunit FixX from Ro. castenholzii (Rc). FixABCX from A. vinelandii (100 μM) was reduced 10 mM Na-dithionite. FixX from Ro. castenholzii (200 μM) was reduced with 5 mM Na-dithionite. Signal intensities were normalized for sample concentration. Experimental (exp) spectra are colored black and simulated (sim) spectra are colored gray. The individual signal components and g- values comprising the overall simulated spectra are shown at the top (blue, axial 2.04; red, rhombic 2.07 magenta, isotropic 2.0; green, axial 2.03). Microwave frequency, 9.38 GHz; microwave power, 1 mW; modulation frequency, 100 kHz; modulation amplitude, 10.0 G; sample temperature, 10 K.

95 REFERENCES

(1) Merkley, E. D., Rysavy, S., Kahraman, A., Hafen, R. P., Daggett, V., and Adkins, J.

N. (2014) Distance restraints from crosslinking mass spectrometry: mining a molecular dynamics simulation database to evaluate lysine-lysine distances. Protein Sci. 23, 747–

759.

(2) Ray, A., Lindahl, E., and Wallner, B. (2012) Improved model quality assessment using ProQ2. BMC Bioinformatics 13, 224.

(3) Buchan, D. W. A., Minneci, F., Nugent, T. C. O., Bryson, K., and Jones, D. T. (2013)

Scalable web services for the PSIPRED Protein Analysis Workbench. Nucleic Acids Res.

41, W349–W357.

(4) Capra, J. A., and Singh, M. (2007) Predicting functionally important residues from sequence conservation. Bioinformatics 23, 1875–1882.

(5) Schmidtke, P., Bidon-Chanal, A., Luque, F. J., and Barril, X. (2011) MDpocket: open-source cavity detection and characterization on molecular dynamics trajectories.

Bioinformatics 27, 3276–3285.

(6) Higurashi, M., Ishida, T., and Kinoshita, K. (2009) PiSite: a database of protein interaction sites using multiple binding states in the PDB. Nucleic Acids Res. 37, D360–

D364.

96 CHAPTER 3

EVIDENCE THAT THE Pi RELEASE EVENT IS THE RATE LIMITING STEP IN

THE NITROGENASE CATALYTIC CYCLE

ABSTRACT

Nitrogenase reduction of dinitrogen (N2) to ammonia (NH3) involves a sequence

of events that occur upon the transient association of the reduced Fe protein containing

two ATP with the MoFe protein that includes electron transfer, ATP hydrolysis, Pi

release, and dissociation of the oxidized, ADP containing Fe protein from the reduced

MoFe protein. Numerous kinetic studies using the non-physiological electron donor

dithionite have suggested that the rate limiting step in this reaction cycle is the

dissociation of the Fe protein from the MoFe protein. Here, we have established the rate

constants for each of the key steps in the catalytic cycle using the physiological reductant

flavodoxin protein in its hydroquinone state. The findings indicate that with this

reductant, the rate limiting step in the reaction cycle is not protein-protein dissociation or

reduction of the oxidized Fe protein, but rather events associated with the Pi release step.

These results allow a revision of the rate limiting step in the nitrogenase Fe protein cycle.

______*Chapter adapted from Zhi-Yong Yang, Rhesa Ledbetter, Sudipta Shaw, Natasha Pence, Monika Tokmina-Lukaszewska, Brian Eilers, Qingjuan Guo, Nilisha Pokhrel, Valerie L. Cash, Dennis R. Dean, Edwin Antony, Brian Bothner, John W. Peters, and Lance C. Seefeldt (2016) Biochemistry, 55:3625-3635. Copyright 2016, American Chemical Society. Reprinted with permission. *Modified version focuses on contributions from Rhesa Ledbetter. The full study can be found in Appendices B and C.

97 INTRODUCTION

Nitrogenase is the catalyst responsible for biological nitrogen fixation, the reduction of

1–3 dinitrogen (N2) to ammonia (NH3). The molybdenum-dependent nitrogenase consists of two catalytic components called the MoFe protein and the Fe protein.4 The Fe protein is a dimer of two identical subunits connected by a single [4Fe-4S] cluster and is responsible for transferring a single electron to the MoFe protein.5,6 The MoFe protein is a heterotetramer, composed of two symmetric αβ units. Each αβ unit contains two unique metal clusters, the electron carrier P-cluster [8Fe-7S] and the active site FeMo-cofactor

[7Fe-9S-Mo-C-homocitrate].7–11 An Fe protein, with two bound MgATP molecules, binds transiently to each αβ unit of the MoFe protein during the electron transfer (ET) event (Figure 3-1).6,12,13 ET is initiated upon association of the Fe protein to the MoFe protein, followed by a conformationally gated one-electron transfer from the P-cluster to

FeMo-cofactor.14,15 This is followed by the one-electron transfer from the reduced Fe protein (Fered) [4Fe-4S] cluster to the oxidized P-cluster (P1+), in what has been called a

“deficit spending” ET process.14 Following ET, the two ATP molecules are hydrolyzed to two ADP and two Pi molecules. Next, the two Pi are released from the complex, followed by the dissociation of the oxidized Fe protein (Feox) with two bound MgADP

ox 16 ox (Fe (ADP)2) from the MoFe protein. The Fe (ADP)2 is then reduced by a reductant and the two ADP molecules are replaced by two ATP molecules.17,18 This cycle, often referred to as the Fe protein cycle, must be repeated a sufficient number of times to accumulate the electrons necessary for substrate reduction in the MoFe protein.1–3,18–20

98

Figure 3-1. Overview of nitrogenase catalysis with a focus on the Fe protein cycle showing dithionite (DT) and flavodoxin (FldHQ) as a non-physiological and physiological reductant, respectively.

Over the past five decades, most of the in vitro mechanistic studies of nitrogenase have been conducted with the non-physiological reductant dithionite (DT), largely because of ease of use.1,3,18 The kinetic parameters for each step in the Fe protein cycle using DT as reductant have been deduced and are summarized in the scheme shown in

Figure 3-2.16–18,21,22 It was concluded from a comparison of the rate constants that the

ox overall rate limiting step in the Fe protein cycle is the dissociation of the Fe (ADP)2 from the MoFe protein with a rate constant of ~6 s-1.16,17 A problem that is often

2- overlooked in such studies is the reversible dissociation of DT (S2O4 ) to generate the

•- actual reductant, the radical anion SO2 , with a Kd of ~1.5 nM and a slow rate constant of

~2 s-1.17 This leads to slow reduction of Feox, with rate constants for this reduction near

ox the rate constant that is reported for dissociation of Fe (ADP)2 from MoFe protein.16,17,22,23

99 [I]

[e-] 2ATP 2ADP ox red red [Fe (ADP)2] [Fe (ADP) ] [Fe (ATP) ] k = 6 s-1 2 fast 2 red red,DT MoFe MoFe MoFe MoFe

4 1

5 -

M

.

M s

4 1

 -

-

-

[

1

1

6 s

1

I

s

s

I d

0

1

= 5

]

-

- e

1

1

7

0

1

s

x s

6

i

a d

A

l

k e

c

t R

i

red ox red v

[MoFe ;Fe (ADP)2] [MoFe;Fe (ATP)2] a ]

t

e

V I

d [ 2[Pi] -1 k = 140-200 s-1 k Pi = 16-22 s ET

-1 red ox k ATP = 50-70 s red ox [MoFe ;Fe (ADP,Pi)2] [MoFe ;Fe (ATP)2]

[III]

Figure 3-2. Fe protein cycle with first order and pseudo-first order kinetic rate constants for each step.

The physiological reductants of Fe protein are known to be the electron carrier proteins flavodoxin (Fld) and/or ferredoxin (Fd).24–30 Because nitrogenase is readily reduced by artificial electrons donors, such as DT, a limited number of kinetic studies using physiological reductants have been conducted.24,25,27,30–35 The diazotroph

Azotobacter vinelandii contains several Flds and Fds. NifF (Fld) has been implicated as an electron donor to nitrogenase26,29,30,36 and has been shown to transfer electrons to Fe protein in vitro.24,25,27,33,34 Further, disruption in the nifF gene results in a 30% decrease in whole-cell acetylene reduction activity. This previous research demonstrates that Fld is a major, but not the sole, electron donor to nitrogenase.26,29 Fld has three different redox states designated as the oxidized quinone (FldQ), one-electron reduced semiquinone

(FldSQ), and two-electron reduced hydroquinone (FldHQ).24,37,38 The pH- and temperature-

100 dependent midpoint redox potentials (Em) for the two redox couples are estimated as -

180 mV (vs the NHE) for the FldQ/FldSQ couple, and -480 mV for the FldSQ/FldHQ couple

(pH 8.5 and 22oC).38 The FldSQ/FldHQ redox couple has sufficient driving force for reduction of the oxidized Fe protein with a midpoint potential of -290 mV without nucleotide bound and -440 mV with ADP bound.25,27,33,34,39 Earlier kinetic studies using

Fld as reductant have yielded results that contradict some aspects of the Fe protein cycle deduced with DT: (i) the one-electron reduction of Fe protein by FldHQ is much faster

ox than that by DT, and this reduction event might happen when Fe (ADP)2 is still bound to the MoFe protein25 and (ii) in the Fe protein cycle, the dissociation step might be faster than what has been widely accepted (~6 s-1).25,31–33,40

In the present work, kinetic studies were performed using both DT and Fld (NifF from A. vinelandii) as reductant to determine the key kinetic parameters in the Fe protein cycle. This work reveals that FldHQ cannot reduce the Fe protein while it is bound to

MoFe protein and the rate limiting step in the Fe protein cycle is not the dissociation of the Fe protein from the MoFe protein, but rather events associated with Pi release.

MATERIALS AND METHODS

General procedures

All chemicals, unless otherwise noted, were obtained from Sigma-Aldrich (St.

Louis, MO) and used without further purification. Hydrogen, acetylene, ethylene, argon, and dinitrogen gases were purchased from Air Liquide America Specialty Gases LLC

(Plumsteadville, PA). The argon and dinitrogen gases were passed through an activated copper-catalyst to remove dioxygen contamination prior to use. A.vinelandii strains

DJ995 (wild type MoFe protein) and DJ884 (wild type Fe protein) were grown, and

101 nitrogenase proteins expressed and purified as previously described.41 Both proteins were greater than 95% pure as confirmed by SDS-PAGE analysis using Coomassie blue staining and fully active (see results and discussion). Proteins and buffers were handled anaerobically in septum-sealed serum vials under an inert atmosphere (argon or dinitrogen), on a Schlenk vacuum line, or anaerobic glove box (Teledyne Analytical

Instruments, MO-10-M, Hudson, NH). The transfer of gases and liquids were done with gastight syringes.

Strain construction and expression, Escherichia coli growth, and NifF purification

The nifF gene from A. vinelandii was PCR amplified and cloned into the NdeI-

BamHI sites of a T7-7 plasmid containing an ampicillin resistance gene for selection. Fld was overexpressed in Escherichia coli (E. coli) BL21 DE3 cells. The cells were grown to an optical density (600 nm) of 0.6-0.8 at 37 oC before the protein expression was induced at 32 oC by adding isopropyl β-D-1-thiogalactopyranoside (IPTG) and flavin mononucleotide (FMN) to a final concentration of 1 mM and 10 mg/L, respectively.

Following a 10 hour induction, cells were harvested.

All steps during the purification of Fld were conducted anaerobically under an argon atmosphere. 100 grams of cell paste were resuspended in 50 mM Tris, pH 8 with 2 mM dithiotheritol (DTT) at a biomass to buffer ratio of 1:5 (w/v). The resuspended cells were lysed in a french pressure cell (SLM Aminco FA-078, Rochester, NY) at 200 MPa.

The cell lysate was centrifuged (Sorvall Lynx 4000, ThermoScientific, Waltham, MA) at

48,000 x g for 30 minutes at 4°C. Following centrifugation, 0.5% (w/v) streptomycin sulfate was added to the supernatant to precipitate nucleic acids. The precipitation was removed via a second centrifugation as described above. The supernatant was loaded

102 onto a 150 ml Q-sepharose column, which was first washed with two column volumes of Buffer B (50 mM Tris, 1 M NaCl, 1 mM DTT, pH 8) and then equilibrated with two column volumes of Buffer A (50 mM Tris, 1 mM DTT, pH 8). A 15-75% salt gradient was run over 5 column volumes. Fractions containing Fld (blue color) were combined and diluted with Buffer A to a final NaCl concentration of less than 100 mM.

FMN was added to a final concentration of 2 mM to increase the percentage of holo-Fld in the presence of 2 mM DT, which slowly changed the color of the protein solution from blue to yellow. The reconstitution was carried out for at least 2 hours at room temperature. The protein was then loaded onto a Q-sepharose column (~30 mL) pre- reduced and equilibrated with Buffer B and Buffer A containing 1 mM DT. After loading, the column was washed with 1 column volume of Buffer A and eluted with

100% Buffer B as a concentrated fraction for loading onto Sephacryl-200 column (~600 ml) equilibrated with an 100 mM HEPES (pH 7.8) with 150 mM NaCl, 0.5 mM DTT, and 1 mM DT. The Fld was followed as a yellow-green band. The protein was concentrated using an Amicon (EMD Millipore, Billerica, MA) concentrator with a

10,000 kDa cutoff membrane and stored in liquid nitrogen. The purity of the Fld was greater than 95% based on SDS-PAGE method described above. The FMN content was determined to be ~70% of the total protein. This was determined by measuring the absorbance of FldQ at 452 nm and using an extinction coefficient of 11.3 mM-1 cm-1.37

The Fld concentration used in this work refers to the concentration of the holo-Fld with

FMN bound.

103 Stopped-flow (SF) spectrophotometry and reduction of Fe protein by DT and FldHQ

SF spectrophotometry was conducted using an AutoSF-120 stopped-flow instrument equipped with a data acquisition system (KinTek Corp., Snow Shoe, PA). The change in absorbance was monitored at 426 nm over time. This wavelength detects a decrease in absorbance as the Fe protein [4Fe-4S] cluster becomes reduced. 426 nm rather than 430 nm was chosen since it is the isosbestic point of Fld. Reactions were conducted at 4°C, unless otherwise stated, with a mixing ratio of 1:1. All samples were prepared in 100 mM MOPS buffer (pH 7.3). Fe protein was oxidized in an anaerobic glove box using ~3-fold molar excess of phenazine methosulfate (PMS). The protein-

PMS mixture was allowed to incubate for 15 minutes prior to separation with Sephadex

G-25 equilibrated with 100 mM MOPS, pH 7.3 with 150 mM NaCl. The oxidized Fe protein was contained in one drive syringe with or without nucleotide. The other drive syringe contained the electron donor, DT or FldHQ-DT mixture, with or without nucleotide. All reagents were used at the following final concentrations and kept under an argon atmosphere: 40 μM oxidized Fe protein, 10 mM DT, 300 μM Fld, 10 mM ATP or

ADP, and 10 mM MgCl2 (only used in the presence of nucleotides). For the reduction of

Feox by DT, the data were fit to a single exponential equation to obtain the pseudo-first order rate constants. Due to the presence of both FldHQ and DT in the experiments, the pseudo-first order rate constants for reduction of Feox by FldHQ were obtained by fitting the data to a double exponential equation.

ox Reduction of Fe (ADP)2 by different reductants in the presence of MoFe protein

The Fe-MoFe dissociation rate constant was determined using SF as described above except that the temperature was 25°C. Fe protein was oxidized (see details above)

104 and MoFe protein was stripped of DT in an anaerobic glove box. DT was removed from MoFe protein using a DOWEX-Sephadex G-25 column equilibrated with 100 mM

MOPS, pH 7.3 containing 150 mM NaCl. One drive syringe contained 80 μM Feox, 80

μM MoFe protein, and 9 mM MgADP. The other syringe contained the reductant mixture: 1) 20 mM DT, 2) 100 μM methyl viologen (MV)-20 mM DT mixture, or 3) 400

μM FldHQ-20 mM DT mixture. All reductants contained 9 mM MgADP. DT data were fit to a single exponential decay equation, and MV-DT and FldHQ-DT data were fit to a double exponential decay equation.

Primary electron transfer

Primary electron transfer from the Fe protein to the MoFe protein in the presence of DT and Fld was measured at 25°C using SF spectrophotometry as described above. All mixtures were prepared in 100 mM MOPS, pH 7.3 and were kept under an argon atmosphere. One syringe contained 80 µM Fe protein, 20 µM MoFe protein, 1 mM DT with or without 200 µM FldHQ. The other syringe was loaded with 1 mM DT and 20 mM

red MgATP. As turnover occurred in the [Fe (MgATP)2-MoFe] complex, the oxidation of the [4Fe-4S] cluster of the Fe protein was monitored by an increase in absorbance at 426 nm. Data were fit to a single exponential curve.

Real-time measurement of inorganic phosphate (Pi) release

Pi release was determined in a stopped-flow (SF) fluorometer (Auto SF-120,

KinTek Corp.) using a coumarin (N-[2-(1-maleimidyl)ethyl]-7-(diethylamino) coumarin-

3-carboxamide) labeled phosphate binding protein (MDCC-PBP) assay.42 Pi was quantified from binding MDCC-PBP that was monitored by an increase in fluorescence

(λ-excitation = 430 nm, λ-emission >450 nm), using a standard curve generated with

105 42 KH2PO4 as described. The experiments were carried out at 25°C in 25 mM HEPES, pH 7.4 containing 1 mM DT. The SF syringes and flow lines were treated with a Pi mop

(25 mM HEPES buffer at pH 7.4 with 320 µM 7-methylguanine (7-meG), and 0.12 U/mL purine nucleoside phosphorylase (PNPase)) before each experiment for 45 minutes to remove contaminating Pi and then rinsed with DT reduced buffer. The same concentration of the Pi mop system was also added to the reaction mixtures. 1 µM MoFe and 16 µM Fe were rapidly mixed with a solution of 25 µM MDCC-PBP, 7.5 mM

MgCl2 and 6 mM ATP and the change in fluorescence was monitored over time. 100 µM

FldHQ was added to the MoFe and Fe protein mixture to monitor the effect of FldHQ on Pi release.

RESULTS AND DISCUSSION

HQ Establishing kcat with DT or Fld

In the two-component catalytic system of nitrogenase, both the Fe protein and the

MoFe protein are catalysts. The rate limiting step for the overall reaction is held to be the dissociation step in the Fe protein cycle (Figure 3-2). This means that the rate constant of the rate limiting step should be the same as the kcat for both Fe protein and MoFe protein cycles in terms of electrons donated/accepted per active site per unit time. Earlier studies established a turnover number (kcat) for electron flow to substrate reduction of between 5-

10 s-1 regardless of substrate.31,43 This value is approximately the same as the first order

ox -1 rate constant for dissociation of the Fe (MgADP)2 from the MoFe protein of ~6 s , leading to the conclusion that the overall reaction rate limiting step is dissociation

(Figure 3-2).1,16–18,31 However, a few previous studies have shown that the rate of

106 dissociation of the Fe-MoFe protein complex using DT is slower than the kcat under saturating (‘high flux’) conditions.40,43

Zhi-Yong Yang determined the Vmax for both Fe protein and MoFe protein cycles using either DT or Fld as reductant by varying the ratio of Fe protein to MoFe protein

(called the “electron flux”). ‘Low-flux’ conditions, which saturate the Fe protein with

-1 MoFe (e.g. 1:1 [Fe:MoFe]) protein, resulted in a Fe protein cycle kcat,Fe of ~6 s with DT

HQ -1 as reductant. When Fld -DT was used as the reductant, a kcat,Fe of 10-11 s was observed (Table 3-1). These results revealed that using Fld as reductant speeds up the overall reaction, suggesting that the rate limiting dissociation step has been accelerated by about 2-fold when Fld is the reductant compared to the dissociation rate constant when

DT is the reductant.40 This enhancement effect, however, was dependent on the electron flux, with lower flux (most representative of physiological conditions) showing the largest effect. When the electron flux was increased to saturate the MoFe protein with Fe protein (‘high flux’, e.g. [Fe]:[MoFe] = 16:1) to measure the MoFe protein cycle, the kcat

-1 was found to be ~10-11 s with both DT and Fld as reductant. The greater kcat,Fe, observed for FldHQ compared to DT as reductant under ‘low flux’ conditions could be explained by an increase in the rate of dissociation of the Fe protein from MoFe protein when FldHQ is the reductant. Based on Yang’s steady-state data, we then determined an

ox apparent dissociation rate constant for Fe (ADP)2 –MoFe protein with either DT or Fld as reductant.

107 Table 3-1. Summary of steady-state kcat values for Fe protein and MoFe protein with different reductants. Data provided by Zhi-Yong Yang.

-1 Protein Reductant kcat (s ) Fe Protein DT 5-6 ‘Low flux’ FldHQ 10-11 MoFe protein DT 10-11 ‘High flux’ FldHQ 10-11

ox Dissociation of the Fe (ADP)2 from the MoFe protein

To determine the apparent dissociation constant using SF spectrophotometry, the

ox Fe (ADP)2-MoFe protein complex was pre-formed in one syringe of the SF spectrophotometer and rapidly mixed against a reductant mixture. Here, such experiments were conducted with either DT or FldHQ in the second syringe. The reduction of the Feox was monitored at 426 nm, an isosbestic point for reversible conversion between FldSQ and FldHQ. As shown in Figure 3-3, the estimated first order rate constant for dissociation

-1 HQ (kobs) with DT as reductant was ~4 s . When Fld was the reductant, the dissociation

ox and reduction of Fe (ADP)2 was much more rapid, with the majority of the reduction

-1 (~65%) occurring in less than 1 ms. A kobs of >760 s was estimated from a fit of the points captured after the dead time. These findings reveal that with FldHQ as reductant,

ox the reduction of Fe is much faster than the observed kcat, implying that dissociation of

ox HQ Fe (ADP)2 from MoFe protein is not the rate limiting step when Fld is used as the reductant. These results are consistent with the previously reported data from similar experiments.25 To further test the effect of reductant on the apparent rate of dissociation, another non-physiological reductant, methyl viologen was used.44 As can be seen in

-1 Figure 3-3, this electron donor gave a kobs of 100 s , again much faster than the kcat.

108

ox HQ Figure 3-3. Reduction of Fe (ADP)2 protein by DT (red), MV (green), or Fld (blue) in ox the presence of MoFe protein. The reduction of Fe (ADP)2 by different reductants was monitored as the decrease of the absorbance at 426 nm as a function of time. The data are displayed as gray dots and were fit to different equations as described in the Materials and Methods section to obtain the pseudo-first order kobs. Syringe 1 contained 20 mM DT with 100 μM MV, or 400 μM FldHQ with 20 mM DT. Syringe 2 contained 80 μM Feox and 80 μM MoFe. MgADP (9 mM) was present in both syringes.

ox HQ The faster reduction of Fe (ADP)2 in the presence of MoFe protein by Fld could be explained by two possible mechanisms: (i) the very rapid dissociation of the

ox HQ Fe (ADP)2 from the MoFe protein when Fld is the reductant, pointing to a different rate limiting step,40 or (ii) FldHQ reduction of the Fe protein while still complexed to the

MoFe protein,45 a mechanism proposed earlier by Haaker et al.25 To test the second

HQ ox model, the ability of Fld to reduce the Fe (ADP)2 while still bound to the MoFe protein was examined.

109 Effect of FldHQ on the primary electron transfer of nitrogenase

The pre-steady state ET from the Fe protein to the MoFe protein offers a straightforward way to monitor the possibility of ET from FldHQ to Feox while still bound to the MoFe protein. When the protein mixture of Fered and MoFe protein is mixed with

MgATP in a SF spectrophotometer, an apparent first order ET event is monitored by the increase in absorbance (oxidation of the Fe protein) that can be fit to a single exponential

-1 to yield a rate constant for ET (kET). With DT as the reductant, a kET of 173 s was observed (Figure 3-4), consistent with literature values that range from 140-200 s-1.14–

16,22,25 Important for this study, the absorbance value plateaus starting at 10 ms and stays roughly flat up to 30 ms. This plateau in absorbance reflects no reduction by DT of the

Feox protein while still in the complex. At much later times (100 ms), the absorbance does change, reflecting a complex set of events as the Fe protein dissociates from the MoFe protein, is reduced, and rebinds to the MoFe protein. When the ET study is conducted with FldHQ as reductant, nearly identical primary ET kinetics are observed (Figure 3-4).

Importantly, no reduction of the Feox protein is observed during the 10 to 30 ms time

ox frame. Given the earlier observation of reduction of Fe (ADP)2 protein-MoFe protein with FldHQ of greater than 760 s-1, an on-complex reduction of the Fe protein in the ET study should have resulted in a significant (if not complete) reduction of the Fe protein before 10 ms. The lack of any observed reduction of the Feox protein in the ET

HQ ox experiment reveals that the Fld protein reduction of the Fe (ADP)2 protein on complex with the MoFe protein must be very slow, as has been reported earlier.45 Thus, these studies, coupled with the dissociation studies presented in the previous section, rule out

FldHQ reduction of the Fe protein while still bound to the MoFe protein,25 and instead

110

red Figure 3-4. Primary ET from Fe (ATP)2 protein to MoFe protein in presence of DT or FldHQ with DT. Syringe 1 contained 80 μM Fered protein, 20 μM MoFe protein, 1 mM DT and 10 mM MgATP. Syringe 2 had 10 mM DT or 200 µM FldHQ with 1 mM DT in the presence of 10 mM MgATP.

ox HQ favor a rapid dissociation of Fe (ADP)2 protein from the MoFe protein when Fld is the reductant.

Reduction of Feox protein by DT and FldHQ

ox The studies, to this point, favor rapid dissociation of the Fe (ADP)2 protein from the MoFe protein when FldHQ is the reductant. The observed lower rate of dissociation when DT is the reductant could indicate slow reduction of Feox protein, rather than slow dissociation of the Fe protein from the MoFe protein.17,23,25 Several previous studies have

•- 17– illustrated low activity for nitrogenase at low concentrations of SO2 as the reductant.

19,22,46 To determine the rates of reduction of Feox protein by different reductants, the pre- steady state studies were conducted using Feox in the absence and presence of nucleotides

111 (ADP or ATP). Feox protein in one syringe was rapidly mixed against either DT or

FldHQ in the other syringe. Additionally, these studies were conducted at 4o C to slow the reactions down enough to observe. Consistent with previous studies, nucleotides significantly slow down the rates of reduction of Feox protein by DT (by ~ 100-fold).22,23

In contrast, with FldHQ as reductant, the rates of reduction remained fast and roughly unchanged with or without nucleotide present (Figure 3-5).25,27

Figure 3-5. Kinetics of the reduction of Feox protein by DT and FldHQ with DT in the presence and absence of nucleotides. ET is monitored by observing the change in absorbance at 426 nm as a function of time. Panel A: Reduction of Fe protein by DT and FldHQ with DT with no nucleotide. Panel B: Reduction of Fe protein by DT and FldHQ with DT in the presence of MgADP. Panel C: Reduction of Fe protein by DT and FldHQ with DT in the presence of MgATP. Syringe 1 contained 80 μM Feox. Syringe 2 contained 20 mM DT or 600 μM FldHQ with 20 mM DT. MgADP and MgATP, when included, were present in both syringes at a final concentration of 10 mM. The kobs values were averaged from two independent experiments except for Fld with MgATP.

112 Taken together, the studies presented so far suggest that when FldHQ is the

ox reductant, the Fe (ADP)2 dissociation from the MoFe protein is not rate limiting and does not correspond to the overall kcat for substrate reduction. Rather, the dissociation

HQ step when Fld is used as the reductant is much more rapid than the kcat. What then is the rate limiting step in the overall reaction cycle?

Kinetic analyses of ATP hydrolysis and Pi release

In the Fe protein cycle using DT as the reductant (Figure 3-2), the ATP

-1 -1 hydrolysis (kATP = 50-70 s ) and the Pi release (kPi = 16-22 s ) steps both are slower than

-1 the ET (kET = 140-200 s ). Sudipta Shaw compared pre-steady state rate constants for

ATP hydrolysis using DT or FldHQ as reductant. A 1:4 ([MoFe]:[Fe]) electron flux

-1 -1 o revealed similar values (kATP,DT and kATP,Fld) of 40 s and 44 s (25 C), respectively (data

-1 not shown). These rate constants were about double the kcat (10-11 s ), which is based on the finding that 2 ATP molecules are hydrolyzed per electron (determined by Sudipta

Shaw, data not shown). Overall, Shaw’s data indicate that ATP hydrolysis is not the rate- limiting step for the overall reaction. Interestingly, Shaw also discovered that the steady state ATP hydrolysis rate constant of 25-27 s-1 is about the same at that reported for the pre-steady state Pi release step (16-22 s-1, Figure 3-2).16,18 Given that two Pi are released per electron transferred, the Pi release rate constant should be double the kcat value (about

20 s-1) if it is the overall rate limiting step.

Pi release was measured using an established fluorometric method42 with either

DT or FldHQ as reductant under high electron flux conditions ([Fe]:[MoFe]: = 16:1). The data showed an initial lag phase followed by a linear phase with both reductants, with the

-1 -1 linear phases having rate constants for Pi release of kPi,DT = 27 s and kPi,Fld = 25 s

113 (Figure 3-6). These rate constants are about double the overall reaction rate constant

-1 kcat of 10-11 s per electron. Two Pi are released for each electron transferred, showing that the overall rate limiting step is likely events associated with Pi release. We have no data on the ADP release event, so the position of this event in the cycle remains unknown.

Establishing the rate-limiting step in the Fe protein cycle

Considering all of the rate constants for the steps in the Fe protein cycle with

HQ -1 -1 Fld as reductant: electron transfer (kET = 173 s ), ATP hydrolysis (kATP ~40-44 s ),

-1 ox -1 Pi release (25-27 s ) and re-reduction of Fe (ADP)2 (kobs >1200 s ), it is concluded

Figure 3-6. Real-time measurement of Pi release during nitrogenase catalysis using DT (gray) or FldHQ with DT (black) as reductant at 25oC. The Pi release was monitored by a fluorescence increase caused by the binding of Pi to MDCC-PBP. The ratio of Pi/MoFe protein is plotted as a function of time. After the initial lag phase (~30 ms), the data collected from 40 to 200 ms were fitted to a linear equation (solid line), giving rate -1 -1 constants of kPi,DT = 27 s and kPi,Fld = 25 s .

114 that the overall rate limiting step for the reaction are events associated with the Pi release step, not complex dissociation. Our findings are consistent with other ATP hydrolyzing systems, including helicases47,48 and myofibrillar ATPases,49–52 where Pi release is the rate limiting step.

Summary

The Fe protein cycle can now be updated with the rate constants determined here using FldHQ as the reductant, as shown in Figure 3-7. The cycle begins with the rapid

red equilibrium docking of Fe (ATP)2 to MoFe protein, which is followed by the conformationally gated ET events. Next occurs ATP hydrolysis, which is followed by events associated with Pi release. It is unknown if the Pi release event is conformationally gated, but it is clear that events associated with this step are rate limiting for the overall

ox Fe protein cycle. The dissociation of the Fe (ADP)2 is fast, with rapid reduction by the

FldHQ. Finally, ATP replaces ADP in the free Fe protein. The order of these last two events is not established, but the rapid reduction by FldHQ suggests that reduction should occur before nucleotide exchange.

115 [I]

[e-] 2ATP 2ADP ox red red [Fe (ADP)2] [Fe (ADP)2] [Fe (ATP)2] -1 fast red k 1200 s

MoFe red,Fld MoFe MoFe 1

- MoFe

5 s

4

M 1

M

.

0 -

-

4

[

1

6 s

1

I

-

1

 7

s

I d

0 5

]

-

s e

1

7

1

 1

- x

1

0

s a

6 s

A

l

i

d e

c k

t R

i

red ox red v

[MoFe ;Fe (ADP)2] [MoFe;Fe (ATP)2] a ]

t

e

V I

d [ 2[Pi] -1 -1 kPi = 25-27 s kET = 173 s

-1 red ox k ATP = 40-44 s red ox [MoFe ;Fe (ADP,Pi)2] [MoFe ;Fe (ATP)2]

[III]

Figure 3-7. Updated Fe protein cycle with FldHQ as reductant.

FUNDING

This work was supported as part of the Biological Electron Transfer and Catalysis Energy

Frontier Research Center funded by the U.S. Department of Energy, Office of Science,

Basic Energy Sciences under award #DE-SC0012518.

ACKNOWLEDGMENTS

The authors thank the members of the Biological Electron Transfer and Catalysis Energy

Frontiers Research Center for helpful discussions.

AUTHOR CONTRIBUTIONS (BASED ON MODIFIED VERSION)

ZYY and RL, Purification of Fld; ZYY and LCS, Steady state kinetics; RL, ZYY, SS, and LCS, Pre-steady state kinetics; SS and LCS, ATP studies; NP and EA, Preparation of reagents for Pi release study; VLC and DRD, Fld genetics; All authors involved in data interpretation, writing, and revisions.

116 REFERENCES

(1) Burgess, B. K., and Lowe, D. J. (1996) Mechanism of molybdenum nitrogenase.

Chem. Rev. 96, 2983–3012.

(2) Seefeldt, L. C., Hoffman, B. M., and Dean, D. R. (2009) Mechanism of Mo- dependent nitrogenase. Annu. Rev. Biochem. 78, 701–722.

(3) Hoffman, B. M., Lukoyanov, D., Yang, Z.-Y., Dean, D. R., and Seefeldt, L. C. (2014)

Mechanism of nitrogen fixation by nitrogenase: the next stage. Chem. Rev. 114, 4041–

4062.

(4) Hageman, R. V., and Burris, R. H. (1978) Nitrogenase and nitrogenase reductase associate and dissociate with each catalytic cycle. Proc. Natl. Acad. Sci. U.S.A. 75, 2699–

2702.

(5) Georgiadis, M. M., Komiya, H., Chakrabarti, P., Woo, D., Kornuc, J. J., and Rees, D.

C. (1992) Crystallographic structure of the nitrogenase iron protein from Azotobacter vinelandii. Science 257, 1653–1659.

(6) Seefeldt, L. C., Hoffman, B. M., and Dean, D. R. (2012) Electron transfer in nitrogenase catalysis. Curr. Opin. Chem. Biol. 16, 19–25.

(7) Chan, M. K., Kim, J., and Rees, D. C. (1993) The nitrogenase FeMo-cofactor and P- cluster pair: 2.2 Å resolution structures. Science 260, 792–794.

(8) Kim, J., and Rees, D. C. (1992) Structural models for the metal centers in the nitrogenase molybdenum-iron protein. Science 257, 1677–1682.

(9) Einsle, O., Tezcan, F. A., Andrade, S. L. A., Schmid, B., Yoshida, M., Howard, J. B., and Rees, D. C. (2002) Nitrogenase MoFe-protein at 1.16 Å resolution: A central ligand in the FeMo-cofactor. Science 297, 1696–1700.

117 (10) Spatzal, T., Aksoyoglu, M., Zhang, L., Andrade, S. L. A., Schleicher, E., Weber,

S., Rees, D. C., and Einsle, O. (2011) Evidence for interstitial carbon in nitrogenase

FeMo cofactor. Science 334, 940–940.

(11) Lancaster, K. M., Roemelt, M., Ettenhuber, P., Hu, Y., Ribbe, M. W., Neese, F.,

Bergmann, U., and DeBeer, S. (2011) X-ray emission spectroscopy evidences a central carbon in the nitrogenase iron-molybdenum cofactor. Science 334, 974–977.

(12) Thorneley, R. N., Lowe, D. J., Eday, R. R., and Miller, R. W. (1979) The coupling of electron transfer in nitrogenase to the hydrolysis of magnesium . Biochem. Soc. Trans. 7, 633–636.

(13) Hageman, R. V., Orme-Johnson, W. H., and Burris, R. H. (1980) Role of magnesium adenosine 5’-triphosphate in the hydrogen evolution reaction catalyzed by nitrogenase from Azotobacter vinelandii. Biochemistry 19, 2333–2342.

(14) Danyal, K., Dean, D. R., Hoffman, B. M., and Seefeldt, L. C. (2011) Electron transfer within nitrogenase: Evidence for a deficit-spending mechanism. Biochemistry 50,

9255–9263.

(15) Danyal, K., Mayweather, D., Dean, D. R., Seefeldt, L. C., and Hoffman, B. M.

(2010) Conformational gating of electron transfer from the nitrogenase Fe protein to

MoFe protein. J. Am. Chem. Soc. 132, 6894–6895.

(16) Duval, S., Danyal, K., Shaw, S., Lytle, A. K., Dean, D. R., Hoffman, B. M., Antony,

E., and Seefeldt, L. C. (2013) Electron transfer precedes ATP hydrolysis during nitrogenase catalysis. Proc. Natl. Acad. Sci. U.S.A. 110, 16414–16419.

118 (17) Thorneley, R. N. F., and Lowe, D. J. (1983) Nitrogenase of Klebsiella pneumoniae. Kinetics of the dissociation of oxidized iron protein from molybdenum-iron protein: Identification of the rate-limiting step for substrate reduction. Biochem. J. 215,

393–403.

(18) Wilson, P. E., Nyborg, A. C., and Watt, G. D. (2001) Duplication and extension of the Thorneley and Lowe kinetic model for Klebsiella pneumoniae nitrogenase catalysis using a MATHEMATICA software platform. Biophys. Chem. 91, 281–304.

(19) Thorneley, R. N. F., and Lowe, D. J. (1985) Kinetics and mechanisms of the nitrogenase enzyme system, in Molybdenum Enzymes (Spiro, T. G., Ed.), pp 221–284.

Wiley, New York.

(20) Seefeldt, L. C., Yang, Z.-Y., Duval, S., and Dean, D. R. (2013) Nitrogenase reduction of carbon-containing compounds. Biochim. Biophys. Acta 1827, 1102–1111.

(21) Lowe, D. J., and Thorneley, R. N. (1984) The mechanism of Klebsiella pneumoniae nitrogenase action: The determination of rate constants required for the simulation of the kinetics of N2 reduction and H2 evolution. Biochem. J. 224, 895.

(22) Ashby, G. A., and Thorneley, R. N. F. (1987) Nitrogenase of Klebsiella pneumoniae: Kinetic studies on the Fe protein involving reduction by sodium dithionite, the binding of MgADP and a conformation change that alters the reactivity of the 4Fe-4S centre. Biochem. J. 246, 455–465.

(23) Wilson, P. E., Bunker, J., Lowery, T. J., and Watt, G. D. (2004) Reduction of nitrogenase Fe protein from Azotobacter vinelandii by dithionite: Quantitative and qualitative effects of nucleotides, temperature, pH and reaction buffer. Biophys. Chem.

109, 305–324.

119 (24) Yates, M. G. (1972) Electron transport to nitrogenase in Azotobacter chroococcum: Azotobacter flavodoxin hydroquinone as an electron donor. FEBS Lett. 27,

63–67.

(25) Duyvis, M. G., Wassink, H., and Haaker, H. (1998) Nitrogenase of Azotobacter vinelandii: Kinetic analysis of the Fe protein redox cycle. Biochemistry 37, 17345–17354.

(26) Bennett, L. T., Jacobson, M. R., and Dean, D. R. (1988) Isolation, sequencing, and mutagenesis of the nifF gene encoding flavodoxin from Azotobacter vinelandii. J. Biol.

Chem. 263, 1364–1369.

(27) Thorneley, R. N., and Deistung, J. (1988) Electron-transfer studies involving flavodoxin and a natural redox partner, the iron protein of nitrogenase: Conformational constraints on protein-protein interactions and the kinetics of electron transfer within the protein complex. Biochem. J. 253, 587–595.

(28) Mortenson, L. E. (1964) Ferredoxin and ATP, requirements for nitrogen fixation in cell-free extracts of Clostridium pasteurianum. Proc. Natl. Acad. Sci. U.S.A. 52, 272–

279.

(29) Martin, A. E., Burgess, B. K., Iismaa, S. E., Smartt, C. T., Jacobson, M. R., and

Dean, D. R. (1989) Construction and characterization of an Azotobacter vinelandii strain with mutations in the genes encoding flavodoxin and ferredoxin I. J. Bacteriol. 171,

3162–3167.

(30) Saeki, K. (2004) Electron transport to nitrogenase: Diverse routes for a common destination, in Genetics and regulation of nitrogen fixation in free-living bacteria (Klipp,

W., Masephol, B., Gallon, J. R., and Newton, W. E., Eds.), pp 257–290. Kluwer

Academic Publishers, Dordrecht.

120 (31) Hageman, R. V., and Burris, R. H. (1978) Kinetic studies on electron transfer and interaction between nitrogenase components from Azotobacter vinelandii.

Biochemistry 17, 4117–4124.

(32) Simpson, F. B., and Burris, R. H. (1984) A nitrogen pressure of 50 atmospheres does not prevent evolution of hydrogen by nitrogenase. Science 224, 1095–1097.

(33) Erickson, J. A., Nyborg, A. C., Johnson, J. L., Truscott, S. M., Gunn, A.,

Nordmeyer, F. R., and Watt, G. D. (1999) Enhanced efficiency of ATP hydrolysis during nitrogenase catalysis utilizing reductants that form the all-ferrous redox state of the Fe protein. Biochemistry 38, 14279–14285.

(34) Lowery, T. J., Wilson, P. E., Zhang, B., Bunker, J., Harrison, R. G., Nyborg, A. C.,

Thiriot, D., and Watt, G. D. (2006) Flavodoxin hydroquinone reduces Azotobacter vinelandii Fe protein to the all-ferrous redox state with a S = 0 spin state. Proc. Natl.

Acad. Sci. U.S.A. 103, 17131–17136.

(35) Fisher, K., Hare, N. D., and Newton, W. E. (2014) Another role for CO with nitrogenase? CO stimulates hydrogen evolution catalyzed by variant Azotobacter vinelandii Mo-Nitrogenases. Biochemistry 53, 6151–6160.

(36) Benemann, J. R., Yoch, D. C., Valentine, R. C., and Arnon, D. I. (1969) The electron transport system in nitrogen fixation by Azotobacter, I: Azotoflavin as an electron carrier. Proc. Natl. Acad. Sci. U.S.A. 64, 1079–1086.

(37) Klugkist, J., Voorberg, J., Haaker, H., and Veeger, C. (1986) Characterization of three different flavodoxins from Azotobacter vinelandii. Eur. J. Biochem. 155, 33–40.

121 (38) Steensma, E., Heering, H. A., Hagen, W. R., and Van Mierlo, C. P. M. (1996)

Redox properties of wild-type, Cys69Ala, and Cys69Ser Azotobacter vinelandii flavodoxin II as measured by cyclic voltammetry and EPR spectroscopy. Eur. J.

Biochem. 235, 167–172.

(39) Lanzilotta, W. N., Ryle, M. J., and Seefeldt, L. C. (1995) Nucleotide hydrolysis and protein conformational changes in Azotobacter vinelandii nitrogenase iron protein:

Defining the function of aspartate 129. Biochemistry 34, 10713–10723.

(40) Jacobs, D., and Watt, G. D. (2013) Nucleotide-assisted [Fe4S4] redox state interconversions of the Azotobacter vinelandii Fe protein and their relevance to nitrogenase catalysis. Biochemistry 52, 4791–4799.

(41) Christiansen, J., Goodwin, P. J., Lanzilotta, W. N., Seefeldt, L. C., and Dean, D. R.

(1998) Catalytic and biophysical properties of a nitrogenase Apo-MoFe protein produced by a nifB-deletion mutant of Azotobacter vinelandii. Biochemistry 37, 12611–12623.

(42) Brune, M., Hunter, J. L., Corrie, J. E. T., and Webb, M. R. (1994) Direct, real-time measurement of rapid inorganic phosphate release using a novel fluorescent probe and its application to actomyosin subfragment 1 ATPase. Biochemistry 33, 8262–8271.

(43) Johnson, J. L., Nyborg, A. C., Wilson, P. E., Tolley, A. M., Nordmeyer, F. R., and

Watt, G. D. (2000) Analysis of steady state Fe and MoFe protein interactions during nitrogenase catalysis. Biochim. Biophys. Acta, Gen. Subj. 1543, 24–35.

(44) Watt, G. D., and Reddy, K. R. N. (1994) Formation of an all ferrous Fe4S4 cluster in the iron protein component of Azotobacter vinelandii nitrogenase. J. Inorg. Biochem. 53,

281–294.

122 (45) Lanzilotta, W. N., and Seefeldt, L. C. (1997) Changes in the midpoint potentials of the nitrogenase metal centers as a result of iron protein−molybdenum-iron protein complex formation. Biochemistry 36, 12976–12983.

(46) Johnson, J. L., Tolley, A. M., Erickson, J. A., and Watt, G. D. (1996) Steady-state kinetic studies of dithionite utilization, component protein interaction, and the formation of an oxidized iron protein intermediate during Azotobacter vinelandii nitrogenase catalysis. Biochemistry 35, 11336–11342.

(47) Wong, E. V., Cao, W., Vörös, J., Merchant, M., Modis, Y., Hackney, D. D.,

Montpetit, B., and De La Cruz, E. M. (2016) Pi release limits the intrinsic and RNA- stimulated ATPase cycles of DEAD-box protein 5 (Dbp5). J. Mol. Biol. 428, 492–508.

(48) Wang, Q., Arnold, J. J., Uchida, A., Raney, K. D., and Cameron, C. E. (2010)

Phosphate release contributes to the rate-limiting step for unwinding by an RNA helicase.

Nucleic Acids Res. 38, 1312–1324.

(49) Cochran, J. C., Krzysiak, T. C., and Gilbert, S. P. (2006) Pathway of ATP hydrolysis by monomeric kinesin Eg5. Biochemistry 45, 12334–12344.

(50) Lionne, C., Iorga, B., Candau, R., Piroddi, N., Webb, M. R., Belus, A., Travers, F., and Barman, T. (2002) Evidence that phosphate release is the rate-limiting step on the overall ATPase of psoas myofibrils prevented from shortening by chemical cross-linking.

Biochemistry 41, 13297–13308.

(51) Lionne, C., Brune, M., Webb, M. R., Travers, F., and Barman, T. (1995) Time resolved measurements show that phosphate release is the rate limiting step on myofibrillar ATPases. FEBS Lett. 364, 59–62.

123 (52) Milic, B., Andreasson, J. O. L., Hancock, W. O., and Block, S. M. (2014)

Kinesin processivity is gated by phosphate release. Proc. Natl. Acad. Sci. U.S.A. 111,

14136–14140.

124 CHAPTER 4

CO2 AND N2 FIXATION BY MICROBIAL ELECTROCATALYSIS

ABSTRACT

The biological and industrial conversions of atmospheric N2 to bioavailable

+ ammonium (NH4 ) are essential for the production of fertilizer in conventional agriculture. However, as Earth’s population continues to grow and we look toward sending humans to Mars, new approaches for the sustainable and economical production of fixed N are necessary. Here, we present a biohybrid approach, in which the phototrophic, nitrogen-fixing bacterium Rhodopseudomonas palustris was grown in association with a platinum electrode. In this system, the cells used light for energy and electrocatalytically-generated H2 as the source of electrons. Initial observations revealed novel features of R. palustris, including its ability to grow photoautotrophically using both incandescent light and infra-red photons (847 nm) with an equivalent generation time of 4 days. Further, cells fixed CO2 under H2-mediated electrocatalytic conditions.

The next step is to grow the cells electrocatalytically under both CO2 and N2 fixing

+ conditions and optimize NH4 production. Using such an approach would offer an

+ alternative route for the microbial electrosynthesis of NH4 that could help make food production more sustainable.

______*Unpublished data with contributions from Rhesa N. Ledbetter, Mathangi Soundararajan, Zhi-Yong Yang, Paul Kusuma, Bruce Bugbee, Scott A. Ensign, Paul W. Ludden, and Lance C. Seefeldt

125 INTRODUCTION

Fixed nitrogen compounds are a critical nutrient for the growth of all food crops.1

+ - These compounds are most often added as fertilizers (i.e. NH3, NH4 , or NO3 ) with the nitrogen being ultimately derived from the reduction of atmospheric dinitrogen (N2). The reduction of N2, which is abundant in the Earth’s lower atmosphere, to ammonia (NH3) is achieved through a process called nitrogen fixation.2–4

There are two major processes that accomplish nitrogen fixation: biological nitrogen fixation in diazotrophic bacteria and industrial nitrogen fixation through the

Haber-Bosch process. Biological nitrogen fixation is achieved through the action of the

5,6 enzyme nitrogenase. This “natural” process has many positive attributes including N2 fixation under benign conditions, but is of limited usefulness in agriculture. In cases where diazotrophic bacteria form symbiotic relationships with plants like legumes

(soybeans, alfalfa, and peanuts), biological nitrogen fixation can provide the nitrogen needs of the crop. Most crops (e.g., corn, wheat, and rice), however, do not associate with diazotrophs.

The invention of the Haber-Bosch industrial process in the early 20th century provided nitrogen fertilizer derived from N2 and was a tremendous boon to food production, with estimates of over half of all nitrogen used coming from this process.2,7

The downside to the Haber-Bosch process is that it requires significant energy inputs derived from fossil fuels. Estimates are that the Haber-Bosch process utilizes around 2%

8–10 of global fossil fuels and is responsible for 3% of the CO2 production worldwide.

There is growing realization that future food production must move to more sustainable methods for achieving N2 reduction. Further, as we look toward the future of deep space

126 travel to Mars, development of technologies to manufacture products, such as fixed nitrogen, independent of resources from Earth is also crucial.

Several alternative approaches11 for fixed nitrogen production have been explored, including the use of transition metal catalysts,12,13 electrocatalysts,10,14 photocatalysts,15–20 diazotrophic bacteria,21–23 and purified nitrogenase.16,24 One area of particular interest is capitalizing on nitrogenase functioning inside bacteria. A challenge with using bacteria as an industrial catalyst is that they need a source of energy to sustain nitrogen fixation, which is usually provided in the form of fixed carbon, such as sugars.

Providing sugars is not an economical or energy efficient way to support nitrogen fixation in bacteria for larger-scale applications. Thus, a more sustainable way to deliver electrons is the critical missing link in developing bacteria as a way to produce fertilizer at scale for non-leguminous crops.

Here, we provide a demonstration of how to couple low-energy, infra-red (IR) light to CO2 fixation under ambient conditions using a biohybrid approach, in which the phototrophic bacterium Rhodopseudomonas palustris is cultured in association with platinum electrodes. For cell growth, electrocatalytically-produced H2 served as a source of electrons and the energy for the cells was derived from long-wavelength IR light that does not compete with wavelengths used by plants. Further, the cells grew electrocatalytically under CO2 fixing conditions eliminating the need for an organic carbon source. The next step is take our collective knowledge and grow cells under both

CO2 and N2 fixing conditions to validate this biohybrid approach as a potential next generation system for sustainable fixed N production for food production.

127 MATERIALS AND METHODS

Microbial media and R. palustris growth conditions

R. palustris TIE-1 was routinely cultured anaerobically under photoheterotrophic growth conditions in photosynthetic medium (PM) with 20 mM acetate as the carbon source.25,26 Photoheterotrophic growths were made anaerobic by filling the tubes completely full with PM medium and then allowing the cells to naturally achieve anaerobicity following consumption of the remaining oxygen. Growth was measured by monitoring the optical density (660 nm) of the cultures.

For photoautotrophic growths, cells were cultured in nitrogen-fixing (NF) medium.27 The overall composition of NF medium is the same as PM, except the nitrogen source is removed; however, for these studies thiosulfate was also eliminated as the electron donor, as H2 was used instead. The NF base medium, lacking both fixed N and thiosulfate, was then supplemented with Wolfe’s vitamin solution (0.5 ml/100 ml),28 1

μM NiSO4, and 200 mM NaCl. The 200 mM NaCl was specifically added to increase conductivity in the electrochemical growths, but was included in all photoautotrophic cultures for consistency. N2 gas was bubbled through to medium to remove any oxygen, and the headspace was equilibrated to 1 atm. An equal volume of H2 gas was included in the headspace to reach a total pressure of 2 atm, and just prior to inoculation, 30 mM

NaHCO3 was added to the degassed medium. Cells grown photoheterotrophically served as the inoculum. To remove any residual fixed N, carbon source, and electron donor, these cells were first washed 3x in NF medium before addition to the sealed vials.

Cultures were then grown with either a 60W incandescent (broad spectrum) or IR light

128 (847nm) (Figure 4-2) at 25°C. Spectral traces of the incandescent light and IR LEDs were made with a spectroradiometer (Apogee Instruments, Model PS-100, Logan, UT).

Bioelectrochemical system set-up and growths

The bioelectrochemical system (BES) consisted of two ~600 ml chambers separated with a 20% polyacrylamide gel plug (4 x 6 cm). For electrolysis, NF medium described above served as the electrolyte solution and platinum wires (16 gauge; length=

~3 cm) acted as the working and counter electrodes. The working chamber (cathode) was degassed with N2 and then 30 mM NaHCO3 was added to both the cathode and anode. R. palustris TIE-1 cells grown photoheterotrophically served as the inoculum. Prior to inoculation, cells were washed 3x in NF medium as described above. The fixed N for cell

+ growth was derived from NH4 production from release of amide groups in the polyacrylamide plug. The cathodic chamber was loosely covered with a stopper to maintain the atmosphere, while minimizing H2 gas pressure build-up. The anodic chamber, however, was open to the atmosphere. Using a Gamry potentiostat

(Warminster, PA), the BES was maintained under a constant current of 5.0 mA for the duration of the experiment. Each day, CO2 was bubbled through the cathode to replenish the carbon source as well as maintain the pH of the medium at ~7.0 (in general 1 min. for each 0.1 change in pH). Following, the addition of CO2, N2 was also replenished by bubbling through the solution for 4 min and the headspace for 2 min. All growths were conducted at 25°C.

129 RESULTS AND DISCUSSION

R. palustris TIE-1 is a purple, non-sulfur phototrophic bacterium with a diverse metabolism, capable of fixing N2 and CO2, as well as accepting electrons from various

29–32 sources (H2, Fe(II), thiosulfate, organic acids, and an electrode). To capitalize on these properties, a biohybrid system was proposed, in which R. palustris TIE-1 is grown in association with platinum electrodes (Figure 4-1), with the electrons for microbial growth derived from H2 produced from electrolysis, and the ATP required for nitrogenase and the generated from light energy. This approach would replace the need for chemicals in the form of sugar required by many bacteria.

e

e

d

d

o

r o N2 t CO r 2

t

c

c

e

l

e +

l 2 H O e 4 H

2

e ATP RuBisCo

m

m

u

u

n -

i

n e

t

i

t a NH3 Biomaterial

l + - a 2 H l 2 O2 + 4 H + 4 e P N ase 2 (N, C)

P

-

-

e

e

d

d

o

o

h

t

n

a

A

C

+ - Potentiostat

Figure 4-1. A microbial electrosynthesis cell for light-driven CO2 and N2 reduction by R. palustris (red oval).

130 Prior to growing the cells electrochemically, however, it was confirmed that

R. palustris could in fact grow in vials, under CO2 and N2 fixing conditions with H2 as the electron donor. In conjunction with this, the response of R. palustris growing with broad-spectrum incandescent light and narrow spectrum IR photons (847 nm) was tested.

While incandescent light is generally used for growing R. palustris in the laboratory, it is well established that purple bacteria harbor photosynthetic pigments known as bacteriochlorophylls (Bchls) in their photosynthetic reaction centers and light-harvesting complexes that have absorption maxima in the IR range (>700-1100 nm).33 The photosynthetic reaction centers in R. palustris absorb light at ~870 nm.34,35 The Bchls are evolutionarily significant, as these organisms are able to grow beneath the plants and cyanobacteria where “leftover” IR photons penetrate.36 Thus, as we look toward

+ biological solutions for NH4 production, especially on Mars, having cells that use lower- energy wavelengths distinct from those of plants would be advantageous. Previous studies have indeed demonstrated the versatility of these cells to grow photoheterotrophically, albeit at different efficiencies, with various wavelengths of

34,37,38 light. In this study, cells were grown under photoautotrophic, N2 fixing conditions using both incandenscent light and IR photons.

The incandescent light (60W) displayed a broad spectrum ranging from ~400 to

>950 nm, and the IR LED presented a narrow emission band with a maximum of ~847 nm (Figure 4-2). A generation time of 4 days was observed for cells grown under photoautotrophic, nitrogen fixing conditions with both the incandescent light and the IR

LED (Figure 4-3). A previous study with R. palustris cells cultivated under

131

Figure 4-2. Spectra of the (A) incandescent light (60W) and (B) IR LED (DMetric IR Illuminator, 96 LED) used in this study.

Figure 4-3.Growth curves of R. palustris grown under phototrophic, nitrogen fixing conditions with incandescent (blue) and IR light (red). CO2 was the carbon source, N2 was the nitrogen source, and H2 was the electron donor. Vials contained a headspace of 50% N2:50% H2 at 2 atm of pressure. The cells had an average generation time of 4 days.

132 photoheterotrophic conditions grew 76% less with IR than incandescent.38 The work presented here, however, cannot be directly compared to the previous study due to different growth parameters. Further, correlations between light treatments in this study are not possible either since the number of photons varied between the light treatments.

Difference in the availability of photons may, in fact, account for the incandescent growth reaching death phase before the IR growth (Figure 4-3). Even without comparison, however, the data presented here demonstrate that cells can exclusively use lower-energy IR light for growth under photoautotrophic, nitrogen-fixing conditions.

Upon revealing the effectiveness of IR light for R. palustris growth, it was subsequently employed as the energy source when examining the growth of the cells under electrochemical conditions. To grow the cells electrochemically, platinum electrodes were connected to a potentiostat, and anodic and cathodic chambers were separated with a polyacrylamide plug. H2 produced electrocatalytically served as the electron source to the bacterial cells. A current of 5 mA was applied to maintain a constant amount of H2 production (2.2 mmol/H2/day) at the cathode. While further experimentation on minimal and maximal currents for growth should be conducted, it was thought that this amount of H2 would be sufficient to support microbial growth as Km values for H2 uptake by hydrogenases are in the low μM range and H2 saturation levels

39,40 are 0.78 mM, well above the Km.

Initially, studies were designed to grow cells electrochemically with CO2 as the carbon source and N2 as the nitrogen source. However, it was demonstrated the

+ acrylamide plug released NH4 over time resulting in fixed N in the extracellular medium. This meant that the study was in fact conducted only under CO2 fixing

133 conditions. Cells did, however, grow under these conditions and exhibited an average generation time of 6 days (Figure 4-4). Cells grown without the electrocatalytically produced H2 did not grow (data not shown), suggesting H2 is serving as the only source of electrons.

While there is no current evidence of N2 fixation under electrochemical conditions, further studies on alternative membranes (e.g. Nafion, dialysis membranes,

+ and agarose) that will not release NH4 are being evaluated. It should not go unnoticed, however, that cells exhibited growth when cultivated under non-electrochemical photoautotrophic, nitrogen fixing conditions (Figure 4-3) and electrochemically under

CO2 fixing conditions (Figure 4-4). Collectively, these results suggest that R. palustris has a high probability of achieving both CO2 and N2 fixing conditions in the electrochemical set-up. Further, H2-mediated electrocatalytic growth under CO2 and N2

Figure 4-4. Growth of R. palustris in a microbial electrosynthesis cell with IR LEDs. A constant current of -5 mA was applied and yielded voltages of approximately -3.2 V. The cells had an average generation time of 6 days.

134 fixing conditions was recently demonstrated in a study that cultivated the autotroph

Xanthobacter autotrophicus.23

As efforts continue in developing a biohybrid approach with R. palustris, Figure

4-5 outlines the strategy of establishing this system as an alternative for production of fixed N for plant growth. First, it will be demonstrated that R. palustris can grow electrochemically under both CO2 and N2 fixing conditions. Then, the accumulated biomass containing the fixed N will be tested as a fertilizer for plant growth (Figure 4-5).

While it is has been demonstrated that some plants can take up microbial cells directly as a source of nutrients,23,41 it is anticipated that the biomass may need to be composted to release the fixed N depending on the plant needs. Further, alternative photoautotrophic, nitrogen-fixing organisms should be investigated, as the generation time of R. palustris grown photoautotrophically is relatively slow.

In conclusion, we have demonstrated that a biohybrid system with R. palustris utilizing IR light exclusively as an energy source and electrocatalytically produced H2 as an electron donor can grow under CO2 fixing conditions. The next step is to establish that the cells can accomplish both CO2 and N2 fixation. Demonstrating this would be particularly useful when considering deep space missions where in situ production of

+ NH4 and minimal infrastructure requirements are a critical consideration for fertilizer production. Even on Earth, this could provide a novel route to fix N2 at ambient temperature and pressure with a much lower dependence on fossil fuels, which the current industrial process depends on heavily. While further research is needed on this

135 system, this study lays some groundwork and insight for further studies focused on capitalizing on biological nitrogen fixation.

N2 CO

e 2

e

d

d o +

r o 4 H

t

r

t c ATP RuBisCo

c

e

l

e

l 2 H O e

2 - e e

m NH3 Biomaterial

m u 2 H

u 2

n N ase

i 2 (N, C)

n

t

i

t

a

l

+ - a

l O2 + 4 H + 4 e P

P

-

-

e

e

d

d

o

o

h

t

n

a

A

C

Compost cells

+ - Potentiostat

Figure 4-5. Development of a biohybrid system for fertilizer production, in which microbial biomass is used a source of fixed N for plant growth.

136 ACKNOWLEDGMENTS

The authors would like to thank Dr. Caroline Harwood at the University of

Washington for providing R. palustris TIE-1, and Dr. Tim Magnuson at Idaho State

University for use of his potentiostat. This work was performed as part of the Center for the Utilization of Biological Engineering in Space (CUBES) and was supported by a

Space Technology Research Institute grant from NASA’s Space Technology Research

Grants Program under grant or cooperative agreement award number

NNXNNX17AJ31G S02.

AUTHOR CONTRIBUTIONS

RNL, MS, LCS, R. palustris growth experiments; RNL, MS, ZYY, Electrochemical set- up; RNL, MS, PK and BB, Incandescent light and IR LED measurements; SAE and

PWL, IR experiment ideas and insight; RNL, MS, ZYY, LCS, BB, Data interpretation, writing, and revisions.

REFERENCES

(1) Tilman, D., Cassman, K. G., Matson, P. A., Naylor, R., and Polasky, S. (2002)

Agricultural sustainability and intensive production practices. Nature.

(2) Smil, V. (2004) Enriching the Earth: Fritz Haber, Carl Bosch, and the transformation of world food production. MIT Press.

(3) Smil, V. (1997) Global population and the nitrogen cycle. Sci. Am. 277, 76–81.

(4) Thomas, R. Q., Brookshire, E. N. J., and Gerber, S. (2015) Nitrogen limitation on land: how can it occur in Earth system models? Glob. Change Biol. 21, 1777–1793.

(5) Burgess, B. K., and Lowe, D. J. (1996) Mechanism of molybdenum nitrogenase.

Chem. Rev. 96, 2983–3012.

137 (6) Patil, B. S., Hessel, V., Seefeldt, L. C., Dean, D. R., Hoffman, B. M., Cook, B. J., and Murray, L. J. (2000) Nitrogen fixation, in Ullmann’s Encyclopedia of Industrial

Chemistry, pp 1–21. Wiley-VCH Verlag GmbH & Co. KGaA.

(7) Smil, V. (1999) Detonator of the population explosion. Nature 400, 415.

(8) Ertl, G. (2008) Reactions at surfaces: from atoms to complexity (Nobel Lecture).

Angew. Chem. Int. Ed Engl. 47, 3524–3535.

(9) Appl, M. (2000) Ammonia, 2. Production Processes, in Ullmann’s Encyclopedia of

Industrial Chemistry. Wiley-VCH Verlag GmbH & Co. KGaA.

(10) Lan, R., Irvine, J. T. S., and Tao, S. (2013) Synthesis of ammonia directly from air and water at ambient temperature and pressure. Sci. Rep. 3, 1145.

(11) Chen, J. G., Crooks, R. M., Seefeldt, L. C., Bren, K. L., Bullock, R. M.,

Darensbourg, M. Y., Holland, P. L., Hoffman, B., Janik, M. J., Jones, A. K., Kanatzidis,

M. G., King, P., Lancaster, K. M., Lymar, S. V., Pfromm, P., Schneider, W. F., and

Schrock, R. R. (2018) Beyond fossil fuel–driven nitrogen transformations. Science 360, eaar6611.

(12) Anderson, J. S., Rittle, J., and Peters, J. C. (2013) Catalytic conversion of nitrogen to ammonia by an iron model complex. Nature 501, 84–87.

(13) Arashiba, K., Miyake, Y., and Nishibayashi, Y. (2011) A molybdenum complex bearing PNP-type pincer ligands leads to the catalytic reduction of dinitrogen into ammonia. Nat. Chem. 3, 120–125.

(14) Jewess, M., and Crabtree, R. H. (2016) Electrocatalytic nitrogen fixation for distributed fertilizer production? ACS Sustain. Chem. Eng. 4, 5855–5858.

138 (15) Chen, X., Li, N., Kong, Z., Ong, W.-J., and Zhao, X. (2018) Photocatalytic fixation of nitrogen to ammonia: state-of-the-art advancements and future prospects.

Mater. Horiz. 5, 9–27.

(16) Brown, K. A., Harris, D. F., Wilker, M. B., Rasmussen, A., Khadka, N., Hamby, H.,

Keable, S., Dukovic, G., Peters, J. W., Seefeldt, L. C., and King, P. W. (2016) Light- driven dinitrogen reduction catalyzed by a CdS:nitrogenase MoFe protein biohybrid.

Science 352, 448–450.

(17) Liu, J., Kelley, M. S., Wu, W., Banerjee, A., Douvalis, A. P., Wu, J., Zhang, Y.,

Schatz, G. C., and Kanatzidis, M. G. (2016) Nitrogenase-mimic iron-containing chalcogels for photochemical reduction of dinitrogen to ammonia. Proc. Natl. Acad. Sci.

113, 5530–5535.

(18) Zhu, D., Zhang, L., Ruther, R. E., and Hamers, R. J. (2013) Photo-illuminated diamond as a solid-state source of solvated electrons in water for nitrogen reduction. Nat.

Mater. 12, 836–841.

(19) Oshikiri, T., Ueno, K., and Misawa, H. (2016) Selective dinitrogen conversion to ammonia using water and visible light through plasmon-induced charge separation.

Angew. Chem. Int. Ed. 55, 3942–3946.

(20) Ali, M., Zhou, F., Chen, K., Kotzur, C., Xiao, C., Bourgeois, L., Zhang, X., and

MacFarlane, D. R. (2016) Nanostructured photoelectrochemical solar cell for nitrogen reduction using plasmon-enhanced black silicon. Nat. Commun. 7, 11335.

(21) Shanmugam, K. T., and Valentine, R. C. (1975) Microbial production of ammonium ion from nitrogen. Proc. Natl. Acad. Sci. U. S. A. 72, 136–139.

139 (22) Barney, B. M., Eberhart, L. J., Ohlert, J. M., Knutson, C. M., and Plunkett, M. H.

(2015) Gene deletions resulting in increased nitrogen release by Azotobacter vinelandii:

Application of a novel nitrogen biosensor. Appl. Environ. Microbiol. AEM.00554-15.

(23) Liu, C., Sakimoto, K. K., Colón, B. C., Silver, P. A., and Nocera, D. G. (2017)

Ambient nitrogen reduction cycle using a hybrid inorganic–biological system. Proc. Natl.

Acad. Sci. 114, 6450–6455.

(24) Milton, R. D., Abdellaoui, S., Khadka, N., Dean, D. R., Leech, D., Seefeldt, L. C., and Minteer, S. D. (2016) Nitrogenase bioelectrocatalysis: heterogeneous ammonia and hydrogen production by MoFe protein. Energy Environ. Sci. 9, 2550–2554.

(25) Kim, M.-K., and Harwood, C. S. (1991) Regulation of benzoate-CoA ligase in

Rhodopseudomonas palustris. FEMS Microbiol. Lett. 83, 199–203.

(26) Ornston, L. N., and Stanier, R. Y. (1966) The conversion of catechol and protocatechuate to β-ketoadipate by Pseudomonas putida. J. Biol. Chem. 241, 3776–

3786.

(27) Oda, Y., Samanta, S. K., Rey, F. E., Wu, L., Liu, X., Yan, T., Zhou, J., and

Harwood, C. S. (2005) Functional genomic analysis of three nitrogenase isozymes in the photosynthetic bacterium Rhodopseudomonas palustris. J. Bacteriol. 187, 7784–7794.

(28) Wolin, E. A., Wolin, M., and Wolfe, R. S. (1963) Formation of methane by bacterial extracts. J. Biol. Chem. 238, 2882–2886.

(29) Jiao, Y., Kappler, A., Croal, L. R., and Newman, D. K. (2005) Isolation and characterization of a genetically tractable photoautotrophic Fe(II)-oxidizing bacterium,

Rhodopseudomonas palustris strain TIE-1. Appl. Environ. Microbiol. 71, 4487–4496.

140 (30) Jiao, Y., and Newman, D. K. (2007) The pio operon is essential for phototrophic

Fe(II) oxidation in Rhodopseudomonas palustris TIE-1. J. Bacteriol. 189, 1765–1773.

(31) Bose, A., Gardel, E. J., Vidoudez, C., Parra, E. A., and Girguis, P. R. (2014)

Electron uptake by iron-oxidizing phototrophic bacteria. Nat. Commun. 5, 3391.

(32) Larimer, F. W., Chain, P., Hauser, L., Lamerdin, J., Malfatti, S., Do, L., Land, M. L.,

Pelletier, D. A., Beatty, J. T., Lang, A. S., Tabita, F. R., Gibson, J. L., Hanson, T. E.,

Bobst, C., Torres, J. L. T. y, Peres, C., Harrison, F. H., Gibson, J., and Harwood, C. S.

(2004) Complete genome sequence of the metabolically versatile photosynthetic bacterium Rhodopseudomonas palustris. Nat. Biotechnol. 22, 55–61.

(33) Hanada, S. (2016) Anoxygenic photosynthesis —A photochemical reaction that does not contribute to oxygen reproduction. Microbes Environ. 31, 1–3.

(34) Hayashi, H., Miyao, M., and Morita, S. (1982) Absorption and fluorescence spectra of light-harvesting bacteriochlorophyll-protein complexes from Rhodopseudomonas palustris in the near-infrared region. J. Biochem. (Tokyo) 91, 1017–1027.

(35) Mizoguchi, T., Isaji, M., Harada, J., and Tamiaki, H. (2012) Isolation and pigment composition of the reaction centers from purple photosynthetic bacterium

Rhodopseudomonas palustris species. Biochim. Biophys. Acta, Bioenerg. 1817, 395–400.

(36) Falkow, S., Rosenberg, E., Schleifer, K.-H., and Stackebrandt, E. (Eds.). (2006) The prokaryotes, Vol. 2: Ecophysiology and biochemistry. Springer Science & Business

Media, Berlin, Germany.

(37) Hartigan, N., Tharia, H. A., Sweeney, F., Lawless, A. M., and Papiz, M. Z. (2002)

The 7.5-Å electron density and spectroscopic properties of a novel low-light B800 LH2 from Rhodopseudomonas palustris. Biophys. J. 82, 963–977.

141 (38) Kuo, F.-S., Chien, Y.-H., and Chen, C.-J. (2012) Effects of light sources on growth and content of photosynthetic bacteria Rhodopseudomonas palustris.

Bioresour. Technol. 113, 315–318.

(39) Yoon, K.-S., Fukuda, K., Fujisawa, K., and Nishihara, H. (2011) Purification and characterization of a highly thermostable, oxygen-resistant, respiratory [NiFe]- hydrogenase from a marine, aerobic hydrogen-oxidizing bacterium Hydrogenovibrio marinus. Int. J. Hydrog. Energy 36, 7081–7088.

(40) Dean, J. A. (Ed.). (1999) Lange’s Handbook of Chemistry 15th ed. McGraw-Hill,

Inc., New York, NY.

(41) Paungfoo-Lonhienne, C., Rentsch, D., Robatzek, S., Webb, R. I., Sagulenko, E.,

Näsholm, T., Schmidt, S., and Lonhienne, T. G. A. (2010) Turning the table: Plants consume microbes as a source of nutrients. PLoS ONE.

142 CHAPTER 5

SUMMARY AND FUTURE DIRECTIONS

INTRODUCTION

Living cells have evolved novel mechanisms to catalyze chemical reactions.

Nitrogenase is an ideal enzyme for study, as it generates a highly reduced product (NH3) under mild conditions. Industry has developed processes to “mimic” biological reactions, but they tend to rely on fossil fuels and often require high temperature and pressure.1

Thus, it is essential to continue building a fundamental knowledge of biological systems to provide additional perspective and inspire more sustainable applications for energy conservation and production.

The research here revealed new observations into the mechanism of electron transfer to nitrogenase, including how low reduction potential reducing equivalents are generated as well as fresh insight into the nitrogenase catalytic cycle using a physiological reductant. Further, this research presented a potential biohybrid (bacterial-

+ electrode) system for NH4 production.

These studies lay a groundwork in three primary areas described below: 1)

Electron bifurcation by FixABCX, 2) Electron transfer to nitrogenase, and 3) Microbial

+ electrosynthesis for NH4 production.

ELECTRON BIFURCATION BY FIXABCX

Chapter 2 showcased the first biochemical evidence of electron bifurcation by

FixABCX to generate reduced flavodoxin or ferredoxin (Fld/Fd) for nitrogen fixation.2

While there have been considerable advances in uncovering processes related to the

143 generation of low reduction potential electrons in the context of electron bifurcation,3,4 they likely just scratch the surface of a world still largely unexplored.

Below, three broad areas of research on FixABCX are proposed, including control of electron transfer through conformational changes, tuning the reduction potential of the bifurcating flavin, and determining the role of FixFd in the FixABCX catalytic cycle.

Uncover conformational changes in FixABCX

An important aspect of electron bifurcation is controlling electron transfer to ensure that the two electrons from NADH are partitioned into both an exergonic and endergonic pathway (overall reaction thermodynamically favorable).5 One element of control could be conformational changes that occur in the enzyme upon substrate binding or upon release of the first bifurcated electron. A recent report on the bifurcating electron transfer flavoprotein/butryl-CoA dehydrogenase provides evidence of a domain rearrangement during electron bifurcation, but a direct role in electron transfer is still ambiguous.6 Other reports on bifurcating NADH-dependent reduced ferredoxin:NADP oxidoreductase (Nfn) complexes have also revealed conformational changes, albeit more subtle.7,8

Figure 5-1 overlays hypothetical electron transfer paths in the FixABCX structural model presented in Chapter 2. Two possible pathways of electron bifurcation are shown. The first pathway denoted by the orange arrows illustrates electron transfer distances from cofactor to cofactor. In this pathway, distances from the a-FAD to b-FAD and b-FAD to c-FAD are >25 Å. The compatible distance for rapid electron transfer is 14

Å.9 Thus, this mode of electron transfer would rely on conformational changes during catalysis. The second pathway represented with black arrows focuses on conserved Tyr

144

C-Try277 11 Å B-Trp215

11 Å

12 Å

28 Å

4 Å FixB 25 Å b-FAD FixC c-FAD A-Tyr37

11 Å

FixA a-FAD 10 Å

FixX

2 [4Fe-4S] 10 Å 10 Å

Figure 5-1. FixABCX structural model illustrating possible electron transfer pathways. Orange arrows: Electron transfer distances from cofactor to cofactor. Black arrows: Electron transfer distances when conserved Trp and Tyr are considered in the electron transfer pathway. FixA=Blue; FixB= Tan; FixC= Green; FixX= red.

and Trp residues that appear to shorten electron transfer distances between cofactors and may also be important in the electron transfer process.

To examine the protein conformational changes in FixABCX, H/D exchange- mass spectrometry (HDX-MS) could be employed. This approach monitors the exchange of amide hydrogens with deuterium. The amount of exchange coupled to the rate can reveal substrate/ligand binding sites as well as conformational dynamics under a given set of conditions.8 For example, FixABCX could be analyzed in both the presence and absence of substrates to reveal changes in protein conformation. Further, the role of the

Trp and Tyr residues could be probed using site-directed mutagenesis.

145 Once a conformational change is confirmed, additional experiments exploring the roles of conformation in electron transfer could be conducted using substrate analogs and/or chemical crosslinking. This in conjunction with transient absorption spectroscopy

(TAS) could resolve rates of electron transfer, ultimately, enhancing understanding of the correlation of electron flow through FixABCX and conformational dynamics.

Elucidate factors that tune the properties of the bifurcating flavin

Mass spectrometry studies on FixABCX revealed the presence of three FADs, two in FixAB and one in FixCX (Chapter 2). However, upon release of flavin from the protein complex, only two FADs could be accounted for in the UV-vis absorption spectrum of the supernatant. The precipitate, however, retained a strong signal at ~425 nm (Figure 5-2). It was discovered that this signature has significant similarity to a signal in the literature identified as a covalently-bound 6-S-cysteinyl flavin.10

Figure 5-2. UV-vis spectrum of FixABCX with non-covalently bound cofactors removed. This spectrum was obtained by subtracting the released flavin spectrum from the spectrum of intact protein.

146 Using molecular modeling, it appears that the FixABCX from Azotobacter vinelandii contains a conserved Cys within 5.6 Å of the C6 position of the proposed bifurcating a-FAD (Figure 5-3). Interestingly, this particular Cys residue is conserved in all diazotrophic bifurcating electron-transfer flavoproteins (ETFs).11 If FixABCX does indeed have a covalently-bound FAD in the bifurcating site, it would be the first known case. The fact that the redox potential of flavins can be modified by the protein environment has been known for a long time,12 thus, we hypothesize that a flavin derivative may tune the redox properties of the flavin to control electron transfer events at the bifurcating site.

In the absence of a crystal structure, evidence of a covalently-bound FAD could be obtained using Raman spectroscopy. Raman is chemical fingerprinting tool that probes

5.6 Å

Figure 5-3. FixABCX structural model illustrating a conserved Cys residue on FixB (tan) within 5.6 Å of the C6 position of the proposed bifurcating FAD in FixA (blue). This Cys may contribute to the formation of a covalent bond between the cofactor and protein.

147 molecular vibrations, where the observed vibrational frequencies represent specific functional groups or groups of bonds. The Raman frequencies of C-S bonds are 580-704 cm-1.13 Using known Raman frequencies as a guide, a fingerprint of FixABCX could be generated and compared to known covalently-bound flavins. Evidence of the C-S bond should be absent from non-covalently bound FAD control samples.

Further, to test our hypothesis that a flavin derivative tunes the cofactor for bifurcation, the Cys could be converted to Ala using site-directed mutagenesis. Assuming the variant binds FAD non-covalently, a suite of biochemical experiments could be conducted to compare the WT and variant proteins. For example, protein activities using diaphorase and bifurcating assays would reveal if the variant is impaired in both NADH oxidation activity, bifurcation activity, or both. Additionally, TAS could be employed to assess the lifetimes of flavin intermediates. Recently, bifurcating flavin sites have been shown to display crossed-potentials, in which the EmSq/Hq is lower than the EmOx/Sq couple.14,15 This means the high-energy semiquinone species does not accumulate, therefore, exhibiting a shorter lifetime detectable by TAS. A flavin not tuned for bifurcation would be expected to have uncrossed potentials and a semiquinone with a longer-lifetime. Finally, a comparison of the reduction potentials of the non-covalently versus covalently bound flavin could be determined using spectroelectrochemistry.

Determine the role of FixFd in the FixABCX catalytic cycle

In addition to genes fixA, fixB, fixC, and fixX, the fix operon of A. vinelandii harbors, two additional genes, fixFd (sometimes referred to as fixP) and Avin_10560

(ORF6). These genes, not commonly found in fix operons,16 open doors for investigation into their characterization and cellular roles.

148 Based on sequence analysis, ORF6 has been designated as “ferritin-like.”

Although this gene product has never been examined, it may serve as an iron reserve for the iron-sulfur containing components of Fix.17 On the other hand, an initial characterization of FixFd was conducted over twenty years ago; however, a direct cellular role of the protein was never discovered. Reyntjens et al. determined FixFd is a

7.2 kDa, 2[4Fe-4S]-containing Fd with the reduction potential of both clusters at -440 mV vs. SHE, a potential with enough driving force to send electrons to nitrogenase.18.

With new understanding of the role of FixABCX in electron transfer to nitrogenase, a study focused on determining the role of FixFd seems a clear direction.

The initial study by Reyntjens purified FixFd from A. vinelandii at its native expression levels, a monumental task as they acquired only ~3 mg FixP per 500 g cell paste.18 In an effort to increase protein yield, we have overexpressed fixP (no tag) in E. coli, and are currently optimizing a purification protocol and ensuring the integrity of the protein (e.g. insertion of the FeS clusters). Once purified to homogeneity, bifurcation assays with

FixABCX could be conducted to determine if FixP can serve as the low-potential acceptor. Furthermore, the bifurcation assay could be coupled to nitrogenase (Figure 5-

4). This would demonstrate, in vitro, if a direct electron transfer chain can be formed between FixABCX, FixP, and nitrogenase. This study on FixP would also offer an opportunity to compare the electron transfer kinetics using NifF Fld (from initial study) and FixP as low-potential acceptors in the bifurcation reaction.

149 NADH CoQ 1 e-

2 e- FixABCX

- NAD+ 1 e FixP Nitrogenase

Monitor acetylene and/or H+ reduction

Figure 5-4. Scheme for coupling the FixABCX electron bifurcation assay to nitrogenase activity.

ELECTRON TRANSFER TO NITROGENASE

Because artificial electron donors such as dithionite (DT) easily reduce nitrogenase, studies focused on physiological electron donors have been limited. Chapter

3 offered insight into electron transfer events taking place in vivo by investigating the nitrogenase catalytic cycle using the physiological reductant Fld and comparing it to DT.

The study revealed the true rate-limiting step of nitrogenase catalysis as events associated with Pi release rather than those with dissociation of the Fe protein from MoFe protein as previously determined with DT.19 This shed light on the importance of investigating native electron donors to nitrogenase. Further, the study on FixABCX (Chapter 2) provided compelling reasons to better understand electron transfer events taking place inside the cell. Included below are two areas of continued investigation into electron transfer to nitrogenase. They include determining the Flds and Fds used for nitrogen fixation in A. vinelandii and investigating the nitrogenase catalytic cycle in vitro using

Fd, rather than Fld, as an electron donor.

150 Identify the flavodoxins and ferredoxins critical for nitrogenase activity in A. vinelandii

Among diazotrophs, there are diverse routes for supplying electrons to nitrogenase.20 Klebsiella pneumonia uses a single Fld electron carrier;21,22 whereas, other microbes like A. vinelandii possess several Flds and Fds (Table 1-1).18,20 Among the electron carrying Flds and Fds, NifF Fld is the only one that has been clearly linked to molybdenum nitrogenase activity in vivo in A. vinelandii.23 Its disruption, however, only results in a 30% decrease in nitrogenase activity suggesting that other electron donors either co-function with NifF or substitute when NifF is unavailable. In addition to NifF, there are three other Fds that seems like logical sources of investigation: 1) FdxN plays a significant role in nitrogen fixation; however, a direct role in electron transfer is not as clear. Disruption of the fdxN gene results in a 60-65% decrease in acetylene reduction both in vivo and in vitro.24 FdxN appears to serve a primary function in cofactor biosynthesis;24,25 however, a role in electron transfer as well cannot be precluded. 2) A role for the Fix-linked FixFd has yet to be established; although, it seems likely to be the primary electron acceptor from the Fix system, since its gene is in the same operon.18 3)

While fdxA, which encodes for FdI, does not seem to be integral for nitrogen-fixation processes in vivo,23 further studies into a possible substitution role should be considered.

While a complex array of Flds and Fds exist in A. vinelandii, future studies could focus on the construction and characterization of A. vinelandii Fld and Fd deletion strains by incorporating antibiotic resistance cassettes into the coding regions. Emphasis could be placed on generating the following strains: ΔnifFΔfdxN, ΔnifFΔfixFd,

ΔnifFΔfdxNΔfixFd. Growth studies measuring the optical density (OD600) could be

151 conducted to determine generation times of WT vs deletion strains grown under both non-nitrogen-fixing and nitrogen-fixing conditions. Further, whole cell acetylene reduction assays using gas chromatography could be performed to measure nitrogenase activity. Additional deletions could be made with fdxA if nitrogenase is still active.

It is expected that the ΔnifFΔfdxNΔfixFd strain would significantly decrease, if not abolish nitrogenase activity. One important aspect to recognize in a study such as this is that Fld or Fd deletions may render a significantly impaired or lethal phenotype as these small redox proteins may play multiple roles in cellular processes; however, for A. vinelandii, it has already been established that individual disruptions in nifF, fdxN, and fdxA are not lethal.23,24 Armed with this information, it is expected that this systematic approach will aid in deducing probable electron donors to nitrogenase in vivo. Further, this research could be expanded upon by also including how these deletions (and others) affect the alternative vanadium (V) and iron-only (Fe-only) nitrogenases.

Investigate Fd as an electron donor to nitrogenase in vitro

Upon establishing which Fds can serve as electron donors to nitrogenase in vivo

(see section above), additional in vitro characterization of their electron transfer capabilities and interactions with Fe protein could be determined. Chapter 3 demonstrated that the observed rate constant of electron transfer is significantly greater in the presence of the physiological reductant Fld compared to DT.19 Although, similar results might be expected with alternative physiological electron donors, it seems plausible that they could exhibit different rates and affinities for Fe protein, especially if they only serve under conditions in which the primary electron donor is not available.

Pre-steady state kinetic experiments using stopped-flow spectroscopy could be done to

152 determine rate constants for electron transfer. Further, interactions between Fe protein and Fd(s) could be explored using limited proteolysis in conjunction with LC/MS. A previous study, focused on the interaction between Fe protein and Fld, revealed nucleotide-induced structural changes in the Fe protein altered its interaction with Fld.26

A similar study with Fd would garner deeper insight into the influence of protein-protein interactions on electron transfer processes.

+ MICROBIAL ELECTROSYNTHESIS FOR NH4 PRODUCTION

In Chapter 4, a new approach was presented for the production of fixed N, in which phototrophic, nitrogen-fixing bacteria are grown in association with electrodes.

The goal of this project is to eventually achieve both CO2 and N2 fixation, with electrons supplied from water oxidation and the energy provided by light. Thus far, it was demonstrated that the phototrophic bacterium, Rhodopseudomonas palustris, can be cultivated exclusively with IR light (~850 nm) and achieve CO2 reduction under electrochemical conditions. Further, it was shown that R. palustris can grow non- electrochemically using H2 as a source of electrons, CO2 as a carbon source, and N2 as a nitrogen source. This proof-of-concept paves the way for achieving N2 reduction in conjunction with CO2 reduction under electrochemical conditions. If successful, this would offer an alternative route for the microbial electrosynthesis of “fertilizer” for

+ sustainable food production with opportunities to optimize NH4 production. Further, experiments directed toward bacterial uptake of electrons directly from the electrode as well as identification of other value-added products, such as hydrocarbons, also offer new avenues of study.

153 + Optimize NH4 production

In addition to using R. palustris biomass as a fertilizer, it could be worth

+ investigating the potential for NH4 release into the extracellular medium. Previous reports in the diazotrophs Rhodospirillum rubrum and Xanthomonas autotrophicus used small molecules to inhibit glutamine synthetase in the ammonium assimilation

27,28 + pathway. Under these conditions NH4 was released from the cells. Similar inhibition studies could be applied to glutamine synthetase in R. palustris.29

+ Another approach to consider is disrupting NH4 transporters in the cells.

Previous studies on AmtB in A. vinelandii and Pseudomonas stutzeri suggested that it

+ 30,31 likely minimizes loss of NH4 in the environment. Further, when disrupted, the cells displayed no growth defects; thus, the homologous transporter genes in R. palustris likely provide a viable target.

+ We hypothesize that disrupting NH4 regulatory and uptake pathways would result in the release of fixed nitrogen by R. palustris. A comparison between electrochemical and non-electrochemical growths as well as combining the glutamate synthetase inhibitor with the transporter disruption(s) could be done and the amount of

+ NH4 in the medium quantified over time. Overall, this study would target specific proteins and begin to identify optimal conditions under which to generate extracellular

+ NH4 .

Grow cells with electrons direct from the electrode

Our initial work growing R. palustris electrochemically utilized mediated electron

+ transfer (MET), in which electrons from water oxidation were used to reduce H to H2.

The cells then used the H2 as a source of electrons. The next step could be to use a direct

154 electron transfer (DET) approach, in which the electrons are supplied to the cells direct from the electrode (Figure 5-5). In comparison to MET, DET offers several advantages. For example, the electrolysis in MET results in significant pH changes in the cathode and anode, which must be adjusted regularly. Further, the MET system must be open so H2 does not build up; thus, the electron donor is being “wasted.” DET would help eliminate these disadvantages.

A recent study by Bose et al. did in fact demonstrate the ability of R. palustris to

32 fix CO2 using electrons from an electrode. Given that success, it is expected that having cells fix both carbon and nitrogen would be feasible. Cells could be cultured using graphite electrodes with CO2 as the carbon source, N2 as the nitrogen source, the

Mediated electron transfer Direct electron transfer

CO2 CO2 + N N2 2 H 2 - e- e

Electrode Fixed C Fixed C H2 NH3 NH3

Figure 5-5. Mediated vs. direct electron transfer from an electrode to a bacterium.

155 electrode as electron source, and light as energy source. If this works, it would be the first demonstration of light-driven carbon and nitrogen fixation by whole cells using electrons from an electrode. The previous work by Bose et al. poised electrodes at +100 mV vs. SHE, because it is the potential of ferrous iron (Fe(II)), which R. palustris can use as an extracellular electron donor.32 An extension of this could be testing a wide range of potentials and examining the affect on cellular growth. This work on extracellular electron uptake could then lead to additional studies on genetic determinants facilitating this ability as well as additional insight into bio-based applications for valuable products.

Identify other value-added products

It has been demonstrated that variants of nitrogenase as well as the wild-type Fe- only nitrogenase can reduce CO2 to hydrocarbons (e.g. CH4, formate, C2 and C3 products).33,34 This observation encourages studies into other hydrocarbon products R. palustris might produce when grown electrochemically. A profile of product formation with cells expressing Mo, V, and Fe-only nitrogenases could be generated. It would be

+ expected that the Mo nitrogenase-expressing cells would produce NH4 and H2, while the

Fe nitrogenase-expressing cells, would produce hydrocarbons, such as CH4 in addition to

+ NH4 and H2. Variant forms of nitrogenase capable of CO2 chemistry could be explored as well. Finally, investigation into whether product distribution could be manipulated by varying the potential of electrons would be extremely valuable as we search for ways to optimize products of interest.

CONCLUSION

This research focused on one of the most important and elegant biological processes on the planet—nitrogen fixation. The beauty of science is that as we discover

156 more answers, it brings up more questions. The cycle never ends. Reflecting on the work here, my hope is that the contributions are a catalyst as scientists continue their quest for understanding and capitalizing on the intricate biological processes that shape us.

ACKNOWLEDGMENTS

Thank you to Anne-Frances Miller for sharing her discovery of the similarity of the FixABCX UV-vis signature to a characterized covalently bound flavin and Cara

Lubner for her discussions and insight on the next generation of FixABCX experiments.

REFERENCES

(1) Smil, V. (2001) Enriching the earth: Fritz Haber, Carl Bosch, and the transformation of world food production. MIT Press, Cambridge, Mass.

(2) Ledbetter, R. N., Garcia Costas, A. M., Lubner, C. E., Mulder, D. W., Tokmina-

Lukaszewska, M., Artz, J. H., Patterson, A., Magnuson, T. S., Jay, Z. J., Duan, H. D.,

Miller, J., Plunkett, M. H., Hoben, J. P., Barney, B. M., Carlson, R. P., Miller, A.-F.,

Bothner, B., King, P. W., Peters, J. W., and Seefeldt, L. C. (2017) The electron bifurcating FixABCX protein complex from Azotobacter vinelandii: Generation of low- potential reducing equivalents for nitrogenase catalysis. Biochemistry 56, 4177–4190.

(3) Buckel, W., and Thauer, R. K. (2013) Energy conservation via electron bifurcating ferredoxin reduction and proton/Na+ translocating ferredoxin oxidation. Biochim.

Biophys. Acta, Bioenerg. 1827, 94–113.

(4) Thauer, R. K., and Buckel, W. (2018) Flavin-based electron bifurcation, ferredoxin, flavodoxin, and anaerobic respiration with protons (Ech) or NAD+ (Rnf) as electron acceptors: A historical review. Front. Microbiol. 9.

157 (5) Peters, J. W., Miller, A. F., Jones, A. K., King, P. W., and Adams, M. W. (2016)

Electron bifurcation. Curr. Opin. Chem. Biol. 31, 146–152.

(6) Demmer, J. K., Pal Chowdhury, N., Selmer, T., Ermler, U., and Buckel, W. (2017)

The semiquinone swing in the bifurcating electron transferring flavoprotein/butyryl-CoA dehydrogenase complex from Clostridium difficile. Nat. Commun. 8, 1577.

(7) Demmer, J. K., Rupprecht, F. A., Eisinger, M. L., Ermler, U., and Langer, J. D.

(2016) Ligand binding and conformational dynamics in a flavin-based electron- bifurcating enzyme complex revealed by Hydrogen-Deuterium Exchange Mass

Spectrometry. FEBS Lett. 590, 4472–4479.

(8) Berry, L., Poudel, S., Tokmina-Lukaszewska, M., Colman, D. R., Nguyen, D. M. N.,

Schut, G. J., Adams, M. W. W., Peters, J. W., Boyd, E. S., and Bothner, B. (2018) H/D exchange mass spectrometry and statistical coupling analysis reveal a role for allostery in a ferredoxin-dependent bifurcating transhydrogenase catalytic cycle. Biochim. Biophys.

Acta, Gen. Subj. 1862, 9–17.

(9) Page, C. C., Moser, C. C., Chen, X., and Dutton, P. L. (1999) Natural engineering principles of electron tunnelling in biological oxidation-reduction. Nature 402, 47–52.

(10) Fujieda, N., Satoh, A., Tsuse, N., Kano, K., and Ikeda, T. (2004) 6-S-cysteinyl flavin mononucleotide-containing histamine dehydrogenase from Nocardioides simplex:

Molecular cloning, sequencing, overexpression, and characterization of redox centers of enzyme. Biochemistry 43, 10800–10808.

(11) Garcia Costas, A., Poudel, S., Miller, A. F., Schut, G., Ledbetter, R., Fixen, K.,

Seefeldt, L. C., Adams, M., Harwood, C., Boyd, E., and Peters, J. (2017) Defining

158 electron bifurcation in the electron-transferring flavoprotein family. J. Bacteriol. 199, e00440-17.

(12) Heuts, D. P. H. M., Scrutton, N. S., McIntire, W. S., and Fraaije, M. W. (2009)

What’s in a covalent bond? FEBS J. 276, 3405–3427.

(13) McCreery, R. L. (2005) Raman spectroscopy for chemical analysis (Winefordner, J.

D., Ed.). Wiley-Blackwell.

(14) Lubner, C. E., Jennings, D. P., Mulder, D. W., Schut, G. J., Zadvornyy, O. A.,

Hoben, J., Tokmina-Lukaszewska, M., Berry, L., Nguyen, D., Lipscomb, G. L., Bothner,

B., Jones, A. K., Miller, A. F., King, P. W., Adams, M. W. W., and Peters, J. W. (2017)

Mechanistic insights into energy conservation by flavin-based electron bifurcation. Nat.

Chem. Biol. 13, 655–659.

(15) Nitschke, W., and Russell, M. J. (2012) Redox bifurcations: Mechanisms and importance to life now, and at its origin. BioEssays 34, 106–109.

(16) Markowitz, V. M., Chen, I. M. A., Palaniappan, K., Chu, K., Szeto, E., Grechkin, Y.,

Ratner, A., Jacob, B., Huang, J., Williams, P., Huntemann, M., Anderson, I., Mavromatis,

K., Ivanova, N. N., and Kyrpides, N. C. (2012) IMG: the Integrated Microbial Genomes database and comparative analysis system. Nucleic Acids Res. 40, D115-122.

(17) Smith, J. L. (2004) The physiological role of ferritin-like compounds in bacteria.

Crit. Rev. Microbiol. 30, 173–185.

(18) Reyntjens, B., Jollie, D. R., Stephens, P. J., Gao-Sheridan, H. S., and Burgess, B. K.

(1997) Purification and characterization of a fixABCX-linked 2[4Fe-4S] ferredoxin from

Azotobacter vinelandii. J. Biol. Inorg. Chem. 2, 595–602.

159 (19) Yang, Z. Y., Ledbetter, R., Shaw, S., Pence, N., Tokmina-Lukaszewska, M.,

Eilers, B., Guo, Q., Pokhrel, N., Cash, V. L., Dean, D. R., Antony, E., Bothner, B.,

Peters, J. W., and Seefeldt, L. C. (2016) Evidence that the Pi release event is the rate- limiting step in the nitrogenase catalytic cycle. Biochemistry 55, 3625–3635.

(20) Saeki, K. (2004) Electron transport to nitrogenase: Diverse routes for a common destination, in Genetics and regulation of nitrogen fixation in free-living bacteria, pp

257–290. Kluwer Academic Publishers.

(21) Roberts, G. P., MacNeil, T., MacNeil, D., and Brill, W. J. (1978) Regulation and characterization of protein products coded by the nif (nitrogen fixation) genes of

Klebsiella pneumoniae. J. Bacteriol. 136, 267–279.

(22) John, R. T. S., Johnston, H. M., Seidman, C., Garfinkel, D., Gordon, J. K., Shah, V.

K., and Brill, W. J. (1975) Biochemistry and genetics of Klebsiella pneumoniae mutant strains unable to fix N2. J. Bacteriol. 121, 759–765.

(23) Martin, A. E., Burgess, B. K., Iismaa, S. E., Smartt, C. T., Jacobson, M. R., and

Dean, D. R. (1989) Construction and characterization of an Azotobacter vinelandii strain with mutations in the genes encoding flavodoxin and ferredoxin I. J. Bacteriol. 171,

3162–3167.

(24) Rodriguez-Quinones, F., Bosch, R., and Imperial, J. (1993) Expression of the nifBfdxNnifOQ region of Azotobacter vinelandii and its role in nitrogenase activity. J.

Bacteriol. 175, 2926–2935.

(25) Jiménez-Vicente, E., Navarro-Rodríguez, M., Poza-Carrión, C., and Rubio, L. M.

(2014) Role of Azotobacter vinelandii FdxN in FeMo-co biosynthesis. FEBS Lett. 588,

512–516.

160 (26) Pence, N., Tokmina-Lukaszewska, M., Yang, Z.-Y., Ledbetter, R. N., Seefeldt,

L. C., Bothner, B., and Peters, J. W. (2017) Unraveling the interactions of the physiological reductant flavodoxin with the different conformations of the Fe protein in the nitrogenase cycle. J. Biol. Chem. 292, 15661–15669.

(27) Weare, N. M., and Shanmugam, K. T. (1976) Photoproduction of ammonium ion from N2 in Rhodospirillum rubrum. Arch. Microbiol. 110, 207–213.

(28) Liu, C., Sakimoto, K. K., Colón, B. C., Silver, P. A., and Nocera, D. G. (2017)

Ambient nitrogen reduction cycle using a hybrid inorganic–biological system. Proc. Natl.

Acad. Sci. 114, 6450–6455.

(29) Arp, D. J., and Zumft, W. G. (1983) L-methionine-SR-sulfoximine as a probe for the role of glutamine synthetase in nitrogenase switch-off by ammonia and glutamine in

Rhodopseudomonas palustris. Arch. Microbiol. 134, 17–22.

(30) Barney, B. M., Eberhart, L. J., Ohlert, J. M., Knutson, C. M., and Plunkett, M. H.

(2015) Gene deletions resulting in increased nitrogen release by Azotobacter vinelandii:

Application of a novel nitrogen biosensor. Appl. Environ. Microbiol. AEM.00554-15.

(31) Zhang, T., Yan, Y., He, S., Ping, S., Alam, K. M., Han, Y., Liu, X., Lu, W., Zhang,

W., Chen, M., Xiang, W., Wang, X., and Lin, M. (2012) Involvement of the ammonium transporter AmtB in nitrogenase regulation and ammonium excretion in Pseudomonas stutzeriA1501. Res. Microbiol. 163, 332–339.

(32) Bose, A., Gardel, E. J., Vidoudez, C., Parra, E. A., and Girguis, P. R. (2014)

Electron uptake by iron-oxidizing phototrophic bacteria. Nat. Commun. 5, 3391.

161 (33) Fixen, K. R., Zheng, Y., Harris, D. F., Shaw, S., Yang, Z.-Y., Dean, D. R.,

Seefeldt, L. C., and Harwood, C. S. (2016) Light-driven carbon dioxide reduction to methane by nitrogenase in a photosynthetic bacterium. Proc. Natl. Acad. Sci. 113,

10163–10167.

(34) Zheng, Y., Harris, D., Fu, Y., Yu, Z., Poudel, S., Ledbetter, R. N., Fixen, K. R.,

Yang, Z. Y., Boyd, E. S., Lidstrom, M. E., Seefeldt, L. C., and Harwood, C. S. A pathway for biological methane production using bacterial iron-only nitrogenase. Nat.

Microbiol.

162

APPENDIX B

FULL MANUSCRIPT

EVIDENCE THAT THE Pi RELEASE EVENT IS THE RATE LIMITING STEP

IN THE NITROGENASE CATALYTIC CYCLE

163 EVIDENCE THAT THE Pi RELEASE EVENT IS THE RATE LIMITING STEP IN

THE NITROGENASE CATALYTIC CYCLE

ABSTRACT

Nitrogenase reduction of dinitrogen (N2) to ammonia (NH3) involves a sequence

of events that occur upon the transient association of the reduced Fe protein containing

two ATP with the MoFe protein that includes electron transfer, ATP hydrolysis, Pi

release, and dissociation of the oxidized, ADP containing Fe protein from the reduced

MoFe protein. Numerous kinetic studies using the non-physiological electron donor

dithionite have suggested that the rate limiting step in this reaction cycle is the

dissociation of the Fe protein from the MoFe protein. Here, we have established the rate

constants for each of the key steps in the catalytic cycle using the physiological reductant

flavodoxin protein in its hydroquinone state. The findings indicate that with this

reductant, the rate limiting step in the reaction cycle is not protein-protein dissociation or

reduction of the oxidized Fe protein, but rather events associated with the Pi release step.

Further, it is demonstrated that: (i) Fe protein transfers only one electron to MoFe protein

in each Fe protein cycle coupled with hydrolysis of two ATP molecules; (ii) the oxidized

Fe protein is not reduced when bound to MoFe protein; (iii) the Fe protein interacts with

______*Coauthored by Zhi-Yong Yang, Rhesa Ledbetter, Sudipta Shaw, Natasha Pence, Monika Tokmina-Lukaszewska, Brian Eilers, Qingjuan Guo, Nilisha Pokhrel, Valerie L. Cash, Dennis R. Dean, Edwin Antony, Brian Bothner, John W. Peters, and Lance C. Seefeldt (2016) Biochemistry, 55:3625-3635. Copyright 2016, American Chemical Society. Reprinted with permission.

164 flavodoxin using the same binding interface that is used with the MoFe protein. These findings allow a revision of the rate limiting step in the nitrogenase Fe protein cycle.

INTRODUCTION

Nitrogenase is the catalyst responsible for biological nitrogen fixation, the reduction

1–3 of dinitrogen (N2) to ammonia (NH3). The molybdenum-dependent nitrogenase consists of two catalytic components called the MoFe protein and the Fe protein.4 The Fe protein is a dimer of two identical subunits connected by a single [4Fe-4S] cluster and is responsible for transferring a single electron to the MoFe protein.5,6 The MoFe protein is a heterotetramer, composed of two symmetric αβ units. Each αβ unit contains two unique metal clusters, the electron carrier P-cluster [8Fe-7S] and the active site FeMo-cofactor

[7Fe-9S-Mo-C-homocitrate].7–11 An Fe protein, with two bound MgATP molecules, binds transiently to each αβ unit of the MoFe protein during the electron transfer (ET) event (Figure B1).6,12,13 ET is initiated upon association of the Fe protein to the MoFe protein, followed by a conformationally gated one-electron transfer from the P-cluster to

FeMo-cofactor.14,15 This is followed by the one-electron transfer from the reduced Fe protein (Fered) [4Fe-4S] cluster to the oxidized P-cluster (P1+), in what has been called a

“deficit spending” ET process.14 Following ET, the two ATP molecules are hydrolyzed to two ADP and two Pi molecules. Next, the two Pi are released from the complex, followed by the dissociation of the oxidized Fe protein (Feox) with two bound MgADP

ox 16 ox (Fe (ADP)2) from the MoFe protein. The Fe (ADP)2 is then reduced by a reductant and the two ADP molecules are replaced by two ATP molecules.17,18 This cycle, often referred to as the Fe protein cycle, must be repeated a sufficient number of times to accumulate the electrons necessary for substrate reduction in the MoFe protein.1–3,18–20

165

Figure B1. Overview of nitrogenase catalysis with a focus on the Fe protein cycle showing dithionite (DT) and flavodoxin (FldHQ) as a non-physiological and physiological reductant, respectively.

Over the past five decades, most of the in vitro mechanistic studies of nitrogenase have been conducted with the non-physiological reductant dithionite (DT), largely because of ease of use.1,3,18 The kinetic parameters for each step in the Fe protein cycle using DT as reductant have been deduced and are summarized in the scheme shown in

Figure B2.16–18,21,22 It was concluded from a comparison of the rate constants that the

ox overall rate limiting step in the Fe protein cycle is the dissociation of the Fe (ADP)2 from the MoFe protein with a rate constant of ~6 s-1.16,17 A problem that is often

2- overlooked in such studies is the reversible dissociation of DT (S2O4 ) to generate the

•- actual reductant, the radical anion SO2 , with a Kd of ~1.5 nM and a slow rate constant of

~2 s-1.17 This leads to slow reduction of Feox, with rate constants for this reduction near

ox the rate constant that is reported for dissociation of Fe (ADP)2 from MoFe protein.16,17,22,23

166 [I]

[e-] 2ATP 2ADP ox red red [Fe (ADP)2] [Fe (ADP) ] [Fe (ATP) ] k = 6 s-1 2 fast 2 red red,DT MoFe MoFe MoFe MoFe

4

1

5

-

M .

M

s 4

1

-

-

-

[

1

1 

6

s

1 I

s

s

I

d

0

1

=

5

]

-

-

e 1

1 7

0

1

s

x

s 6

i

a

d

A

l

k

e

c

t

R

i

red ox red v

[MoFe ;Fe (ADP)2] [MoFe;Fe (ATP)2] a

]

t

e

V

I

d [ 2[Pi] -1 k = 140-200 s-1 k Pi = 16-22 s ET

-1 red ox k ATP = 50-70 s red ox [MoFe ;Fe (ADP,Pi)2] [MoFe ;Fe (ATP)2]

[III] Figure B2. Fe protein cycle with pseudo-first order kinetic rate constants for each step.

The physiological reductants of Fe protein are known to be the electron carrier proteins flavodoxin (Fld) and/or ferredoxin (Fd).24–30 Because nitrogenase is readily reduced by artificial electrons donors, such as DT, a limited number of kinetic studies using physiological reductants have been conducted.24,25,27,30–35 The diazotroph,

Azotobacter vinelandii, contains several Flds and Fds. NifF (Fld) has been implicated as an electron donor to nitrogenase26,29,30,36 and has been shown to transfer electrons to Fe protein in vitro.24,25,27,33,34 Further, disruption in the nifF gene results in a 30% decrease in whole-cell acetylene reduction activity. This previous research demonstrates that Fld is a major, but not the sole, electron donor to nitrogenase.26,29 Fld has three different redox states designated as the oxidized quinone (FldQ), one-electron reduced semiquinone

(FldSQ), and two-electron reduced hydroquinone (FldHQ).24,37,38 The pH- and temperature- dependent midpoint redox potentials (Em) for the two redox couples are estimated as -180

167 mV (vs the NHE) for the FldQ/FldSQ couple, and -480 mV for the FldSQ/FldHQ couple

(pH 8.5 and 22oC).38 The FldSQ/FldHQ redox couple has sufficient driving force for reduction of the oxidized Fe protein with a midpoint potential of -290 mV without nucleotide bound and -440 mV with ADP bound.25,27,33,34,39 Earlier kinetic studies using

Fld as reductant have yielded results that contradict some aspects of the Fe protein cycle deduced with DT: (i) the one-electron reduction of Fe protein by FldHQ is much faster

ox than that by DT, and this reduction event might happen when Fe (ADP)2 is still bound to the MoFe protein;25 (ii) Fe protein can be reduced by two electrons by FldHQ to an all ferrous state, possibly allowing two electrons to be transferred per two ATP hydrolyzed

(a ratio of 1ATP:1e-);33,34 and (iii) in the Fe protein cycle, the dissociation step might be faster than what has been widely accepted (~6 s-1).25,31–33,40

In the present work, kinetic studies were performed using both DT and Fld (NifF from A. vinelandii) as reductant to deduce the key kinetic parameters in the Fe protein cycle. This work reveals that FldHQ cannot reduce the Fe protein while it is bound to

MoFe protein, the ratio of ATP hydrolyzed per electron transferred remains at 2:1 for a wide range of substrates, and the rate limiting step in the Fe protein cycle is not the dissociation of the Fe protein from the MoFe protein, but rather events associated with Pi release.

MATERIALS AND METHODS

General procedures

All chemicals, unless otherwise noted, were obtained from Sigma-Aldrich (St.

Louis, MO) and used without further purification. Hydrogen, acetylene, ethylene, argon, and dinitrogen gases were purchased from Air Liquide America Specialty Gases LLC

168 (Plumsteadville, PA). The argon and dinitrogen gases were passed through an activated copper-catalyst to remove dioxygen contamination prior to use. A.vinelandii strains DJ995 (wild type MoFe protein) and DJ884 (wild type Fe protein) were grown, and nitrogenase proteins expressed and purified as previously described.41 Both proteins were greater than 95% pure as confirmed by SDS-PAGE analysis using Coomassie blue staining and fully active (see results and discussion). Proteins and buffers were handled anaerobically in septum-sealed serum vials under an inert atmosphere (argon or dinitrogen), on a Schlenk vacuum line, or anaerobic glove box (Teledyne Analytical

Instruments, MO-10-M, Hudson, NH). The transfer of gases and liquids were done with gastight syringes.

Strain construction and expression, Escherichia coli growth, and NifF purification

The nifF gene from A. vinelandii was PCR amplified and cloned into the NdeI-

BamHI sites of a T7-7 plasmid containing an ampicillin resistance gene for selection. Fld was overexpressed in Escherichia coli (E. coli) BL21 DE3 cells. The cells were grown to an optical density (600 nm) of 0.6-0.8 at 37 oC before the protein expression was induced at 32 oC by adding isopropyl β-D-1-thiogalactopyranoside (IPTG) and flavin mononucleotide (FMN) to a final concentration of 1 mM and 10 mg/L, respectively.

Following a 10 hour induction, cells were harvested.

All steps during the purification of Fld were conducted anaerobically under an argon atmosphere. 100 grams of cell paste were resuspended in 50 mM Tris, pH 8 with 2 mM dithiotheritol (DTT) at a biomass to buffer ratio of 1:5 (w/v). The resuspended cells were lysed in a french pressure cell (SLM Aminco FA-078, Rochester, NY) at 200 MPa.

The cell lysate was centrifuged (Sorvall Lynx 4000, ThermoScientific, Waltham, MA) at

169 48,000 x g for 30 minutes at 4°C. Following centrifugation, 0.5% (w/v) streptomycin sulfate was added to the supernatant to precipitate nucleic acids. The precipitation was removed via a second centrifugation as described above. The supernatant was loaded onto a 150 ml Q-sepharose column, which was first washed with two column volumes of

Buffer B (50 mM Tris, 1 M NaCl, 1 mM DTT, pH 8) and then equilibrated with two column volumes of Buffer A (50 mM Tris, 1 mM DTT, pH 8). A 15-75% salt gradient was run over 5 column volumes. Fractions containing Fld (blue color) were combined and diluted with Buffer A to a final NaCl concentration of less than 100 mM. FMN was added to a final concentration of 2 mM to increase the percentage of holo-Fld in the presence of 2 mM DT, which slowly changed the color of the protein solution from blue to yellow. The reconstitution was carried out for at least 2 hours at room temperature.

The protein was then loaded onto a Q-sepharose column (~30 mL) pre-reduced and equilibrated with Buffer B and Buffer A containing 1 mM DT. After loading, the column was washed with 1 column volume of Buffer A and eluted with 100% Buffer B as a concentrated fraction for loading onto Sephacryl-200 column (~600 ml) equilibrated with an 100 mM HEPES (pH 7.8) with 150 mM NaCl, 0.5 mM DTT, and 1 mM DT. The Fld was followed as a yellow-green band. The protein was concentrated using an Amicon

(EMD Millipore, Billerica, MA) concentrator with a 10,000 kDa cutoff membrane and stored in liquid nitrogen. The purity of the Fld was greater than 95% based on SDS-

PAGE method described above. The FMN content was determined to be ~70% of the total protein. This was determined by measuring the absorbance of FldQ at 452 nm and using an extinction coefficient of 11.3 mM-1 cm-1.37 The Fld concentration used in this work refers to the concentration of the holo-Fld with FMN bound.

170 Steady-state substrate reduction assays for determination of kcat for Fe protein and

MoFe protein

Substrate reduction assays were carried out in a 9.4-mL sealed serum vials with a liquid volume of 1 mL in an assay buffer consisting of an MgATP regeneration system (5 mM MgCl2, 22 mM phosphocreatine, 4 mM ATP, 0.2 mg/mL creatine phosphokinase, 1 mg/mL BSA) in 100 mM MOPS buffer (pH 7.3) with 10 mM DT or 600 μM FldHQ-10 to

12 mM DT mixture. After solutions were made anaerobic, the headspace in the reaction vials was adjusted to proper partial pressures for the various gaseous substrates, such as acetylene and N2. The MoFe protein was then added to the designated final concentration. Each reaction vial was pre-incubated in a 30°C water bath for 1 minute before initiation of the reaction by the addition of Fe protein. Reactions were incubated for 30 s at 30°C with a shaking rate of 130 rpm before being quenched by the addition of

500 μL of 400 mM EDTA at pH 8.0 or 500 μL of 1 M formic acid. The products (H2,

C2H4, and NH3) from different substrate reduction assays were quantified according to published methods.42 Refer to substrate reduction tables and figures for detailed information on assay conditions and concentrations.

Determination of ATP/e- values for different substrate reduction reactions

The total number of electrons transferred for product formation (H2, C2H4, and

NH3) from different substrate reduction reactions was determined. The reaction mixture contained 8.5 mM ATP, 9.7 mM MgCl2, and 1 mg BSA in 100 mM MOPS buffer (pH

7.3) with either 10 mM DT or 712 μM FldHQ-10 mM DT. General substrate reduction assays procedures are described above. To quantify the amount of hydrolyzed ATP as formation of ADP, a trace amount of [α-32P]ATP with final concentration of about 0.006

171 µM was added to the reaction mixture. The reactions were then initiated by addition of a mixture of MoFe and Fe protein at a designated molar ratio. An aliquot of 25 μL was taken out of the reaction and quenched with 50 μL of 1 M formic acid. 1 μL of the quenched mixture was spotted onto a silicon-gel thin-layer chromatography (TLC) plate and developed in 0.6 M potassium phosphate buffer, pH 3.4 for 70 min. The plate was dried and exposed overnight to a phosphor screen. The [α-32P]ATP and the [α-32P]ADP were detected with a Storm PhosphorImager and quantified using the ImageQuant software (Molecular Dynamics). The amount of hydrolyzed ATP or produced ADP were quantified based on the density ratio of [α-32P]ATP and [α-32P]ADP spots from each experiment after subtracting no protein controls.

Stopped-flow (SF) spectrophotometry and reduction of Fe protein by DT and FldHQ

SF spectrophotometry was conducted using an AutoSF-120 stopped-flow instrument equipped with a data acquisition system (KinTek Corp., Snow Shoe, PA). The change in absorbance was monitored at 426 nm over time. This wavelength detects a decrease in absorbance as the Fe protein [4Fe-4S] cluster becomes reduced. 426 nm rather than 430 nm was chosen since it is the isosbestic point of Fld (Figure B6).

Reactions were conducted at 4°C, unless otherwise stated, with a mixing ratio of 1:1. All samples were prepared in 100 mM MOPS buffer (pH 7.3). Fe protein was oxidized in an anaerobic glove box using ~3-fold molar excess of phenazine methosulfate (PMS). The protein-PMS mixture was allowed to incubate for 15 minutes prior to separation with

Sephadex G-25 equilibrated with 100 mM MOPS, pH 7.3 with 150 mM NaCl. The oxidized Fe protein was contained in one drive syringe with or without nucleotide. The other drive syringe contained the electron donor, DT or FldHQ-DT mixture, with or

172 without nucleotide. All reagents were used at the following final concentrations and kept under an argon atmosphere: 40 μM oxidized Fe protein, 10 mM DT, 300 μM Fld, 10 mM ATP or ADP, and 10 mM MgCl2 (only used in the presence of nucleotides). For the reduction of Feox by DT, the data were fit to a single exponential equation to obtain the pseudo-first order rate constants. Due to the presence of both FldHQ and DT in the experiments, the pseudo-first order rate constants for reduction of Feox by FldHQ were obtained by fitting the data to a double exponential equation.

ox Reduction of Fe (ADP)2 by different reductants in the presence of MoFe protein

The Fe-MoFe dissociation rate constant was determined using SF as described above except that the temperature was 25°C. Fe protein was oxidized (see details above) and MoFe protein was stripped of DT in an anaerobic glove box. DT was removed from

MoFe protein using a DOWEX-Sephadex G-25 column equilibrated with 100 mM

MOPS, pH 7.3 containing 150 mM NaCl. One drive syringe contained 80 μM Feox, 80

μM MoFe protein, and 9 mM MgADP. The other syringe contained the reductant mixture: 1) 20 mM DT, 2) 100 μM methyl viologen (MV)-20 mM DT mixture, or 3) 400

μM FldHQ-20 mM DT mixture. All reductants contained 9 mM MgADP. DT data were fit to a single exponential decay equation, and MV-DT and FldHQ-DT data were fit to a double exponential decay equation.

Primary electron transfer

Primary electron transfer from the Fe protein to the MoFe protein in the presence of DT and Fld was measured at 25°C using SF spectrophotometry as described above. All mixtures were prepared in 100 mM MOPS, pH 7.3 and were kept under an argon atmosphere. One syringe contained 80 µM Fe protein, 20 µM MoFe protein, 1 mM DT

173 with or without 200 µM FldHQ. The other syringe was loaded with 1 mM DT and 20

red mM MgATP. As turnover occurred in the [Fe (MgATP)2-MoFe] complex, the oxidation of the [4Fe-4S] cluster of the Fe protein was monitored by an increase in absorbance at

426 nm. Data were fit to a single exponential curve.

Quench-flow studies for ATP hydrolysis

Pre-steady state ATP hydrolysis experiments were performed using a rapid chemical quench-flow (KinTek, Snow Shoe, PA) in a Coy chamber (Grass Lake, MI) under a dinitrogen atmosphere. Mixtures were prepared in 50 mM MOPS, pH 7.4.

Syringe A contained 80 μM Fe and 20 μM MoFe proteins with either 10 mM DT or 500

HQ µM Fld -10 mM DT mixture. Syringe B contained 6 mM ATP, 8 mM MgCl2 with [α-

32P]ATP. 18 µl from syringe A was mixed with 18 µl from syringe B and then rapidly quenched with 45 μL of 0.75 M formic acid contained in syringe C. Aliquots (1 µl) of the quenched mixture were plated on TLC plates and the ratio of [α-32P]ATP vs [α-32P]ADP formed were analyzed as described above for the steady state ATP hydrolysis experiments.

Real-time measurement of inorganic phosphate (Pi) release

Pi release was determined in a stopped-flow (SF) fluorometer (Auto SF-120,

KinTek Corp.) using a coumarin (N-[2-(1-maleimidyl)ethyl]-7-(diethylamino) coumarin-

3-carboxamide) labeled phosphate binding protein (MDCC-PBP) assay.43 Pi was quantified from binding MDCC-PBP that was monitored by an increase influorescence

(λ-excitation = 430 nm, λ-emission >450 nm), using a standard curve generated with

43 KH2PO4 as described. The experiments were carried out at 25°C in 25 mM HEPES, pH

7.4 containing 1 mM DT. The SF syringes and flow lines were treated with a Pi mop (25

174 mM HEPES buffer at pH 7.4 with 320 µM 7-methylguanine (7-meG), and 0.12 U/mL purine nucleoside phosphorylase (PNPase)) before each experiment for 45 minutes to remove contaminating Pi and then rinsed with DT reduced buffer. The same concentration of the Pi mop system was also added to the reaction mixtures. 1 µM MoFe and 16 µM Fe were rapidly mixed with a solution of 25 µM MDCC-PBP, 7.5 mM

MgCl2 and 6 mM ATP and the change in fluorescence was monitored over time. 100 µM

FldHQ was added to the MoFe and Fe protein mixture to monitor the effect of FldHQ on Pi release.

In silico docking study of protein-protein interactions

In silico protein-protein docking simulations were performed using the computational docking program ClusPro 2.0. ClusPro 2.0 uses PIPER, a Fast Fourier

Transform-based protein docking program with pairwise potentials to derive the structure model. Flavodoxin II (PDB ID 1YOB) was used as the ligand and Fe protein (PDB ID

1FP6) as the receptor. The final docking model was chosen based on agreement of the electrostatic potentials of the bound complex, which were generated with PyMOL.

Time resolved proteolysis of flavodoxin and Fe proteins

Limited proteolysis experiments were performed on Fe protein, Fld and the Fe protein-Fld complex. Protein ratios of 1:1, 1:2, and 2:1 (Fe protein:Fld) were digested with Trypsin Gold (Promega, Madison, WI). The reactions were performed in sealed vials in a 50 mM ammonium bicarbonate, pH 8 buffer with1 mM sodium dithionite in a total volume of 120 μL and a protease-to-protein ratio of 1:1000 (w/w) at room temperature. 15 μL samples were taken at 0, 5, 10, 20, 40, 60 and 240 minutes. Reactions were quenched with 1μl of 10% formic acid. 10 μL of each sample was used for analysis

175 of the trypsin digestion pattern by SDS-PAGE using a 4-20% linear gradient gel

(Mini-Protean TGX, Bio-Rad, Hercules, CA).

Mapping of proteolytic cleavage sites

The molecular weight of released peptides was determined using an Autoflex III

MALDI-TOF/TOF mass spectrometry (Bruker Daltonics, Billerica, MA). 1 μL of each sample was co-crystallized with a saturated solution of α-cyano-4-hydroxycinnamic acid matrix in 50% acetonitrile (Thermo-Fisher Scientific, Waltham, MA) containing 0.1% formic acid. Peptide mass spectra were acquired in positive, reflectron mode with an acceleration voltage of 20 kV. Spectra were an accumulation of 1000 laser shots. Spectra were averaged from three spots for each time point. Tryptic peptides were mapped to the protein sequence using the Protein Analysis Worksheet (PAWS) software package

(ProteoMetrics, LLC). Results were mapped onto the ClusPro 2.0 Docking model.

Chemical cross-linking

Chemical cross-linking experiments were performed with Fe protein (20 μM) and

Fld (Fe protein:Fld ratios were 2:1, 1:1, 1:2) using 10mM glutaraldehyde in 50mM

HEPES pH 7.5, 150 mM NaCl, and 1mM sodium dithionite. Reaction conditions were optimized to limit formation of higher order aggregates. Samples were incubated with glutaraldehyde for 10 minutes, at room temperature. The reaction was quenched with 100 mM Tris buffer, pH 8.0. To preserve the native state and conformation of proteins and proteins in complex with their cofactors, all steps were performed under strict anaerobic conditions. Complex formation and crosslinking were initiated at the same time.

Cross-linked samples and controls were analyzed by SDS-PAGE (4-20% linear gradient Mini-Protean TGX, Bio-Rad, Hercules, CA). Protein bands corresponding to

176 monomers and dimers were digested with trypsin as previously described44 prior to mass spectrometry analysis. LC/MS/MS measurements were performed using a maXis

Impact UHR-QTOF (Bruker Daltonics, Billerica, CA) interfaced with an Agilent 1100

HPLC nanoflow system (Agilent Technologies, Santa Clara, CA). Peptide mixtures were separated on a Dionex column (20 mm x 100 μm, 5 μm, 100Å, Acclaim PepMap100,

C18, Dionex), kept at 40°C, and eluted with flow of 800nl/min in solvents A (0.1% formic acid) and B (acetonitrile/0.1% formic acid) gradient: 3–30% B in 16 min, followed by 3 min 30-95% B and 5 min 95% B column wash step. The column was equilibrated with 3% B for 2 minutes prior to the next injection. Electrospray conditions in both MS and auto MS/MS modes were: drying gas flow 4.0 L/min at 100°C, capillary

1600 V. Data was collected over the m/z range of 300–1700 at 2 spectra/s acquisition rate for MS and MS/MS, respectively. A linear voltage gradient depending on mass to charge ratio was applied for peptides fragmented in auto MS/MS experiments with a decreasing order of preference +2> +3> +4 > +1 charged parent ions. Raw data were converted to mgf format using MS Convert and uploaded to SearchGUI, (version 1.26.6.) for sequence identification and visualized and validated in PeptideShaker (version 0.40.1).45 Validated peptides with the identification confidence higher than 95% were selected for further analysis.

RESULTS AND DISCUSSION

HQ Establishing kcat with DT or Fld

In the two-component catalytic system of nitrogenase, both the Fe protein and the

MoFe protein are catalysts. The rate limiting step for the overall reaction is held to be the dissociation step in the Fe protein cycle (Figure B2). This means that the rate constant of

177 the rate limiting step should be the same as the kcat for both Fe protein and MoFe protein cycles in terms of electrons donated/accepted per active site per unit time. Earlier studies established a turnover number (kcat) for electron flow to substrate reduction of between 5-10 s-1 regardless of substrate.31,46 This value is approximately the same as the

ox first order rate constant for dissociation of the Fe (MgADP)2 from the MoFe protein of

~6 s-1, leading to the conclusion that the overall reaction rate limiting step is dissociation

(Figure B2).1,16–18,31 However, a few previous studies have shown that the rate of dissociation of the Fe-MoFe protein complex using DT is slower than the kcat under saturating (‘high flux’) conditions.40,46

Here, we determined the Vmax for both Fe protein and MoFe protein cycles using either DT or Fld as reductant by varying the ratio of Fe protein to MoFe protein (called the “electron flux”). The Vmax values for the Fe protein cycle were determined under low flux conditions with a molar ratio of 1:1([Fe]:[MoFe]) for acetylene reduction and 1:2 for proton reduction using either DT or FldHQ-DT as a reductant. These ‘low-flux’ conditions saturated the Fe protein with MoFe protein and resulted in a Fe protein cycle kcat,Fe of ~6

-1 HQ -1 s with DT as reductant. When Fld -DT was used as the reductant, a kcat,Fe of 10-11 s was observed (Table B1 and C1, Figures C1-C4). These results reveal that using Fld as reductant speeds up the overall reaction, suggesting that the rate limiting dissociation step has been accelerated by about 2-fold when Fld is the reductant compared to the dissociation rate constant when DT is the reductant.40 This enhancement effect on the

Vmax and kcat of both Fe and MoFe protein is dependent on the electron flux, with lower flux showing the largest effect (Figures C2-C5). When the electron flux is increased to

178 Table B1. Steady-state kcat values for Fe protein and MoFe protein with different reductants.

-1 a Protein Reductant kcat (s ) DT 5-6 Fe Protein FldHQ 10-11 DT 10-11 MoFe protein FldHQ 10-11 a The steady-state kcat values for Fe protein and MoFe protein were estimated based on the Vmax of Fe protein (MW≈ 64,000 Da) under low electron flux conditions and Vmax of MoFe protein (MW≈ 240,000 Da) under high flux conditions. For detailed reaction conditions refer to the figure legends for Figures C1-C4 and Table C1.

saturate the MoFe protein with Fe protein (‘high flux’, [Fe]:[MoFe] = 16-20:1) to

-1 measure the MoFe protein cycle, the kcat is found to be ~10-11 s with both DT and Fld

HQ as reductant. The increase in kcat,Fe observed for Fld compared to DT as reductant could be explained by an increase in the rate of dissociation of the Fe protein from MoFe protein when FldHQ is the reductant under low flux condition. To test this, we determined

ox an apparent dissociation rate constant for Fe (ADP)2 –MoFe protein with either DT or

Fld as reductant.

ox Dissociation of the Fe (ADP)2 from the MoFe protein

To determine the apparent dissociation constant using SF spectrophotometry, the

ox Fe (ADP)2-MoFe protein complex was pre-formed in one syringe of the SF spectrophotometer and rapidly mixed against a reductant mixture. Here, such experiments were conducted with either DT or FldHQ in the second syringe. The reduction of the Feox was monitored at 426 nm, an isosbestic point for reversible conversion between FldSQ and FldHQ (Figure C6). As shown in Figure B3, the estimated first order rate constant for

179

ox HQ Figure B3. Reduction of Fe (ADP)2 protein by DT (red), MV (green), or Fld (blue) in ox the presence of MoFe protein. The reduction of Fe (ADP)2 by different reductants was monitored as the decrease of the absorbance at 426 nm as a function of time. The data are displayed as gray dots and were fit to different equations as described in the Materials and Methods section to obtain the pseudo-first order kobs. Syringe 1 contained 20 mM DT with 100 μM MV, or 400 μM FldHQ with 20 mM DT. Syringe 2 contained 80 μM Feox and 80 μM MoFe. MgADP (9 mM) was present in both syringes.

-1 HQ dissociation (kobs) with DT as reductant was ~4 s . When Fld was the reductant, the

ox dissociation and reduction of Fe (ADP)2 was much more rapid, with the majority of the

-1 reduction (~65%) occurring in less than 1 ms. A kobs of >760 s was estimated from a fit of the points captured after the dead time. These findings reveal that with FldHQ as

ox reductant, the reduction of Fe is much faster than the observed kcat, implying that

ox HQ dissociation of Fe (ADP)2 from MoFe protein is not the rate limiting step when Fld is used as the reductant. These results are consistent with the previously reported data from similar experiments.25 To further test the effect of reductant on the apparent rate of dissociation, another non-physiological reductant, methyl viologen was tested.47 As can

180 -1 be seen in Figure B3, this electron donor gave a kobs of 100 s , again much faster than the kcat.

ox HQ The faster reduction of Fe (ADP)2 in the presence of MoFe protein by Fld could be explained by two possible mechanisms: (i) the very rapid dissociation of the

ox HQ Fe (ADP)2 from the MoFe protein when Fld is the reductant, pointing to a different rate limiting step,40 or (ii) FldHQ reduction of the Fe protein while still complexed to the

MoFe protein,48 a mechanism proposed earlier by Haaker et al.25 To test the second

HQ ox model, the ability of Fld to reduce the Fe (ADP)2 while still bound to the MoFe protein was examined.

Effect of FldHQ on the primary electron transfer of nitrogenase

The pre-steady state ET from the Fe protein to the MoFe protein offers a straightforward way to monitor the possibility of ET from FldHQ to Feox while still bound to the MoFe protein. When the protein mixture of Fered and MoFe protein is mixed with

MgATP in a SF spectrophotometer, an apparent first order ET event is monitored by the increase in absorbance (oxidation of the Fe protein) that can be fit to a single exponential

-1 to yield a rate constant for ET (kET). With DT as the reductant, a kET of 173 s was observed (Figure B4), consistent with literature values that range from 140-200 s-1.14–

16,22,25 Important for this study, the absorbance value plateaus starting at 10 ms and stays roughly flat up to 30 ms. This plateau in absorbance reflects no reduction by DT of the

Feox protein while still in the complex. At much later times (100 ms), the absorbance does change, reflecting a complex set of events as the Fe protein dissociates from the MoFe protein, is reduced, and rebinds to the MoFe protein. When the ET study is conducted with FldHQ as reductant, nearly identical primary ET kinetics are observed (Figure B4).

181

red Figure B4. Primary ET from Fe (ATP)2 protein to MoFe protein in presence of DT or FldHQ with DT. Syringe 1 contained 80 μM Fered protein, 20 μM MoFe protein, 1 mM DT and 10 mM MgATP. Syringe 2 had 10 mM DT or 200 µM FldHQ with 1 mM DT in the presence of 10 mM MgATP.

Importantly, no reduction of the Feox protein is observed during the 10 to 30 ms time

ox frame. Given the earlier observation of reduction of Fe (ADP)2 protein-MoFe protein with FldHQ of greater than 760 s-1, an on-complex reduction of the Fe protein in the ET study should have resulted in a significant (if not complete) reduction of the Fe protein before 10 ms. The lack of any observed reduction of the Feox protein in the ET

HQ ox experiment reveals that the Fld protein reduction of the Fe (ADP)2 protein on complex with the MoFe protein must be very slow, as has been reported earlier.48 Thus, these studies, coupled with the dissociation studies presented in the previous section, rule out

FldHQ reduction of the Fe protein while still bound to the MoFe protein,25 and instead

182 ox HQ favor a rapid dissociation of Fe (ADP)2 protein from the MoFe protein when Fld is the reductant.

Reduction of Feox protein by DT and FldHQ

ox The studies, to this point, favor rapid dissociation of the Fe (ADP)2 protein from the MoFe protein when FldHQ is the reductant. The observed lower rate of dissociation when DT is the reductant could indicate slow reduction of Feox protein, rather than slow dissociation of the Fe protein from the MoFe protein.17,23,25 Several previous studies have

•- 17– illustrated low activity for nitrogenase at low concentration of SO2 as the reductant.

19,22,49 To determine the rates of reduction of Feox protein by different reductants, the pre- steady state studies were conducted using Feox in the absence and presence of nucleotides

(ADP or ATP). Feox protein in one syringe was rapidly mixed against either DT or FldHQ in the other syringe. Additionally, these studies were conducted at 4o C to slow the reactions down enough to observe. Consistent with previous studies, nucleotides significantly slow down the rates of reduction of Feox protein by DT (by ~ 100-fold).22,23

In contrast, with FldHQ as reductant, the rates of reduction remained fast and roughly unchanged with or without nucleotide present (Figure B5).25,27

Taken together, the studies presented so far suggest that when FldHQ is the

ox reductant, the Fe (ADP)2 dissociation from the MoFe protein is not rate limiting and does not correspond to the overall kcat for substrate reduction. Rather, the dissociation

HQ step when Fld , is used as the reductant is much more rapid than the kcat.

183

Figure B5. Kinetics of the reduction of Feox protein by DT and FldHQ with DT in the presence and absence of nucleotides. ET is monitored by observing the change in absorbance at 426 nm as a function of time. Panel A: Reduction of Fe protein by DT and FldHQ with DT with no nucleotide. Panel B: Reduction of Fe protein by DT and FldHQ with DT in the presence of MgADP. Panel C: Reduction of Fe protein by DT and FldHQ with DT in the presence of MgATP. Syringe 1 contained 80 μM Feox. Syringe 2 contained 20 mM DT or 600 μM FldHQ with 20 mM DT. MgADP and MgATP, when included, were present in both syringes at a final concentration of 10 mM. The kobs values were averaged from two independent experiments except for Fld with MgATP.

Protein-protein interaction between Fe protein and Fld

The results presented so far indicate that FldHQ does not reduce the Fe protein while still complexed to the MoFe protein, suggesting that reduction only occurs once the

Fe protein is free from the MoFe protein. We next employed molecular modeling to predict where the Fld would bind to the Fe protein to achieve ET. The docking model produced by ClusPro 2.0 predicts the binding of 1 Fld monomer to a dimer of Fe protein

(Figure B6, Panel A). The distance between the [4Fe-4S] cluster at the active site of the

184 Fe protein and the FMN of the Fld measured in PyMOL was within electron transfer distance (<10Å) (Figure B6, Panel B).

Analysis of electrostatic potentials performed in PyMOL demonstrated qualitative agreement with the ClusPro 2.0 docking model. The highly negative patch at the FMN cofactor of Fld sits on top of the highly positive patch at the location of the [4Fe-4S] cluster of the Fe protein (Figure C7). Salt bridge interactions were identified in the proposed complex binding region on either side of the FMN cofactor of Fld and the [4Fe-

4S] cluster of the Fe protein (Figure B7, Panel A). The identified residues are Arg100 on one of the Fe protein subunits and Glu104 of Fld and between Arg140 of the Fe protein and Asp154 of Fld. These salt bridge interactions further support the analysis that electrostatic interactions between the Fe protein and Fld enable complex formation.

Figure B6. Interaction of Fe protein and Fld for ET. Panel A: The docking model from ClusPro 2.0 predicts the binding of one Fld monomer to a dimer of Fe protein. The complex structure shows the nitrogenase Fe protein with subunits in yellow and green (PDB ID: 1FP6) and Fld (NifF) is in blue (PDB ID: 1YOB). Panel B: Close-up of the docked proteins. The distance between the [4Fe-4S] cluster at the active site of the iron protein and the FMN cofactor at the active site of the flavodoxin is predicted to be 6.4Å.

185 Interestingly, the Fe protein Arg100 residue is the site of ADP ribosylation, which controls association of the Fe protein with the MoFe protein in response to N status of the cell.50,51

Time-resolved limited proteolysis experiments were subsequently performed to test the docking model. These experiments involve incubation of proteins with proteases for varying lengths of time and then mapping the observed cleavages on the protein sequences. This approach facilitates identification of kinetically favored sites of protease cleavage. By comparing the peptide fragments before and after formation of a complex, regions that change conformation or are protected in the complex can be identified.

Reactions were performed at Fe protein/Fld ratios of 1:1, 1:2, and, 2:1, and with the two proteins alone. Sites of proteolysis were identified using MALDI-TOF and LC/MSMS and mapped onto the docking model. Cleavage sites on Fe protein were located on solvent exposed surfaces under all conditions (Figure B7). Fewer sites were found on

Fld, however, upon formation of the complex, Fld cleavage sites directly adjacent to the proposed binding surfaces (Figure B7, Panel B) were protected. Specifically, Fld tryptic fragment Lys15-Lys22 was no longer present when bound to Fe protein (Figure B7,

Panel B). Furthermore, regions near the [4Fe-4S] cluster of the Fe protein and the FMN cofactor of Fld were not observed, indicating that these regions were also protected from proteolysis (Figure B7, Panel B). The proteolysis data are consistent with the docking model with respect to the surfaces involved in mediating protein-protein interactions.

186

Figure B7. Mapping of the Fe protein-Fld interaction site. Panel A: Salt bridge interactions were identified in the proposed binding region on either side of the FMN cofactor of Fld (blue) and the [4Fe-4S] cluster of the Fe protein (FeP)(green and yellow). The identified residues were Arg-100 on one of the Fe protein subunits (green) and Glu- 104 of flavodoxin and between Arg-140 on the other Fe protein subunit (yellow) and Asp-154 of Fld. Panel B: Time resolved proteolysis experiments show changes in Fld upon binding to Fe protein. Fe protein and Fld were digested with trypsin before (Panel B, left) and after formation of a complex (Panel B, right). Mapping of the kinetically favored sites of cleavage revealed different patterns. One of the favored sites on Fld, which mapped near the binding surface with Fe protein (black arrow), was protected from cleavage in the complex (red arrow). Panel C: Addition of cross-linking reagent to Fe protein-Fld complex (1:1 w/w protein ratio) resulted predominantly in formation of monomers (intraprotein linkages) and dimers (interprotein linkages between Fe protein- Fe protein and Fe protein-Fld) (Panel C, left). In addition to cross-linked complex, the following controls were run: Fe protein, Fld, and Fe protein-Fld non cross-linked, and Fe protein and Fld cross-linked individually. To identify interacting domains, corresponding monomers and dimmers were digested with trypsin and identified peptides (red) were mapped onto the ClusPro 2.0 Docking model (Panel C, middle and right show Fld and Fld in complex with Fe protein after exposure to glutaraldehyde, respectively). Fragments colored in orange denote peptides underrepresented in cross-linked sample with respect to unlabeled protein. Red arrow, points to this part of Fld structure which was absent after crosslinking.

187 While implied by our data and previous models, the stoichiometry of the solution phase complex has not been directly addressed. To investigate this, chemical cross-linking experiments were performed. Fe protein alone, Fld alone, and after complex formation were covalently cross-linked using 20 mM glutaraldehyde (GA). The proteinswere exposed to GA for 10 min before the reaction was quenched. Analysis of the samples by SDS-PAGE showed that the predominant species present was consistent with an Fe protein-Fld binary complex (Figure B7, Panel C, left). Mass spectrometry analysis confirmed the presence of both proteins in the gel band of interest. However, some peptide fragments were no longer detectable or were present in much lower quantity when Fe protein-Fld were engaged in complex and exposed to GA (Figure B7).

For example, based on Fld (Figure B7, Panel C, middle and left), there are three distinguished regions: (1) colored in red, this part of the protein chain was easily digested by trypsin and was identify by LC/MSMS with high confidence. One of the fragments,

Ile125-Lys146, was always detectable regardless of GA treatment, suggesting lack of involvement in any kind of interactions (within Fld itself and/or between Fld and Fe protein); (2) colored in orange, these parts of protein sequence were also identified, however, they were found in significantly lower concentrations in cross-linked samples what implies more structured regions (within Fld itself) that were in close proximity and were linked with GA and/or protein sections potentially engaged in interactions with Fe protein; colored in blue, this part of Fld chain was not detected in any sample. Taken together our observations from chemical labeling, Fld peptide fragment Phe147-Lys160 is directly involved in contact with Fe protein, however more Fld sequence parts might be involved in complex formation through subtle conformation adjustments.

188 The results from the in silico modeling, proteolysis, and cross-linking experiments all support the predicted 1:1 stoichiometry for the Fe protein-Fld interaction with an interface bringing the two cofactors ([4Fe-4S] cluster and FMN) close together to favor the ET (Figures B6, B7, and C7). The interaction site on the Fe protein surface with Fld is the same as that with MoFe protein. To test this prediction, we carried out the proton reduction assay under both low and high electron flux conditions. The results display a significant inhibitory effect of Fld on the proton reduction activity (Figure C8) in the presence of 2.4 mM FldHQ, compared to those in the presence of 600 μM FldHQ

(Figure C1). The observed inhibition from Fld further supports the transient interaction between the Fe protein and Fld during nitrogenase catalysis.27

Efficiency of ATP hydrolysis per electron transferred for substrate reduction

It has been reported that FldHQ can reduce the Feox protein by 1 electron to the

[4Fe-4S]1+ state25,27 or by two electrons to the all ferrous [4Fe-4S]0 state.33,34 From the all ferrous state, the Fe protein could deliver two electrons to the MoFe protein per association, resulting in 2 ATP hydrolyzed/2 e- transferred.33,34,40 This possibility is contrary to the consensus that each association event results in 2 ATP hydrolyzed per 1 e- transferred. Using DT as the reductant, the ATP/e- ratio was found to be ~2:1 with three different substrates and under high or low electron flux conditions (Table B2, Figure

C9). This result is consistent with the previously reported data. Using FldHQ as the reductant, the ratio of ATP/e- ratio was also found to be ~2:1 under high and low electron flux with three substrates. The results clearly indicate that ~2 ATP were hydrolyzed per electron transferred (ATP/e-= 2, Table B2), which was independent of substrate, reductant (DT or FldHQ), and electron flux condition. These findings are consistent with a

189 Table B2. Number of ATP hydrolyzed per electron transferred for reduction of different substrates by nitrogenase under different electron flux conditions.

d Reaction ATP/e- Ratio Reductant a e f Conditions Proton (Ar) N2 (1 atm) Acetylene (0.11 atm) DT 1.87  0.04 1.85 2.17  0.07 Low Fluxb FldHQ 1.91  0.05 1.91 2.26  0.12 DT 1.63  0.12 1.82 1.89  0.10 High Fluxc FldHQ 1.96  0.17 1.78 2.74  0.21 aAll assays for product quantification were done in a buffer containing 8.5 mM MgATP without a regeneration system with 12 mM DT or 712 μM FldHQ plus 10 mM DT. For ADP quantification assays, trace concentration of -32P-ATP was added to the reaction mixtures. bLow flux assays were done using 1.95 μM of Fe protein with 2.08 μM of MoFe protein for proton and acetylene reduction, and 4.16 μM of MoFe protein for N2 reduction. cHigh flux assays were done with 0.41 μM MoFe protein and 8 μM Fe protein. dThe ATP/e- ratios for proton and acetylene reduction assays were averaged from the data at three different reaction times (Figure S1) with the standard deviation shown. N2 reduction values were calculated from assays with a reaction time of 60 sec. eTotal f electrons for H2 and NH3 production were counted. Only electrons for ethylene production were counted as proton reduction was below the detection limit.

single ET per association event coupled to the hydrolysis of two ATP molecules even when FldHQ is the reductant.1,52

Pre-steady state and steady state analysis of ATP hydrolysis

Based on the findings presented to this point, it is concluded that the Fe protein-

MoFe protein dissociation step is not rate-limiting. What then is the rate limiting step in the overall reaction cycle? In the Fe protein cycle using DT as the reductant (Figure B2),

-1 -1 the ATP hydrolysis (kATP = 50-70 s ) and the Pi release (kPi = 16-22 s ) steps both are

-1 slower than the ET (kET = 140-200 s ).

Comparing rate constants for ATP hydrolysis from pre-steady state studies with

DT or FldHQ as reductant and a 1:4 ([MoFe]:[Fe]) electron flux revealed similar values

190 -1 -1 o (kATP,DT and kATP,Fld) of 40 s and 44 s (25 C), respectively (Figure B8). These rate

-1 constants are about double the kcat (10-11 s ) when considering that 2 ATP molecules are hydrolyzed per electron, indicating that ATP hydrolysis is not the rate-limiting step for the overall reaction. The steady state (linear) rates of ATP hydrolysis are found to be

-1 -1 o kATP,DT = 10 s and kATP,Fld = 20 s (25 C), consistent with the kcat values reported here.

The slow steady-state ATP hydrolysis rate using DT matches with the slow reduction rate

ox ox -1 of Fe (ADP)2 and Fe (ATP)2 (kobs = 5-6 s ) by DT. Whereas, the faster kATP,Fld is

ox ox HQ consistent with the faster reduction of the Fe (ADP)2 and Fe (ATP)2 states by Fld

Figure B8. Time course of pre-steady state and steady state ATP hydrolysis during nitrogenase catalysis under Ar with DT (open squares) or FldHQ with DT (closed squares) as reductant at 25oC. The hydrolysis of ATP was monitored with [α-32P]ATP as a tracing reagent. The ratio of ADP/MoFe protein was plotted as a function of time. The data were fitted to two phases: a burst exponential followed by a linear steady state for both DT and HQ Fld with DT. Pre-steady state burst phase gave first order rate constants of kATP,DT = 40 -1 -1 s and kATP,Fld = 44 s , while the steady state linear phase gave rate constants of kATP,DT = -1 -1 10 s and kATP,Fld = 20 s .

191 (Figure B5).22,23 The ATP hydrolysis rate constant obtained under high flux conditions ([MoFe]:[Fe] = 1:16) showed no difference in the pre-steady state rate constants for ATP hydrolysis with DT or Fld as reductant (data not shown). However, the steady state rate constant for ATP hydrolysis using DT is significantly increased to 25 s-1, which is about the same as the value when using FldHQ (27 s-1) under same conditions. It is interesting that the steady-state ATP hydrolysis rate constant (25-27 s-1) is about the same as those previously reported for the pre-steady state Pi release step (16-22 s-1,

Figure B2).16,18 Given that two Pi are released per electron transferred, the Pi release rate

-1 constant should be double the kcat value (about 20 s ) if this step is the overall rate limiting step.

Pi release events

Pi release was measured using an established fluorometric method43 with either

DT or FldHQ as reductant under high electron flux conditions ([Fe]:[MoFe]: = 16:1). The data showed an initial lag phase followed by a linear phase with both reducatants, with

-1 -1 the linear phases having rate constants for Pi release of kPi,DT = 27 s and kPi,Fld = 25 s

(Figure B9). These rate constants are about double the overall reaction rate constant kcat of 10-11 s-1 per electron. Two Pi are released for each electron transferred, showing that the overall rate limiting step is likely events associated with Pi release. We have no data on the ADP release event, so the position of this event in the cycle remains unknown.

Establishing the rate-limiting step in the Fe protein cycle

Considering all of the rate constants for the steps in the Fe protein cycle with

HQ -1 -1 Fld as reductant: electron transfer (kET = 173 s ), ATP hydrolysis (kATP ~40-44 s ),

192

Figure B9. Real-time measurement of Pi release during nitrogenase catalysis using DT (gray) or FldHQ with DT (black) as reductant at 25oC. The Pi release was monitored by a fluorescence increase caused by the binding of Pi to MDCC-PBP. The ratio of Pi/MoFe protein is plotted as a function of time. After the initial lag phase (~30 ms), the data collected from 40 to 200 ms were fitted to a linear equation (solid line), giving rate -1 -1 constants of kPi,DT = 27 s and kPi,Fld = 25 s .

-1 ox -1 Pi release (25-27 s ) and re-reduction of Fe (ADP)2 (kobs >1200 s ), it is concluded that the overall rate limiting step for the reaction are events associated with the Pi release step, not complex dissociation. Our findings are consistent with other ATP hydrolyzing systems, including helicases53,54 and myofibrillar ATPases,55–58 where Pi release is the rate limiting step.

Summary

The Fe protein cycle can now be updated with the rate constants determined here using FldHQ as the reductant, as shown in Figure B10. The cycle begins with the rapid

red equilibrium docking of Fe (ATP)2 to MoFe protein, which is followed by the

193 conformationally gated ET events. Next occurs ATP hydrolysis, which is followed by events associated with Pi release. It is unknown if the Pi release event is conformationally gated, but it is clear that events associated with this step are rate limiting for the overall

ox Fe protein cycle. The dissociation of the Fe (ADP)2 is fast, with rapid reduction by the

FldHQ. Finally, ATP replaces ADP in the free Fe protein. The order of these last two events is not established, but the rapid reduction by FldHQ suggests that reduction should occur before nucleotide exchange.

[I]

[e-] 2ATP 2ADP ox red red [Fe (ADP)2] [Fe (ADP)2] [Fe (ATP)2] -1 fast red k 1200 s MoFe red,Fld MoFe MoFe

1

- MoFe

5

s

4

M

1

M .

0

-

-

4

[

1

6

s

1 I

-

1 

7

s

I

d

0

5

]

-

s

e

1 7

1

1

-

x 1

0

s

a

6

s

A

l

i

d

e

c

k

t

R

i

red ox red v

[MoFe ;Fe (ADP)2] [MoFe;Fe (ATP)2] a

]

t

e

V

I

d [ 2[Pi] -1 -1 kPi = 25-27 s kET = 173 s

-1 red ox k ATP = 40-44 s red ox [MoFe ;Fe (ADP,Pi)2] [MoFe ;Fe (ATP)2]

[III]

Figure B10. Updated Fe protein cycle with FldHQ as reductant.

194 FUNDING

This work was supported as part of the Biological Electron Transfer and Catalysis Energy

Frontier Research Center funded by the U.S. Department of Energy, Office of Science,

Basic Energy Sciences under award #DE-SC0012518.

ACKNOWLEDGMENTS

The authors thank the members of the Biological Electron Transfer and Catalysis Energy

Frontiers Research Center for helpful discussions.

REFERENCES

(1) Burgess, B. K., and Lowe, D. J. (1996) Mechanism of molybdenum nitrogenase.

Chem. Rev. 96, 2983–3012.

(2) Seefeldt, L. C., Hoffman, B. M., and Dean, D. R. (2009) Mechanism of Mo- dependent nitrogenase. Annu. Rev. Biochem. 78, 701–722.

(3) Hoffman, B. M., Lukoyanov, D., Yang, Z.-Y., Dean, D. R., and Seefeldt, L. C. (2014)

Mechanism of nitrogen fixation by nitrogenase: the next stage. Chem. Rev. 114, 4041–

4062.

(4) Hageman, R. V., and Burris, R. H. (1978) Nitrogenase and nitrogenase reductase associate and dissociate with each catalytic cycle. Proc. Natl. Acad. Sci. U.S.A. 75, 2699–

2702.

(5) Georgiadis, M. M., Komiya, H., Chakrabarti, P., Woo, D., Kornuc, J. J., and Rees, D.

C. (1992) Crystallographic structure of the nitrogenase iron protein from Azotobacter vinelandii. Science 257, 1653–1659.

(6) Seefeldt, L. C., Hoffman, B. M., and Dean, D. R. (2012) Electron transfer in nitrogenase catalysis. Curr. Opin. Chem. Biol. 16, 19–25.

195 (7) Chan, M. K., Kim, J., and Rees, D. C. (1993) The nitrogenase FeMo-cofactor and

P-cluster pair: 2.2 Å resolution structures. Science 260, 792–794.

(8) Kim, J., and Rees, D. C. (1992) Structural models for the metal centers in the nitrogenase molybdenum-iron protein. Science 257, 1677–1682.

(9) Einsle, O., Tezcan, F. A., Andrade, S. L. A., Schmid, B., Yoshida, M., Howard, J. B., and Rees, D. C. (2002) Nitrogenase MoFe-protein at 1.16 Å resolution: A central ligand in the FeMo-cofactor. Science 297, 1696–1700.

(10) Spatzal, T., Aksoyoglu, M., Zhang, L., Andrade, S. L. A., Schleicher, E., Weber, S.,

Rees, D. C., and Einsle, O. (2011) Evidence for interstitial carbon in nitrogenase FeMo cofactor. Science 334, 940–940.

(11) Lancaster, K. M., Roemelt, M., Ettenhuber, P., Hu, Y., Ribbe, M. W., Neese, F.,

Bergmann, U., and DeBeer, S. (2011) X-ray emission spectroscopy evidences a central carbon in the nitrogenase iron-molybdenum cofactor. Science 334, 974–977.

(12) Thorneley, R. N., Lowe, D. J., Eday, R. R., and Miller, R. W. (1979) The coupling of electron transfer in nitrogenase to the hydrolysis of magnesium adenosine triphosphate. Biochem. Soc. Trans. 7, 633–636.

(13) Hageman, R. V., Orme-Johnson, W. H., and Burris, R. H. (1980) Role of magnesium adenosine 5’-triphosphate in the hydrogen evolution reaction catalyzed by nitrogenase from Azotobacter vinelandii. Biochemistry 19, 2333–2342.

(14) Danyal, K., Dean, D. R., Hoffman, B. M., and Seefeldt, L. C. (2011) Electron transfer within nitrogenase: Evidence for a deficit-spending mechanism. Biochemistry 50,

9255–9263.

196 (15) Danyal, K., Mayweather, D., Dean, D. R., Seefeldt, L. C., and Hoffman, B. M.

(2010) Conformational gating of electron transfer from the nitrogenase Fe protein to

MoFe protein. J. Am. Chem. Soc. 132, 6894–6895.

(16) Duval, S., Danyal, K., Shaw, S., Lytle, A. K., Dean, D. R., Hoffman, B. M., Antony,

E., and Seefeldt, L. C. (2013) Electron transfer precedes ATP hydrolysis during nitrogenase catalysis. Proc. Natl. Acad. Sci. U.S.A. 110, 16414–16419.

(17) Thorneley, R. N. F., and Lowe, D. J. (1983) Nitrogenase of Klebsiella pneumoniae.

Kinetics of the dissociation of oxidized iron protein from molybdenum-iron protein:

Identification of the rate-limiting step for substrate reduction. Biochem. J. 215, 393–403.

(18) Wilson, P. E., Nyborg, A. C., and Watt, G. D. (2001) Duplication and extension of the Thorneley and Lowe kinetic model for Klebsiella pneumoniae nitrogenase catalysis using a MATHEMATICA software platform. Biophys. Chem. 91, 281–304.

(19) Thorneley, R. N. F., and Lowe, D. J. (1985) Kinetics and mechanisms of the nitrogenase enzyme system, in Molybdenum Enzymes (Spiro, T. G., Ed.), pp 221–284.

Wiley, New York.

(20) Seefeldt, L. C., Yang, Z.-Y., Duval, S., and Dean, D. R. (2013) Nitrogenase reduction of carbon-containing compounds. Biochim. Biophys. Acta 1827, 1102–1111.

(21) Lowe, D. J., and Thorneley, R. N. (1984) The mechanism of Klebsiella pneumoniae nitrogenase action: The determination of rate constants required for the simulation of the kinetics of N2 reduction and H2 evolution. Biochem. J. 224, 895.

(22) Ashby, G. A., and Thorneley, R. N. F. (1987) Nitrogenase of Klebsiella pneumoniae: Kinetic studies on the Fe protein involving reduction by sodium dithionite,

197 the binding of MgADP and a conformation change that alters the reactivity of the

4Fe-4S centre. Biochem. J. 246, 455–465.

(23) Wilson, P. E., Bunker, J., Lowery, T. J., and Watt, G. D. (2004) Reduction of nitrogenase Fe protein from Azotobacter vinelandii by dithionite: Quantitative and qualitative effects of nucleotides, temperature, pH and reaction buffer. Biophys. Chem.

109, 305–324.

(24) Yates, M. G. (1972) Electron transport to nitrogenase in Azotobacter chroococcum:

Azotobacter flavodoxin hydroquinone as an electron donor. FEBS Lett. 27, 63–67.

(25) Duyvis, M. G., Wassink, H., and Haaker, H. (1998) Nitrogenase of Azotobacter vinelandii: Kinetic analysis of the Fe protein redox cycle. Biochemistry 37, 17345–17354.

(26) Bennett, L. T., Jacobson, M. R., and Dean, D. R. (1988) Isolation, sequencing, and mutagenesis of the nifF gene encoding flavodoxin from Azotobacter vinelandii. J. Biol.

Chem. 263, 1364–1369.

(27) Thorneley, R. N., and Deistung, J. (1988) Electron-transfer studies involving flavodoxin and a natural redox partner, the iron protein of nitrogenase: Conformational constraints on protein-protein interactions and the kinetics of electron transfer within the protein complex. Biochem. J. 253, 587–595.

(28) Mortenson, L. E. (1964) Ferredoxin and ATP, requirements for nitrogen fixation in cell-free extracts of Clostridium pasteurianum. Proc. Natl. Acad. Sci. U.S.A. 52, 272–

279.

(29) Martin, A. E., Burgess, B. K., Iismaa, S. E., Smartt, C. T., Jacobson, M. R., and

Dean, D. R. (1989) Construction and characterization of an Azotobacter vinelandii strain

198 with mutations in the genes encoding flavodoxin and ferredoxin I. J. Bacteriol. 171,

3162–3167.

(30) Saeki, K. (2004) Electron transport to nitrogenase: Diverse routes for a common destination, in Genetics and regulation of nitrogen fixation in free-living bacteria (Klipp,

W., Masephol, B., Gallon, J. R., and Newton, W. E., Eds.), pp 257–290. Kluwer

Academic Publishers, Dordrecht.

(31) Hageman, R. V., and Burris, R. H. (1978) Kinetic studies on electron transfer and interaction between nitrogenase components from Azotobacter vinelandii. Biochemistry

17, 4117–4124.

(32) Simpson, F. B., and Burris, R. H. (1984) A nitrogen pressure of 50 atmospheres does not prevent evolution of hydrogen by nitrogenase. Science 224, 1095–1097.

(33) Erickson, J. A., Nyborg, A. C., Johnson, J. L., Truscott, S. M., Gunn, A.,

Nordmeyer, F. R., and Watt, G. D. (1999) Enhanced efficiency of ATP hydrolysis during nitrogenase catalysis utilizing reductants that form the all-ferrous redox state of the Fe protein. Biochemistry 38, 14279–14285.

(34) Lowery, T. J., Wilson, P. E., Zhang, B., Bunker, J., Harrison, R. G., Nyborg, A. C.,

Thiriot, D., and Watt, G. D. (2006) Flavodoxin hydroquinone reduces Azotobacter vinelandii Fe protein to the all-ferrous redox state with a S = 0 spin state. Proc. Natl.

Acad. Sci. U.S.A. 103, 17131–17136.

(35) Fisher, K., Hare, N. D., and Newton, W. E. (2014) Another role for CO with nitrogenase? CO stimulates hydrogen evolution catalyzed by variant Azotobacter vinelandii Mo-Nitrogenases. Biochemistry 53, 6151–6160.

199 (36) Benemann, J. R., Yoch, D. C., Valentine, R. C., and Arnon, D. I. (1969) The electron transport system in nitrogen fixation by Azotobacter, I: Azotoflavin as an electron carrier. Proc. Natl. Acad. Sci. U.S.A. 64, 1079–1086.

(37) Klugkist, J., Voorberg, J., Haaker, H., and Veeger, C. (1986) Characterization of three different flavodoxins from Azotobacter vinelandii. Eur. J. Biochem. 155, 33–40.

(38) Steensma, E., Heering, H. A., Hagen, W. R., and Van Mierlo, C. P. M. (1996) Redox properties of wild-type, Cys69Ala, and Cys69Ser Azotobacter vinelandii flavodoxin II as measured by cyclic voltammetry and EPR spectroscopy. Eur. J. Biochem. 235, 167–172.

(39) Lanzilotta, W. N., Ryle, M. J., and Seefeldt, L. C. (1995) Nucleotide hydrolysis and protein conformational changes in Azotobacter vinelandii nitrogenase iron protein:

Defining the function of aspartate 129. Biochemistry 34, 10713–10723.

(40) Jacobs, D., and Watt, G. D. (2013) Nucleotide-assisted [Fe4S4] redox state interconversions of the Azotobacter vinelandii Fe protein and their relevance to nitrogenase catalysis. Biochemistry 52, 4791–4799.

(41) Christiansen, J., Goodwin, P. J., Lanzilotta, W. N., Seefeldt, L. C., and Dean, D. R.

(1998) Catalytic and biophysical properties of a nitrogenase Apo-MoFe protein produced by a nifB-deletion mutant of Azotobacter vinelandii. Biochemistry 37, 12611–12623.

(42) Barney, B. M., Igarashi, R. Y., Dos Santos, P. C., Dean, D. R., and Seefeldt, L. C.

(2004) Substrate interaction at an iron-sulfur face of the FeMo-cofactor during nitrogenase catalysis. J. Biol. Chem. 279, 53621–53624.

(43) Brune, M., Hunter, J. L., Corrie, J. E. T., and Webb, M. R. (1994) Direct, real-time measurement of rapid inorganic phosphate release using a novel fluorescent probe and its application to actomyosin subfragment 1 ATPase. Biochemistry 33, 8262–8271.

200 (44) Maaty, W. S., Selvig, K., Ryder, S., Tarlykov, P., Hilmer, J. K., Heinemann, J.,

Steffens, J., Snyder, J. C., Ortmann, A. C., Movahed, N., Spicka, K., Chetia, L., Grieco,

P. A., Dratz, E. A., Douglas, T., Young, M. J., and Bothner, B. (2012) Proteomic analysis of Sulfolobus solfataricus during Sulfolobus turreted icosahedral virus infection. J.

Proteome Res. 11, 1420–1432.

(45) Vaudel, M., Burkhart, J. M., Zahedi, R. P., Oveland, E., Berven, F. S., Sickmann, A.,

Martens, L., and Barsnes, H. (2015) PeptideShaker enables reanalysis of MS-derived proteomics data sets. Nat. Biotechnol. 33, 22–24.

(46) Johnson, J. L., Nyborg, A. C., Wilson, P. E., Tolley, A. M., Nordmeyer, F. R., and

Watt, G. D. (2000) Analysis of steady state Fe and MoFe protein interactions during nitrogenase catalysis. Biochim. Biophys. Acta 1543, 24–35.

(47) Watt, G. D., and Reddy, K. R. N. (1994) Formation of an all ferrous Fe4S4 cluster in the iron protein component of Azotobacter vinelandii nitrogenase. J. Inorg. Biochem. 53,

281–294.

(48) Lanzilotta, W. N., and Seefeldt, L. C. (1997) Changes in the midpoint potentials of the nitrogenase metal centers as a result of iron protein−molybdenum-iron protein complex formation. Biochemistry 36, 12976–12983.

(49) Johnson, J. L., Tolley, A. M., Erickson, J. A., and Watt, G. D. (1996) Steady-state kinetic studies of dithionite utilization, component protein interaction, and the formation of an oxidized iron protein intermediate during Azotobacter vinelandii nitrogenase catalysis. Biochemistry 35, 11336–11342.

201 (50) Lowery, R. G., and Ludden, P. W. (1988) Purification and properties of dinitrogenase reductase ADP-ribosyltransferase from the photosynthetic bacterium

Rhodospirillum rubrum. J. Biol. Chem. 263, 16714–16719.

(51) Moure, V. R., Danyal, K., Yang, Z.-Y., Wendroth, S., Müller-Santos, M., Pedrosa,

F. O., Scarduelli, M., Gerhardt, E. C. M., Huergo, L. F., Souza, E. M., and Seefeldt, L. C.

(2013) The nitrogenase regulatory enzyme dinitrogenase reductase ADP- ribosyltransferase (DraT) is activated by direct interaction with the signal transduction protein GlnB. J. Bacteriol. 195, 279–286.

(52) Seefeldt, L. C., and Dean, D. R. (1997) Role of nucleotides in nitrogenase catalysis.

Acc. Chem. Res. 30, 260–266.

(53) Wong, E. V., Cao, W., Vörös, J., Merchant, M., Modis, Y., Hackney, D. D.,

Montpetit, B., and De La Cruz, E. M. (2016) Pi release limits the intrinsic and RNA- stimulated ATPase cycles of DEAD-box protein 5 (Dbp5). J. Mol. Biol. 428, 492–508.

(54) Wang, Q., Arnold, J. J., Uchida, A., Raney, K. D., and Cameron, C. E. (2010)

Phosphate release contributes to the rate-limiting step for unwinding by an RNA helicase.

Nucl. Acids Res. 38, 1312–1324.

(55) Cochran, J. C., Krzysiak, T. C., and Gilbert, S. P. (2006) Pathway of ATP hydrolysis by monomeric kinesin Eg5. Biochemistry 45, 12334–12344.

(56) Lionne, C., Iorga, B., Candau, R., Piroddi, N., Webb, M. R., Belus, A., Travers, F., and Barman, T. (2002) Evidence that phosphate release is the rate-limiting step on the overall ATPase of psoas myofibrils prevented from shortening by chemical cross-linking.

Biochemistry 41, 13297–13308.

202 (57) Lionne, C., Brune, M., Webb, M. R., Travers, F., and Barman, T. (1995) Time resolved measurements show that phosphate release is the rate limiting step on myofibrillar ATPases. FEBS Lett. 364, 59–62.

(58) Milic, B., Andreasson, J. O. L., Hancock, W. O., and Block, S. M. (2014) Kinesin processivity is gated by phosphate release. Proc. Natl. Acad. Sci. U.S.A. 111, 14136–

14140.

203

APPENDIX C

SUPPORTING INFORMATION

EVIDENCE THAT THE Pi RELEASE EVENT IS THE RATE LIMITING STEP

IN THE NITROGENASE CATALYTIC CYCLE

204 Table C1. Electron distribution for H2 and NH3 production under different electron flux conditions with DT or Fld as reductants.

a 1 atm N2 Reaction Conditions - - - e for H2 (nmol) e for NH3 (nmol) % of e for NH3 DT 241.68 ± 4.19 231.94 ± 5.05 48.97 ± 1.26 Low Fluxb FldHQ 329.72 ± 2.59 292.78 ± 2.55 47.03 ± 0.49 DT 172.17 ± 8.71 171.83 ± 1.29 49.95 ± 1.33 High Fluxc FldHQ 157.79 ± 4.36 130.47 ± 5.76 45.26 ± 2.30 aAll reactions were done in a buffer containing a MgATP regeneration system. For assays using dithionite as the reductant, 10 mM DT was used. And for assays using flavodoxin as the reductant, 600 µM of flavodoxin and 10 mM DT were used. All reactions were done at 30oC for 30 sec. bProtein concentrations were 1.95 µM of Fe protein and 2.08 µM of MoFe protein ([Fe]:[MoFe] ≈ 1:1). cProtein concentrations were 8 µM of Fe protein and 0.41 µM of MoFe protein ([Fe]:[MoFe] ≈ 20:1).

205

Figure C1. Acetylene reduction by wild-type Mo-nitrogenase using DT or Fld as reductants under different electron flux conditions. Low flux: 1.95 µM of Fe protein and 2.08 µM of MoFe protein ([Fe]:[MoFe] ≈ 1:1) in the reaction mixture. Units for specific activity of acetylene reduction are nmol ethylene/min/mg Fe protein. High flux: 8 µM of Fe protein and 0.41 µM of MoFe protein ([Fe]:[MoFe] ≈ 20:1). Units for specific activity of acetylene reduction are nmol ethylene/min/mg MoFe protein. All assays were done in the presence of 0.1 atm of acetylene and 0.9 atm of argon. Reaction mixtures contained 12 mM DT or 600 µM FldHQ with 12 mM DT as reductants.

206

Figure C2. Electron flux dependence on proton reduction activity by wild-type Mo- nitrogenase using DT or Fld as reductants. The assays were done in the presence of 1.95 µM of Fe protein with varied MoFe protein concentrations under argon. See Figure B3-1 for additional information on reaction conditions.

207

Figure C3. Electron flux dependence on N2 reduction by wild-type Mo-nitrogenase using DT or Fld as reductants. Specific activities are plotted in terms of NH3 (left panel) and H2 (right panel) production. Assays were done at 1 atm N2. See Figure B3-1 for additional information on reaction conditions.

208

Figure C4. Electron flux dependence on the total activity of proton and N2 reduction by wild-type Mo-nitrogenase using DT or Fld as reductants under N2. The total activity of Fe protein with different MoFe protein concentration was calculated by adding the electrons diverted to NH3 production to that for H2 production in Figure B3-3, and is plotted here as H2 evolution activity.

209

Figure C5. Electron flux dependence on MoFe protein activity for N2 reduction using DT or Fld as reductants. The assays were done in the presence of 0.41 µM of MoFe protein with varied Fe protein concentrations under 1 atm N2. Other conditions are the same as that used in Figure B3-1.

210

Figure C6. Determination of the isosbestic point in UV-vis spectrum for the reversible conversion of FldHQ to FldSQ. The spectra were recorded at a scan frequency of 1 min-1 HQ from the oxidation of Fld by O2 with the blue trace showing the first no O2 spectrum (time 0 min) and the red trace showing the last spectrum after incremental oxidation by

O2 for 16 min.

211

Figure C7. Left: Front and side view of electrostatic potentials for the bound complex between the Fe protein and Fld generated in PyMOL. Right: Side and top view of electrostatic potentials for the individual proteins with complex docking sites shown by the arrows.

212

Figure C8. Inhibition of nitrogenase activity by high concentration of flavodoxin under different electron flux conditions. The electron flux condition and protein concentrations for low flux and high flux experiments are the same as in the Figure B3-1 legend. Reductants are 12 mM DT or 2.4 mM Fld plus 12 mM DT. The experiments were carried out for proton reduction under argon, and the total amount of H2 after a 60 second incubation at 30o C were plotted.

213

Figure C9. Time course study of ATP hydrolysis and electron utilization for proton reduction under low electron flux (panel A, [Fe]:[MoFe] ≈ 1:1) and high electron flux (panel B, [Fe]:[MoFe] ≈ 20:1) conditions using DT or Fld as reductants. The amounts of produced ADP and electrons for H2 evolution under argon are plotted versus reaction time. The reaction conditions are shown in Table 3-2.

214 0300 OLD MAIN HILL • LOGAN, UTAH 84322 PHONE: 208.406.3546 • E- MAIL: [email protected] RHESA LEDBETTER

SUMMARY

Dedicated and enthusiastic scientist and teacher. Experience in the areas of biochemistry, microbiology, general biology, and general chemistry. Innovative and dynamic instructor with solid teaching experience at the collegiate level, including classroom instruction and training undergraduate and graduate researchers. Applies a number of different strategies to enhance student experience and performance. Strong science communication skills, serving as a science reporter and editor. Self-starter who finds reward in focusing on task, yet remains resilient and adaptable to changing needs of the job. A team player with well-developed communication and management skills. International work experience in Australia and China. Distinguished recipient of several awards and grants.

EDUCATION

Ph.D., Biochemistry May 2018 Utah State University, Logan, Utah Dissertation title: Electron flow and management in living systems: Advancing understanding of electron transfer to nitrogenase M.S., Microbiology August 2007 Idaho State University, Pocatello, Idaho Thesis title: ‘Microbial life signatures’: Novel biogenic mineral production and enzyme detection in microorganisms B.S., Microbiology May 2005 Idaho State University, Pocatello, Idaho Senior thesis title: Isolation and identification of a novel arsenic-metabolizing organism from the Alvord Basin hydrothermal system A.S., Natural Science May 2003 College of Southern Idaho, Twin Falls, Idaho

TEACHING AND COORDINATING EXPERIENCE

Utah State University, Logan, Utah 2013-2018 Department of Chemistry and Biochemistry Guest Lecturer, 2016-2018

• Lead Introductory Biochemistry Lectures Graduate Teaching Assistant, 2013-2014

• Instructed Chemical Principles Laboratory II

215 Mohave Community College, Bullhead City, Arizona 2011-2013 Resident Science Faculty

• Designed course materials and taught Introductory Biology and Microbiology Lectures and Laboratories • Produced curriculum for summer forensic science course aimed at high school students • Received a $10,000 National Park Teachers’ Grant in collaboration with the National Park Service to enhance student field microbiology experience Duke Talent Identification Program, Sarasota, Florida 2012 Summer Program Instructor

• Developed curriculum and taught an Introduction to Laboratory Sciences course for gifted junior high school students Jiaxing University, Jiaxing, China 2010 Department of Foreign Languages Lecturer

• Taught English as a Second Language Question Master, Jiaxing University English Speaking Contest and 5th Annual English Speaking Contest of Zhejiang Province • Evaluated ESL students on prepared and impromptu speeches Idaho State University, Pocatello, Idaho 2005-2009 Department of Biological Sciences Assistant Lecturer, 2007-2009

• Taught Introductory Microbiology Laboratory, General Microbiology Laboratory, Microbial Physiology Lecture and Laboratory, and Special Topics in Microbiology Microbiology Lab Coordinator, 2007-2009

• Supervised Microbiology Laboratory teaching assistants • Developed and updated Microbiology Laboratory course curricula on a continuous basis • Maintained live culture collection • Ordered and organized Introductory and General Microbiology Laboratory materials Graduate Teaching Assistant, 2005-2007

• Facilitated the Introductory Microbiology and Microbial Physiology Laboratories

RESEARCH EXPERIENCE

Utah State University, Logan, Utah 2013-2018 Department of Chemistry and Biochemistry Graduate Research Assistant, Advisor Dr. Lance Seefeldt

• Investigates electron flow and management in living systems • Produces value-added products from yeast grown on industrial effluent • Mentors undergraduate and new graduate students • Writes and contributes scientific data to peer-reviewed manuscripts Cold Spring Harbor Laboratory, Cold Spring Harbor, New York 2014 Summer Course Participant, Instructors Dr. Maitreya Dunham, Dr. Marc Gartenberg, and Dr. Grant Brown

• Performed experiments focused on classical and modern yeast genetics techniques

216 Idaho State University, Pocatello, Idaho 2003-2007 Department of Biological Sciences Graduate Research Assistant, Advisors Dr. Tim Magnuson and Dr. Gene Scalarone, 2005-2007

• Characterized a novel arsenic-metabolizing bacterium • Developed new methods to detect enzyme activities in microbial biofilm samples • Mentored undergraduate students working in the laboratory Undergraduate Research Assistant, Advisor Dr. Tim Magnuson, 2003-2005

• Discovered a novel arsenic-metabolizing bacterium • Assisted research scientists with media preparation and other basic tasks University of Melbourne, Melbourne, Australia 2006 Department of Microbiology and Immunology Summer Research Assistant, Advisor Dr. Peter Janssen

• Conducted studies on how to increase the culturability of soil microorganisms Savannah River Ecology Laboratory, Aiken, South Carolina 2004 Summer Research Assistant, Advisors Dr. Andrew Neal and Dr. Christopher Bagwell

• Researched the impacts of metal contamination on microbial populations at the Savannah River Site

SCIENCE COMMUNICATION EXPERIENCE

Utah Public Radio 2016-2018 National Public Radio Affiliate Volunteer Science Reporter

• Shares science occurring throughout the state of Utah with the general public through written articles and radio stories Frontiers in Energy Research 2015 Department of Energy Volunteer Writer and Editor

• Showcased the accomplishments and activities of Energy Frontier Research Centers to the scientific community and general public through written articles

PEER-REVIEWED PUBLICATIONS

Poudel, S., D.R. Colman, K.R. Fixen, R.N. Ledbetter, Y. Zheng, N. Pence, L.C. Seefeldt, J.W. Peters, C.S. Harwood, and E. Boyd. 2018. Electron transfer to nitrogenase in different genomic and metabolic backgrounds. Journal of Bacteriology, 200:e00757-17. Zheng, Y., D. Harris, Y. Fu, Z. Yu, S. Poudel, R.N. Ledbetter, K.R. Fixen, Z.Y. Yang, E.S. Boyd, M.E. Lidstrom, L.C. Seefeldt, C.S. Harwood. 2018. A pathway for biological methane production using bacterial iron-only nitrogenase. Nature Microbiology, 3:281-286. Garcia Costas, A.M., S. Poudel, A.F. Miller, G.J. Schut, R.N. Ledbetter, K.R. Fixen, L.C. Seefeldt, M.W.W. Adams, C.S. Harwood, E.S. Boyd, and J.W. Peters. 2017. Defining electron bifurcation in the Electron-Transferring Flavoprotein Family. Journal of Bacteriology, 199:e00440-17.

217 Pence, N., M. Tokmina-Lukaszewska, Z.Y. Yang, R.N. Ledbetter, L.C. Seefeldt, B. Bothner, and J.W. Peters. 2017. Unraveling the interactions of the physiological reductant flavodoxin with the different conformations of the Fe protein in the nitrogenase cycle. Journal of Biological Chemistry, 292:15661-15669. Ledbetter, R.N., A.M. Garcia Costas, C.E. Lubner, D.W. Mulder, M. Tokmina-Lukaszewska, J.H. Artz, A. Patterson, T.S. Magnuson, Z.J Jay, H.D. Duan, J. Miller, M.H. Plunkett, B.M. Barney, R.P. Carlson, A.F. Miller, B. Bothner, P.W. King, J.W. Peters, and L.C. Seefeldt. 2017. Electron bifurcation by the FixABCX complex of Azotobacter vinelandii: Generation of low- potential reducing equivalents for nitrogenase catalysis. Biochemistry, 56:4177-4190. Yang, Z.Y., R. Ledbetter, S. Shaw, N. Pence, M. Tokmina-Lukaszewska, B. Eilers, Q. Guo, B. Bothner, J. Peters, and L.C. Seefeldt. 2016. Evidence that the Pi release event is the rate-limiting step in the nitrogenase catalytic cycle. Biochemistry, 55:3625-3635. Jena, U., A.T. McCurdy, A. Warren, H.M. Summers, R.N. Ledbetter, S.K. Hoekman, L.C. Seefeldt, and J.C. Quinn. 2015. Oleaginous yeast platform for producing biofuels via co-solvent hydrothermal liquefaction. Biotechnology for Biofuels, 8:167-186. Summers, H., R.N. Ledbetter, A.T. McCurdy, M.R. Morgan, L.S. Seefeldt, U. Jena, S.K. Hoekman, and J.C. Quinn. 2015. Techno-economic feasibility and life cycle assessment of dairy effluent to renewable diesel via hydrothermal liquefaction. Bioresource Technology, 196:431- 440. Ledbetter, R.N., S.A. Connon, A.L. Neal, A. Dohnalkova, and T.S. Magnuson. 2007. Biogenic mineral production by a novel arsenic-metabolizing thermophilic bacterium from the Alvord Basin, Oregon, USA. Applied and Environmental Microbiology, 73:5928-5936.

BOOK CHAPTERS

Magnuson, T.S. and R.N. Ledbetter. 2016. Experimental geomicrobiology: From field to laboratory. In: Yates, M., C. Nakatsu, R. Miller, and S. Pillai (Eds.). Manual of Environmental Microbiology, 4th edition. American Society for Microbiology (ASM) Press, Washington, DC. 5.2.2-1- 5.2.2-7. Ledbetter, R.N. and T.S. Magnuson. 2009. Geomicrobiology of arsenic. In: Barton, L., M. Mandl, and A. Loy (Eds.). Geomicrobiology: Molecular and environmental perspective. Springer Publishing, New York. 147-168.

NEWS ARTICLES AND RADIO STORIES

Ledbetter, R.N. 2018. A dynamic duo: Partnership between musicians and public land agencies inspires ‘music for nature’. Utah Public Radio. Ledbetter, R.N. 2017. Crossing borders: A mother and her transgender son reflect. Utah Public Radio. Ledbetter, R.N. 2017. A state of addiction: The search for alternative painkillers. Utah Public Radio. Ledbetter, R.N. 2017. Crossing borders: Merrell bootmaker now gets people back on their feet…literally. Utah Public Radio.

218 Ledbetter, R.N. 2017. ‘They’re built on love’: Raising children with Williams syndrome. Utah Public Radio. Ledbetter, R.N. 2017. Sowing seeds of goodness: Utah couple uses their passions to raise money for public radio. Utah Public Radio. Ledbetter, R.N. 2017. Shower the National Weather Service with data by being a backyard weather observer. Utah Public Radio. Ledbetter, R.N. 2017. Dr. Scott the Paleontologist keeps on inspiring young minds with new book. Utah Public Radio. Ledbetter, R.N. 2017. Invisible companions: Scientists begin to uncover how bacteria strike up friendly relationships. Utah Public Radio. Ledbetter, R.N. 2017. Nature is calling: Satisfy your curiosity and contribute to scientific research using iNaturalist. Utah Public Radio. Ledbetter, R.N. 2017. Southern Utah, a geological hot spot: Dazzling uranium-rich minerals discovered in Red Canyon. Utah Public Radio. Ledbetter, R.N. 2017. A dose of hope: Individualized treatments lend promise to those battling lung cancer. Utah Public Radio. Ledbetter, R.N. 2017. Protecting teens from HPV: The vaccine and what one parent learned on her journey to know more. Utah Public Radio. Ledbetter, R.N. 2016. Religion and the brain: What happens when Mormons “feel the spirit?” Utah Public Radio. Ledbetter, R.N. 2016. Fueling the future with…our waste. Utah Public Radio. Ledbetter, R.N. 2016. Got a sweet tooth? Unlock the magic of customized, 3D printed candy. Utah Public Radio. Ledbetter, R.N. 2016. Did dinosaurs really sound like the movies? Scientists offer insight based on ancient vocal organ. Utah Public Radio. Ledbetter, R.N. 2016. Ancient creature ‘arm’ed in unusual fashion: Reptile displayed unique forelimbs with giant claw. Utah Public Radio. Ledbetter, R.N. 2016. Alan Alda takes a leading role, inspires scientists to be effective communicators. Utah Public Radio. Ledbetter, R.N. 2016. Nature’s teaching tools: Monarch butterflies brighten up a Utah classroom. Utah Public Radio. Ledbetter, R.N. 2016. Snail insulin picks up speed for diabetic community. Utah Public Radio. Ledbetter, R.N. 2016. Cache Valley’s newest mountain trail: A community endeavor. Utah Public Radio. Ledbetter, R.N. 2016. Back-to-school critters: Combating treatment-resistant ‘super lice’. Utah Public Radio. Ledbetter, R.N. 2016. Feast or famine: Why did hunter-gatherers start farming anyway? Utah Public Radio. Ledbetter, R.N. 2016. Scientists use the power of predication to save species from human space invasion. Utah Public Radio.

219 Ledbetter, R.N. 2016. Soaking up science: Students, educators, and scientists conduct Utah water research. Utah Public Radio. Ledbetter, R.N. 2016. A sustainable education: Integrating global thinking into classroom curriculum. Utah Public Radio. Ledbetter, R.N. 2016. Getting fired up for the next wave of space exploration. Utah Public Radio. Ledbetter, R.N. 2016. Memorials matter: The impressions history leaves behind. Utah Public Radio. Ledbetter, R.N. 2016. Desert survivor: Moss quenches thirst using optimized water collection system. Utah Public Radio. Ledbetter, R.N. 2016. Infection detection: Scientists unveil new tool for diagnosing infectious diseases. Utah Public Radio. Ledbetter, R.N. 2016. Digging up our prehistoric past: New horned dinosaur discovered in Southern Utah. Utah Public Radio.

Ledbetter, R.N. 2016. H2Oath: Utahns encouraged to take water conservation pledge. Utah Public Radio. Ledbetter, R.N. 2016. The story behind the scientist: One who beat the odds. Frontiers in Energy Research. Ledbetter, R.N. 2015. Getting to know scientists undertaking energy challenges. Frontiers in Energy Research.

ORAL PRESENTATIONS

Ledbetter, R.N. 2014. Untapped Treasure: Upgrading industrial effluent into value-added products. Cold Spring Harbor Laboratory, Cold Spring Harbor, NY. Ledbetter, R.N. 2012. Bringing instruction to life. Mohave Community College All Staff Day Meeting, Lake Havasu City, AZ. Ledbetter, R.N., A.L. Neal, and C.E. Bagwell. 2004. Effects of metal contamination on bacterial populations in a riparian environment. Savannah River Ecology Laboratory, Aiken, SC.

POSTER PRESENTATIONS

Zheng, Y., R.N. Ledbetter, L.C. Seefeldt, D.F. Harris, S. Poudel, K.R. Fixen, Z.Y. Yang, P.W. King, E.S. Boyd, and C.S. Harwood. 2017. FeFe nitrogenase: C-H and N-H bond formation. Energy Frontier Research Centers-Energy Innovation Hubs-Computation Materials Sciences Principle Investigators’ Meeting. Washington, DC. Ledbetter, R.N., K. Fixen, Y. Zheng, D. Harris, S. Shaw, Z. Yang, Z. Jay, R. Carlson, L.C. Seefeldt, and C. Harwood. 2015. Light driven C-H bond formation. Energy Frontier Research Centers Principle Investigators’ Meeting. Washington, DC. Ledbetter, R.N., A.T. McCurdy, M.R. Morgan, and L.C. Seefeldt. 2014. Single-cell protein production from yeast grown on industrial effluent. American Society for Microbiology, Boston, MA.

220 Ledbetter, R.N., A.T. McCurdy, M.R. Morgan, H. Summers, J.C. Quinn, and L.C. Seefeldt. 2014. Using low-cost effluent to make value-added products. UNLV-USU Symposium on Biotechnology, Renewable Energy, and Novel Materials, Las Vegas, NV. Ledbetter, R.N., A.T. McCurdy, M.R. Morgan, and L.C. Seefeldt. 2014. The effect of physiologic conditions on yeast-based lipid production. National Biodiesel Conference and Expo, San Diego, CA. Pharis, J., M.J. Meehan, R.N. Ledbetter, and T.S. Magnuson. 2009. Selenium-transforming bacteria from river sediments and thermal areas. 109th General Meeting, American Society for Microbiology, Philadelphia, PA. Ledbetter, R.N., L.A. Deobald, A.J. Paszczynski, and T.S. Magnuson. 2007. Directed proteomics applied to the detection and characterization of arsenic-transforming enzymes from complex communities in the Alvord Basin hydrothermal system. 17th Annual Goldschmidt Conference, Cologne, Germany. Ledbetter, R.N., M. James, A.L. Neal, and T.S. Magnuson. 2006. Microbial ‘Life Signatures’: Novel biogenic mineral production and enzyme detection in microorganisms. 16th Annual Goldschmidt Conference, Melbourne, Australia. James, M., R.N. Ledbetter, A.L. Neal, and T.S. Magnuson. 2006. Biogenic mineral production as evidence for life. Idaho State University Undergraduate Research Symposium, Pocatello, ID. Ledbetter, R.N., S.A. Connon, A.L. Neal, and T.S. Magnuson. 2005. Isolation of a novel arsenic-metabolizing organism from the Alvord Basin hydrothermal system. The Joint International Symposia for Subsurface Microbiology (ISSM 2005) and Environmental Biogeochemistry (ISEB XVII), Jackson Hole, WY. Clawson, J.R., R.N. Ledbetter, S.A. Connon, and T.S. Magnuson. 2004. Microbial sulfate reduction in a hydrothermal environment. Idaho Academy of Science, Pocatello, ID.

PROFESSIONAL DEVELOPMENT

ComSciCon-Rocky Mountain West, University of Colorado, Boulder, 2017. Science Communication Workshop, Alan Alda Center for Communicating Science, 2016. Cold Spring Harbor Laboratory Yeast Genetics and Genomics Course, Cold Spring Harbor Laboratory, 2014. The Teaching Professor Workshop: Infusing Learner-Centered Strategies, Austin, Texas. 2013. Park Teachers’ Workshop, Everglades National Park, 2012.

SERVICE

Energy Frontier Research Center Early Career Network Representative, Department of Energy, 2016-2017. Biological Electron Transfer and Catalysis Progress Review Panel, Department of Energy, 2016-2018.

221 SCIENCE GRANTS, AWARDS, AND HONORS

Thomas F. Emery Outstanding Graduate Student in Biochemistry, Utah State University, 2018. Student Travel Award, American Society for Microbiology, 2014. Next Generation Scientists for Biodiesel Scholarship, National Biodiesel Board, 2014. Park Teachers’ Grant, National Park Foundation, 2012. Goldschmidt Travel Grant, Geochemical Society and the National Science Foundation, 2007. Goldschmidt Travel Grant, Geochemical Society and the National Science Foundation, 2006. Graduate Research Fellowship, NASA Idaho Space Grant Consortium, 2006. East Asia and Pacific Summer Institutes for U.S. Graduate Students, University of Melbourne, National Science Foundation, 2006. Graduate Research Fellowship, NASA Idaho Space Grant Consortium and NASA Idaho EPSCoR Fellowship Program, 2005. Research Experience for Undergraduates, Savannah River Ecology Laboratory, University of Georgia, National Science Foundation, 2004.

RADIO AWARDS

Mark of Excellence Award: Radio Feature, 1st place (Intermountain West) for “‘They’re built on love’: Raising children with Williams syndrome,” Society of Professional Journalists, 2018. Mark of Excellence Award: Radio In-depth Reporting, 1st place (Intermountain West) for “A state of addiction: The search for alternative painkillers,” Society of Professional Journalists, 2018. Mark of Excellence Award: Radio Feature, Finalist (Intermountain West) for “Crossing borders: A mother and her transgender son reflect,” Society of Professional Journalists, 2018. Radio Education Reporting, 2nd place for “Monarch butterflies brighten up a Utah classroom,” Society of Professional Journalists, Utah Chapter, 2017.