Advertisement
Editors' Picks| Volume 292, ISSUE 16, P6838-6850, April 21, 2017

A unique structural domain in Methanococcoides burtonii ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco) acts as a small subunit mimic

  • Laura H. Gunn
    Correspondence
    To whom correspondence should be addressed: Dept. of Cell and Molecular Biology, Uppsala University, Husargatan 3, Box 596, S-751 24 Uppsala, Sweden. Tel.: 46-18-4716698; Fax: 46-18-511755;.
    Affiliations
    Department of Cell and Molecular Biology, Uppsala University, S-751 24 Uppsala, Sweden
    Search for articles by this author
  • Karin Valegård
    Affiliations
    Department of Cell and Molecular Biology, Uppsala University, S-751 24 Uppsala, Sweden
    Search for articles by this author
  • Inger Andersson
    Affiliations
    Department of Cell and Molecular Biology, Uppsala University, S-751 24 Uppsala, Sweden
    Search for articles by this author
  • Author Footnotes
    2 The abbreviations used are: RuBPribulose 1,5-bisphosphateRubiscoRuBP carboxylase/oxygenase3-PGA3-phosphoglycerateLSuRubisco large subunitSSuRubisco small subunitMbRM. burtonii RubiscoRADRubisco assembly domain2-CABP2-carboxyarabinitol 1,5-bisphosphate4-CABP4-carboxyarabinitol 1,5-bisphosphateXuBPxylulose 1,5-bisphosphatePDBProtein Data Bankr.m.s.d.root mean square deviationEexpect valueRLPRubisco-like proteinMhRM. hollandica RubiscoH6-UbHis6-ubiquitinESRFEuropean Synchrotron Research FacilityBicineN,N-bis(2-hydroxyethyl)glycineIMACimmobilized metal affinity chromatography.
Open AccessPublished:January 30, 2017DOI:https://doi.org/10.1074/jbc.M116.767145
      The catalytic inefficiencies of the CO2-fixing enzyme ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco) often limit plant productivity. Strategies to engineer more efficient plant Rubiscos have been hampered by evolutionary constraints, prompting interest in Rubisco isoforms from non-photosynthetic organisms. The methanogenic archaeon Methanococcoides burtonii contains a Rubisco isoform that functions to scavenge the ribulose-1,5-bisphosphate (RuBP) by-product of purine/pyrimidine metabolism. The crystal structure of M. burtonii Rubisco (MbR) presented here at 2.6 Å resolution is composed of catalytic large subunits (LSu) assembled into pentamers of dimers, (L2)5, and differs from Rubiscos from higher plants where LSus are glued together by small subunits (SSu) into hexadecameric L8S8 enzymes. MbR contains a unique 29-amino acid insertion near the C terminus, which folds as a separate domain in the structure. This domain, which is visualized for the first time in this study, is located in a similar position to SSus in L8S8 enzymes between LSus of adjacent L2 dimers, where negatively charged residues coordinate around a Mg2+ ion in a fashion that suggests this domain may be important for the assembly process. The Rubisco assembly domain is thus an inbuilt SSu mimic that concentrates L2 dimers. MbR assembly is ligand-stimulated, and we show that only 6-carbon molecules with a particular stereochemistry at the C3 carbon can induce oligomerization. Based on MbR structure, subunit arrangement, sequence, phylogenetic distribution, and function, MbR and a subset of Rubiscos from the Methanosarcinales order are proposed to belong to a new Rubisco subgroup, named form IIIB.

      Introduction

      Ribulose-1,5-bisphosphate (RuBP)
      The abbreviations used are: RuBP
      ribulose 1,5-bisphosphate
      Rubisco
      RuBP carboxylase/oxygenase
      3-PGA
      3-phosphoglycerate
      LSu
      Rubisco large subunit
      SSu
      Rubisco small subunit
      MbR
      M. burtonii Rubisco
      RAD
      Rubisco assembly domain
      2-CABP
      2-carboxyarabinitol 1,5-bisphosphate
      4-CABP
      4-carboxyarabinitol 1,5-bisphosphate
      XuBP
      xylulose 1,5-bisphosphate
      PDB
      Protein Data Bank
      r.m.s.d.
      root mean square deviation
      E
      expect value
      RLP
      Rubisco-like protein
      MhR
      M. hollandica Rubisco
      H6-Ub
      His6-ubiquitin
      ESRF
      European Synchrotron Research Facility
      Bicine
      N,N-bis(2-hydroxyethyl)glycine
      IMAC
      immobilized metal affinity chromatography.
      carboxylase/oxygenase (Rubisco) catalyzes the carboxylation of RuBP to yield two molecules of 3-phosphoglycerate (3-PGA). In the majority of organisms harboring Rubisco, this is the initial and rate-limiting reaction in the reductive pentose phosphate cycle of photosynthesis leading to the incorporation of CO2 into the biosphere. Rubisco exhibits a slow CO2 fixation rate and poorly discriminates between substrate CO2 and O2. Fixation of O2 produces the toxic compound 2-phosphoglycolate whose recycling consumes energy and releases fixed CO2. Discerning between CO2 and O2 at the molecular level is fundamentally a difficult task and was unnecessary for ancestral Rubisco that arose under relatively sparse atmospheric O2 concentrations. Rubisco's catalytic inefficiencies often limit plant growth and resource-use-efficiency (
      • Andersson I.
      Catalysis and regulation in Rubisco.
      ,
      • Long S.P.
      • Zhu X.-G.
      • Naidu S.L.
      • Ort D.R.
      Can improvement in photosynthesis increase crop yields?.
      ). However, strategies to engineer more efficient plant Rubiscos have been thwarted by an inherent catalytic trade-off between turnover and specificity suggesting that higher plant Rubisco catalysis may have reached an evolutionary maximum (
      • Tcherkez G.G.
      • Farquhar G.D.
      • Andrews T.J.
      Despite slow catalysis and confused substrate specificity, all ribulose bisphosphate carboxylases may be nearly perfectly optimized.
      ). Catalytic improvements may be further complicated by a heavy reliance on co-evolved molecular partners (
      • Durão P.
      • Aigner H.
      • Nagy P.
      • Mueller-Cajar O.
      • Hartl F.U.
      • Hayer-Hartl M.
      Opposing effects of folding and assembly chaperones on evolvability of Rubisco.
      ) and an inability using current computational approaches to predict the assembly capacity and catalytic performance resulting from alterations in sequence and 3D structure (
      • Whitney S.M.
      • Houtz R.L.
      • Alonso H.
      Advancing our understanding and capacity to engineer nature's CO2-sequestering enzyme, Rubisco.
      ). However, different Rubisco lineages have faced different selection pressures, particularly Rubisco isoforms from non-photosynthetic organisms (
      • Wilson R.H.
      • Alonso H.
      • Whitney S.M.
      Evolving Methanococcoides burtonii archaeal Rubisco for improved photosynthesis and plant growth.
      ). Understanding Rubisco's evolutionary history may free us from the bias of the evolutionary constraints placed upon higher plant Rubisco isoforms when developing and designing alternative CO2-fixing enzymes.
      Catalysis in Rubisco occurs at the interface of two 50–52-kDa large subunits (LSu). This L2 dimer is the functional unit and contains two active sites. Form II (bacterial) and form III (archaeal) Rubiscos exist as such L2 dimers or as oligomers of dimers, (L2)n. In form I (higher plant, cyanobacterial, algal, and most proteobacterial) Rubiscos, four L2 dimers are assembled around a 4-fold axis by two tetramers of ∼15-kDa small subunits (SSu) that cap each end of the (L2)4 barrel. These ∼550-kDa hexadecameric L8S8 Rubiscos exhibit far superior catalytic efficiency than their (L2)n counterparts. The SSu is thought to have evolved before the divergence of proteobacteria and cyanobacteria, but after the transfer of an archaeal form III Rubisco to eubacteria (
      • Tabita F.R.
      • Hanson T.E.
      • Li H.
      • Satagopan S.
      • Singh J.
      • Chan S.
      Function, structure, and evolution of the RubisCO-like proteins and their RubisCO homologs.
      ), and may share a common ancestor with a protein involved in Rubisco compartmentalization in cyanobacteria (
      • Price G.D.
      • Howitt S.M.
      • Harrison K.
      • Badger M.R.
      Analysis of a genomic DNA region from the cyanobacterium Synechococcus sp. strain PCC7942 involved in carboxysome assembly and function.
      ).
      Form III Rubiscos do not contribute to photosynthetic processes in vivo, but because the catalytic mechanism is conserved, they can functionally substitute for photosynthetic Rubiscos (
      • Finn M.W.
      • Tabita F.R.
      Synthesis of catalytically active form III ribulose 1,5-bisphosphate carboxylase/oxygenase in archaea.
      ), albeit with a low specificity factor and carboxylation rate (
      • Alonso H.
      • Blayney M.J.
      • Beck J.L.
      • Whitney S.M.
      Substrate-induced assembly of Methanococcoides burtonii d-ribulose-1,5-bisphosphate carboxylase/oxygenase dimers into decamers.
      ). The Rubisco from the Antarctic-dwelling methanogenic archaeon Methanococcoides burtonii likely acts as part of the AMP/ADP-recycling pathway by scavenging the RuBP by-product (
      • Finn M.W.
      • Tabita F.R.
      Modified pathway to synthesize ribulose 1,5-bisphosphate in methanogenic archaea.
      ,
      • Sato T.
      • Atomi H.
      • Imanaka T.
      Archaeal type III RuBisCOs function in a pathway for AMP metabolism.
      ). M. burtonii Rubisco (MbR) has many intriguing characteristics. Five L2 dimers assemble into a rare (L2)5 structure observed exclusively in thermostable archaeal Rubiscos (
      • Kitano K.
      • Maeda N.
      • Fukui T.
      • Atomi H.
      • Imanaka T.
      • Miki K.
      Crystal structure of a novel-type archaeal Rubisco with pentagonal symmetry.
      ,
      • Maeda N.
      • Kanai T.
      • Atomi H.
      • Imanaka T.
      The unique pentagonal structure of an archaeal Rubisco is essential for its high thermostability.
      ). Unlike other (L2)5 enzymes, MbR oligomerization is substrate-stimulated (
      • Alonso H.
      • Blayney M.J.
      • Beck J.L.
      • Whitney S.M.
      Substrate-induced assembly of Methanococcoides burtonii d-ribulose-1,5-bisphosphate carboxylase/oxygenase dimers into decamers.
      ). MbR contains a unique 26–30-amino acid insertion between α6 and β7 at the bottom of the βα-barrel (
      • Tabita F.R.
      • Hanson T.E.
      • Li H.
      • Satagopan S.
      • Singh J.
      • Chan S.
      Function, structure, and evolution of the RubisCO-like proteins and their RubisCO homologs.
      ,
      • Alonso H.
      • Blayney M.J.
      • Beck J.L.
      • Whitney S.M.
      Substrate-induced assembly of Methanococcoides burtonii d-ribulose-1,5-bisphosphate carboxylase/oxygenase dimers into decamers.
      ), and deleting this “bonus” sequence yields enzymes that are catalytically competent, albeit with altered kinetic properties, and are incapable of assembling into (L2)5 enzymes (
      • Witte B.H.
      ). MbR exhibits higher sequence homology to proteobacterial form II Rubiscos, and on this basis it has been classified as a form II Rubisco (
      • Tabita F.R.
      • Hanson T.E.
      • Li H.
      • Satagopan S.
      • Singh J.
      • Chan S.
      Function, structure, and evolution of the RubisCO-like proteins and their RubisCO homologs.
      ), despite being of archaeal origin and exhibiting archaeon-like decameric assembly.
      A lack of structural information hinders MbR classification and prevents mechanistic insights into the role of the bonus sequence in substrate-induced oligomerization. Here, we describe the crystal structure of (L2)5 MbR, and by comparing MbR sequence and structure elements with other Rubisco lineages, we propose a re-evaluation of the current Rubisco classification system, establishing a new Rubisco group, form IIIB, which contains a subset of methanogenic archaeal Rubiscos with well defined sequence and phylogenetic characteristics. Biochemical approaches complement 3D structural analyses to describe the role of the unique bonus sequence as an inbuilt Rubisco SSu.

      Results

      Overall structure of MbR

      The MbR model presented here is essentially complete. Of the 474 residues, amino acids 1–473 could be modeled with the exception of residue 473 in chain C and residue 1 in chain E, which were not included in the model because of weak supporting electron density. The completeness of the MbR model is thus higher than previous structures of form II or form III Rubisco enzymes, including the Rhodopseudomonas palustris search model (Table 1). The MbR asymmetric unit contains five LSus. Applying the crystallographic symmetry results in the L10 biological unit (Fig. 1A) consistent with previous biochemical observations (
      • Alonso H.
      • Blayney M.J.
      • Beck J.L.
      • Whitney S.M.
      Substrate-induced assembly of Methanococcoides burtonii d-ribulose-1,5-bisphosphate carboxylase/oxygenase dimers into decamers.
      ).
      Table 1Comparison of available form II and form III Rubisco structures
      Rubisco formSource organismQuarternary structureNo. of amino acids per LSuNo. of amino acids per LSu in structureSequence similarity to MbRSuperposition with MbR LSuSuperposition with R. rubrum LSu
      No. of aligned residuesr.m.s.d.No. of aligned residuesr.m.s.d.
      %
      Form IIR. rubrumL2490436393511.40
      R. palustrisL6481457404241.303640.95
      Gallionellaceae sp.L6479461384241.253680.92
      Form II (?)M. burtoniiL2, L4 … L104744733521.42
      Form IIIT. kodakarensisL10444427333951.593391.67
      P. horikoshiiL8430424343851.453421.63
      Figure thumbnail gr1
      Figure 1MbR has its own inbuilt SSu that concentrates LSu dimers. A, comparison of the position of the Rubisco SSu and MbR RAD in relation to Rubisco LSus: top and side views of the 3D crystal structures of Spinacia oleracea (spinach) form I L8S8 Rubisco (LSus and SSus green and yellow, respectively; PDB code 8RUC); T. kodakarensis form II L10 Rubisco (LSus, pink; PDB code 1GEH); MbR L10 Rubisco (LSus, blue; RAD, red). In the top MbR structure, five L2 dimers are arranged around a non-crystallographic 5-fold axis perpendicular to the page. B, structure of the MbR RAD rainbow-colored from blue at the C terminus to red at the N terminus. The RAD was defined from structural alignments as residues 361–389. C, potential locking mechanism between αJ of the RAD (red) and α1 of a neighboring Rubisco LSu (blue) involves the coordination of four negatively charged side chains and a solvent molecule around a magnesium ion. D, RAD (red) packs against its own LSu (dark blue) and interacts with a neighboring LSu (light blue). The interface is formed between loop residues in the RAD and residues in α1 and α2 of the neighboring LSu. Lock site, ionic, and hydrogen-bonding interactions at the interface are shown as green, yellow, and black dashed lines, respectively. The top and side views correspond to the views shown in A.
      The overall MbR LSu structure possesses the characteristic Rubisco fold, consisting of an N-terminal domain and the C-terminal 8-stranded βα-barrel domain (see supplemental Figs. S1 and S2 for an explanation of the naming conventions for Rubisco secondary structure). The MbR structure (half of the biological unit) contains 10 Mg2+ ions, 5 Cl ions, and 294 water molecules. As expected, Mg2+ binds to the active site (5 per half-L10 decamer), stabilizing the carbamate formed at the active site lysine (
      • Lorimer G.H.
      • Badger M.R.
      • Andrews T.J.
      The activation of ribulose-1,5-bisphosphate carboxylase by carbon dioxide and magnesium ions. Equilibria, kinetics, a suggested mechanism, and physiological implications.
      ). A second binding site for Mg2+ was also inferred from difference electron density at the interface of the L2 dimers (two Mg2+ per interface and 5 per half-L10 decamer). This Mg2+ ion is coordinated by a structural motif unique to MbR, which adopts a loop-helix-loop conformation (Fig. 1B). The motif, which forms a separate domain, is encoded by a C-terminal 29-amino acid insertion in MbR (residues 361–389) and corresponds to a β-hairpin motif predicted from sequence data (
      • Tabita F.R.
      • Hanson T.E.
      • Satagopan S.
      • Witte B.H.
      • Kreel N.E.
      Phylogenetic and evolutionary relationships of RubisCO and the RubisCO-like proteins and the functional lessons provided by diverse molecular forms.
      ). The helix of the motif (named αJ) is located between helix α6 and strand β7 of the βα-barrel (supplemental Figs. S1 and S2). A search using the Dali server (
      • Holm L.
      • Rosenström P.
      Dali server: conservation mapping in 3D.
      ) did not identify structures that exhibit structural similarity to this new motif. Although the search model was at the lower size limit for the Dali server, these results suggest that the structure of this motif is unique. Because of its strategic location and its clear functional role (see below), we named it the Rubisco assembly domain (RAD). The Mg2+ ion is ligated by oxygen donor atoms provided by the side chains of Asp-366 and Asp-368 located at the N-terminal end of the RAD helix αJ (supplemental Fig. S1) and Glu-179 and Asp-183 located in helix α1 of the βα-barrel in the LSu of the neighboring L2 dimer (Fig. 1C and supplemental Fig. S1). A solvent molecule donates a further oxygen atom such that the lock site exhibits near-octahedral coordination geometry with a coordination number of 6 as expected for a magnesium ion (Fig. 1C) (
      • Harding M.M.
      Geometry of metal-ligand interactions in proteins.
      ). Whereas the carboxylates of Glu-179 and Asp-368 are monodentate, the carboxylates of Asp-183 and Asp-366 tend toward bidentate in the equatorial plane; it is likely that at least one is bidentate at any one instant. The coordination of ligands to the Mg2+ ion from two adjacent subunits potentially distorts the coordination sphere, which may be considered somewhat between octahedral and trigonal bipyramidal (
      • Zheng H.
      • Chordia M.D.
      • Cooper D.R.
      • Chruszcz M.
      • Müller P.
      • Sheldrick G.M.
      • Minor W.
      Validation of metal-binding sites in macromolecular structures with the CheckMyMetal web server.
      ). The minor deviations in geometry observed may be a result of the limited resolution or caused by crystal packing forces. However, more likely, the location at the interface between two LSus is the cause of the deviation from ideal magnesium coordination geometry. In addition to the Mg2+ lock site, the interaction surface between the RAD and the neighboring dimer is formed between residues 385 and 389 in the loop region of the RAD and residues in helices α1 and α2 in a neighboring LSu. This interface is likely stabilized by (i) salt bridges between Asp-385 and Lys-219 and Lys-222 in helix α2 and (ii) hydrogen bonding between the side chains of RAD residue Ser-387 and residue Glu-215 in helix α2. Aside from its role as a ligand to the lock site Mg2+ ion, residue Glu-179 is also engaged in hydrogen bonding to residues Trp-388 and Arg-389 of the RAD (Fig. 1D).
      The modeling of solvent and metal ions at the current resolution (2.6 Å) is non-trivial. We also considered Na+ (
      • Zheng H.
      • Chordia M.D.
      • Cooper D.R.
      • Chruszcz M.
      • Müller P.
      • Sheldrick G.M.
      • Minor W.
      Validation of metal-binding sites in macromolecular structures with the CheckMyMetal web server.
      ), which was present during crystallization, but favor Mg2+ on the grounds of geometry and chemistry. M. burtonii has a magnesium transporter, CorA, for active uptake of Mg2+ (
      • Allen M.A.
      • Lauro F.M.
      • Williams T.J.
      • Burg D.
      • Siddiqui K.S.
      • De Francisci D.
      • Chong K.W.
      • Pilak O.
      • Chew H.H.
      • De Maere M.Z.
      • Ting L.
      • Katrib M.
      • Ng C.
      • Sowers K.R.
      • Galperin M.Y.
      • et al.
      The genome sequence of the psychrophilic archaeon, Methanococcoides burtonii: the role of genome evolution in cold adaptation.
      ,
      • Smith R.L.
      • Gottlieb E.
      • Kucharski L.M.
      • Maguire M.E.
      Functional similarity between archaeal and bacterial CorA magnesium transporters.
      ), and it requires greater than 0.01 m MgSO4 or MgCl2 for growth.

      MbR assembly domain, an inbuilt SSu mimic

      The MbR RAD is located between neighboring L2 dimers in a similar position as the SSu in L8S8 Rubiscos (Fig. 1A). This indicates that it may play a similar role to the SSu, concentrating LSu dimers. This type of concentrating structure is lacking in the archaeal Thermococcus kodakarensis L10 Rubisco (Fig. 1A).
      Rubisco from T. kodakarensis is known to be more thermostable as a L10 decamer than a L2 dimer (
      • Maeda N.
      • Kanai T.
      • Atomi H.
      • Imanaka T.
      The unique pentagonal structure of an archaeal Rubisco is essential for its high thermostability.
      ), but there is as yet no molecular explanation for this stability. There is no MbR-like Mg2+-facilitated lock site in T. kodakarensis Rubisco, and packing at the T. kodakarensis Rubisco dimer-dimer interface appears to be primarily stabilized by hydrogen bonding and salt bridges. In general, ionic interactions are insufficient to maintain the structure of thermostable proteins because the pKa values of positively charged residues are sensitive to temperature fluctuations (
      • Danson M.J.
      • Hough D.W.
      • Russell R.J.
      • Taylor G.L.
      • Pearl L.
      Enzyme thermostability and thermoactivity.
      ). Therefore, the reason L10 T. kodakarensis Rubisco is so thermostable remains unclear. The optimal temperature for MbR catalysis is 55 °C (
      • Alonso H.
      • Blayney M.J.
      • Beck J.L.
      • Whitney S.M.
      Substrate-induced assembly of Methanococcoides burtonii d-ribulose-1,5-bisphosphate carboxylase/oxygenase dimers into decamers.
      ), but it is also not known whether an L10 Rubisco arrangement is intrinsically more thermostable.
      Analysis of the electrostatic surface charges of the MbR and T. kodakarensis L2 dimers indicates that the interface between L2 dimers (dimer-dimer interface) in T. kodakarensis includes a large number of charged residues positioned such that an extensive network of ionic interactions could facilitate dimer-dimer packing (Fig. 2B, solid circles). Although there are compatible charges at the MbR dimer-dimer interface (Fig. 2A), these are not as densely concentrated as in T. kodakarensis Rubisco, and the most highly charged areas are the negatively charged patches at the proposed lock site (Fig. 2A, solid circles). The solvent channel (interior) of L10 T. kodakarensis is highly negatively charged, which is not observed in the MbR structure (Fig. 2), and it may somehow contribute to T. kodakarensis thermostability. Furthermore, whereas salt bridges may act to stabilize the MbR L10 complex, the coordination of glutamic and aspartic acid residues (whose pKa are not drastically affected by temperature) to Mg2+ may represent the major locking mechanism of MbR L2 dimers, tethering them against one another to form higher oligomeric complexes.
      Figure thumbnail gr2
      Figure 2Comparison of the electrostatic surface potential of L2 Rubisco dimers from T. kodakarensis and M. burtonii. Electrostatic surface potential at the interface between L2 dimers and at the surface that lies within the Rubisco solvent channel in MbR (A) and T. kodakarensis (B). Electrostatic surfaces are colored blue in positive regions and red in negative regions. The regions corresponding to the MbR dimer-dimer lock site and the complementary charges at the T. kodakarensis Rubisco dimer-dimer interface are indicated by solid circles. The location of the RAD is indicated by dashed circles in A.

      MbR oligomerization

      Site-directed mutagenesis was undertaken to provide biochemical confirmation of the proposed α1-αJ lock site (Fig. 1C). The negatively charged residues Glu-179, Asp-183, Asp-366, and Asp-368, which appear to facilitate the RAD-locking mechanism, were mutated to alanine. The positively charged residue Lys-367, which is situated between the putative Asp-366 and Asp-368 lock residues, was also mutated to alanine to determine whether this residue is important for stabilizing the lock site. Higher order oligomer formation was only impeded by the E179A mutation (Fig. 3A). Residues Asp-183, Asp-366, and Asp-368 lie in an equatorial plane around the magnesium ion with the potential to contribute up to five bonds to the lock site (Fig. 1C), and thus disrupting any one of these three residues could be compensated for by the enzyme. In contrast, Glu-179 is the only axial protein ligand to Mg2+ (the other is provided by solvent), which could explain why the E179A mutation could disrupt the lock site and impede higher MbR oligomer formation (Fig. 1C). The long positively charged side chain of Lys-367 does not seem to be important for correct positioning of the neighboring Asp-366 and Asp-368.
      Figure thumbnail gr3
      Figure 3Oligomerization potential of MbR harboring site-specific mutations. Non-denaturing PAGE analyses of IMAC-purified single-MbR mutants (A) and double-MbR mutants (B) incubated with a 10× molar concentration of 2-CABP (relative to the number of Rubisco active sites). C, 2-CABP-bound wild-type MbR and putative lock site single mutants from A were incubated with increasing concentrations of EDTA. 5 μg of protein was loaded per lane. Lane m, protein molecular mass marker, sizes shown in kDa; lane C, pHUE empty-vector negative control; lane WT, wild-type MbR (positive control); lane WT-tag cleaved, wild-type MbR without a H6-Ub tag. Protein bands corresponding to distinct MbR oligomeric states are indicated.
      Mutants E179A, D183A, D366A, and D368A MbR on the dimer-dimer interface exhibited altered migration on non-denaturing PAGE (Fig. 3A). In contrast, the K367A mutant migrated like unmodified MbR. Clearly, disrupting any one of the negative charges of the protein side chains at the lock site alters the charge to mass ratio and/or the hydrodynamic size of L2 MbR. However, no difference in L2 (or higher oligomer) migration was observed in the presence of 2-carboxy-arabinitol 1,5-bisphosphate (2-CABP). These data suggest that a structural rearrangement is induced in both the assembly domain and helix α1 upon substrate binding, even if the mutant enzyme is unable to oligomerize.
      In the presence of 2-CABP, MbR with the D183A or D366A mutations was predominantly in the L10 form (Fig. 3A) rather than a mixture of oligomeric assemblies observed in wild-type MbR and the D368A mutant or predominantly L2 as observed in the E179A mutant. The destabilizing effect caused by mutating Asp-183 and Asp-366 (i.e. the lock site residues that tend toward bidentate) could thus be regarded as intermediate between that of the E179A and D368A mutations.
      To evaluate how tightly the magnesium ion is bound at the dimer-dimer lock site in higher order MbR complexes, the 2-CABP-bound MbR and the single mutant variants (Fig. 3A) were incubated with increasing concentrations of EDTA. Non-denaturing PAGE analyses indicated that EDTA was unable to pull apart higher order MbR complexes, even in the presence of 2-fold molar concentrations of EDTA (relative to the Mg2+ concentration in the buffer used for Rubisco activation and 2-CABP binding; Fig. 3C). The inability of EDTA to chelate MbR-bound Mg2+ indicates that the complex formed between Glu-179, Asp-183, Asp-366, and Asp-368 and Mg2+ is stable. Similarly, EDTA cannot chelate Mg2+ bound to the carbamylated lysine at the Rubisco active site (
      • Zhu G.
      • Jensen R.G.
      Status of the substrate binding sites of ribulose bisphosphate carboxylase as determined with 2-C-carboxyarabinitol 1,5-bisphosphate.
      ).
      The double mutations D183A/D366A, D183A/D366A, and D366A/D368A were also introduced into MbR. All double mutants impeded oligomerization of MbR dimers (Fig. 3B), confirming the importance of negatively charged side chains at the lock site for MbR oligomerization.

      Substrate-induced oligomerization; the ligand is important

      It has been previously demonstrated that MbR oligomerization could be induced by substrate RuBP and the reaction-intermediate analog 2-CABP, whereas the Calvin-Benson-Bassham cycle intermediates 6-phosphogluconate, ribose 5-phosphate, fructose 6-phosphate, or fructose 1,6-bisphosphate failed to induce oligomerization (
      • Alonso H.
      • Blayney M.J.
      • Beck J.L.
      • Whitney S.M.
      Substrate-induced assembly of Methanococcoides burtonii d-ribulose-1,5-bisphosphate carboxylase/oxygenase dimers into decamers.
      ). MbR was incubated with the ligands 4-carboxyarabinitol 1,5-bisphosphate (4-CABP), xylulose 1,5-bisphosphate (XuBP), and 3-PGA to further probe what triggers MbR oligomerization. 4-CABP and XuBP are the C3 epimers of 2-CABP and RuBP, respectively (Fig. 4A); 4-CABP is a close mimic of the reaction intermediate 2-carboxy-3-keto-d-arabinitol 1,5-bisphosphate, and XuBP is a misfire product of RuBP that is formed at carbamylated catalytic sites on Rubisco (
      • Zhu G.
      • Jensen R.G.
      Fallover of ribulose 1,5-bisphosphate carboxylase/oxygenase activity: decarbamylation of catalytic sites depends on pH.
      ). Both XuBP and 4-CABP de-carbamylate the activator lysine in Rubisco during or after binding to the active site in the spinach enzyme (
      • Zhu G.
      • Jensen R.G.
      Fallover of ribulose 1,5-bisphosphate carboxylase/oxygenase activity: decarbamylation of catalytic sites depends on pH.
      ,
      • Taylor T.C.
      • Fothergill M.D.
      • Andersson I.
      A common structural basis for the inhibition of ribulose 1, 5-bisphosphate carboxylase by 4-carboxyarabinitol 1, 5-bisphosphate and xylulose 1, 5-bisphosphate.
      ). Given their stereochemistry, the presence of these sugar compounds at the Rubisco active site precludes the presence of a metal ion (
      • Taylor T.C.
      • Fothergill M.D.
      • Andersson I.
      A common structural basis for the inhibition of ribulose 1, 5-bisphosphate carboxylase by 4-carboxyarabinitol 1, 5-bisphosphate and xylulose 1, 5-bisphosphate.
      ). The product of Rubisco catalysis, 3-PGA, stabilizes loop 6 (supplemental Fig. S1) that folds over the active site during catalysis, but it does not induce complete closure of the loop (
      • Taylor T.C.
      • Andersson I.
      Structure of a product complex of spinach ribulose-1,5-bisphosphate carboxylase/oxygenase.
      ).
      Figure thumbnail gr4
      Figure 4Ligand binding to MbR. A, ligands used in this study. B, non-denaturing PAGE protein separation. Purified and activated MbR were incubated with 1 or 10× molar concentrations of 2-CABP, 4-CABP, and XuBP and 1, 10, and 1000× molar concentrations of 3-PGA. 14 μg of protein was loaded per lane. Lane m, protein molecular mass marker, sizes shown (in kDa); lane C, purified MbR incubated with crystallization buffer.
      Unlike 2-CABP, the ligands 4-CABP, XuBP, and 3-PGA did not induce MbR oligomerization above the small background level observed in negative controls (Fig. 4B). The inability of 10-fold molar concentrations (80 μm) of 4-CABP and XuBP to stimulate MbR oligomerization is striking, considering their similarity to the stereoisomers 2-CABP and RuBP, respectively, and their ability to bind to the active site of other Rubisco isoforms resulting in loop 6 closure (
      • Taylor T.C.
      • Fothergill M.D.
      • Andersson I.
      A common structural basis for the inhibition of ribulose 1, 5-bisphosphate carboxylase by 4-carboxyarabinitol 1, 5-bisphosphate and xylulose 1, 5-bisphosphate.
      ). We conclude that only 6-carbon molecules that exhibit a particular stereochemistry at the C3 carbon induced MbR oligomerization (Fig. 4).
      To examine differences between XuBP and 2-CABP binding to MbR, an MbR crystal structure was obtained to 2.8 Å resolution in the presence of XuBP (data not shown); however, poor data quality limited structural interpretation. Despite ligand binding assays indicating that L2 MbR does not oligomerize into higher order states in the presence of XuBP (Fig. 4B), the subunit arrangement in XuBP-incubated MbR crystals was L10. This L10 arrangement may be an artifact of crystallization. Just like the 2-CABP-bound MbR structure, the dimer-dimer lock site was held together by Mg2+ present in the crystallization buffer (data not shown). XuBP was in fact not bound, and the catalytic lysine was carbamylated. This suggests that either XuBP (i) cannot decarbamylate MbR or (ii) binds MbR so weakly that MbR was readily re-carbamylated by Mg2+ and NaHCO3 present in the crystallization buffer. Similarly, 4-CABP may not bind strongly, if at all, to MbR active sites.
      Comparison of the 2-CABP-bound MbR structure with available Rubisco structures with XuBP and 4-CABP bound (PDB coordinates 1RCO and 1RBO of spinach Rubisco and 1RSC of Synechococcus elongatus Rubisco) revealed no dramatic differences in active site architecture, and the identity of catalytically significant residues is conserved in MbR but indicates that the stereochemistry at C3 of XuBP and 4-CABP may not be compatible with the geometry of the active site in the activated enzyme. Similar to the situation in Rubisco from higher plants, reversal of the configuration at C3 brings the C3 hydroxyl of the inhibitor closer to the position of the carbamoyl group of the activator lysine 193 (data not shown), but instead of decarbamylating the active site lysine, as observed in Rubisco from higher plants and cyanobacteria, MbR appears to expel XuBP from the active site. The reason for the different behavior in MbR is unclear, but it may be due to a low affinity of XuBP for MbR as was observed for form II and other non-higher plant Rubiscos. Future studies should determine the affinity of XuBP for the MbR active site.
      A 1000-fold molar excess (8 mm) of 3-PGA was insufficient to induce MbR oligomerization (Fig. 4B). Assuming that MbR behaves the same in the presence of 3-PGA as other Rubisco isomers, this may indicate that loop 6 stabilization is not enough to induce oligomerization but that the complete closure of loop 6 is required. Given that only RuBP and 2-CABP can trigger MbR oligomerization, it seems likely that oligomerization is triggered by ligand binding at the active site and/or loop 6 closure.

      How may substrate binding affect the lock site?

      The RAD is linked to the catalytic site through helix α1 of the βα-barrel, which functions as a scaffold for the active site (Fig. 5). Loop 6 residue Lys-330, whose side chain can contact O7 or one of the carboxyl oxygens (O3P) of 2-CABP, is indirectly connected to the N terminus of the RAD via a 29-residue-long β-hairpin and α-helix segment. Similarly, residue Ser-399, whose main chain carbonyl contacts one of the P5 phosphate oxygens (O5P) of 2-CABP, connects to the C-terminal end of the RAD via a 9-amino acid extended structure. Residue Lys-167, located in the 8-amino acid long loop 1 of the βα-barrel that links to the N-terminal end of helix α1, binds 2-CABP at O1 and O6. The Mg2+ ion that stabilizes the carbamylated catalytic Lys-193 coordinates to O2, O3, and O6 of 2-CABP upon substrate binding. Lys-193 links via the 4-amino acid-long loop 2 to the C-terminal end of helix α1. The firm link of the active site to the RAD and/or α1 suggests that substrate binding could easily communicate some structural rearrangement to regions involved in the MbR lock site.
      Figure thumbnail gr5
      Figure 5Relative locations of the active and lock sites in MbR. A, surface representations of L10 MbR. LSus are colored dark/light blue, and the assembly domain is highlighted in red. Left and center, side view of MbR showing lock sites at the interface between MbR dimers. Right, top view showing five lock sites. Lock and active sites are indicated by white and yellow arrowheads, respectively. B, close-up view of the green boxed region in A, right, showing the relative location of the lock sites (sticks) to the active site. Mg2+ ions are shown as green spheres, and 2-CABP bound at the active site is shown as ball and sticks. The helices αJ and α1 (ribbon representation) are linked to active site residues Lys-330, Ser-399, Lys-167, and Lys-193 (sticks) through defined structural elements (schematic representation). Distances between the active- and lock- site magnesium ions are indicated.

      Structural comparison with other Rubisco enzymes of bacterial and archaeal origin

      The MbR structure was compared with available structures of form II and III Rubiscos as follows: the proteobacterial form II Rubiscos from R. palustris, Rhodospirillum rubrum, and Gallionellaceae sp., and archaeal form III Rubiscos from T. kodakarensis and Pyrococcus horikoshii. These form II and form III Rubisco structures exhibit low sequence similarity to MbR (33–40%; Table 1). R. rubrum functions as L2 dimers, whereas the R. palustris and Gallionaceae sp. Rubiscos form L6 hexamers. Unlike the form II Rubiscos, the archaeal form III P. horikoshii (L8) and T. kodakarensis (L10) Rubiscos function in non-photosynthetic pathways (Table 2).
      Table 2Rubisco classification
      The overall MbR LSu structure is highly similar to form II and form III Rubisco structures. The root mean square deviations (r.m.s.d.) between the Cα atoms of the MbR LSu and the LSus from the form II/III structures indicate that the MbR LSu is structurally more similar to the form II (mean r.m.s.d. 1.3) than the form III Rubisco LSus (mean r.m.s.d. 1.5), although the differences are small (Table 1). However, given the limited availability of structural data for form II and form III Rubiscos, the relatively low resolution of the majority of the available structures, and the small structural differences, overall, the LSu structural divergence as evaluated by r.m.s.d. is insufficient to convincingly classify MbR as a form II or form III Rubisco.
      The most prominent difference between MbR and other Rubisco structures is the presence of the assembly domain, which is unique to MbR. Apart from an additional N-terminal helix, unique to MbR, the MbR N and C termini are otherwise structurally very similar to the form III structures (supplemental Fig. S2). Residues between αB and βA in the N-terminal domain, which fold as a helix in the proteobacterial structures, form an unstructured loop in the MbR and archaeal structures (region i in supplemental Figs. S1 and S2), and the form II Rubisco structures are extended by two short extra helices at the C terminus that are not present in the form III or MbR structures (region viii supplemental Figs. S1 and S2).
      Subtle, yet consistent, differences between the form II and form III Rubisco structures may be significant: variation in the lengths and conformations of the loops that connect secondary structure elements are evident in both sequence and structural alignments (regions i–viii in supplemental Figs. S1 and S2). The MbR LSu exhibits distinctly form II-like structure in βB-βC, α3-β4, αI-βE, and βF-βG (supplemental Fig. S2F). Loop regions in which the MbR LSu assumes a form III-like structure include the loop before β1 in the C-terminal domain and ηA-αE (supplemental Fig. S2F). Loop 6 assumes different conformations in the different structures, presumably as a consequence of the character of the ligand bound in the active site in the various structures. We conclude that the MbR LSu structure exhibits the loop length and positioning characteristic of either the form II or form III Rubiscos in different loops. The archaeal ancestry of MbR is distinct, and MbR appears to occupy an intermediate sequence and structural position between the proteobacterial form II and archaeal form III Rubiscos.
      In general, R. rubrum Rubisco is structurally more similar to the form II than form III Rubiscos, as illustrated by the structural superposition (Table 1 and supplemental Fig. S2). Whereas the N-terminal domain and the bulk of the βα-barrel in the LSu of the neighboring barrel superimpose almost perfectly, the C-terminal region of R. rubrum Rubisco is shifted with respect to both form II and form III structures; this shift includes helix 7 and helix 8 of the βα-barrel and the remaining C-terminal region. In this respect, the R. rubrum structure appears to be a structural outlier within the form II/III groups. A similar observation was also made in a comparison with the form I spinach enzyme (
      • Schneider G.
      • Knight S.
      • Andersson I.
      • Brändén C.I.
      • Lindqvist Y.
      • Lundqvist T.
      Comparison of the crystal structures of L2 and L8S8 Rubisco suggests a functional role for the small subunit.
      ).

      Distribution of the assembly domain sequence

      The length of the RAD sequence (29 amino acids) complicates the identification of sequences that exhibit statistically convincing homology. A BLAST search of the NCBI database yields no protein sequences with Expect values (E) below 0.1 that exhibit similarity to the assembly motif sequence from MbR, apart from a subset of Rubiscos from the Methanosarcinales order (sequences marked with an asterisk in supplemental Table S2 and a gray arrow in supplemental Fig. S3). The MbR RAD exhibited the highest sequence identity (90%) to the RAD from Methanococcoides methylutens. Although there were many hits to the full MbR sequence in metagenomic databases (data not shown), no significant hits were obtained for the assembly motif sequence alone (all hits with E values <10 are listed in supplemental Table S1), apart from one hit (E = 0.046) to the assembly motif from Methanosaeta concilii (see supplemental Table S2). The evolutionary acquisition of the assembly domain sequence thus remains unclear.
      The phylogenetic tree showing the evolutionary history of the Rubisco LSu in Fig. 6 is congruous with the Rubisco distribution reported in previous studies (
      • Tabita F.R.
      • Hanson T.E.
      • Li H.
      • Satagopan S.
      • Singh J.
      • Chan S.
      Function, structure, and evolution of the RubisCO-like proteins and their RubisCO homologs.
      ,
      • Tabita F.R.
      • Hanson T.E.
      • Satagopan S.
      • Witte B.H.
      • Kreel N.E.
      Phylogenetic and evolutionary relationships of RubisCO and the RubisCO-like proteins and the functional lessons provided by diverse molecular forms.
      ,
      • Delwiche C.F.
      Tracing the thread of plastid diversity through the tapestry of life.
      ). Form I Rubiscos and their subgroups cluster as a lineage distinct from other Rubisco forms. Although there are bootstrap values less than 95% for branch points leading to the form II, form III, and Rubisco-like protein (RLP) divisions, individual clades within these divisions are statistically convincing (supplemental Fig. S3). Despite overall uncertainty about the evolutionary history of the form III Rubiscos, the form II and RAD Rubisco lineages clearly diverge from the form III sequences.
      Figure thumbnail gr6
      Figure 6Unrooted minimum evolution phylogenetic tree of Rubisco LSu sequences. The optimal unrooted Rubisco LSu tree with the sum of branch length = 12.58 is shown. The tree is drawn to scale, with branch lengths in the same units as those of the evolutionary distances used to infer the phylogenetic tree. The evolutionary distances are in the units of the number of amino acid differences per site. The analysis included 15 representative sequences from each of the forms IA–ID, form II, form III, and form IV (RLPs) Rubisco groups and subgroups and all available Rubisco sequences from the Methanosarcinales order. The sequences used for phylogenetic reconstruction, and their homology to MbR, are included in . Bootstrap values ≥95% (** = 100%, * = 95–99%) obtained after 2000 bootstrap iterations are plotted at branch points. A black arrow indicates the location of the MbR sequence.
      All Rubisco sequences within the Methanosarcinales order that contains an assembly domain sequence (labeled Assembly motif in Fig. 6) share a common ancestor with the form II Rubiscos, accounting for the RAD Rubiscos exhibiting the highest sequence similarity to form II Rubiscos. However, the RAD Rubisco clade is distinct from the form II Rubiscos. Because all positions in the multiple sequence alignment that contained gaps were deleted prior to phylogenetic analyses, the assembly domain sequence does not contribute to the tree construction, and thus the phylogenetic relationship between the RAD Rubiscos runs much deeper than simply the presence of the additional sequence.
      Most Rubisco sequences lacking the assembly domain sequence from the order Methanosarcinales cluster together in a clade within the form III Rubiscos (Fig. 6, gray arrows, also see supplemental Fig. S3) with 100% bootstrap confidence. The exceptions are the Rubisco sequences from Methanoseta harundinacea and Methanoseta thermophila, which cluster together nearby with Thermoplasmatales and Methanomassiliicoccales form III Rubiscos. The Candidatus Methanoperedens nitroreducens Rubisco sequence clusters with Methanomicrobiales form III Rubiscos. It is not surprising that closely related organisms may have divergent Rubisco sequences, especially given the rampant horizontal gene transfer of Rubisco-encoding genes. Rubisco phylogeny is often incongruent with phylogenies created using non-Rubisco-encoding genes (
      • Delwiche C.F.
      Tracing the thread of plastid diversity through the tapestry of life.
      ,
      • Delwiche C.F.
      • Palmer J.D.
      Rampant horizontal transfer and duplication of rubisco genes in eubacteria and plastids.
      ).
      Methanomethylovorans hollandica appears to harbor two distinct Rubisco-encoding genes annotated in this study as “M. hollandica (A)” (MhR-A) and “M. hollandica (B)” (MhR-B) (supplemental Table S2). MhR-B contains an assembly domain sequence, whereas MhR-A does not. The full MhR-A and MhR-B sequences exhibit only 32% identity. When the RAD sequences are excluded from analyses, MhR-A and MhR-B exhibit 31% sequence identity. Thus, just like Methanosarcinales Rubiscos with and without the RAD sequence (supplemental Fig. S3), the MhR-A and MhR-B sequences differ much more fundamentally than just the presence/absence of the assembly domain sequence. MhR-A branches early from all the Methanosarcinales Rubiscos lacking the RAD sequence with 100% bootstrap confidence. The function of both Rubisco isoforms present in M. hollandica have not been studied, and thus it is not known whether MhR-A and MhR-B have distinct functional roles or how M. hollandica acquired two distinct Rubisco isoforms. Further work is required to find the evolutionary origin of the RAD Rubiscos. It is conceivable that a duplication event and subsequent divergence of one of the copies of Rubisco in an archaeal species could have given rise to the RAD Rubiscos.

      Rubisco sequence identity and Rubisco classification

      MbR exhibits up to 41% identity to form II Rubiscos and up to 37% to the form III Rubiscos (supplemental Table S2). However, using this small difference in sequence similarity is not a very convincing basis for Rubisco classification as a form II Rubisco. In fact, Methanosarcinales RAD Rubiscos share similar amino acid identities with the form II proteobacterial (36–41% identity) and the form III archaeal Rubiscos (33–37% identity), but those without a RAD sequence are more similar to the form III (35–52% identity) than form II Rubiscos (25–33% identity; supplemental Table S2). The MbR sequence was compared with all archaeal Rubisco sequences available in the NCBI database (831 sequences). MbR is much more similar to RAD Rubiscos, exhibiting ≥74% sequence identity with archaeal Rubiscos that do contain an assembly domain sequence and ≤38% identity with those that do not contain an assembly domain sequence (data not shown).

      Discussion

      Re-evaluating Rubisco classification

      It is not possible to clearly classify MbR as a form II or form III Rubisco. MbR's initial classification as a form II enzyme based on sequence homology (
      • Tabita F.R.
      • Hanson T.E.
      • Li H.
      • Satagopan S.
      • Singh J.
      • Chan S.
      Function, structure, and evolution of the RubisCO-like proteins and their RubisCO homologs.
      ,
      • Mueller-Cajar O.
      • Badger M.R.
      New roads lead to Rubisco in archaebacteria.
      ) was appropriate given the available information, but the validity of this classification has been questioned by other researchers (
      • Alonso H.
      • Blayney M.J.
      • Beck J.L.
      • Whitney S.M.
      Substrate-induced assembly of Methanococcoides burtonii d-ribulose-1,5-bisphosphate carboxylase/oxygenase dimers into decamers.
      ) and warrants reconsideration especially now that new structural information has been gathered.
      Under the current classification system, form I and II Rubiscos are photosynthetic; RLPs are strictly non-photosynthetic; and form III Rubiscos are found in non-photosynthetic organisms but can functionally substitute for form I and II Rubiscos in vivo (Table 2). MbR scavenges the RuBP by-product of purine/pyrimidine metabolism in M. burtonii but can also catalyze the addition of CO2 to RuBP when transplanted into photosynthetic organisms. Thus, despite the MbR amino acid sequence more closely resembling certain form II Rubiscos, the function of MbR is characteristic of form III Rubiscos. MbR kinetics are intermediate between the form II and form III Rubiscos. Different kinetic properties of MbR are either more similar to the form II (e.g. affinity for CO2) or form III Rubiscos (e.g. a much higher affinity for substrate RuBP consistent with its function recycling RuBP present in low abundance (
      • Alonso H.
      • Blayney M.J.
      • Beck J.L.
      • Whitney S.M.
      Substrate-induced assembly of Methanococcoides burtonii d-ribulose-1,5-bisphosphate carboxylase/oxygenase dimers into decamers.
      )). Uniquely, MbR exhibits high affinity for O2 and a much lower rate of oxygenation, which may be explained by the anoxygenic environment of M. burtonii.
      MbR's subunit arrangement as an oligomer of dimers is common to form II and form III Rubiscos (Table 2). MbR has a unique assembly domain, which corresponds to a unique structural variation (Fig. 1). Unique structural features can be used to clearly distinguish between Rubisco forms, e.g. a βE-βF loop is only present in the SSu of “red” Rubiscos (form IC and ID, see Table 2). Similarly, common structural variation can be used to distinguish between distinct Rubisco subgroups, i.e. the length of the βA-βB loop that varies in the Rubisco SSu of form I Rubiscos (
      • Knight S.
      • Andersson I.
      • Brändén C.I.
      Crystallographic analysis of ribulose 1,5-bisphosphate carboxylase from spinach at 2.4 A resolution. Subunit interactions and active site.
      ,
      • Newman J.
      • Gutteridge S.
      The X-ray structure of Synechococcus ribulose-bisphosphate carboxylase/oxygenase-activated quaternary complex at 2.2-A resolution.
      ,
      • Taylor T.C.
      • Backlund A.
      • Bjorhall K.
      • Spreitzer R.J.
      • Andersson I.
      First crystal structure of Rubisco from a green alga, Chlamydomonas reinhardtii.
      ).
      Sequence comparisons of Methanosarcinales Rubiscos with and without the RAD suggest that the forms lacking the RAD are clearly form III, but the Rubiscos containing the RAD cannot be clearly determined to be form II or form III Rubiscos based on sequence homology alone. Furthermore, MbR is phylogenetically distinct from both the archaeal and proteobacterial Rubiscos (Figs. 6 and supplemental Fig. S3).
      We propose that, based on MbR sequence, subunit arrangement and structure, and phylogenetic distribution and function, MbR and a subset of Rubiscos from the Methanosarcinales order belong to a new Rubisco subgroup, form IIIB. We favor introducing a new subclass of form III Rubiscos (rather than form II) on the basis of function.

      Rubisco and inhibitory ligands

      The inhibition of form I Rubisco activity upon binding of XuBP and other 6-carbon molecules is well studied (
      • Parry M.A.
      • Keys A.J.
      • Madgwick P.J.
      • Carmo-Silva A.E.
      • Andralojc P.J.
      Rubisco regulation: a role for inhibitors.
      ). XuBP is a more potent inhibitor of higher plant Rubiscos than other form I Rubiscos (

      Smrcka, A. V. (1990) Structural and Evolutionary Factors Affecting the Tight Binding of Inhibitors by Ribulose 1,5-Bisphosphate Carboxylase/Oxygenase, Ph.D. thesis, University of Arizona

      ). XuBP can also weakly bind R. rubrum Rubisco (
      • Lee E.H.
      • Harpel M.R.
      • Chen Y.R.
      • Hartman F.C.
      Perturbation of reaction-intermediate partitioning by a site-directed mutant of ribulose-bisphosphate carboxylase/oxygenase.
      ) and non-photosynthetic RLPs (
      • Saito Y.
      • Ashida H.
      • Sakiyama T.
      • de Marsac N.T.
      • Danchin A.
      • Sekowska A.
      • Yokota A.
      Structural and functional similarities between a ribulose-1,5-bisphosphate carboxylase/oxygenase (RuBisCO)-like protein from Bacillus subtilis and photosynthetic RuBisCO.
      ). No data for XuBP binding to any form III Rubiscos are available. Photosynthetic Rubiscos, in particular those from higher plants, may in fact be adapted to bind inhibitors, i.e. the Rubisco active site binding inhibitory sugars is not promiscuity on behalf of Rubisco, but rather is a selected adaptation. Tight binding of inhibitory sugars to higher plant Rubiscos necessitates a heavy reliance on complicated regulatory pathways to remove these compounds (
      • Parry M.A.
      • Keys A.J.
      • Madgwick P.J.
      • Carmo-Silva A.E.
      • Andralojc P.J.
      Rubisco regulation: a role for inhibitors.
      ). Rubisco inhibitors are produced under conditions where it is less than optimal for high rates of Rubisco catalysis, for example in the dark in the absence of energy production to drive the Calvin-Benson-Bassham cycle. In organisms with a form I Rubisco, an ATP-driven enzyme, Rubisco activase, actively removes inhibitory sugars (
      • Larson E.M.
      • O'Brien C.M.
      • Zhu G.
      • Spreitzer R.J.
      • Portis A.R.
      Specificity for activase is changed by a Pro-89 to Arg substitution in the large subunit of ribulose-1,5-bisphosphate carboxylase/oxygenase.
      ,
      • Li C.
      • Salvucci M.E.
      • Portis A.R.
      Two residues of Rubisco activase involved in recognition of the Rubisco substrate.
      ). Thus, binding of inhibitory sugars results in tighter control of Rubisco activity to ensure that Rubisco is only active when it is required.

      RAD as a Rubisco concentrating mechanism

      MbR has its own inbuilt SSu mimic that concentrates L2 dimers by the coordination of residues with negatively charged side chains around Mg2+. This unique structural variation has not been observed in any Rubisco crystal structure to date. Mechanisms that concentrate Rubisco are widespread in nature, suggesting that they confer some advantage, e.g. SSus are required for optimal activity of form I Rubiscos, higher order form III Rubisco complexes exhibit increased thermostability, and the essential pyrenoid component 1 protein concentrates and localizes Rubisco to Chlamydomonas pyrenoids (
      • Mackinder L.C.
      • Meyer M.T.
      • Mettler-Altmann T.
      • Chen V.K.
      • Mitchell M.C.
      • Caspari O.
      • Freeman Rosenzweig E.S.
      • Pallesen L.
      • Reeves G.
      • Itakura A.
      • Roth R.
      • Sommer F.
      • Geimer S.
      • Mühlhaus T.
      • Schroda M.
      • et al.
      A repeat protein links Rubisco to form the eukaryotic carbon-concentrating organelle.
      ). Rubisco SSus sequester and concentrate LSu dimers, and carbon concentrating mechanisms, such as carboxysomes in cyanobacteria and altered leaf physiology in C4 plants, not only concentrate but also act to compartmentalize Rubisco such that the relative concentrations of substrate CO2 and inhibitory O2 can be controlled. Although the Rubisco SSu shows sequence homology to one of the carboxysome shell proteins, the origin of the MbR assembly motif sequence is not clear; this sequence exhibits statistically weak homology to proteins of bacterial, viral, and archaeal origin. The RAD may have evolved completely independently of other Rubisco-concentrating mechanisms.
      It is feasible that substrate binding at the Rubisco active site could induce structural rearrangements bringing the αJ helix into closer proximity to helix α1 for the lock site to form. Although the structural data presented here provides clues as to how these messages could be communicated, there is yet insufficient data to propose a mechanism. The structure of the assembly domain in dimeric L2 MbR is unknown. The RAD could be flexible and undergo a large structural change upon substrate binding or the conformational differences between the L2 and L10 form could be more subtle.
      The structural characterization of the assembly domain from MbR offers exciting opportunities to use this structural element in engineering strategies; the domain could be used as a protein glue to force protein units to associate, particularly in a Mg2+ (or other ion)-dependent manner. Recently, there has been a lot of interest in introducing carboxysomes into higher plant chloroplasts to concentrate Rubisco and occlude inhibitory O2 to improve the catalytic performance of Rubisco (
      • Hanson M.R.
      • Lin M.T.
      • Carmo-Silva A.E.
      • Parry M.A.
      Towards engineering carboxysomes into C3 plants.
      ). The assembly domain could provide another route to concentrate Rubisco LSus, particularly in variants engineered from the more primitive form II or form III Rubisco templates via rational design or directed evolution approaches.

      Experimental procedures

      DNA cloning

      The pHUE-MbiiL plasmid (
      • Alonso H.
      • Blayney M.J.
      • Beck J.L.
      • Whitney S.M.
      Substrate-induced assembly of Methanococcoides burtonii d-ribulose-1,5-bisphosphate carboxylase/oxygenase dimers into decamers.
      ) that encodes the M. burtonii Rubisco LSu immediately downstream of His6-ubiquitin (H6-Ub) in the pHue vector (
      • Catanzariti A.-M.
      • Soboleva T.A.
      • Jans D.A.
      • Board P.G.
      • Baker R.T.
      An efficient system for high-level expression and easy purification of authentic recombinant proteins.
      ) was used as the template to introduce single site-specific mutations into MbR. The following primers were designed to introduce these mutations using the QuikChange II (Agilent) mutagenesis kit: E179A (5′-CCTCAGCGGAATATGCGGCAGTATGTTATGATTTCTG-3′); E183A (5′-GGAATATGCGGAAGTATGTTATGCTTTCTGGGTAGGTGG-3′); D366A (5′-GTCAAAGATAATGGATACCGCCAAGGATGTCATCAACCTTG-3′); K367A (5′-TGGTCAAAGATAATGGATACCGACGCCGATGTCATCAACCTTGTTAATGAG-3′); and D368A (5′-GATAATGGATACCGACAAGGCTGTCATCAACCTTGTTAATG-3′). Similarly, to introduce double mutations into MbR, the D183A primer was used to introduce a D183A mutation into the pHUE-MbiiL-D366A and pHUE-MbiiL-D368A plasmids created as described above. The primer D366A_D368A (5′-GATAATGGATACCGACAAGGCTGTCATCAACCTTGTTAATG-3′) was used to introduce a D368A mutation into the pHUE-MbiiL-D366A plasmid. All changes are underlined. Incorporation of the desired mutations within the coding region of modified plasmids was verified by BigDye terminator sequencing with a forward primer near the 3′ end of Ub, a pET-reverse primer (
      • Baker R.T.
      • Catanzariti A.M.
      • Karunasekara Y.
      • Soboleva T.A.
      • Sharwood R.
      • Whitney S.
      • Board P.G.
      Using deubiquitylating enzymes as research tools.
      ), and an additional internal sequencing primer, MbRaF (5′-GATGGGACTTACCTCAGCGG-3′). DNA sequencing was performed by SciLifeLab (Uppsala, Sweden).

      Protein expression

      The pHUE-MbiiL plasmid and mutated derivative plasmids were used to express MbR in Escherichia coli BL21 cells (Thermo Fisher Scientific). E. coli cells were grown in 4 liters (pHUE-MbiiL for crystallization) or 50 ml (mutant constructs for oligomerization experiments) of 2× YT medium (1.6% tryptone, 1% yeast extract, 0.5% NaCl) with 200 μg ml−1 ampicillin at 37 °C. Cells were cultured to mid-log phase and chilled on ice, and then protein expression was induced by addition of 0.5 mm isopropyl β-d-1-thiogalactopyranoside and incubation for 4 h at 30 °C. H6-Usp2cc was expressed and purified as described previously (
      • Catanzariti A.-M.
      • Soboleva T.A.
      • Jans D.A.
      • Board P.G.
      • Baker R.T.
      An efficient system for high-level expression and easy purification of authentic recombinant proteins.
      ).

      Protein purification

      MbR for crystallographic study and ligand-binding experiments was purified using a multistep purification procedure that involved immobilized metal affinity chromatography (IMAC), H6-Ub tag cleavage, and removal and size-exclusion chromatography.
      Cells were resuspended in cell lysis buffer (50 mm Tris-Cl, pH 8.0, 300 mm NaCl, 10 mm imidazole, 2 mm MgCl2, protease inhibitor tablets (Pierce), 1 μg ml−1 DNase I, 5% glycerol) and lysed by sonication. Soluble protein was obtained by centrifugation (48,000 × g, Beckman JL-25.50), applied to 5 ml of ProfinityTM IMAC resin (Bio-Rad), and allowed to enter the resin by gravity flow. The resin was washed extensively with Wash buffer (50 mm Tris-Cl, pH 8.0, 300 mm NaCl, 10 mm imidazole), and protein was eluted using Elution buffer (Lysis buffer supplemented with 240 mm imidazole). Peak fractions were pooled, and the H6-Ub tag was cleaved with H6-Usp2cc, as described previously (
      • Baker R.T.
      • Catanzariti A.M.
      • Karunasekara Y.
      • Soboleva T.A.
      • Sharwood R.
      • Whitney S.
      • Board P.G.
      Using deubiquitylating enzymes as research tools.
      ), where precise H6-Ub tag cleavage yields an unmodified MbR protein product.
      Protein was buffer exchanged overnight at 4 °C into Superdex buffer (100 mm Bicine, pH 8.0, 50 mm NaCl, 10 mm MgCl2, 1 mm EDTA), and then the H6-Ub tag, uncleaved protein, and H6-Ub-Usp2 were selectively removed by passage through Superdex buffer-equilibrated ProfinityTM IMAC.
      A final purification step was performed at 4 °C by size-exclusion chromatography using a HiLoad 26/60 Superdex 200 (GE Healthcare) column with a flow rate of 2 ml min−1 attached to a NGC chromatography system (Bio-Rad). Peak fractions were evaluated by non-denaturing and SDS-PAGE, pooled, and dialyzed into Crystallization buffer (100 mm HEPES-OH, pH 8.0, 10 mm MgCl2, 1 mm EDTA, 10 mm NaHCO3).
      Mutated MbR protein for oligomerization analyses was purified by IMAC as described above, except with only 1 ml of ProfinityTM IMAC resin (Bio-Rad). Purified Rubisco was activated by addition of 10 mm NaHCO3 (in addition to the 10 mm MgCl2 already present in the crystallization buffer) and incubated for 30 min at room temperature.

      Oligomerization experiments

      The oligomerization capacity of MbR was determined in the presence of different ligands and in MbR harboring site-specific substitutions. Aliquots of activated purified soluble wild-type MbR were incubated with 1- (8 μm), 10- (80 μm), or 1000-fold (8 mm) molar concentrations (relative to the number of Rubisco active sites) of the ligands 3-PGA, XuBP, 4-CABP, or 2-CABP at room temperature for 1 h. Mutant MbR proteins were analyzed in the presence of a 10-fold molar excess of 2-CABP only. Addition of crystallization buffer was used as a negative control. 2-CABP was prepared as described previously (
      • Pierce J.
      • Tolbert N.E.
      • Barker R.
      Interaction of ribulosebisphosphate carboxylase/oxygenase with transition-state analogs.
      ). 4-CABP and XuBP were gifts from G. H. Lorimer (Department of Chemistry and Biochemistry, University of Maryland). To chelate the Mg2+ bound at the dimer-dimer lock site, 2-CABP-bound wild-type and mutant MbR were incubated with equimolar (20 mm) or 2× molar (40 mm) concentrations of EDTA (relative to Mg2+ concentration in the buffer) for 30 min at room temperature.

      PAGE analyses

      Proteins were separated on 4–15% Mini-Protean TGX Stain-Free gels (Bio-Rad) in a Mini-Protean Tetra Vertical Electrophoresis Cell (Bio-Rad). Non-denaturing (native)-PAGE was run in native running buffer (25 mm Tris, 0.19 m glycine, pH 8.3). For SDS-PAGE separation, the native running buffer was supplemented with 1% SDS. PAGE-separated proteins were visualized by staining with AcquaStain (Bulldog Bio). Precision Plus Protein Unstained Protein Marker (Bio-Rad) or High Molecular Weight Native Marker (GE Healthcare) were included for size comparisons.

      Crystallization

      A 10-fold molar excess of 2-CABP was added to purified and activated MbR and incubated for 1 h at room temperature to induce oligomerization (
      • Alonso H.
      • Blayney M.J.
      • Beck J.L.
      • Whitney S.M.
      Substrate-induced assembly of Methanococcoides burtonii d-ribulose-1,5-bisphosphate carboxylase/oxygenase dimers into decamers.
      ). (L2)5-MbR was then concentrated to 5 mg ml−1 (Vivaspin, Sartorius Stedim Biotech) in Crystallization buffer (100 mm HEPES-OH, pH 8.0, 10 mm MgCl2, 1 mm EDTA, 10 mm NaHCO3). Equal volumes of (L2)5-MbR were mixed with the reservoir solution (0.1 m Tris, pH 7.8, 0.2 m KBr, 2.9% polyglutamic acid (w/v), 4.8% PEG 20K (w/v)). A purified and concentrated aliquot of MbR was also incubated with a 10-fold molar excess of XuBP and diluted to 4 mg ml−1 in Crystallization buffer before equal volumes of L2-MbR were mixed with the reservoir solution (0.1 m Tris, pH 7.0, 0.2 m l-arginine, 15% PEG 3350 (w/v)). Crystals were obtained by the hanging-drop vapor-diffusion method after equilibration at room temperature for 1–3 weeks. In preparation for data collection, crystals were transferred to a nylon loop (Hampton Research), soaked in cryo-protectant solution (reservoir solution containing 25% ethylene glycol), and then flash-cooled in liquid nitrogen.

      2-CABP-bound MbR data collection and structure determination

      X-ray diffraction data were collected on beamline ID-29 of the European Synchrotron Radiation Facility (ESRF), Grenoble, France, at 100 K with a Pilatus 6M pixel detector. Data were processed using XDS (
      • Kabsch W.
      XDS.
      ) and scaled and merged using AIMLESS (
      • Evans P.R.
      • Murshudov G.N.
      How good are my data and what is the resolution?.
      ,
      • Winn M.D.
      • Ballard C.C.
      • Cowtan K.D.
      • Dodson E.J.
      • Emsley P.
      • Evans P.R.
      • Keegan R.M.
      • Krissinel E.B.
      • Leslie A.G.
      • McCoy A.
      • McNicholas S.J.
      • Murshudov G.N.
      • Pannu N.S.
      • Potterton E.A.
      • Powell H.R.
      • et al.
      Overview of the CCP 4 suite and current developments.
      ), with 5% of reflections set aside for calculation of free R values. Calculation of the Matthews coefficient (VM = 3.96 Å3 Da−1 (
      • Matthews B.W.
      Solvent content of protein crystals.
      )), assuming an LSu molecular mass of 52,857 Da (
      • Alonso H.
      • Blayney M.J.
      • Beck J.L.
      • Whitney S.M.
      Substrate-induced assembly of Methanococcoides burtonii d-ribulose-1,5-bisphosphate carboxylase/oxygenase dimers into decamers.
      ), predicted the presence of half a decamer (L2)5, per asymmetric unit, corresponding to a solvent content of 69%. The Diffraction Anisotropy Server at UCLA (
      • Strong M.
      • Sawaya M.R.
      • Wang S.
      • Phillips M.
      • Cascio D.
      • Eisenberg D.
      Toward the structural genomics of complexes: crystal structure of a PE/PPE protein complex from Mycobacterium tuberculosis.
      ) indicated that there was almost no anisotropy. Recommended resolution limits where Fσ was above 3.0 along a, b, and c were 2.6, 2.6, and 2.7 Å, respectively. Despite high Rmerge and low CC1/2, I/σ remained above 1.0 down to a resolution of 2.6 Å. Reflections were therefore used from 48.3 to 2.6 Å. Phases were obtained by molecular replacement using Phaser (
      • McCoy A.J.
      • Grosse-Kunstleve R.W.
      • Adams P.D.
      • Winn M.D.
      • Storoni L.C.
      • Read R.J.
      Phaser crystallographic software.
      ,
      • Adams P.D.
      • Afonine P.V.
      • Bunkóczi G.
      • Chen V.B.
      • Davis I.W.
      • Echols N.
      • Headd J.J.
      • Hung L.-W.
      • Kapral G.J.
      • Grosse-Kunstleve R.W.
      • McCoy A.J.
      • Moriarty N.W.
      • Oeffner R.
      • Read R.J.
      • Richardson D.C.
      • et al.
      PHENIX: a comprehensive Python-based system for macromolecular structure solution.
      ) with one L2 dimer from the X-ray crystal structure of R. palustris Rubisco (PDB code 4LF1) as the search model. Structures were obtained by alternating between refinement using BUSTER (
      • Bricogne G.
      • Blanc E.
      • Brandl M.
      • Flensburg C.
      • Keller P.
      • Paciorek W.
      • Roversi P.
      • Sharff A.
      • Smart O.S.
      • Vonrhein C.
      • Womack T.O.
      ) and manual building in Coot (
      • Emsley P.
      • Lohkamp B.
      • Scott W.G.
      • Cowtan K.
      Features and development of Coot.
      ) and O (
      • Jones T.A.
      • Zou J.Y.
      • Cowan S.W.
      • Kjeldgaard M.
      Improved methods for building protein models in electron density maps and the location of errors in these models.
      ). Waters were inserted by alternating between automated water addition using Phenix refine (
      • Adams P.D.
      • Afonine P.V.
      • Bunkóczi G.
      • Chen V.B.
      • Davis I.W.
      • Echols N.
      • Headd J.J.
      • Hung L.-W.
      • Kapral G.J.
      • Grosse-Kunstleve R.W.
      • McCoy A.J.
      • Moriarty N.W.
      • Oeffner R.
      • Read R.J.
      • Richardson D.C.
      • et al.
      PHENIX: a comprehensive Python-based system for macromolecular structure solution.
      ,
      • Afonine P.V.
      • Grosse-Kunstleve R.W.
      • Echols N.
      • Headd J.J.
      • Moriarty N.W.
      • Mustyakimov M.
      • Terwilliger T.C.
      • Urzhumtsev A.
      • Zwart P.H.
      • Adams P.D.
      Towards automated crystallographic structure refinement with phenix.refine.
      ) and manual evaluation in Coot. The final model has an Rwork and Rfree of 0.1890 and 0.2250, respectively, for all data between 48.37 and 2.6 Å. Data collection and refinement statistics are summarized in Table 3. Ramachandran plots were calculated using the PDB Server (
      • Berman H.
      • Henrick K.
      • Nakamura H.
      Announcing the worldwide Protein Data Bank.
      ) to assess model geometry and indicated that 1.0% of the residues lie in disallowed regions. These residues were located in well defined density and exhibit no direct or obvious involvement in MbR catalysis or assembly. The mean coordinate error was 0.340 Å as calculated from a Luzzati plot.
      Table 3Data collection and refinement statistics for MbR
      Data collection
       BeamlineID29, ESRF
       Wavelength (Å)0.9763
       Space groupP321
       Unit cell parameters (Å, °)a = b = 273.8, c = 96.7, γ = 120.0
       Resolution range (Å)48.4–2.6 (2.69–2.60)
       Total no. of observations1,293,072 (49,370)
       No. of unique reflections126,745
      Rmeas
      Data are as defined by Diederichs and Karplus (69).
      0.238 (2.059)
      I/σ(I)10.8 (1.3)
      CC1/2
      CC1/2 was calculated by first randomly splitting the unmerged data into two, and then calculating the Pearson correlation coefficient between the average intensities of these two data sets (30).
      0.994 (0.231)
       Completeness (%)99.7 (93.7)
        Multiplicity10.2 (8.5)
      Refinement
       Resolution range (Å)48.4–2.6 (2.65–2.60)
       No. of reflections12,6739
      Rcryst
      Rcryst = Σhkl‖Fobs| − ‖Fcalc‖/Σhkl |Fobs|, where Fobs and Fcalc are the observed and calculated structure factor amplitudes, respectively. Rfree was calculated from 5% of randomly selected unique reflections.
      0.189
       Rfree
      Rcryst = Σhkl‖Fobs| − ‖Fcalc‖/Σhkl |Fobs|, where Fobs and Fcalc are the observed and calculated structure factor amplitudes, respectively. Rfree was calculated from 5% of randomly selected unique reflections.
      0.225
       Residues in modelA1–473, B1–473, C1–472, D1–473, E2–473
       No. of atoms19,003
       Protein18,589
       Waters294
       Mg2+10
       Cl5
       2-CABP105
       Average B-values (Å2)
       Estimated from Wilson plot78.8
       r.m.s. deviations from ideal values
       Bond lengths (Å)0.010
       Bond angles (°)1.29
       Ramachandran analysis
      From PDB Server (58).
       Outliers (%)1.0
      a Data are as defined by Diederichs and Karplus (
      • Diederichs K.
      • Karplus P.A.
      Improved R-factors for diffraction data analysis in macromolecular crystallography.
      ).
      b CC1/2 was calculated by first randomly splitting the unmerged data into two, and then calculating the Pearson correlation coefficient between the average intensities of these two data sets (
      • Karplus P.A.
      • Diederichs K.
      Linking crystallographic model and data quality.
      ).
      c Rcryst = ΣhklFobs| − ‖Fcalc‖/Σhkl |Fobs|, where Fobs and Fcalc are the observed and calculated structure factor amplitudes, respectively. Rfree was calculated from 5% of randomly selected unique reflections.
      d From PDB Server (
      • Berman H.
      • Henrick K.
      • Nakamura H.
      Announcing the worldwide Protein Data Bank.
      ).

      XuBP-bound MbR data collection and structure determination

      X-ray diffraction data were collected on beam line ID30A-3 of the ESRF at 100 K with a Pilatus3X 2M pixel detector. Data processing, scaling, and merging was performed as described for the 2-CABP-bound structure above. The 2-CABP-bound MbR crystal structure was used as the search model to obtain the XuBP-incubated MbR phases.

      Sequence and structure comparison

      The full MbR amino acid sequence (containing the assembly motif, residues 361–389) was compared with other sequences using the “Protein Sequences for Metagenomes” database from NCBI. Also searched were the ALV, AMB, BHA, BKD, DMB, EKJ, FRI, GAI, GCH, GCM, HFG, NXV, TXW, UGW, and WFB databases from MetagenomesOnline (
      • Wommack K.E.
      • Bhavsar J.
      • Polson S.W.
      • Chen J.
      • Dumas M.
      • Srinivasiah S.
      • Furman M.
      • Jamindar S.
      • Nasko D.J.
      VIROME: a standard operating procedure for analysis of viral metagenome sequences.
      ). Databases were selected to cover a range of different environments, organisms, and geographical locations. Sequence alignments were generated using BlastP 2.3.0+ (
      • Altschul S.F.
      • Madden T.L.
      • Schäffer A.A.
      • Zhang J.
      • Zhang Z.
      • Miller W.
      • Lipman D.J.
      Gapped BLAST and PSI-BLAST: a new generation of protein database search programs.
      ) employing composition-based statistics (
      • Schäffer A.A.
      • Aravind L.
      • Madden T.L.
      • Shavirin S.
      • Spouge J.L.
      • Wolf Y.I.
      • Koonin E.V.
      • Altschul S.F.
      Improving the accuracy of PSI-BLAST protein database searches with composition-based statistics and other refinements.
      ) using the default search strategy (e.g. BLOSUM62 matrix) for full Rubisco sequences and the preset search strategy optimized for short sequences (e.g. PAM30 matrix) for assembly motif alignments.
      A structural superposition of MbR with bacterial Rubisco structures in the PDB was performed with the least squares superposition function in O (
      • Jones T.A.
      • Zou J.Y.
      • Cowan S.W.
      • Kjeldgaard M.
      Improved methods for building protein models in electron density maps and the location of errors in these models.
      ). The default distance cutoff limit of 3.8 Å was used. Amino acid sequence alignments were created using ClustalOmega (
      • Sievers F.
      • Wilm A.
      • Dineen D.
      • Gibson T.J.
      • Karplus K.
      • Li W.
      • Lopez R.
      • McWilliam H.
      • Remmert M.
      • Söding J.
      • Thompson J.D.
      • Higgins D.G.
      Fast, scalable generation of high-quality protein multiple sequence alignments using Clustal Omega.
      ) and manually adjusted to reflect the structural alignment obtained in O. The graphical output was created in ESPript (
      • Robert X.
      • Gouet P.
      Deciphering key features in protein structures with the new ENDscript server.
      ). The MbR structure was also compared with other structures available in the PDB using the Dali server (
      • Holm L.
      • Rosenström P.
      Dali server: conservation mapping in 3D.
      ). The assembly domain was defined as residues 360–390 for Dali structural analysis because this resource requires a query structure with a minimum of 30 amino acids.

      Phylogenetic analyses

      A Rubisco LSu multiple sequence alignment was generated in COBALT (
      • Papadopoulos J.S.
      • Agarwala R.
      COBALT: constraint-based alignment tool for multiple protein sequences.
      ), and evolutionary analyses were conducted in MEGA6 (
      • Tamura K.
      • Stecher G.
      • Peterson D.
      • Filipski A.
      • Kumar S.
      MEGA6: molecular evolutionary genetics analysis, version 6.0.
      ). The analysis involved 128 amino acid sequences. All positions containing gaps and missing data were eliminated. There were a total of 282 positions in the final dataset. The evolutionary history of the Rubisco LSu was inferred using the Minimum Evolution method (
      • Rzhetsky A.
      • Nei M.
      A simple method for estimating and testing minimum-evolution trees.
      ). The evolutionary distances were computed using the p-distance method (
      • Nei M.
      • Kumar S.
      ). The Minimum Evolution tree was searched using the Close-Neighbor-Interchange algorithm (
      • Nei M.
      • Kumar S.
      ) at a search level of 1. The Neighbor-joining algorithm (
      • Saitou N.
      • Nei M.
      The neighbor-joining method: a new method for reconstructing phylogenetic trees.
      ) was used to generate the initial tree. Bootstrap values were obtained from 2000 bootstrap iterations.

      Electrostatic surface analysis

      Electrostatic potentials were calculated in the PyMOL Molecular Graphics System (Version 1.7.4, Schrödinger, LLC) using the generate vacuum electrostatics function. Analyses were performed on the coordinates of one Rubisco L2 dimer from T. kodakarensis (PDB code 1GEH) and MbR. All solvent molecules were removed before analysis.

      Other software

      Figures were produced using PyMOL.

      Author contributions

      L. H. G. and I. A. conceived the study, wrote the paper, and analyzed the data. L. H. G., I. A., and K. V. undertook crystal structure refinement and interpretation of the electron density. I. A. performed sequence-structural alignments. L. H. G. designed and conducted the experiments. All authors reviewed the results and approved the final version of the manuscript.

      Acknowledgments

      We acknowledge ESRF/EMBL, Grenoble, France, for providing beam time and data collection facilities. We also thank Spencer Whitney and Robert Wilson from The Australian National University for providing the pUSP2cc and pHUE-MbiiL plasmids, respectively. We are grateful to Kenta Okamoto from Uppsala University for assistance with accessing metagenomic databases.

      Supplementary Material

      Author Profile

      References

        • Andersson I.
        Catalysis and regulation in Rubisco.
        J. Exp. Bot. 2008; 59: 1555-1568
        • Long S.P.
        • Zhu X.-G.
        • Naidu S.L.
        • Ort D.R.
        Can improvement in photosynthesis increase crop yields?.
        Plant Cell Environ. 2006; 29: 315-330
        • Tcherkez G.G.
        • Farquhar G.D.
        • Andrews T.J.
        Despite slow catalysis and confused substrate specificity, all ribulose bisphosphate carboxylases may be nearly perfectly optimized.
        Proc. Natl. Acad. Sci. U.S.A. 2006; 103: 7246-7251
        • Durão P.
        • Aigner H.
        • Nagy P.
        • Mueller-Cajar O.
        • Hartl F.U.
        • Hayer-Hartl M.
        Opposing effects of folding and assembly chaperones on evolvability of Rubisco.
        Nat. Chem. Biol. 2015; 11: 148-155
        • Whitney S.M.
        • Houtz R.L.
        • Alonso H.
        Advancing our understanding and capacity to engineer nature's CO2-sequestering enzyme, Rubisco.
        Plant Physiol. 2011; 155: 27-35
        • Wilson R.H.
        • Alonso H.
        • Whitney S.M.
        Evolving Methanococcoides burtonii archaeal Rubisco for improved photosynthesis and plant growth.
        Sci. Rep. 2016; 6: 22284
        • Tabita F.R.
        • Hanson T.E.
        • Li H.
        • Satagopan S.
        • Singh J.
        • Chan S.
        Function, structure, and evolution of the RubisCO-like proteins and their RubisCO homologs.
        Microbiol. Mol. Biol. Rev. 2007; 71: 576-599
        • Price G.D.
        • Howitt S.M.
        • Harrison K.
        • Badger M.R.
        Analysis of a genomic DNA region from the cyanobacterium Synechococcus sp. strain PCC7942 involved in carboxysome assembly and function.
        J. Bacteriol. 1993; 175: 2871-2879
        • Finn M.W.
        • Tabita F.R.
        Synthesis of catalytically active form III ribulose 1,5-bisphosphate carboxylase/oxygenase in archaea.
        J. Bacteriol. 2003; 185: 3049-3059
        • Alonso H.
        • Blayney M.J.
        • Beck J.L.
        • Whitney S.M.
        Substrate-induced assembly of Methanococcoides burtonii d-ribulose-1,5-bisphosphate carboxylase/oxygenase dimers into decamers.
        J. Biol. Chem. 2009; 284: 33876-33882
        • Finn M.W.
        • Tabita F.R.
        Modified pathway to synthesize ribulose 1,5-bisphosphate in methanogenic archaea.
        J. Bacteriol. 2004; 186: 6360-6366
        • Sato T.
        • Atomi H.
        • Imanaka T.
        Archaeal type III RuBisCOs function in a pathway for AMP metabolism.
        Science. 2007; 315: 1003-1006
        • Kitano K.
        • Maeda N.
        • Fukui T.
        • Atomi H.
        • Imanaka T.
        • Miki K.
        Crystal structure of a novel-type archaeal Rubisco with pentagonal symmetry.
        Structure. 2001; 9: 473-481
        • Maeda N.
        • Kanai T.
        • Atomi H.
        • Imanaka T.
        The unique pentagonal structure of an archaeal Rubisco is essential for its high thermostability.
        J. Biol. Chem. 2002; 277: 31656-31662
        • Witte B.H.
        Taming the Wild RubisCO: Explorations in Functional Metagenomics. The Ohio State University, Columbus2012 (Ph.D. thesis)
        • Lorimer G.H.
        • Badger M.R.
        • Andrews T.J.
        The activation of ribulose-1,5-bisphosphate carboxylase by carbon dioxide and magnesium ions. Equilibria, kinetics, a suggested mechanism, and physiological implications.
        Biochemistry. 1976; 15: 529-536
        • Tabita F.R.
        • Hanson T.E.
        • Satagopan S.
        • Witte B.H.
        • Kreel N.E.
        Phylogenetic and evolutionary relationships of RubisCO and the RubisCO-like proteins and the functional lessons provided by diverse molecular forms.
        Philos. Trans. R. Soc. Lond. B Biol. Sci. 2008; 363: 2629-2640
        • Holm L.
        • Rosenström P.
        Dali server: conservation mapping in 3D.
        Nucleic Acids Res. 2010; 38: W545-W549
        • Harding M.M.
        Geometry of metal-ligand interactions in proteins.
        Acta Crystallogr. D Biol. Crystallogr. 2001; 57: 401-411
        • Zheng H.
        • Chordia M.D.
        • Cooper D.R.
        • Chruszcz M.
        • Müller P.
        • Sheldrick G.M.
        • Minor W.
        Validation of metal-binding sites in macromolecular structures with the CheckMyMetal web server.
        Nat. Protoc. 2014; 9: 156-170
        • Allen M.A.
        • Lauro F.M.
        • Williams T.J.
        • Burg D.
        • Siddiqui K.S.
        • De Francisci D.
        • Chong K.W.
        • Pilak O.
        • Chew H.H.
        • De Maere M.Z.
        • Ting L.
        • Katrib M.
        • Ng C.
        • Sowers K.R.
        • Galperin M.Y.
        • et al.
        The genome sequence of the psychrophilic archaeon, Methanococcoides burtonii: the role of genome evolution in cold adaptation.
        ISME J. 2009; 3: 1012-1035
        • Smith R.L.
        • Gottlieb E.
        • Kucharski L.M.
        • Maguire M.E.
        Functional similarity between archaeal and bacterial CorA magnesium transporters.
        J. Bacteriol. 1998; 180: 2788-2791
        • Danson M.J.
        • Hough D.W.
        • Russell R.J.
        • Taylor G.L.
        • Pearl L.
        Enzyme thermostability and thermoactivity.
        Protein Eng. 1996; 9: 629-630
        • Zhu G.
        • Jensen R.G.
        Status of the substrate binding sites of ribulose bisphosphate carboxylase as determined with 2-C-carboxyarabinitol 1,5-bisphosphate.
        Plant Physiol. 1990; 93: 244-249
        • Zhu G.
        • Jensen R.G.
        Fallover of ribulose 1,5-bisphosphate carboxylase/oxygenase activity: decarbamylation of catalytic sites depends on pH.
        Plant Physiol. 1991; 97: 1354-1358
        • Taylor T.C.
        • Fothergill M.D.
        • Andersson I.
        A common structural basis for the inhibition of ribulose 1, 5-bisphosphate carboxylase by 4-carboxyarabinitol 1, 5-bisphosphate and xylulose 1, 5-bisphosphate.
        J. Biol. Chem. 1996; 271: 32894-32899
        • Taylor T.C.
        • Andersson I.
        Structure of a product complex of spinach ribulose-1,5-bisphosphate carboxylase/oxygenase.
        Biochemistry. 1997; 36: 4041-4046
        • Schneider G.
        • Knight S.
        • Andersson I.
        • Brändén C.I.
        • Lindqvist Y.
        • Lundqvist T.
        Comparison of the crystal structures of L2 and L8S8 Rubisco suggests a functional role for the small subunit.
        EMBO J. 1990; 9: 2045-2050
        • Delwiche C.F.
        Tracing the thread of plastid diversity through the tapestry of life.
        Am. Nat. 1999; 154: S164-S177
        • Karplus P.A.
        • Diederichs K.
        Linking crystallographic model and data quality.
        Science. 2012; 336: 1030-1033
        • Delwiche C.F.
        • Palmer J.D.
        Rampant horizontal transfer and duplication of rubisco genes in eubacteria and plastids.
        Mol. Biol. Evol. 1996; 13: 873-882
        • Mueller-Cajar O.
        • Badger M.R.
        New roads lead to Rubisco in archaebacteria.
        Bioessays. 2007; 29: 722-724
        • Knight S.
        • Andersson I.
        • Brändén C.I.
        Crystallographic analysis of ribulose 1,5-bisphosphate carboxylase from spinach at 2.4 A resolution. Subunit interactions and active site.
        J. Mol. Biol. 1990; 215: 113-160
        • Newman J.
        • Gutteridge S.
        The X-ray structure of Synechococcus ribulose-bisphosphate carboxylase/oxygenase-activated quaternary complex at 2.2-A resolution.
        J. Biol. Chem. 1993; 268: 25876-25886
        • Taylor T.C.
        • Backlund A.
        • Bjorhall K.
        • Spreitzer R.J.
        • Andersson I.
        First crystal structure of Rubisco from a green alga, Chlamydomonas reinhardtii.
        J. Biol. Chem. 2001; 276: 48159-48164
        • Parry M.A.
        • Keys A.J.
        • Madgwick P.J.
        • Carmo-Silva A.E.
        • Andralojc P.J.
        Rubisco regulation: a role for inhibitors.
        J. Exp. Bot. 2008; 59: 1569-1580
      1. Smrcka, A. V. (1990) Structural and Evolutionary Factors Affecting the Tight Binding of Inhibitors by Ribulose 1,5-Bisphosphate Carboxylase/Oxygenase, Ph.D. thesis, University of Arizona

        • Lee E.H.
        • Harpel M.R.
        • Chen Y.R.
        • Hartman F.C.
        Perturbation of reaction-intermediate partitioning by a site-directed mutant of ribulose-bisphosphate carboxylase/oxygenase.
        J. Biol. Chem. 1993; 268: 26583-26591
        • Saito Y.
        • Ashida H.
        • Sakiyama T.
        • de Marsac N.T.
        • Danchin A.
        • Sekowska A.
        • Yokota A.
        Structural and functional similarities between a ribulose-1,5-bisphosphate carboxylase/oxygenase (RuBisCO)-like protein from Bacillus subtilis and photosynthetic RuBisCO.
        J. Biol. Chem. 2009; 284: 13256-13264
        • Larson E.M.
        • O'Brien C.M.
        • Zhu G.
        • Spreitzer R.J.
        • Portis A.R.
        Specificity for activase is changed by a Pro-89 to Arg substitution in the large subunit of ribulose-1,5-bisphosphate carboxylase/oxygenase.
        J. Biol. Chem. 1997; 272: 17033-17037
        • Li C.
        • Salvucci M.E.
        • Portis A.R.
        Two residues of Rubisco activase involved in recognition of the Rubisco substrate.
        J. Biol. Chem. 2005; 280: 24864-24869
        • Mackinder L.C.
        • Meyer M.T.
        • Mettler-Altmann T.
        • Chen V.K.
        • Mitchell M.C.
        • Caspari O.
        • Freeman Rosenzweig E.S.
        • Pallesen L.
        • Reeves G.
        • Itakura A.
        • Roth R.
        • Sommer F.
        • Geimer S.
        • Mühlhaus T.
        • Schroda M.
        • et al.
        A repeat protein links Rubisco to form the eukaryotic carbon-concentrating organelle.
        Proc. Natl. Acad. Sci. U.S.A. 2016; 113: 5958-5963
        • Hanson M.R.
        • Lin M.T.
        • Carmo-Silva A.E.
        • Parry M.A.
        Towards engineering carboxysomes into C3 plants.
        Plant J. 2016; 87: 38-50
        • Catanzariti A.-M.
        • Soboleva T.A.
        • Jans D.A.
        • Board P.G.
        • Baker R.T.
        An efficient system for high-level expression and easy purification of authentic recombinant proteins.
        Protein Sci. 2004; 13: 1331-1339
        • Baker R.T.
        • Catanzariti A.M.
        • Karunasekara Y.
        • Soboleva T.A.
        • Sharwood R.
        • Whitney S.
        • Board P.G.
        Using deubiquitylating enzymes as research tools.
        Methods Enzymol. 2005; 398: 540-554
        • Pierce J.
        • Tolbert N.E.
        • Barker R.
        Interaction of ribulosebisphosphate carboxylase/oxygenase with transition-state analogs.
        Biochemistry. 1980; 19: 934-942
        • Kabsch W.
        XDS.
        Acta Crystallogr. D Biol. Crystallogr. 2010; 66: 125-132
        • Evans P.R.
        • Murshudov G.N.
        How good are my data and what is the resolution?.
        Acta Crystallogr. D Biol. Crystallogr. 2013; 69: 1204-1214
        • Winn M.D.
        • Ballard C.C.
        • Cowtan K.D.
        • Dodson E.J.
        • Emsley P.
        • Evans P.R.
        • Keegan R.M.
        • Krissinel E.B.
        • Leslie A.G.
        • McCoy A.
        • McNicholas S.J.
        • Murshudov G.N.
        • Pannu N.S.
        • Potterton E.A.
        • Powell H.R.
        • et al.
        Overview of the CCP 4 suite and current developments.
        Acta Crystallogr. D Biol. Crystallogr. 2011; 67: 235-242
        • Matthews B.W.
        Solvent content of protein crystals.
        J. Mol. Biol. 1968; 33: 491-497
        • Strong M.
        • Sawaya M.R.
        • Wang S.
        • Phillips M.
        • Cascio D.
        • Eisenberg D.
        Toward the structural genomics of complexes: crystal structure of a PE/PPE protein complex from Mycobacterium tuberculosis.
        Proc. Natl. Acad. Sci. U.S.A. 2006; 103: 8060-8065
        • McCoy A.J.
        • Grosse-Kunstleve R.W.
        • Adams P.D.
        • Winn M.D.
        • Storoni L.C.
        • Read R.J.
        Phaser crystallographic software.
        J. Appl. Crystallogr. 2007; 40: 658-674
        • Adams P.D.
        • Afonine P.V.
        • Bunkóczi G.
        • Chen V.B.
        • Davis I.W.
        • Echols N.
        • Headd J.J.
        • Hung L.-W.
        • Kapral G.J.
        • Grosse-Kunstleve R.W.
        • McCoy A.J.
        • Moriarty N.W.
        • Oeffner R.
        • Read R.J.
        • Richardson D.C.
        • et al.
        PHENIX: a comprehensive Python-based system for macromolecular structure solution.
        Acta Crystallogr. D Biol. Crystallogr. 2010; 66: 213-221
        • Bricogne G.
        • Blanc E.
        • Brandl M.
        • Flensburg C.
        • Keller P.
        • Paciorek W.
        • Roversi P.
        • Sharff A.
        • Smart O.S.
        • Vonrhein C.
        • Womack T.O.
        BUSTER, version 2.10.0. Global Phasing Ltd., Cambridge, UK2011
        • Emsley P.
        • Lohkamp B.
        • Scott W.G.
        • Cowtan K.
        Features and development of Coot.
        Acta Crystallogr. D Biol. Crystallogr. 2010; 66: 486-501
        • Jones T.A.
        • Zou J.Y.
        • Cowan S.W.
        • Kjeldgaard M.
        Improved methods for building protein models in electron density maps and the location of errors in these models.
        Acta Crystallogr. A. 1991; 47: 110-119
        • Afonine P.V.
        • Grosse-Kunstleve R.W.
        • Echols N.
        • Headd J.J.
        • Moriarty N.W.
        • Mustyakimov M.
        • Terwilliger T.C.
        • Urzhumtsev A.
        • Zwart P.H.
        • Adams P.D.
        Towards automated crystallographic structure refinement with phenix.refine.
        Acta Crystallogr. D Biol. Crystallogr. 2012; 68: 352-367
        • Berman H.
        • Henrick K.
        • Nakamura H.
        Announcing the worldwide Protein Data Bank.
        Nat. Struct. Biol. 2003; 10: 980
        • Wommack K.E.
        • Bhavsar J.
        • Polson S.W.
        • Chen J.
        • Dumas M.
        • Srinivasiah S.
        • Furman M.
        • Jamindar S.
        • Nasko D.J.
        VIROME: a standard operating procedure for analysis of viral metagenome sequences.
        Stand. Genomic Sci. 2012; 6: 427-439
        • Altschul S.F.
        • Madden T.L.
        • Schäffer A.A.
        • Zhang J.
        • Zhang Z.
        • Miller W.
        • Lipman D.J.
        Gapped BLAST and PSI-BLAST: a new generation of protein database search programs.
        Nucleic Acids Res. 1997; 25: 3389-3402
        • Schäffer A.A.
        • Aravind L.
        • Madden T.L.
        • Shavirin S.
        • Spouge J.L.
        • Wolf Y.I.
        • Koonin E.V.
        • Altschul S.F.
        Improving the accuracy of PSI-BLAST protein database searches with composition-based statistics and other refinements.
        Nucleic Acids Res. 2001; 29: 2994-3005
        • Sievers F.
        • Wilm A.
        • Dineen D.
        • Gibson T.J.
        • Karplus K.
        • Li W.
        • Lopez R.
        • McWilliam H.
        • Remmert M.
        • Söding J.
        • Thompson J.D.
        • Higgins D.G.
        Fast, scalable generation of high-quality protein multiple sequence alignments using Clustal Omega.
        Mol. Syst. Biol. 2011; 7: 539
        • Robert X.
        • Gouet P.
        Deciphering key features in protein structures with the new ENDscript server.
        Nucleic Acids Res. 2014; 42: W320-W324
        • Papadopoulos J.S.
        • Agarwala R.
        COBALT: constraint-based alignment tool for multiple protein sequences.
        Bioinformatics. 2007; 23: 1073-1079
        • Tamura K.
        • Stecher G.
        • Peterson D.
        • Filipski A.
        • Kumar S.
        MEGA6: molecular evolutionary genetics analysis, version 6.0.
        Mol. Biol. Evol. 2013; 30: 2725-2729
        • Rzhetsky A.
        • Nei M.
        A simple method for estimating and testing minimum-evolution trees.
        Mol. Biol. Evol. 1992; 9: 945
        • Nei M.
        • Kumar S.
        Molecular Evolution and Phylogenetics. Oxford University Press, New York2000: 33-35 (99–103)
        • Saitou N.
        • Nei M.
        The neighbor-joining method: a new method for reconstructing phylogenetic trees.
        Mol. Biol. Evol. 1987; 4: 406-425
        • Diederichs K.
        • Karplus P.A.
        Improved R-factors for diffraction data analysis in macromolecular crystallography.
        Nat. Struct. Biol. 1997; 4: 269-275