<<

1 TITLE: Structural basis of interprotein electron transfer in bacterial sulfite

2 oxidation.

3 AUTHORS: Aaron P. McGratha,1, Elise M. Lamingb,2, G. Patricia Casas Garciac,

4 Marc Kvansakulc, J. Mitchell Gussb, Jill Trewhellab, Benoit Calmesd,e, Paul V.

5 Bernhardtd,e, Graeme R. Hansond,f,3, Ulrike Kapplerd,e,4, Megan J. Maherb,4

6

7 AFFILIATIONS:

8 aStructural Biology Program, Centenary Institute, Locked Bag 6, NSW 2042,

9 Australia

10 bSchool of Molecular Bioscience, University of Sydney, NSW, 2006, Australia.

11 cLa Trobe Institute for Molecular Science, La Trobe University, Melbourne, 3086,

12 Australia.

13 dCentre for Metals in Biology, eSchool of Chemistry and Molecular Biosciences and

14 fCentre for Advanced Imaging, The University of Queensland, Brisbane, QLD 4072,

15 Australia.

16 1Present address: Skaggs School of Pharmacy and Pharmaceutical Sciences,

17 University of California at San Diego, La Jolla, CA, 92037, USA.

18 2Present address: The Victor Chang Cardiac Research Institute, Sydney, NSW, 2010,

19 Australia.

20 3Deceased

21 4Correspondence: [email protected] (MJM)

22 [email protected] (UK)

23 COMPETING INTERESTS: The authors declare no financial or non-financial 24 competing interests.

1 25 ABSTRACT

26 Interprotein electron transfer underpins the essential processes of life and relies on the

27 formation of specific, yet transient protein-protein interactions. In biological systems,

28 the detoxification of sulfite is catalyzed by the sulfite-oxidizing (SOEs),

29 which interact with an electron acceptor for catalytic turnover. Here, we report the

30 structural and functional analyses of the SOE SorT from Sinorhizobium meliloti and

31 its cognate electron acceptor SorU. Kinetic and thermodynamic analyses of the

32 SorT/SorU interaction showed the complex is dynamic in solution, and that the

33 proteins interact with Kd = 13.5 ± 0.8 μM. The crystal structures of the oxidized SorT

34 and SorU both in isolation and in complex, reveal the interface to be remarkably

35 electrostatic, with an unusually large number of direct hydrogen bonding interactions.

36 The assembly of the complex is accompanied by an adjustment in the structure of

37 SorU and conformational sampling provides a mechanism for dissociation of the

38 SorT/SorU assembly.

39

2 40 INTRODUCTION

41 Although electron transfer reactions are key biochemical events, which underpin

42 fundamental processes, such as respiration and photosynthesis, the study of the

43 molecular details of the interprotein interactions at their core can be largely

44 intractable. In particular, atomic resolution crystal structures of electron transfer

45 complexes rare, due to their fundamentally transient nature (1). Electron transfer

46 pathways are made up of chains of proteins, which provide a path for the

47 controlled flow of electrons and rely on efficient docking of protein redox partners

48 through noncovalent, dynamic protein-protein interfaces (2). Complementary

49 electrostatic surfaces, hydrophobic interactions and dynamics at the protein-protein

50 interface have all been proposed to contribute to efficient interprotein electron transfer

51 (19), with a strong correlation between the driving force for the reaction, the distance

52 between redox centers and the rate of electron transfer (2, 20).

53

54 Interprotein electron transfer processes are central to the redox conversions of cellular

55 compounds, which are an evolutionarily ancient type of metabolism that has

56 existed as long as cellular life (3, 4). Sulfur-containing compounds mediate many

57 crucial reactions in the cell (for example, in , sulfur containing amino

58 acids or ), but their reactivity also makes them potentially toxic (5). Sulfite

59 in particular, is a highly reactive sulfur compound that can cause damage to proteins,

60 DNA and lipids, resulting in oxidative stress and irreversible cellular damage (6). In

61 most cells, the detoxification of sulfite by oxidation to (Eq.(1)) is catalyzed by

62 sulfite oxidizing enzymes (SOEs) (7).

2- 2- + - 63 SO3 + H2O  SO4 + 2H + 2e (1)

3 64 SOEs from plants, higher animals and bacteria have been characterized and they all

65 catalyze the same fundamental reaction. However, their cellular functions, catalytic

66 properties (5, 6, 8-10) and the identities of their natural electron acceptors vary

67 significantly. Some SOEs transfer electrons to (11), while others interact with

68 redox proteins such as c (7, 8, 12-16) or as yet unknown cellular

69 components (5, 12, 13, 17). To date, three unique crystal structures of SOEs have

70 been reported: from chicken, plant and bacteria, which differ significantly in their

71 domain architectures and redox compositions. However, none of these

72 studies show details of a SOE in complex with its external electron acceptor (7, 11,

73 21). At present, no structural information on the molecular interactions of any of these

74 enzymes with their respective electron acceptors is available and the determinants that

75 dictate the type of electron acceptor individual SOEs employ, while maintaining the

76 efficiency of the basic reaction are open questions. (5, 18)

77

78 Here, we have investigated an electron transfer complex involving the periplasmic

79 SorT from the α-Proteobacterium Sinorhizobium meliloti,

80 which represents a structurally uncharacterized type of SOE, and its electron acceptor,

81 the c-type cytochrome SorU (12, 17). In S. meliloti the SorT sulfite dehydrogenase is

82 part of a sulfite detoxification system that is induced in response to the degradation of

83 sulfur containing substrates such as organosulfonate taurine (17). Electrons derived

84 from sulfite oxidation are passed on to the SorU cytochrome, and likely then to

85 cytochrome oxidase, as S. meliloti is capable of sulfite respiration (12). Here we

86 report the crystal structures of both the isolated SorT and SorU proteins and the

87 biochemical and structural analyses of the SorT/SorU electron transfer complex. This

4 88 is the first time that a crystal structure of a enzyme in complex with its

89 external electron acceptor has been solved.

90

91 RESULTS 92

93 The interactions between SorT and SorU are highly dynamic and efficient in

94 solution.

95 Sulfite-oxidizing enzymes, particularly those from bacteria, are known to be highly

96 efficient catalysts (5, 22). Previous work has established that SorT is able to transfer

97 electrons to the SorU cytochrome that is encoded on the same operon, however no

98 kinetic details of the interaction were reported (12). With the artificial electron

99 acceptor ferricyanide, SorT was shown to have turnover number of 338 ± 3 s-1 (12,

100 17). Employing SorU as the , we analyzed the kinetics of the interaction

101 between SorT and SorU and found the interaction to be fast and highly specific, with

-1 102 a KM(SorU) of 32 ± 5 μM and a kcat of 140 ± 11 s , confirming that SorU is the natural

103 electron acceptor of SorT. Measurement of the thermodynamics of the SorT/SorU

104 interaction by isothermal titration calorimetry (ITC) revealed a dissociation constant

105 of Kd = 13.5 ± 0.8 μM with a determined stoichiometry of 0.8 ± 0.2. These values are

106 in the range observed for other electron transfer complexes (23, 24) and match a

107 model where SorT sequentially transfers two electrons, derived from sulfite oxidation,

108 to two SorU molecules. In other words, the SorT/SorU complex must form twice

109 (with two different ferric SorU protein molecules) to complete the oxidative half

110 reaction of SorT.

111

112 The KM of SorT for SorU is very close to its affinity for sulfite (KM=15.5 ± 1.9 μM

113 (17)), and the turnover number in the SorU-based assay is approx. 40% of that seen

5 114 with ferricyanide as the electron acceptor (17), where no significant reorientation and

115 docking of the electron acceptor is required. These data reveal interesting details

116 about the formation of the SorT/SorU complex, which appears to form with similar

117 affinities between the two proteins when both are oxidized as in the ITC experiments

118 and in a system where SorT constantly undergoes oxidation and reduction (SorU-

119 based enzyme assay). In addition, the similarity between the determined values of Kd

120 and KM indicates that the affinity of SorT for SorU is unaffected by the presence of

121 substrate or . However, the kinetic parameters for this interaction are clearly

122 distinct from those for other sulfite-oxidizing enzymes, such as the bacterial SorAB

123 enzyme or chicken (CSO), both of which have much higher affinities

124 for their respective electron accepting c (both with KM(Cyt c) ca. 2 μM)

-1 125 (25). The catalytic turnover of CSO is relatively slow (kcat = 47.5 ± 1.9 s ), due to a

126 mechanism requiring internal rearrangement, while turnover of SorAB with its natural

127 electron acceptor (334 ± 11 s-1) is significantly faster than that for SorT (22, 25).

128

129 The crystal structure of the SorT homodimer reveals a head to tail subunit

130 arrangement

131 In order to investigate whether there are any structural reasons for these differences,

132 we solved the crystal structure of SorT by molecular replacement and refined it to 2.4

133 Å resolution. The structure shows two homodimeric assemblies per asymmetric unit

134 (Table 1), which is in agreement with the quaternary structure as determined by

135 MALLS (12, 17). Unexpectedly however, within the SorT homodimer the protomers

136 are oriented in a head-to-tail orientation (Figure 1A), a subunit arrangement that has

137 not been previously observed in structures of sulfite-oxidizing enzymes. In keeping

138 with the nomenclature applied to other structurally-characterized SOEs, a

6 139 ‘dimerization’ domain typically defines the interface between the two monomers (7,

140 11), but the structure of the SorT dimer does not follow this paradigm.

141

142 Nevertheless, the fold of the SorT monomers is similar to those of other SOEs (7, 11,

143 21), comprising a central ‘SUOX-fold’ domain (26) that harbors the Mo

144 and a ‘dimerization’ domain (Figure 1A). The SorT active site has a square-

145 pyramidal geometry seen in all other SOE structures with a five coordinate

146 molybdenum atom and a single tricyclic pyranopterin cofactor (Figure 1A, Table 2)

147 (27).

148

149 Single electron reduction of SorT to its EPR active MoV form was achieved using a

150 combination of Ti(III)citrate and a suite of organic redox mediators. The MoV EPR

151 spectrum is similar to the so-called ‘high pH’ EPR signature of SOEs (Figure 2A,

152 Table 3, Appendix 1). An additional feature is the presence of superhyperfine

153 coupling between the unpaired electron on the Mo ion and two nearby I = ½ nuclei.

154 This implies that the equatorially coordinated O-donor is an aqua ligand at pH 8

155 (Figure 2B(ii)) or that a hydroxido ligand is hydrogen bonded with a water molecule

156 whose proximal H-atom is coupled with the electron spin on Mo (Figure 2B(i),

157 Appendix 1).

158

159 Within the SorT dimer, the subunit interface involves both the SUOX and the

160 ‘dimerization’ domains (Figure 1A), resulting in a buried surface area of ca. 1280 Å2

161 per monomer, which is approximately 9% of the solvent-accessible surface of each

162 monomer. The functional consequences and structural origins of these significant

163 differences among the quaternary structures of SOEs are at this point unknown, as all

7 164 of these enzymes are highly catalytically active, have similar active site structures and

165 no known kinetic that would imply a functional role for the different

166 oligomeric assemblies (8, 10, 28).

167

168 In complex, SorT and SorU form a SorU/SorT2/SorU assembly that reveals a

169 pathway for electron transfer.

170 Despite the dynamic nature of the SorT/SorU interaction, it was possible to co-

171 crystallize SorT with SorU, resulting in the crystal structure of the SorT/SorU

172 complex, where a single SorT/SorU entity is present per asymmetric unit, and the

173 application of crystallographic 2-fold symmetry reveals a SorU/SorT2/SorU assembly

174 (Figure 3A, Table 1). Small angle X-ray scattering (SAXS) with a sample prepared

175 as a stoichiometric mixture of SorT and SorU (Figure 3B, Table 4, Appendix Figure

176 1) confirmed that the structure observed in the crystal is preserved in solution.

177

178 The central assembly within the SorU/SorT2/SorU complex is the SorT homodimer,

179 which is identical to the structure of the SorT homodimer alone (Figures 1A and 3A,

180 Table 1). The structure of the SorU protein, both within the SorT/SorU complex and

181 when crystallized in isolation (Figure 1B, Table 1) is predominantly α-helical with

182 three major α-helices arranged to form a bundle that frames the -

183 (Figure 1B) (C47XXC50H, axial ligands: His 51 and Met 87).

184

185 In the SorT/SorU complex, the SorU protein docks within a pocket adjacent to the

186 SorT active site, with the heme cofactor located at the protein-protein interface

187 (Figures 3A and 4A). The shortest ‘edge-to-edge’ distance between the SorT Mo

188 atom and the propionate group from the SorU heme c cofactor is 8.2 Å, which is well

8 189 within the distance for fast electron transfer through the protein medium (31).

190 PATHWAY analysis (32) (Table 5) further indicates that the dominant electron

191 tunneling pathway from SorT to SorU proceeds from the Mo atom, via the

- 192 coordinating H2O/OH , to the guanidinium group of SorT residue Arg 78 and across

193 the protein-protein interface to the heme propionate group and to the pyrrole ring of

194 heme c to the heme iron (Figure 4B, Table 5).

195

196 Conformational change reminiscent of an ‘induced fit’ mechanism facilitates

197 docking and electron transfer between SorT and SorU.

198 Specific structural adaptations of the SorU protein take place when the SorT/SorU

199 assembly is formed (Figure 5B). A surface loop on SorU (residues 82-93) moves

200 away from the SorT/SorU complex interface, leading to a reorientation of the heme

201 ligand residue Met 87 so that a different Met 87 rotamer coordinates the iron in the

202 SorU structures within and outside of the complex (Figure 5B). This change in the

203 structure of SorU is required to allow a Mo-heme edge-to-edge distance of ‘closest

204 approach’ of ca. 8 Å within the SorT/SorU assembly. Without this adjustment (for

205 example, if the SorU residue 82-93 loop structure remained rigid) the closest

206 approach for the redox cofactors would be ca. 10 Å.

207

208 Changes in the conformation of axial heme ligands are known to alter the orbital

209 interactions between the iron atom and the ligand, which can change the redox

210 properties of the heme group (33). However, this does not appear to be the case here

211 as the redox potential of the SorU heme was determined by optical

212 spectroelectrochemistry to be +108(±10) and +111(±10) mV vs. NHE (pH 8.0),

213 respectively, in the presence and absence of SorT (Table 6, Figure 6B).

9 214

215 The MoVI/V and MoV/IV redox potentials of SorT (+110(±10) mV and -18(±10) mV vs.

216 NHE (pH 8)) were determined by an EPR-monitored redox titration where the initial

217 EPR-silent MoVI form is reduced to the EPR-active MoV state (Figure 2A), which

218 then gives way to the EPR-silent MoIV form at low potential, resulting in a bell-

219 shaped curve (Table 6, Figure 6A). The MoVI/V redox potential at pH 8 matches that

220 of the ferric/ferrous SorU couple.

221

222 The structural change in SorU indicates that an ‘induced fit’ mechanism is responsible

223 for the formation of a productive SorT/SorU electron transfer complex. This type of

224 mechanism has in the past been used to describe electron transfer complexes

225 (involving electron transfer flavoproteins or ferredoxin reductases) where the protein

226 partners include mobile domains, and where conformational change is necessary for

227 the creation of high affinity protein-protein interfaces (34, 35). However, although

228 facilitating redox interactions, these systems are distinct from the docking mechanism

229 seen for the SorT and SorU proteins which accompanies modifications to the structure

230 of SorU and allows the two redox centers to attain positions of closest approach for

231 fast electron transfer.

232

233 The SorT/SorU interface features extensive electrostatic interactions.

234 The SorT/SorU complex interface shows significant charge complementarity, with the

235 negative charge on SorU correlating with a concentration of positive charges at the

236 SorU binding site on SorT (Figures 3C, 3D). Unlike other known cytochromes c that

237 can act as electron acceptors to SOEs (12, 30, 36), the electrostatic surface of SorU

238 has an overall negative charge (Figures 3C, 3D). The positive charge on the SorT

10 239 surface therefore explains the low catalytic activity of SorT with horse heart

240 cytochrome c (7 U/mg, pI ~10), the natural electron acceptor for vertebrate SOEs,

241 compared to the high activity observed with SorU (212 U/mg; pI ~4) (5, 12, 17). In its

242 electrostatic nature, the interaction surface between SorT and SorU is unusual.

243 Structures of other cytochrome-containing electron transfer complexes (37-40) show

244 binding interfaces characterized by a ‘ring’ of electrostatic interactions that

245 encompass contact surfaces that are predominantly hydrophobic. In fact, the ‘steering’

246 of electron transfer partners by electrostatic interactions, accompanied by ‘tuning’ via

247 hydrophobic interactions is a dominant observation for protein-protein electron

248 transfer complexes (37-40).

249

250 An unusually large number of hydrogen bonds and salt bridges characterize the

251 SorT/SorU interface.

252 In addition to the electrostatic interactions that support the formation of the

253 SorT/SorU complex, there are six hydrogen bonds found at the SorT/SorU protein-

254 protein interface, as well as a salt bridge between the SorT active site residue Arg 78

255 and a propionate group of the SorU heme moiety (Table 7, Figure 7). This is an

256 unusually large number in comparison with structures of other cytochrome-containing

257 electron transfer complexes (37-40), which tend to have fewer hydrogen bonds and

258 lack salt bridges. In fact, direct hydrogen bonds between electron transfer proteins are

259 generally considered unfavorable for a transient interaction because of energetically

260 disadvantageous desolvation (41). Also significant is the observation that no

261 intermolecular interactions at the interface are mediated by water molecules (Figure

262 7) (37, 42). In fact, very little ordered water is observed in proximity to the interfacing

263 region of the SorT molecule (a total of 2 water molecules only and these are

11 264 hydrogen-bonded to the SorT molecule rather than mediating the SorT/SorU

265 interaction). However, the current analysis is limited by the moderate resolution of the

266 current structure (2.6 Å), which as a consequence, includes only ca. 0.15 modeled

267 water molecules per residue.

268

269 Compared with the subunit interface of the SorAB bacterial sulfite dehydrogenase,

270 which contains 30 hydrogen-bonds and 2 salt bridges (21), supporting a permanent,

271 heterodimeric complex of a heme c subunit (SorB) and a Mo cofactor containing

272 (SorA) catalytic subunit (21) (Table 7), the extent of the subunit interactions in the

273 SorT/SorU structure is modest. The difference between the permanent SorAB and the

274 transient SorT/SorU complex is also illustrated by calculations of the shape

275 complementarity and the buried surface areas between the protomers (43), with the

276 latter being about twice as large for the SorAB assembly than for SorT/SorU (Table

277 7). Interestingly, much of the additional contact area between molecules in the SorAB

278 structure derives from the SorB N-terminal structure (residues B501-B518, PDB

279 2BLF), which extends away from the core of the subunit, wraps around the SorA

280 ‘SUOX’ fold domain and contributes one salt-bridge and 6 hydrogen-bonding

281 interactions (Table 7) (21). This feature is absent from the SorU structure.

282

283 The dynamic SorT/SorU interaction observed in solution is reflected in the

284 crystal.

285 Despite the intricate assembly of interactions at the protein-protein interface, but in

286 agreement with the kinetics and thermodynamics of the SorT/SorU interaction in

287 solution, the SorT/SorU contact is dynamic, as illustrated by a temperature factor

288 analysis of the complex structure. Within the SorT/SorU complex, the SorU protein

12 289 shows a significantly increased average atomic temperature factor (reflecting

290 significant flexibility) relative to the structure of the SorT protomer (1.5 versus 0.9 Å2,

291 respectively; Table 7). Furthermore, the relative temperature factors per residue for

292 the SorU molecule increase with increasing distance from the SorT/SorU interface

293 (Figure 5A), indicating that the SorU molecule is dynamic relative to SorT within the

294 crystalline lattice. In contrast, the ‘static’ SorAB complex shows uniform, low

295 temperature factors for both redox subunits (Table 7, Figure 5A). This observation is

296 an exquisite illustration of ‘conformational sampling’ within the SorT/SorU electron

297 transfer complex, which results from the conformational flexibility of one protein

298 redox partner relative to the other and both facilitates electron exchange by accessing

299 the optimal orientations of each redox partner and promotes fast dissociation of the

300 complex following transfer (19, 44).

301

302 DISCUSSION

303 By describing the structure of the SorT/SorU complex in this work, we report the first

304 example of a structure of an SOE in complex with its external electron acceptor; all

305 previous structures of SOEs being of the enzymes and their internal heme domains or

306 subunits only (7, 11, 21). The structure of the SorT/SorU complex therefore allows

307 insights into electron transfer in what is thought to be a highly prevalent type of

308 bacterial SOE (12) and into protein-protein electron transfer in general. While the

309 complex shows dynamic adaptations similar to those demonstrated previously for

310 electron transfer complexes and has a dissociation constant of the right order of

311 magnitude, it also shows some features that have not been seen in electron transfer

312 complexes, namely an interface that is stabilized by a relatively large number of

313 hydrogen bonds and salt bridges, and an induced fit docking mechanism.

13 314

315 The structure of the protein-protein interface in the SorT/SorU structure is particularly

316 intriguing. The relative lack of bound water molecules, mirrors more the observations

317 made of permanent heterodimeric complexes than transient interactions (42). Previous

318 investigations into the factors that influence protein-protein docking for electron

319 transfer have shown that the strength of the protein-protein interaction correlates

320 linearly with the product of the total charges on the protein partners (45, 46). In this

321 way, the affinity between the SorT/SorU proteins (Kd = 13.5 ± 0.8 μM) correlates

-1 322 with the fast measured turnover rates (kcat of 140 ± 11 s ) and with the predominantly

323 electrostatic nature of the protein-protein interface.

324

325 The turnover number for SorT with SorU as the electron acceptor is also in the range

326 of values measured for the human sulfite oxidase (HSO) and CSO (25.0 ± 1.3 s-1 and

327 47.5 ± 1.9 s-1, respectively), but significantly slower than that observed for the

328 permanent SorAB complex (345 ± 11 s-1) (25, 28, 30). Importantly, in the structure of

329 SorAB, the docking site of the heme subunit (SorB) with the SorA is almost

330 identically positioned to that seen for SorT/SorU, with major differences existing only

331 in the number of hydrogen bonds and salt bridges in the protein-protein interface. For

332 the CSO and HSO enzymes, docking of the mobile domain near the Mo active

333 site has been proposed to be similar to that seen for SorT/SorU (29). It should be

334 noted, however, that the docking events in CSO (and HSO) and between SorT/SorU

335 serve fundamentally different purposes: for CSO and HSO, domain docking enables

336 intramolecular electron transfer and involves a heme domain that is an intrinsic part of

337 the enzyme. This is a step that precedes interactions with the external electron

338 acceptor for these enzymes. In contrast, and despite the fact that it is occupying a

14 339 similar docking site to that predicted for CSO and HSO (29), SorU is the external

340 electron acceptor for SorT and the electron transfer is intermolecular.

341

342 The SorT/SorU complex described here thus represents an elegant compromise

343 between the requirements for fast and efficient electron transfer and reaction

344 specificity. It also illustrates new aspects for highly dynamic protein-protein

345 interactions: (i) A relatively large number of hydrogen bonds and salt bridges may be

346 required to form the initial stable protein complex, but this does not preclude a

347 dynamic protein – protein interaction; (ii) relatively subtle structural adjustments in

348 one redox partner (SorU) can facilitate electron transfer by ideally locating the redox

349 active cofactors in close proximity; (iii) the comparatively complex binding interface

350 in SorT/SorU can be counterbalanced by the conformational sampling of one protein

351 relative to the other, which enables the rapid dissolution of the complex following

352 electron exchange.

353

354 It remains to be seen whether these principles apply to other SOE – external electron

355 acceptor interactions. Future work should focus on investigating interaction interfaces

356 in currently little studied SOEs where new types of interactions may be present as for

357 many of these enzymes currently no external electron acceptor is known.

358

15 359 MATERIALS AND METHODS.

360 Protein overexpression, purification, data collection and structure solution.

361 Recombinant SorT and SorU proteins were overproduced and purified as previously

362 described (12), with minor modifications. SorT was crystallized by hanging drop

363 vapor diffusion with drops consisting of equal volumes (2 μL) of protein and

364 crystallization solution (0.1 M HEPES pH 7.5, 8% ethylene glycol, 0.1 M manganese

365 (II) chloride tetrahydrate and 17.5% PEG 10,000) at 20°C. Crystals were

366 cryoprotected in reservoir solution with 30% glycerol before flash cooling in liquid

367 nitrogen. Small (ca. 20 × 10 × 10 μ) crystals of SorU were grown in drops containing

368 equal volumes (2 μL) of protein and reservoir solution (1.8 M tri-sodium citrate, pH

369 5.5, 0.1 M glycine), which were harvested and flash-cooled in liquid nitrogen without

370 additional cryoprotection. Purified SorT (20 mM Tris pH 7.8, 2.5% glycerol) and

371 SorU (20 mM Tris pH 7.8, 150 mM NaCl) were mixed and incubated on ice at a

372 molar ratio of 2:1 (SorU:SorT; total protein concentration 8 mgmL-1) before

373 crystallization via hanging-drop vapor diffusion with a reservoir solution containing

374 0.2 M sodium formate, 0.1 M Bis-Tris propane pH 7.5 and 20% PEG 3350. Crystals

375 grew to a maximum size of ca. 150 x 100 x 20 μ in 4 days at 20°C and were flash

376 cooled in liquid nitrogen after brief soaking in mother liquor containing 30% glycerol.

377 All diffraction data were collected on an ADSC Quantum 315r detector at the

378 Australian Synchrotron on beamline MX2 at 100 K and were processed with

379 HKL2000 (47). Unit cell parameters and data collection statistics are presented in

380 Table 1.

381

382 The crystal structure of SorT was solved by molecular replacement using PHASER

383 (48) with a search model generated with CHAINSAW (49) from the SorA portion of

16 384 the SorAB crystal structure (29.0% sequence identity, entry 2BLF

385 (21)) as a template (50). The resulting model was refined by iterative cycles of

386 amplitude based twin refinement (using twin operators H, K, L and –H, -K, L with

387 estimated twin fractions of 0.495 and 0.505 respectively) within REFMAC (51),

388 interspersed with manual inspection and correction against calculated electron density

389 maps using COOT (52). The refinement of the model converged with residuals R =

390 0.208 and Rfree = 0.239 (Table 1). The structure of the SorT/SorU complex was

391 solved by molecular replacement using PHASER (48), with the refined SorT structure

392 as a search model. Initial rounds of refinement yielded a difference Fourier electron

393 density map, which clearly showed positive difference density for the location of one

394 molecule of SorU per asymmetric unit, which was manually built using COOT (52).

395 Refinement was carried out with REFMAC5 (51) and PHENIX (53) and converged

396 with residuals R = 0.211 and Rfree = 0.260 (Table 1). The refined SorU model, from

397 the SorT/SorU complex structure, was used as a search model to solve the SorU

398 structure by molecular replacement using PHASER (48). Refinement was carried out

399 with REFMAC5 and PHENIX (53) and converged with residuals R = 0.192 and Rfree

400 = 0.240. All structures were judged to have excellent geometry as determined by

401 MOLPROBITY (54)(Table 1).

402

403 Small Angle X-ray Scattering (SAXS).

404 SAXS analysis of the SorT/SorU complex was performed in a buffer of 20 mM Tris

405 pH 7.8, 2.5% v/v glycerol. Purified SorU and SorT were mixed and incubated on

406 ice at a molar ratio of 2:1 (SorU:SorT), generating two samples of total protein

407 concentrations 2.75 and 6.25 mgmL-1. SAXS data were measured as described

408 previously (55) with the data collection parameters listed in Table 4. Data were

17 409 reduced to I(q) vs q (q=4πsinθλ, where q=4sin2θ is the scattering angle) using the

410 program SAXSquant that includes corrections for sample absorbance, detector

411 sensitivity, and the slit geometry of the instrument. Intensities were placed on an

412 absolute scale using the known scattering from H2O. Protein scattering was obtained

413 by subtraction of the scattering from the matched solvents (20 mM Tris pH 7.8, 2.5%

414 v/v glycerol obtained from the flow-through after protein concentration by centrifugal

415 ultrafiltration). Molecular weight (Mr) estimates for the proteins were made using the

416 equation from Orthaber (62): Mr=NAI(0)CΔρM2 where NA is Avogadro’s number, C

417 is the protein concentration and ΔρM=Δρυ, where Δρ is the protein contrast and υ the

418 partial specific volume, both of which were determined using the program MULCh

419 (63).

420 The ATSAS program package (56) was used for data analysis and modeling, with the

421 specific programs used detailed in Table 4, along with the data ranges and results of

422 each of the calculations. Further detail on data interpretation and analysis for these

423 experiments is detailed in Appendix 2.

424

425 Electron Paramagnetic Resonance (EPR) spectroscopy

426 Continuous-wave X-band (ca. 9 GHz) (CW) electron paramagnetic resonance (EPR)

427 spectra were recorded with a Bruker Elexsys E580 CW/pulsed EPR spectrometer

428 fitted with a super high Q resonator; the microwave frequency and magnetic field

429 were calibrated with a Bruker microwave frequency counter and a Bruker ER 036TM

430 Teslameter, respectively. A microwave power of 20 mW was used and optimal

431 spectral resolution was obtained by keeping the modulation amplitude to a 1/10 of the

432 linewidth. A flow-through cryostat in conjunction with a Eurotherm (B-VT-2000)

18 433 variable temperature controller provided temperatures of 127-133 K at the sample

434 position in the cavity.

435

436 Bruker’s Xepr (version 2.6b.45) software was used to control the data acquisition

437 including, spectrometer tuning, signal averaging, temperature control and

438 visualization of the spectra. Computer simulation of the EPR spectra were performed

439 with the following spin Hamiltonian (Eq. 2)

2 440 H = βB⋅ g⋅S + S ⋅ A()95,97Mo ⋅ I − g β B ⋅ I + ()S ⋅ A()1H ⋅ I − g β B⋅ I (2) n n i=1 n n

441 using the XSophe-Sophe-XeprView (version 1.1.4) computer simulation software

442 suite (64, 65) on a personal computer, running the Mandriva Linux v2010.2 operating

443 system. Further detail on data interpretation and analysis for these experiments is

444 detailed in Appendix 1.

445

446 EPR-monitored redox potentiometry

447 The MoIV/V and MoV/VI redox potentials of SorT were determined by an EPR

448 monitored redox titration carried out in a Belle Technology anaerobic box. The

449 protein solution (1.5 mL, 40-90 μM in Tris-HCl, pH 8.0 and 10% glycerol) also

450 contained the following redox mediators at concentrations of ~50 µM: diaminodurol

451 (2,3,5,6-tetramethylphenylene-1,4-diamine, Em,7 +276 mV),

452 dichlorophenolindophenol (Em,7 +217 mV), 2,6-dimethylbenzoquinone (Em,7 +180

453 mV), phenazine methosulfate (Em,7 +80 mV), 2,5-dihydroxybenzoquinone (Em,7 –60

454 mV), indigo trisulfonate (Em,7 –90 mV), 2-hydroxy-1,4-naphthoquinone (Em,7 -152

455 mV) and anthraquinone 2-sulfonate (Em,7 -230 mV). The reductant was Ti(III) citrate

456 and the oxidant was NaS2O8 (both ~100 mM). After addition of titrant and

457 equilibration (15-30 min), the equilibrium potential was measured with a combination

19 458 Pt wire/AgAgCl redox electrode attached to a Hanna 8417 meter calibrated against

459 the quinhydrone redox couple (Eo′ (pH 7) = +284 mV vs NHE). A 100 μL aliquot of

460 protein was withdrawn and transferred to an EPR tube (in the anaerobic box) which

461 was then sealed and then carefully frozen in liquid nitrogen (outside the box).

462 Potentials for all experiments were measured with a combination Pt wire-Ag/AgCl

463 electrode attached to a Hanna 8417 meter. The intensity of the MoV signal (I) was

464 recorded as a function of measured potential (E), Figure 6.

465

466 Optical Spectroelectrochemistry

467 Spectroelectrochemistry of SorU in isolation and the SorT:SorU complex was

468 performed with a Bioanalytical Systems BAS100B/W potentiostat connected to a

469 Bioanalytical Systems thin layer spectroelectrochemical cell (0.5 or 1 mm pathlength)

470 bearing a transparent Au mini-grid working electrode, a Pt wire counter and Ag/AgCl

471 reference electrode. Redox mediators (all Fe complexes) used in the experiment were

472 employed at concentrations of 50 μM (Scheme (2)).

473 (2)

474 None exhibit any significant absorption in the spectral range of interest at micromolar

475 concentrations. The total solution volume was ca. 700 μL. The buffer was 20 mM

476 Tris (pH 8) containing 200 mM NaCl as supporting electrolyte. The SorU

477 concentration was ca. 50 μM while experiments on the SorU:SorT complex used

478 approximately equal concentrations of both proteins (50 μM). Spectra were acquired

479 within a Belle Technology anaerobic box with an Ocean Optics USB2000 fibre optic

480 spectrophotometer. Initially the cell potential was poised at ca. -100 mV and the

20 481 system was allowed to equilibrate until no further spectral changes were apparent

482 (fully reduced SorU). The potentials were then increased in 50 mV increments and the

483 spectrum was measured when no further changes were seen (5-10 min). When the

484 protein was fully oxidized the potential was scanned in the reverse direction in 50 mV

485 intervals to establish reversibility. Data were fitted using the program ReactLab

486 Redox (58).

487

488 Isothermal Titration Calorimetry (ITC)

489 Affinity measurements were conducted using a Microcal ITC200 system (GE

490 Healthcare) at 25°C using SorT and SorU in buffer (25 mM HEPES pH 7.5, 150 mM

491 NaCl and 2.5 % glycerol) at final concentrations of 300 μM and 30 μM, respectively.

492 SorT at a concentration of 300 μM was titrated with eighteen injections (2.0 μl each)

493 of SorU. All affinity measurements were performed in triplicate and fitted using a

494 single site mode. Protein concentrations were estimated using Bicinchoninic acid

495 (BCA) protein assay kit (ThermoScientific).

496

497

498 Sulfite dehydrogenase enzyme assays were carried out as described previously (12, 17,

499 36). The reduced – oxidized extinction coefficient for SorU at 550 nm was 17.486

500 mM-1 cm-1 as determined by spectroelectrochemistry. Data fitting was carried out

501 using Sigmaplot 12 (Systat).

502

21 503 ACKNOWLEDGEMENTS.

504 This work was supported by a La Trobe Institute for Molecular Science (LIMS)

505 Senior Research Fellowship to MJM an Australian Research Council grant and

506 Fellowship (DP0878525) to UK and an NHMRC CJ Martin Postdoctoral Research

507 Fellowship to APM. PVB also acknowledges financial support from the Australian

508 Research Council (DP1211465). Aspects of this research were undertaken on the

509 Macromolecular Crystallography beamline at the Australian Synchrotron (Victoria,

510 Australia) and we thank the beamline staff for their enthusiastic and professional

511 support. A/Prof Matthew Perugini (La Trobe University) is thanked for very helpful

512 discussions. Assistance with EPR spectroscopy from Dr Jeff Harmer and Dr Chris

513 Noble (Centre for Advanced Imaging, University of Queensland) is also gratefully

514 acknowledged.

515

516

517

518

519

22 520 REFERENCES.

521 1. Antonyuk SV, Han C, Eady RR, Hasnain SS. Structures of protein-protein

522 complexes involved in electron transfer. Nature. 2013;496(7443):123-6.

523 2. Moser CC, Keske JM, Warncke K, Farid RS, Dutton PL. Nature of biological

524 electron transfer. Nature. 1992;355(6363):796-802.

525 3. Schidlowski M. Antiquity and evolutionary status of bacterial sulfate

526 reduction: sulfur isotope evidence. Origins of life. 1979;9(4):299-311.

527 4. Kappler U, Bernhardt PV, Kilmartin J, Riley MJ, Teschner J, McKenzie KJ, et

528 al. SoxAX cytochromes, a new type of heme copper protein involved in bacterial

529 energy generation from sulfur compounds. J Biol Chem. 2008;283(32):22206-14.

530 5. Kappler U. Bacterial sulfite-oxidizing enzymes. Biochim Biophys Acta.

531 2011;1807(1):1-10.

532 6. Feng C, Tollin G, Enemark JH. Sulfite oxidizing enzymes. Biochim Biophys

533 Acta. 2007;1774(5):527-39.

534 7. Kisker C, Schindelin H, Pacheco A, Wehbi WA, Garrett RM, Rajagopalan KV,

535 et al. Molecular basis of sulfite oxidase deficiency from the structure of sulfite

536 oxidase. Cell. 1997;91(7):973-83.

537 8. Kappler U, Wilson JJ. Sulfite oxidation in Sinorhizobium meliloti. Biochimica

538 Et Biophysica Acta-Bioenergetics. 2009;1787(12):1516-25.

539 9. Hille R. Molybdenum and tungsten in biology. Trends Biochem Sci.

540 2002;27(7):360-7.

541 10. Hansch R, Lang C, Rennenberg H, Mendel RR. Significance of plant sulfite

542 oxidase. Plant biology. 2007;9(5):589-95.

23 543 11. Schrader N, Fischer K, Theis K, Mendel RR, Schwarz G, Kisker C. The

544 crystal structure of plant sulfite oxidase provides insights into sulfite oxidation in

545 plants and animals. Structure. 2003;11(10):1251-63.

546 12. Low L, Kilmartin JR, Bernhardt PV, Kappler U. How are "Atypical" Sulfite

547 Dehydrogenases Linked to Cell Metabolism? Interactions between the SorT Sulfite

548 Dehydrogenase and Small Redox Proteins. Frontiers in microbiology. 2011;2:58.

549 13. Bailey S, Rapson T, Johnson-Winters K, Astashkin AV, Enemark JH, Kappler

550 U. Molecular basis for enzymatic sulfite oxidation: how three conserved active site

551 residues shape enzyme activity. J Biol Chem. 2009;284(4):2053-63.

552 14. Cohen HJ, Fridovich I. Hepatic sulfite oxidase. Purification and properties. J

553 Biol Chem. 1971;246(2):359-66.

554 15. Cohen HJ, Fridovich I. Hepatic sulfite oxidase. The nature and function of the

555 heme prosthetic groups. J Biol Chem. 1971;246(2):367-73.

556 16. Cohen HJ, Fridovich I, Rajagopalan KV. Hepatic sulfite oxidase. A functional

557 role for molybdenum. J Biol Chem. 1971;246(2):374-82.

558 17. Wilson JJ, Kappler U. Sulfite oxidation in Sinorhizobium meliloti. Biochim

559 Biophys Acta. 2009;1787(12):1516-25.

560 18. Kappler U. Bacterial sulfite dehydrogenases - enzymes for chemolithtrophs

561 only? Microbial . Berlin: Springer; 2008. p. 151-69.

562 19. Leys D, Scrutton NS. Electrical circuitry in biology: emerging principles from

563 protein structure. Curr Opin Struct Biol. 2004;14(6):642-7.

564 20. Marcus RA, Sutin N. Electron transfer in chemistry and biology. Biochim

565 Biophys Acta. 1985;811:265-316.

566 21. Kappler U, Bailey S. Molecular basis of intramolecular electron transfer in

567 sulfite-oxidizing enzymes is revealed by high resolution structure of a heterodimeric

24 568 complex of the catalytic subunit and a c-type cytochrome subunit. J

569 Biol Chem. 2005;280(26):24999-5007.

570 22. Kappler U, Enemark JH. Sulfite-oxidizing enzymes. Journal of biological

571 inorganic chemistry : JBIC : a publication of the Society of Biological Inorganic

572 Chemistry. 2015;20(2):253-64.

573 23. Dell'acqua S, Pauleta SR, Monzani E, Pereira AS, Casella L, Moura JJ, et al.

574 Electron transfer complex between nitrous oxide reductase and cytochrome c552 from

575 Pseudomonas nautica: kinetic, nuclear magnetic resonance, and docking studies.

576 Biochemistry. 2008;47(41):10852-62.

577 24. Pettigrew GW, Pauleta SR, Goodhew CF, Cooper A, Nutley M, Jumel K, et al.

578 Electron transfer complexes of cytochrome c peroxidase from Paracoccus

579 denitrificans containing more than one cytochrome. Biochemistry.

580 2003;42(41):11968-81.

581 25. Kappler U, Bailey S, Feng C, Honeychurch MJ, Hanson GR, Bernhardt PV, et

582 al. Kinetic and structural evidence for the importance of Tyr236 for the integrity of

583 the Mo active site in a bacterial sulfite dehydrogenase. Biochemistry.

584 2006;45(32):9696-705.

585 26. Workun GJ, Moquin K, Rothery RA, Weiner JH. Evolutionary persistence of

586 the molybdopyranopterin-containing sulfite oxidase protein fold. Microbiology and

587 molecular biology reviews : MMBR. 2008;72(2):228-48, table of contents.

588 27. George GN, Pickering IJ, Kisker C. X-ray Absorption SPectroscopy of

589 Chicken Sulfite Oxidase Crystals. Inorganic chemistry. 1999;38:2539-40.

590 28. Wilson HL, Rajagopalan KV. The role of tyrosine 343 in substrate binding

591 and by human sulfite oxidase. J Biol Chem. 2004;279(15):15105-13.

25 592 29. Utesch T, Mroginski MA. Three-Dimensional Structural Model of Chicken

593 Liver Sulfite Oxidase in its Activated Form. J Phys Chem Lett. 2010;1(14):2159-64.

594 30. Brody MS, Hille R. The kinetic behavior of chicken liver sulfite oxidase.

595 Biochemistry. 1999;38(20):6668-77.

596 31. Page CC, Moser CC, Chen X, Dutton PL. Natural engineering principles of

597 electron tunnelling in biological oxidation-reduction. Nature. 1999;402(6757):47-52.

598 32. Onuchic JN, Beratan DN, Winkler JR, Gray HB. Pathway analysis of protein

599 electron-transfer reactions. Annu Rev Biophys Biomol Struct. 1992;21:349-77.

600 33. Tai H, Tonegawa K, Shibata T, Hemmi H, Kobayashi N, Yamamoto Y.

601 Inversion of the stereochemistry around the sulfur atom of the axial side

602 chain through alteration of side chain packing in Hydrogenobacter

603 thermophilus cytochrome C552 and its functional consequences. Biochemistry.

604 2013;52(28):4800-9.

605 34. Senda M, Kishigami S, Kimura S, Fukuda M, Ishida T, Senda T. Molecular

606 mechanism of the redox-dependent interaction between NADH-dependent ferredoxin

607 reductase and Rieske-type [2Fe-2S] ferredoxin. J Mol Biol. 2007;373(2):382-400.

608 35. Toogood HS, Leys D, Scrutton NS. Dynamics driving function: new insights

609 from electron transferring flavoproteins and partner complexes. The FEBS journal.

610 2007;274(21):5481-504.

611 36. Kappler U, Bennett B, Rethmeier J, Schwarz G, Deutzmann R, McEwan AG,

612 et al. Sulfite:Cytochrome c from Thiobacillus novellus. Purification,

613 characterization, and molecular biology of a heterodimeric member of the sulfite

614 oxidase family. J Biol Chem. 2000;275(18):13202-12.

26 615 37. Nojiri M, Koteishi H, Nakagami T, Kobayashi K, Inoue T, Yamaguchi K, et al.

616 Structural basis of inter-protein electron transfer for nitrite reduction in denitrification.

617 Nature. 2009;462(7269):117-20.

618 38. Axelrod HL, Abresch EC, Okamura MY, Yeh AP, Rees DC, Feher G. X-ray

619 structure determination of the cytochrome c2: reaction center electron transfer

620 complex from Rhodobacter sphaeroides. J Mol Biol. 2002;319(2):501-15.

621 39. Pelletier H, Kraut J. Crystal structure of a complex between electron transfer

622 partners, cytochrome c peroxidase and cytochrome c. Science. 1992;258(5089):1748-

623 55.

624 40. Solmaz SR, Hunte C. Structure of complex III with bound cytochrome c in

625 reduced state and definition of a minimal core interface for electron transfer. J Biol

626 Chem. 2008;283(25):17542-9.

627 41. Miyashita O, Onuchic JN, Okamura MY. Continuum electrostatic model for

628 the binding of cytochrome c2 to the photosynthetic reaction center from Rhodobacter

629 sphaeroides. Biochemistry. 2003;42(40):11651-60.

630 42. Gray HB, Winkler JR. Long-range electron transfer. Proceedings of the

631 National Academy of Sciences of the United States of America. 2005;102(10):3534-9.

632 43. Lawrence MC, Colman PM. Shape complementarity at protein/protein

633 interfaces. J Mol Biol. 1993;234(4):946-50.

634 44. van Amsterdam IM, Ubbink M, Einsle O, Messerschmidt A, Merli A,

635 Cavazzini D, et al. Dramatic modulation of electron transfer in protein complexes by

636 crosslinking. Nature structural biology. 2002;9(1):48-52.

637 45. Trana EN, Nocek JM, Knutson AK, Hoffman BM. Evolving the [myoglobin,

638 cytochrome b(5)] complex from dynamic toward simple docking: charging the

639 electron transfer reactive patch. Biochemistry. 2012;51(43):8542-53.

27 640 46. Xiong P, Nocek JM, Vura-Weis J, Lockard JV, Wasielewski MR, Hoffman

641 BM. Faster interprotein electron transfer in a [myoglobin, b(5)] complex with a

642 redesigned interface. Science. 2010;330(6007):1075-8.

643 47. Otwinowski Z, Minor W. Processing of X-Ray Diffraction Data Collected in

644 Oscillation Mode. Methods in Enzymology. 1997;276(Part A):307-26.

645 48. Mccoy AJ, Grosse-Kunstleve RW, Adams PD, Winn MD, Storoni LC, Read

646 RJ. Phaser crystallographic software. Journal of Applied Crystallography.

647 2007;40:658-74.

648 49. Stein N. CHAINSAW: a program for mutating pdb files used as templates in

649 molecular replacement. Journal of Applied Crystallography. 2008;41:641-3.

650 50. Larkin MA, Blackshields G, Brown NP, Chenna R, McGettigan PA,

651 McWilliam H, et al. Clustal W and Clustal X version 2.0. Bioinformatics.

652 2007;23(21):2947-8.

653 51. Murshudov GN, Skubak P, Lebedev AA, Pannu NS, Steiner RA, Nicholls RA,

654 et al. REFMAC5 for the refinement of macromolecular crystal structures. Acta

655 Crystallogr D Biol Crystallogr. 2011;67(Pt 4):355-67.

656 52. Emsley P, Cowtan K. Coot: model-building tools for molecular graphics. Acta

657 Crystallographica Section D-Biological Crystallography. 2004;60:2126-32.

658 53. Adams PD, Grosse-Kunstleve RW, Hung LW, Ioerger TR, McCoy AJ,

659 Moriarty NW, et al. PHENIX: building new software for automated crystallographic

660 structure determination. Acta Crystallographica Section D-Biological Crystallography.

661 2002;58:1948-54.

662 54. Chen VB, Arendall WB, 3rd, Headd JJ, Keedy DA, Immormino RM, Kapral

663 GJ, et al. MolProbity: all-atom structure validation for macromolecular

28 664 crystallography. Acta crystallographica Section D, Biological crystallography.

665 2010;66(Pt 1):12-21.

666 55. Jeffries CM, Lu Y, Hynson RM, Taylor JE, Ballesteros M, Kwan AH, et al.

667 Human cardiac myosin binding protein C: structural flexibility within an extended

668 modular architecture. J Mol Biol. 2011;414(5):735-48.

669 56. Volkov VV, Svergun DI. Uniqueness of ab initio shape determination in

670 samll-angle scattering. Journal of Applied Crystallography. 2003;36:860-4.

671 57. Beratan DN, Onuchic JN, Winkler JR, Gray HB. Electron-tunneling pathways

672 in proteins. Science. 1992;258(5089):1740-1.

673 58. M. Maeder and P. King P, WA http://www.jplusconsulting.com/. ReactLab-

674 Redox.

675 59. Winn MD, Ballard CC, Cowtan KD, Dodson EJ, Emsley P, Evans PR, et al.

676 Overview of the CCP4 suite and current developments. Acta Crystallogr 2011;D67(Pt

677 4):235-42.

678 60. CCP4. Collaborative Computational Project no. 4. Acta Crystallographica.

679 1994;D50:760-3.

680 61. Krissinel E, Henrick K. Inference of macromolecular assemblies from

681 crystalline state. J Mol Biol. 2007;372(3):774-97.

682 62. Orthaber D, Bergmann A, Glatter O. SAXS experiments on absolute scale

683 with Kratky systems using water as a secondary standard. Journal of Applied

684 Crystallography. 2000;33:218-25.

685 63. Whitten AE, Cai SZ, Trewhella J. MULCh: modules for the analysis of small-

686 angl neutron contrast variation data from biomolecular assemblies. Journal of Applied

687 Crystallography. 2008;39:277-86.

29 688 64. Hanson GR, Gates KE, Noble CJ, Griffin M, Mitchell A, Benson S. XSophe-

689 Sophe-XeprView. A computer simulation software suite (v. 1.1.3) for the analysis of

690 continuous wave EPR spectra. Journal of inorganic biochemistry. 2004;98(5):903-16.

691 65. Hanson GR, Noble CJ, Benson SS. XSophe - Sophe - XeprView and

692 Molecular Sophe Computer simulation software suites for the analysis of continuous

693 wave and pulsed EPR and ENDOR spectra. In: Lund A, Shiotani M, editors. EPR of

694 Free Radicals in Solids: Trends in Methods and Applications (2nd Edition). Progress

695 in Theoretical Chemistry and Physics. Heidelberg, Germany: Springer; 2013. p. 223-

696 83.

697 66. Drew SC, Hill JP, Lane I, Hanson GR, Gable RW, Young CG. Synthesis,

698 structural characterization, and multifrequency electron paramagnetic resonance

699 studies of mononuclear thiomolybdenyl complexes. Inorganic chemistry.

700 2007;46(7):2373-87.

701 67. Drew SC, Young CG, Hanson GR. A density functional study of the electronic

702 structure and spin Hamiltonian parameters of mononuclear thiomolybdenyl

703 complexes. Inorganic chemistry. 2007;46(7):2388-97.

704 68. Drew SC, Hanson GR. Determination of the metal-dithiolate fold angle in

705 mononuclear molybdenum(V) centers by EPR spectroscopy. Inorganic chemistry.

706 2009;48(5):2224-32.

707 69. Enemark JH, Astashkin AV, Raitsimring AM. High resolution EPR

708 spectroscopy of Mo enzymes. Sulfite oxidases: structural and functional implications.

709 In: Hanson GR, Berliner LJ, editors. Metals in Biology: Applications of High

710 Resolution EPR to Metalloenzymes. Biological Magnetic Resonance. New York,

711 USA: Springer; 2010. p. 121-68.

30 712 70. Doonan CJ, Wilson HL, Bennett B, Prince RC, Rajagopalan KV, George GN.

713 MoV electron paramagnetic resonance of sulfite oxidase revisited: the low-pH

714 chloride signal. Inorganic chemistry. 2008;47(6):2033-8.

715 71. Lamy MT, Gutteridge S, Bary RC. Electron-paramagnetic-resonance

716 parameters of molybdenum(V) in sulphite oxidase from chicken liver. Biochem J.

717 1980;185(2):397-403.

718 72. Astashkin AV, Raitsimring AM, Feng C, Johnson JL, Rajagopalan KV,

719 Enemark JH. The Mo-OH proton of the low-pH form of sulfite oxidase: Comparison

720 of the hyperfine interactions obtained from pulsed ENDOR, CW-EPR and ESEEM

721 measurements. Appl Magn Reson. 2002;22(3):421-30.

722 73. Klein EL, Astashkin AV, Raitsimring AM, Enemark JH. Applications of

723 pulsed EPR spectroscopy to structural studies of sulfite oxidizing enzymes. Coordin

724 Chem Rev. 2013;257(1):110-8.

725 74. Enemark JH, Astashkin AV, Raitsimring AM. Investigation of the

726 coordination structures of the molybdenum(v) sites of sulfite oxidizing enzymes by

727 pulsed EPR spectroscopy. Dalton T. 2006(29):3501-14.

728 75. Astashkin AV, Hood BL, Feng C, Hille R, Mendel RR, Raitsimring AM, et al.

729 Structures of the Mo(V) forms of sulfite oxidase from Arabidopsis thaliana by pulsed

730 EPR spectroscopy. Biochemistry. 2005;44(40):13274-81.

731 76. Astashkin AV, Klein EL, Ganyushin D, Johnson-Winters K, Neese F, Kappler

732 U, et al. Exchangeable in the vicinity of the molybdenum center of the high-

733 pH form of sulfite oxidase and sulfite dehydrogenase. Physical chemistry chemical

734 physics : PCCP. 2009;11(31):6733-42.

31 735 77. Raitsimring AM, Kappler U, Feng C, Astashkin AV, Enemark JH. Pulsed

736 EPR studies of a bacterial sulfite-oxidizing enzyme with pH-invariant hyperfine

737 interactions from exchangeable protons. Inorganic chemistry. 2005;44(21):7283-5.

738

32 739 FIGURE TITLES AND LEGENDS.

740 Figure 1. The crystal structures of the SorT and SorU proteins in isolation. Panel

741 A: The structure of the SorT homodimer. Molecule A in blue/yellow; molecule B in

742 gray (with transparent surface). For molecule A, the ‘SUOX-fold domain’ and

743 ‘dimerization domain’ are represented in blue and yellow, respectively. The

744 molybdopterin cofactor is shown as sticks within the SUOX-fold domain. The

745 corresponding domains of the opposing protomer (shown in molecular surface

746 representation), which constitute the dimer interface are coloured to highlight the

747 ‘head to tail’ dimer arrangement. INSET: a closer view of the molybdenum binding-

748 site: the molybdenum atom (green sphere) is coordinated by two dithioline ligands

749 from the molybdopterin (yellow spheres), residue Cys 127, an axial oxo ligand and an

750 equatorial hydroxo or water ligand (red spheres); Panel B: The structure of SorU. The

751 main three helices are labeled and the heme cofactor is shown in red. INSET: the

752 heme binding site with the heme cofactor, coordinating residues, covalent links to Cys

753 50 and 57 and hydrophobic residues lining binding site: Phe 39, Val 62, Val 76, Val

754 80, Val 101 highlighted.

755

756 Figure 2. EPR analysis of the SorT protein. Panel A. X-band EPR spectra of the

757 Mo(V) center in SorT. (a) First and (b) second derivative EPR spectra of SorT at +50

758 mV vs NHE in Tris-HCl pH 8.0, ν= 9.43462 GHz, T = 136.3 K. (c) Computer

759 simulation of the second derivative spectrum with the spin Hamiltonian parameters

760 listed in Table 3; (d,e) Expansion of spectra (b) and (c), respectively. Panel B.

761 Schematic structures of the (a) high and (b) low pH forms of sulfite oxidase.

762

33 763 Figure 3. The structure of the SorT/SorU electron transfer complex. Panel A.

764 The asymmetric unit from the crystal structure of the SorT/SorU complex contains the

765 functional electron transfer complex. The SorU/SorT2/SorU complex is revealed by

766 the application of crystallographic symmetry operators. The positions of the redox

767 active molybdenum (SorT) and heme c (SorU) cofactors are indicated. Panel B. Two

768 views of an overlay of the SorU/SorT2/SorU crystal structure with the averaged and

769 filtered dummy atom model from 10 ab initio reconstructions as revealed by SAXS

770 analyses. Panel C. ‘Open-book unfolding’ of SorT/SorU complex (SorT is shown in

771 blue, SorU in red) indicating the ‘footprint’ of interfacing residues from each protein.

772 Panel D. The same view as Panel C, showing the charge complementarity of the

773 SorT/SorU interface (areas of positive charge in blue, negative charge in red and

774 neutral in white).

775

776 Figure 4. Orientation of the redox cofactors in the crystal structure of the

777 SorT/SorU electron transfer complex. Panel A. Electron density map in the region

778 of the SorT/SorU interface. The SorT molecule is represented in blue and the SorU

779 molecule in red. The 2Fo-Fc electron density map is shown as a blue net and the

780 redox cofactors (molybdenum and heme) are colored according to the representation

781 in Panel B. Panel B. Pathway for electron transfer (57).

782

783 Figure 5. Comparisons of (A) the SorT/SorU and SorAB structures and (B) the

784 structures of SorT and SorU within and outside of the electron transfer complex.

785 Panel A. Structures of the SorT/SorU (left) and SorAB (right) complexes, where the

786 Cα traces of the heme-containing protomers are colored according to temperature

787 factor. Panel B. Superposition of the SorT and SorU structures within and outside of

34 788 the electron transfer complex, highlighting conformational changes that were

789 observed to accompany complex formation. The crystal structures of SorT and SorU

790 within the SorT/SorU complex are shown in gray, and the superposed structures of

791 SorT and SorU determined alone are shown in blue and red respectively.

792

793 Figure 6. Redox analyses of the SorT protein and the SorT/SorU complex. Panel

V 794 A. Plot of EPR intensity (Ip) at 343 mT (from Mo form of SorT) as a function of

795 solution redox potential (E mV vs NHE). The solid line is a fit to the equation

I p = VI/V 796 IE() −− using the potentials E1 = Mo = +110(±10) 110++()/59(EE12 10 E E )/59

V/IV 797 mV and E2 Mo -18(±10) mV vs NHE). Panel B. Electronic spectra of ferric and

798 ferrous SorU obtained from spectroelectrochemistry. Inset: plot of absorbance at 550

799 nm (ferrous α-band) and 406 nm (ferric Soret band) as a function of applied potential.

800 The solid lines are theoretical curves based on the equation

(10εε(')/59EE− + ) 801 Abs= ox red c where the extinction coefficients refer to the 110+ (')/59EE− tot

802 oxidized and reduced forms of the protein and Abs is the absorbance at this same

803 wavelength. ctot is the total protein concentration. The redox potential (E' = +111 mV

804 vs NHE) was obtained by global analysis of all potential dependent spectra across all

805 wavelengths with the program ReactLab Redox (58).

806

807 Figure 7. The bonding network at the interface of SorT/SorU. Panel A: An open-

808 book representation depicts residues involved in forming stable bonds at the interface

809 between SorT and SorU as corresponding color patches mapped onto the molecular

810 surface. Panel B: Stereoview of the interface between SorT and SorU. Bonding

35 811 residues are shown as sticks with bonds shown as dashes between atoms. SorT is

812 shown in light grey and SorU is shown in magenta.

813

814 Appendix Figure 1. SAXS Data and Interpretation. Panel A. Guinier plot of the

815 desmeared I(q) versus q; Panel B. log:log plot of the measured (slit smeared) I(q)

816 versus q with the P(r)- model I(q) (black line) and smeared P(r)-model I(q) (red line)

817 fits; Panel C. P(r) versus r for the P(r) model in (B), dmax is 110 Å; Panel D.

818 superposition of the desmeared I(q) versus q with that calculated from the crystal

819 structure of the SorT/SorU2/SorT complex (red line) and the SorT dimer (green line).

820

821

36 Table 1. Data collection and refinement statistics

SorT SorU SorT/SorU complex

Data collection

Space Group P21 F222 P21212

Cell dimensions

a, b, c (Å) 96.0, 92.2, 109.4 70.9, 129.2, 197.0 109.6, 95.8, 49.9

α, β, γ (°) 90, 89.7, 90 90, 90, 90 90, 90, 90

X-ray source AUS MX2 AUS MX2 AUS MX2

λ (Å) 0.950 0.954 0.954

Detector ADSC Quantum 315r ADSC Quantum 315r ADSC Quantum 315r

Resolution range (Å) 50-2.4 (2.43-2.35)a 50-2.2 (2.28-2.20) 50-2.5 (2.50-2.59)

Observed reflections 240521 96340 64273

Unique reflections 77971 23453 18883

Completeness (%) 98.4 (99.4) 99.9 (100) 99.3 (99.6)

Multiplicity 3.1 (3.1) 4.1 (4.1) 3.4 (3.4)

6.7 (2.1) 8.9 (1.6) 8.8 (1.7)

b Rmerge (%) 15.9 (66.7) 13.3 (76.6) 13.9 (76.6)

Refinement

Reflections in working 74024 22113 17781 set

Reflections in test set 3927 1201 960

Protomers per ASU 4 4 1

Total atoms (non-H) 11378 2941 3422

Protein atoms 10890 2542 3290

37 Metal atoms 4 4 2

Water atoms 376 227 63

Other atoms 108 168 67

c Rwork (%) 20.8 (31.7) 19.2 (30.8) 21.1 (30.2)

d Rfree (%) 23.9 (34.7) 24.0 (34.3) 26.0 (36.5)

Rmsd bond lengths (Å) 0.008 0.006 0.012

Rmsd bond angles (deg) 1.08 0.91 1.41

(Å2)e 32.5 20.6 38.0

Cruickshank's DPI 0.07 0.23 0.49

PDB ID 4PW3 4PWA 4PW9

aValues in parenthesis are for highest-resolution shell

b Rmerge = ∑hkl ∑i⏐Ii (hkl) - I(hkl)⏐/∑hkl ∑i Ii (hkl)

c Rwork = Σh ⏐Fobs – Fcalc⏐/ΣhFobs

d Calculated as for Rwork using 10% of the diffraction data that had

been excluded from the refinement

eAs calculated by BAVERAGE (59, 60)

822

823

38 824 Table 2. Mo coordination geometry in the active site of SorT.

Bond Distance (Å)

Mo-S1 (pterin) 2.4

Mo-S2 (pterin) 2.4

Mo-S (Cys 127) 2.3

Mo=O 1.7

Mo-OH/H2O 1.9

825

826

39 827 Table 3. Spin Hamiltonian parameters for the Mo(V) center of SorT and various

828 low and high pH forms of human, avian, plant and bacterial sulfite oxidases.

829

Species Parameter X Y Z βo Ref

SorT g 1.94930 1.95997 1.98632 -

A(95Mo)b 20.5 36.0 53.9 26

A(1H)b,c 3.5 4.0 4.8 0

SorAc g 1.9541 1.9661 1.9914 -d 56

Human SO g Low pH 1.9646 1.9723 2.0023 - 52

A(1H)b 11.47 7.10 7.71 - 52

Chicken SO g (Low pH) 1.9658 1.9720 2.0037 - 51

A(1H)b 11.93 7.37 7.95 - 51

g (High pH) 1.9531 1.9641 1.9872 51

A. Thaliana g (Low pH) 1.963 1.974 2.005 - 57

SO

A(1H)b 11.9 9.2 10.3 -- 57

g (High pH) 1.956 1.964 1.989 - 57

830

831 a Non coincident angle between g and A (rotation about x axis). bUnits 10-4 cm-1. c

832 Two magnetically equivalent protons (I=1/2) were included in the computer simulated

c 95 833 spectra. Mo hyperfine couplings were unresolved and the shoulders on gz were

834 incorrectly attributed to 95Mo hyperfine resonances. dEuler angles were not

835 determined.

836

40 837 Table 4. Data collection and processing parameters for analysis of the SorT/SorU

838 complex in solution by Small Angle X-ray Scattering (SAXS).

839

Data collection parameters

Instrument SAXSess (Anton Paar)

Beam geometry 10 mm slit

AH, LH (Å-1), GNOM beam geometry definition 0.28, 0.12

q-range measured (Å-1) 0.01-0.400

Exposure time (min) 60 (4 x 15)

-1 SorT2SorU2 concentration range (mg mL ) 2.75-5.5

Temperature (ºC) 10

Structural parameters*

-1 Rg (Å), I(0) (cm ) 30.8 ± 0.4, 0.223 ± 0.002

from Guinier (desmeared data) q*Rg < 1.3

-1 Rg (Å), I(0) (cm ) 32.0 ± 0.3, 0.235 ± 0.002

from P(r) (q-range 0.01 – 0.25 Å-1)

dmax (Å) from P(r) 110

Molecular mass determination*

Molecular mass Mr from Guinier I(0) (ratio with 108741 (0.984)

expected)

Molecular mass Mr from P(r) I(0) (ratio with expected) 114593 (1.037)

SorT2SorU2 parameters calculated from sequence and chemical composition

Molecular volume (Å3) 134385

Molecular weight Mr (Da) 110556

41 Partial specific volume (cm3 g-1) 0.732

Contrast (X-rays) (Δρ x 1010 cm-2) 2.895

Modeling results and validation

Crystal structure Rg, dmax (Å)

SorT/SorU2/SorT 31.3, 108

SorT 27.9, 99

Crystal structure compare to desmeared I(q) (χ-value)

-1 SorT/SorU2/SorT (q-range 0.01 – 0.15 Å ) 1.7

SorT (q-range 0.01 – 0.15 Å-1) 2.3

Results from 10 ab initio shape restorations. P1 symmetry:

Average molecular volume (Å3) 140800

Normalised spatial distribution (NSD) and NSD 0.508 (0.008) variation

χ value for fit to desmeared data 1.8

Software employed

Calculation of expected Mr, Δρ and υ values MULCh

Primary data reduction, I(q) vs q SAXSQuant 1D

Desmearing SAXSQuant

Guinier analysis PRIMUS

P(r) analysis GNOM

Model I(q) from crystal coordinates CRYSOL ab initio shape restorations DAMMIN

42 3D graphics representations PYMOL

840 *Reported for 2.75 mg ml-1 measurement.

841

43 842 Table 5. Electron transfer parameters between SorT (Mo) and SorU (Fe) (32)

Distance (Mo-Fe, Å) 16.5 Å

Atomic packing density (ρ) 0.97

Average decay exponential (β) 0.97

-4 Electronic coupling (HDA) 3.4 x 10

Maximum ET rate (s-1) 1.2 x 107

843

844

44 845 Table 6. Redox potential values for SorT and SorU at pH8a.

Protein Couple Eº (mV vs NHE)

SorT MoVI/V +110(±10)

MoV/IV -18(±10)

SorU FeIII/II +108 (±10)

SorU FeIII/II +111 (±10)

(in the presence of SorT)

846 aRedox potentials of SorT were determined by redox potentiometry, and SorU redox

847 potentials by optical spectroelectrochemistry.

848

45 849

850 Table 7. Comparison of the protein-protein interfaces in the SorT/SorU and

851 SorAB structures.

852

Parameter SorT/SorUa SorABb

SorT SorU SorA SorB

Average relative B factorc 0.9 1.5 1.0 1.1

(Å2)

Buried surface area (Å2)d 644 696 1254 1380

Interfacing residuesd 31 21 46 33

Hydrogen-bonds 6 30e

Salt-bridges 1 2e

Shape complementarity 0.63 0.77

statisticf

853 a This work

854 b PDB code 2BLF (21)

855 c Calculated as the average for the protomer of interest divided by the average for the

856 entire complex structure.

857 d (61)

858 e Taken from (21)

859 f (43)

860

861

862

863

46 864 APPENDIX 1

865 SUPPLEMENTAL MATERIAL ON THE INTERPRETATION OF EPR DATA

866 The optimum potential to measure the continuous wave (CW) EPR spectrum of the

867 Mo(V) center in SorT was found to be +50 mV from the redox potentiometry

868 experiments (Figure 6A). The CW EPR spectrum at 0 mV (Figure 2A(a)) arises

869 from a single rhombically distorted Mo(V) center.

870

871 Naturally abundant Mo consists of a mixture of isotopes (95,97Mo, I=5/2, 25.5 %

872 abundance; 92,94,96,98,100Mo, I=0, 74.5 % abundance) and consequently the spectrum

873 consists of three I=0 resonances corresponding to the principal directions of the g-

874 matrix and six (2I+1) satellite resonances, each with an intensity of ~4%. Increased

875 spectral resolution was obtained by numerically differentiating the spectrum and

876 carefully Fourier filtering (Hamming function) the spectrum to remove the high

877 frequency noise without distorting the spectrum (Figure 2A(b)). Interestingly, a close

878 examination of the gz resonance around 339.35 mT reveals a triplet (Figure 2A) (64,

879 65) in an approximate ratio of 1:2:1. Computer simulation of the CW EPR spectrum

880 with a monoclinic (Cs symmetry) spin Hamiltonian (Eq. (2)) incorporating two

881 magnetically equivalent I=1/2 nuclei and the spin Hamiltonian parameters listed in

882 Table 3 produces the spectrum shown in Figure 2A(c).

883

884 Previously we have shown in a study of model Mo(V) complexes that the non

885 coincident angle, β (rotation of Ay,z from gy,z about x) can be accurately determined

886 utilizing multifrequency CW EPR in conjunction with computer simulation studies

887 (66, 67). Herein, the increased spectral resolution in the second derivative spectrum

888 and the narrow line widths enable an accurate determination of the non coincident

47 889 angle β without resorting to the use of multiple microwave frequencies. Computer

890 simulation of the second derivative spectrum showed that the 95,97Mo hyperfine

891 resonant field positions along the ‘z’ and ‘y’ directions were highly sensitive to the

892 magnitude of these hyperfine couplings and the non coincident angle β. The excellent

893 agreement between the simulated and experimental second derivative spectra (Figure

894 2A(b)(c)) gives confidence in the values of the spin Hamiltonian parameters. We have

895 also shown through a systematic density functional theory study that the non

896 coincident angle β can be correlated to the pterin fold angle (68) and for β=26o, the

897 predicted pterin fold angle would be 5.6o, which is in good agreement with that

898 determined (1.9o±0.2o) from the X-ray crystal structure of SorT.

899

900 CW and pulsed EPR spectra of the Mo(V) center in human, avian, plant and bacterial

901 sulfite have been extensively studied (Table 3) (36, 69-75). The Mo(V) CW EPR

902 spectra of SO from humans, birds and plants are pH dependent. The CW EPR spectra

903 arise from a Mo(V) center with rhombic or lower symmetry and at low pH the

904 resonances are split into a doublet arising from a strongly coupled proton (Table 3).

905 CW and pulsed EPR spectra in conjunction with isotope enrichment (17O, 2H) studies

906 have identified the origin of the proton as an equatorial hydroxo ligand. Loss of the

907 proton superhyperfine coupling at high pH in the CW EPR spectrum, was shown

908 through orientation selective multifrequency electron spin echo envelope modulation

909 (ESEEM) studies to be a rotation of the hydroxo moiety out of the xy plane (rather

910 than deprotonation), thereby reducing the unpaired electron spin density on the 1H

911 nucleus of the hydroxo ligand (69, 73, 76). 17O ESEEM studies also revealed two

912 other weakly coupled oxygen ligands, the axially coordinated Mo=O moiety and a

48 913 weakly coupled OH- moiety hydrogen bonded to the equatorial hydroxyo ligand (73,

914 76).

915

916 In contrast, the Mo(V) EPR spectra of SOEs from bacteria (SorA (36) and SorT) are

917 pH independent. A comparison of the gz resonances from SorA (36) and SorT

918 (Figure 2) shows their lineshapes to be very similar, both exhibiting shoulders.

919 Herein we have clearly shown through computer simulation of the Mo(V) CW EPR

920 spectrum that the shoulders do not arise from 95,97Mo hyperfine coupling (Figure 2,

921 Table 3), but arise from two weakly coupled 1H nuclei. While the computer

922 simulation assumed that the two protons were magnetically equivalent, a slight

923 magnetic inequivalence of the two 1H superhhyperfine couplings cannot be ruled out.

924 The origin of the 1H superhyperfine couplings is likely to be either (i) an equatorial

925 aqua ligand with the protons lying outside of the equatorial plane, thereby reducing

926 the overlap with Mo based dxy orbital containing the unpaired electron or (ii) an

927 equatorial hydroxo ligand which is hydrogen bonded to another hydroxyl moiety

928 (Figure 2B) where both protons lie outside of the equatorial plane. The proton

929 superhyperfine couplings are similar to those found from pulsed EPR studies of SorA

930 (69, 77).

931

932

933

934

935

936

937

49 938 APPENDIX 2

939 SUPPLEMENTAL MATERIAL FOR THE INTERPRETATION OF SAXS

940 DATA

941 The scattering data from the sample prepared as a stoichiometric mixture of SorT and

942 SorU is presented in Figure 3B and Figure 3-figure supplement 1, while Table 4

943 summarizes the structural parameters derived from those data. Significantly the Mr

944 value determined from I(0) is in excellent agreement with that expected for a 2:2

945 complex (within 2-4% for calculations based on Guinier or P(r)derived I(0)). Further,

946 the scattering data fit the crystal structure of the SorT/SorU complex reasonably well,

947 yielding a χ value of 1.7 for the overall fit and good agreement with the crystal

948 structure Rg value. This agreement is not expected to be perfect as the crystal structure

949 is missing a total of 56 residues: 33 from the N-terminus of the SorT, 2 from the N-

950 terminus of SorU and 21 from the C-terminus of SorU; accounting for 12% of the

951 total molecular mass. By comparison, a significantly worse χ value of 2.3 is obtained

952 by fitting only the SorT dimer, which also predicts a significantly smaller Rg value

953 than was observed (28 Å compared to 31 Å).

954

955 Dummy atom reconstructions yielded shapes that had the expected molecular volume

956 and in projection had the expected shape for the SorU/SorT2/SorU assembly, though

957 the shapes were consistently a somewhat flatter disk-like structure than that observed

958 in the crystal structure (Figure 3B). This result is not unexpected as this class of

959 structure (flattened anisotropic particles) is known to present some difficulty on ab

960 initio shape reconstructions (56).

50