Abstract
Rubisco catalyses the first rate-limiting step in CO2 fixation and is responsible for the vast majority of organic carbon present in the biosphere. The function and regulation of Rubisco remain an important research topic and a longstanding engineering target to enhance the efficiency of photosynthesis for agriculture and green biotechnology. The most abundant form of Rubisco (Form I) consists of eight large and eight small subunits, and is found in all plants, algae, cyanobacteria, and most phototrophic and chemolithoautotrophic proteobacteria. Although the active sites of Rubisco are located on the large subunits, expression of the small subunit regulates the size of the Rubisco pool in plants and can influence the overall catalytic efficiency of the Rubisco complex. The small subunit is now receiving increasing attention as a potential engineering target to improve the performance of Rubisco. Here we review our current understanding of the role of the small subunit and our growing capacity to explore its potential to modulate Rubisco catalysis using engineering biology approaches.
Keywords: Algae, carboxylation, oxygenation, photosynthesis, plants
A review of the role of the small subunit of Rubisco in the evolution, assembly, activity, and regulation of Rubisco, and engineering opportunities to enhance crop performance.
Introduction
Carboxylation by Rubisco has been the dominant biological driving force for inorganic carbon sequestration since the early beginnings of life on our planet. In almost all prokaryotic and eukaryotic autotrophs, Rubisco catalyses the addition of CO2 to ribulose 1,5-bisphosphate (RuBP) to form two molecules of 3-phosphoglyceric acid (3PGA) as the initial step of inorganic carbon fixation through the Calvin–Benson–Basham (CBB) cycle (Raven, 2009). However, Rubisco can also catalyse the oxygenation of RuBP to generate 2-phosphoglycolate (2PG), a toxic metabolite that inhibits two enzymes in the CBB cycle (i.e. triose phosphate isomerase and sedoheptulose 1,7-bisphosphate phosphatase) and requires recycling back to 3PGA through the photorespiratory salvage pathway (Fernie and Bauwe, 2020). While photorespiration may play a regulatory role in carbon and nitrogen metabolism (Flügel et al., 2017; Busch et al., 2018; Busch, 2020; Timm, 2020; Shi and Bloom, 2021), it is generally considered an energetically wasteful process that results in the loss of previously fixed CO2, and reduces the overall efficiency of photosynthesis and crop yield potential (Zhu et al., 2010). As such, Rubisco and the proteins associated with its assembly and regulation (the so-called ‘Rubiscosome’) have been key targets for crop improvement for several decades (Parry et al., 2013; Erb and Zarzycki, 2018; Orr and Parry, 2020; von Caemmerer, 2020; Taylor et al., 2022).
Form I Rubisco is the dominant form found today and is considered the most abundant enzyme in the living world (Ellis, 1979; Tabita et al., 2007; Bar-On and Milo, 2019; Hayer-Hartl and Hartl, 2020). It is characterized by the presence of eight small subunit (RbcS; ~12–18 kDa) protomers that hold together four large subunit (RbcL; ~50–55 kDa) dimers to form a hexadecameric L8S8 complex (~530–550 kDa) in the shape of a cylinder with a diameter and height of ~110 Å and 100 Å, respectively (Fig. 1) (Bracher et al., 2017). As each RbcL dimer forms two active sites, RbcL has long been the primary focus of attempts to engineer improvements in Rubisco performance (for reviews, see Carmo-Silva et al., 2015; Sharwood, 2017). In comparison, no RbcS residues interact directly with the active sites, and as such the RbcS has received relatively less attention. Although the RbcS is considered to play a structural role in stabilizing the L8 complex, experimental evidence has shown that the presence of the RbcS is also important for efficient assembly and maximal catalytic activity in all Form I Rubiscos (Morell et al., 1997), expression of nuclear-encoded rbcS gene(s) (i.e. in most eukaryotes) plays a key role in regulating overall Rubisco levels (Rodermel, 1999; Wostrikoff and Stern, 2007; Wietrzynski et al., 2021), and the composition of the RbcS peptide can have a significant impact on the catalytic parameters of the Rubisco complex (Box 1). Nevertheless, a comprehensive understanding of how the RbcS influences Rubisco activity remains elusive, and the structural role and functional importance of the RbcS in the evolution of Form I Rubisco are still far from fully understood (Banda et al., 2020).
Fig. 1.
Evolutionary origins and diversification of Form I Rubisco. Differently coloured nodes represent the four broad types of Form I Rubisco, and their positions on the phylogram reflect currently hypothesized evolutionary trajectories. The biophysical carbon-concentrating mechanisms (CCMs) in which the various Form I Rubisco lineages take part are shown, with ‘C’ and ‘P’ signifying carboxysomal and pyrenoid-based mechanisms, respectively. Note, pyrenoid formation has been gained and lost several times in different organisms, so not all extant Form IB (M) and ID Rubisco molecules form pyrenoids. The representative crystal structures of Form Iʹ and I Rubisco are from Promineofilum breve [Protein Data Bank (PDB) code 6URA] and Spinacia oleracea (PDB: 1RCO), respectively.
Box 1. Common catalytic parameters of Rubiscos.
Net CO2 fixation by all Rubisco forms is determined by the difference between the rates of carboxylation and oxygenation, which are ultimately determined by the maximum rates of carboxylation (Vc) and oxygenation (Vo), the carboxylation turnover rate per active site (kcatc), the Michaelis–Menten constants for CO2 (Kc and Kcair in the absence and presence of O2, respectively), and O2 (Ko), and the concentrations of CO2 and O2 at the Rubisco active site (Genkov and Spreitzer, 2009; Flamholz et al., 2019). Further parameters include the specificity of Rubisco for CO2 versus O2, often called the specificity factor (Sc/o or Ω), which can be derived from the catalytic efficiency of carboxylation relative to the catalytic efficiency of oxygenation (i.e. VcKo/VoKc) (Parry et al., 1989), Rubisco carboxylation efficiency (kcatc/Kcair) (Orr et al., 2016), and the initial response of the rate of carboxylation to the concentration of CO2 (Vc/Kcair) (Shih et al., 2016).
Although Rubisco has been the subject of a considerable number of reviews over the past years, the last dedicated review for the RbcS was produced almost 20 years ago (Spreitzer, 2003), before techniques to modify RbcS genes in land plants were well established. Several excellent reviews have since discussed newer aspects of RbcS-related research (e.g. Bracher et al., 2017; Sharwood, 2017), while advances in engineering biology [e.g. clustered regularly interspaced short palindromic repeats (CRISPR)/CRISPR-associated peptide (Cas)] have now made it possible to edit partial or entire RbcS gene families (Donovan et al., 2020; Martin-Avila et al., 2020; Matsumura et al., 2020). Furthermore, the capacity to express plant Rubiscos in Escherichia coli has opened up new opportunities to screen RbcS families, ancestral forms of RbcS, and synthetic RbcS variants (Aigner et al., 2017; Lin et al., 2020, 2022). Although progress is being made, chloroplast engineering is still only well established in a small number of plant species (Ruf et al., 2019; Yu et al., 2020). Thus, strategies to engineer RbcS are arguably more advanced and may be easier to implement broadly in crop plants than those for RbcL. Developing our understanding of the extent to which the RbcS could enhance Rubisco performance should be a critical goal going forward in efforts to engineer improvements in photosynthetic capacity (Long et al., 2019). This review is aimed as an update to Spreitzer (2003), and will include a discussion of our current understanding of the evolutionary significance of the RbcS, the role of RbcS in assembly, catalytic properties, and biophysical CO2-concentrating mechanisms, RbcS families in different plant species, and RbcS as an engineering target to enhance Rubisco performance (Long et al., 2018; Atkinson et al., 2020; Oltrogge et al., 2020).
The origins of Form I Rubisco and the small subunit
There are three known groups of Rubisco found in nature, Form I, II, and III, which differ in terms of structure and sequence (Tabita et al., 2008), and a fourth group of Rubisco-like proteins (RLPs, or Form IV) that cannot catalyse carboxylation but instead function in various bacterial pathways, including sulfur metabolism and sugar degradation (Hanson and Tabita, 2001; Ashida et al., 2003; Erb et al., 2012; Zhang et al., 2016). The most ancient of the three functional forms of Rubisco (Form III) could have emerged up to 3.5 billion years ago (bya) when the atmosphere was anoxygenic (Andersson and Backlund, 2008; Iñiguez et al., 2020), and may have evolved from the enolase enzyme family in a non-CO2-fixing archaeal ancestor (Erb and Zarzycki, 2018). Form III Rubiscos are found mainly in anaerobic archaea and are typically associated with nucleotide and nucleoside metabolism rather than CO2 fixation through the CBB cycle, with a few exceptions (Frolov et al., 2019). The discovery of Form II/III intermediates in the archaeal order Methanosarcinales has suggested that the original functional role of Rubisco may not have been to capture CO2, and that Rubisco-based autotrophy via the CBB evolved either in an ancient archaeon or later, possibly during the transfer of Form III-type Rubisco from archaea to eubacteria and the subsequent evolution of Form II and Form I Rubiscos (Tabita et al., 2007; Wrighton et al., 2016). The CBB cycle is speculated to have arisen subsequently from a primitive carbon metabolic pathway utilizing Rubisco, such as the archaeal reductive hexulose-phosphate pathway (Kono et al., 2017; Erb and Zarzycki, 2018).
All functional Rubiscos share a common core structural component of two RbcL peptides that assemble head to tail into an L2 dimer to form two surface-exposed active sites, with each active site located at the interface of the C-terminal domain of one subunit and the N-terminal domain of the other. Form II Rubiscos can consist of one or more L2 dimers, while Form III Rubiscos are assembled in oligomeric arrays of 3–5 dimers (Kitano et al., 2001). As both Form II and III Rubiscos lack RbcS, oligomerization must be facilitated by dimer–dimer interactions. For example, a 29 residue C-terminal structural domain in the Form III-type RbcL of the archaeon Methanococcoides burtonii acts as a small subunit ‘mimic’ that assists in the transition between dimeric and decameric states when substrate (i.e. RuBP) is present (Gunn et al., 2017). Oligomerization probably acts to concentrate the Rubisco active sites and increases carboxylation efficiency, which is also considered one of the key contributions of the RbcS in Form I Rubisco. Based on the available catalytic data, extant Form II and Form III Rubiscos can be fast (i.e. have high kcatc values), but both are highly sensitive to oxygen (O2), with Sc/o values ranging from 1 to 15 (Badger and Bek, 2008; Liu et al., 2017).
Form I Rubiscos bearing RbcS most probably originated after the transfer of Form III Rubisco to eubacteria but prior to the divergence of proteobacteria and cyanobacteria following the evolution of oxygenic photosynthesis ~2.9 bya in the late Archaean (Nisbet et al., 2007; Tabita et al., 2007). Thus, the atmosphere would still have been rich in CO2 (Zang et al., 2021). Phylogenetic analyses have indicated distinct periods of RbcL amino acid substitutions associated with adaptation to rising O2 stress during the transition from anaerobic Form III Rubisco to aerobic Form I Rubisco, which probably pre-date the acquisition of Form I Rubisco by fully derived cyanobacterial clades (Kacar et al., 2017). Nevertheless, the subsequent global transition from a reducing to an oxidizing atmosphere during the so-called ‘Great Oxygenic Event’ (2.3–2.5 bya) was due to the success of early cyanobacteria that had acquired Form I Rubisco (Flamholz and Shih, 2020).
The discovery of Form Iʹ Rubisco
The emergence of the RbcS is still an area of ongoing research and debate (Shih et al., 2016). Recent work has uncovered a new clade of Form I-like, or Form Iʹ, Rubiscos in the non-phototrophic Anaerolineales order of the diverse Chloroflexi phylum that oligomerize into an octameric L8 complex, akin to the hexadecameric (L8S8) complex, but lack RbcS peptides (Fig. 1) (Banda et al., 2020). This intriguing study highlights that RbcL can form octamers in the absence of RbcS via a fortified network of additional hydrogen bonds and salt bridges at the interface between L2 dimers. Although Form Iʹ may have evolved from a Form I Rubisco that subsequently lost its RbcS, the most parsimonious inference is that both Form I and Form Iʹ diverged from a common ancestor pre-dating cyanobacteria that lacked RbcS. Extant Form Iʹ Rubisco from the mesophilic Chloroflexi species ‘Candidatus Promineofilum breve’ has a low Sc/o value (36) compared with its close cyanobacterial relative Synechococcus sp. PCC 6301 (56), which has a typical RbcS-bearing Form I Rubisco. This suggests that the L8 Form Iʹ, like Form II and III, may have a limited capacity to evolve a higher specificity for CO2 to adapt to environments with elevated O2. However, the existence of a stable L8 core could have presented an opportunistic platform for the docking of ancestral RbcS-type peptides. Overall, the RbcS seems to have appeared as an evolutionary response to increasing O2 that afforded Form I Rubiscos a greater capacity to diversify their catalytic range (e.g. in terms of kcatc and SC/O), while providing a scaffolding function to concentrate eight active sites (Spreitzer, 2003; Erb and Zarzycki, 2018). Modelling studies have also indicated that residues on the RbcS may bind CO2 and potentially act as a ‘reservoir’ to increase local CO2 availability, but this has yet to be experimentally verified (Van Lun et al., 2014).
It remains unclear what kind of ancestral protein RbcS might have evolved from. The apparent sequence similarity between RbcS and CcmM35, a linker protein involved in Rubisco compartmentalization in cyanobacterial β-carboxysomes, has previously led to speculation that RbcS emerged from an early carboxysome-like bacterial microcompartment (Spreitzer, 2003). However, carboxysomes emerged a considerable time after Form I Rubisco evolved (Iñiguez et al., 2020), and probably after the primary endosymbiotic event, as eukaryotic autotrophs do not appear to possess genes that code for carboxysome-like proteins (Shih et al., 2016; Price et al., 2019). More recent sequence analyses have indicated that the resemblance of CcmM35 to RbcS reflects small, coincidental local similarities rather than a close evolutionary relationship (Ryan et al., 2019). CcmM35 also lacks key motifs used by RbcS to bind with RbcL. Cryo-EM analysis has confirmed that the RbcS-like domains of CcmM35 bind to Rubisco in a groove between two RbcL subunits and the adjacent RbcS, and do not use the RbcS-binding region or displace RbcS (Wang et al., 2019; Zang et al., 2021). Thus, it is likely that the RbcS has a different and possibly more ancient origin.
RbcS diversity, structure, and interactions
Form I Rubiscos show considerable diversification and are divided into four subgroups consisting of the ‘green-types’ Forms IA and IB and ‘red-types’ Forms IC and ID (Fig. 2A) (Tabita et al., 2008). Form IA is found in proteobacteria and cyanobacteria, and is further subdivided into Forms IAc and IAq based on differences in RbcS sequence and gene arrangements (Badger and Bek, 2008). The RbcL and RbcS genes for Form IAc Rubisco are found near an α-carboxysome operon, while Form IAq Rubisco genes are associated with the presence of the putative Rubisco chaperonin gene cbbQ, additional genes for Form II Rubisco, and the lack of an α-carboxysome operon. Form IB is the largest group and includes proteobacteria, cyanobacteria, green algae, and higher plants. It is subdivided into IB and IBc to indicate the Form IBc in cyanobacteria associated with β-carboxysomes (Badger et al., 2002). Forms IC occurs in proteobacteria and the Chloroflexi, while Form ID can be found in proteobacteria and non-green algae, such as the Rhodophyta and Haptophyta. In prokaryotes, Form I Rubisco is found as an RbcL–RbcS operon, while in most eukaryotes RbcL is found on the chloroplast genome and RbcS has been transferred to the nuclear genome [with some exceptions, including the Rhodophyta and some Chromophyta (Valentin and Zetsche, 1989; Kostrzewa et al., 1990)], and proliferated into an RbcS gene family.
Fig. 2.
Gene maps and structural alignments of Form I Rubisco small subunits. (A) Gene maps of Form I Rubisco isoforms. Forms IAq, IAc, Ibc, and IC are encoded in bacterial genomic DNA on a single operon [adapted from Badger and Bek (2008); Multiple Rubisco forms in proteobacteria: their functional significance in relation to CO2 acquisition by the CBB cycle. Journal of Experimental Botany 59, 1525–154, by permission of the Society for Experimental Biology]. The Form ID Rubisco operon is located in the chloroplast genome. The Form IB Rubisco RbcS is nuclear encoded, while the RbcL is chloroplast encoded. (B) Superpositions of RbcS from Nicotiana tabacum (T-type NtRbcS-T) in green (predicted using AlphaFold) (Laterre et al., 2017), Chlamydomonas reinhardtii (CrRbcS1) in purple (PDB: 1GK8), Synechocystis sp. PCC 6803 in cyan (AlphaFold code AF-P54206-F1), Halothiobacillus neapolitanus in grey (PDB: 1SVD), Alcaligenes eutrophus in yellow (PDB: 1BXN), and Galdieria partita in orange (1BWV) onto AtRbcS3B from Arabidopsis thaliana (At) in magenta (PDB: 5IU0). Arrows indicate secondary structural features.
The amino acid sequences of RbcL in Form I Rubiscos are relatively well conserved. For example, the RbcL sequences of Form IB Rubiscos in plants are generally 90% identical (Liu et al., 2017). In contrast, RbcS are characterized by much greater sequence diversity, with only ≥30% similarity observed across different species (Bracher et al., 2017). The core structure of each RbcL subunit is characterized by a short N-terminal domain consisting of a four-stranded β-sheet and two α-helices, and a longer C-terminal domain that forms an eight-stranded β/α-barrel (Whitney et al., 2011; Valegård et al., 2018). The conserved catalytic residues reside within the β/α-barrel, which form an active site together with residues from the N-terminal domain of the adjacent RbcL in each L2 dimer.
All RbcS also fold into a common core structure of a four-stranded antiparallel β-sheet, consisting of β-strands A–D, which are covered on one side by two α-helices (i.e. α-helices A and B) (Fig. 2B) (Knight et al., 1990). The differences between RbcS isoforms in Forms IA–D Rubiscos include the length of the loop sequence between β-strand A and B (i.e. the βA–βB loop) and the length of the C-terminus, which in red-type Rubiscos contain an additional β hairpin consisting of β-strands E and F (i.e. the βE–βF hairpin) (Spreitzer, 2003; Andersson and Backlund, 2008). Furthermore, red-type Rubiscos can have a slightly longer loop between β-strands C and D (i.e. the βC–βD loop).
The variability of the βA–βB loop
The βA–βB loop of each RbcS faces inwards towards the central solvent channel, or pore, formed by the (L2)4 assembly, while the two α-helices are solvent exposed. The length of the βA–βB loop is the most variable structural feature of the RbcS and is thought to regulate the width of the aperture of the solvent channel (Spreitzer, 2003). Prokaryotes and non-green algae are characterized by a short loop of only ~10 residues, plants have an average of 22 residues, while green algae tend to have longer βA–βB loops ranging from 20 to 31 residues. The length of the βA–βB loop has been used as a phylogenetic marker, for example to distinguish between Chlorophyte and Streptophyte algae (Goudet et al., 2020). In red-type Rubiscos, which have RbcS with shorter βA–βB loops, the C-terminal βE–βF hairpins of the four RbcS on each end of the Rubisco complex come together to form a central β-barrel around the entrance of the solvent channel (Andersson and Backlund, 2008). This fills the space typically occupied by the βA–βB loop from green-type Rubiscos, although the contacts between the βA–βB loops of each RbcS in green-type Rubiscos are less extensive.
Interactions between RbcS and RbcL
Each RbcS sits in the groove between two L2 dimers at the top and the bottom of the L8S8 complex and makes polar contacts with three RbcL and two neighbouring RbcS subunits (Fig. 3A, B; Table 1). However, for each RbcS, the majority of polar interactions occur with the β/α-barrel domains of two RbcL subunits, with the RbcS N-terminal region forming extensive interactions with one RbcL subunit that accounts for at least two-thirds of the overall interaction energy (Knight et al., 1990; Ryan et al., 2019). The second RbcL subunit interacts primarily with the βA–βB loop region of the RbcS. There are relatively fewer contacts with other subunits. For example, in spinach (Spinacia oleracea) Rubisco, two residues in the RbcS βC–βD loop interact with one of the RbcL N-terminal domains in a neighbouring L2 dimer, and L7 in the N-terminal domain and T46 in the βA–βB loop interact with an adjacent RbcS.
Fig. 3.
Interactions of the small subunit of Rubisco within the Rubisco complex and with Rubisco linker proteins. (A) Polar interactions of a given RbcS (shown in red) with adjacent subunits, including three RbcL (RbcL1, RbcL2, and RbcL3 shown in yellow, orange, and magenta, respectively) and two RbcS (RbcS1 and RbcS2 shown in green and blue, respectively) subunits are shown in an L3S3 representation of Spinacia oleracea Rubisco (PDB: 1RCO). Details of interactions with each subunit are shown in insets. (B) Sequence of SoRbcS1 from Spinacia oleracea with interacting residues highlighted [colours match subunits in (A)]. Highly and strictly conserved residues in higher plants are marked with a black dot or underlined, respectively. R65 interacts with both RbcL1 and RbcL3 (indicated with an asterisk, *). (C) Interactions between RbcS and RbcL with Rubisco linker proteins from Chlamydomonas reinhardtii (EPYC1, PDB: 7JFO), Halothiobacillus neapolitanus (CSoS2, PDB: 6UEW), and Synechococcus elongatus PCC 7942 (CcmM35, PDB: 6HBC), respectively, are shown using L2S2 representations (Wang et al., 2019; He et al., 2020; Oltrogge et al., 2020). Inset views highlight the interfaces between EPYC1, CSoS2, and CcmM35 with Rubisco subunits. Interacting residues are listed in Table 1.
Table 1.
Amino acid residue interactions with the small subunit of Rubisco
![]() |
Interactions are shown for a given RbcS with adjacent subunits in the surrounding Rubisco complex from spinach (Spinacia oleracea, PDB: 1RCO) (Fig. 3A) and between RbcS or RbcL and the Rubisco linker EPYC1 from Chlamydomonas reinhardtii (PDB: 7JFO), CsoS2 from Halothiobacillus neapolitanus (PDB: 6UEW), and CcmM35 from Synechococcus elongatus PCC 7942 (PDB: 6HBC) (from Fig. 3C). Only polar interactions are shown between the RbcS and adjacent Rubisco subunits as determined using the InterfaceResidues.py script (https://pymolwiki.org/index.php/InterfaceResidues) in PyMOL (v.2.3.2). Interactions between the linker proteins and RbcS or RbcL include salt bridge, hydrogen bond, and cation-π interactions and, for EPYC1, residues involved in the formation of a hydrophic pocket. Amino acids and structural locations shown in bold represent the focal RbcS (in red in Fig. 3A); those not in bold represent the adjacent subunit. See Fig. 3 for matching illustrations and additional information.
The role of RbcS in biophysical CO2-concentrating mechanisms
RbcS is also an important component in the assembly of Rubisco-containing micro-compartments associated with biophysical CO2-concentrating mechanisms (CCMs), namely the carboxysomes in cyanobacteria and pyrenoids in algae and hornworts (see recent reviews by Hennacy and Jonikas, 2020; Barrett et al., 2021; Borden and Savage, 2021). In both α- and β-carboxysomes, condensation of Rubisco by liquid–liquid phase separation appears to play a key part in carboxysome biogenesis (Wang et al., 2019; Oltrogge et al., 2020; Zang et al., 2021). Condensation is mediated by low affinity, multivalent interactions between Rubisco and the linker proteins CsoS2 and CcmM35 in α- and β-carboxysomes, respectively. Both proteins interact with residues on RbcL and RbcS, either by intrinsically disordered, linear motifs in CsoS2 or by folded domains in CcmM35 (Fig. 3C; Table 1). Similarly, condensation of the Rubisco matrix within the pyrenoid of the green alga Chlamydomonas reinhardtii (hereafter Chlamydomonas) is mediated by the linker protein EPYC1 via five conserved motifs that interact exclusively with the α-helices of the RbcS (Mackinder et al., 2016; Atkinson et al., 2019; He et al., 2020). This motif is a common feature of many other proteins found within the Rubisco matrix, which indicates that the RbcS plays a key role in mediating pyrenoid assembly in Chlamydomonas (Meyer et al., 2020). Nevertheless, sequences of the RbcS α-helices differ significantly between species with pyrenoids (Goudet et al., 2020), and EPYC1 is not broadly conserved. This suggests that the peptide sequences required for interactions between Rubisco and linker proteins are highly variable, which may relate to the wide diversity of different pyrenoid architectures and that pyrenoids appear to have evolved several times (Villarreal and Renner, 2012; Barrett et al., 2021).
Putative RbcS chaperones and the phenomenon of RbcS homogeneity in Rubisco
Our present understanding of the ancillary chaperones required for RbcL folding and Form I Rubisco assembly in prokaryotes and eukaryotes has been reviewed in detail recently (see Bracher et al., 2017; Hayer-Hartl and Hartl, 2020). In prokaryotes, current models show that the (L2)4 assembly is held in place by the chaperones RbcX and RAF1, which are then displaced by RbcS to form the L8S8 complex (Xia et al., 2020). In eukaryotes, an additional step is involved, where the stromal protein BSD2 displaces RbcX and RAF1, and in turn is displaced by RbcS (Aigner et al., 2017). The mechanisms regulating RbcS binding and chaperone displacement are not yet fully understood. Unlike RbcL, RbcS is able to fold spontaneously and remains monomeric when expressed heterologously in E. coli (Andrews and Ballment, 1983; Andrews, 1988), but in vivo RbcS folding might require chaperones. In eukaryotes, RbcS may require assistance for folding following import and maturation into the chloroplast, possibly by the stromal chaperone Hsp70 (Wilson et al., 2019). In some plant species, the N-terminal methionine of the mature RbcS peptide is methylated, although the functional significance remains unclear (Ying et al., 1999). Furthermore, evidence in maize (Zea mays) and Arabidopsis thaliana (hereafter Arabidopsis) suggests that RbcS might form transient complexes with RAF1, RAF2, and BSD2, which in turn facilitate docking of RbcS with the (L2)4 assembly (Feiz et al., 2014; Fristedt et al., 2018). Notably, the genomes of organisms with red-type Rubisco lack several assembly chaperones, such as RbcX and RAF1, and instead rely on the plastid-encoded RbcS for assembly, with the C-terminal β-hairpin extension of the red-type RbcS playing a critical role (Joshi et al., 2015). It is possible that the red-type RbcS may more closely represent the ancestral form of RbcS prior to the evolution of additional assembly chaperones.
In eukaryotes with green-type Rubiscos, the composition of RbcS isoforms in each Rubisco complex in vivo is still unclear (Yamada et al., 2019). One intriguing phenomenon, based on current empirical evidence, is that eukaryotic Rubiscos may assemble with only one RbcS isoform per L8S8 complex, despite the presence of a family of different nuclear-encoded RbcS isoforms in most species. Multiple examples of crystal structures exist for plant Rubiscos, with eight identical RbcS isoforms, albeit with a few potential exceptions (Shibata et al., 1996; Loewen et al., 2013). More recently, Valegård et al. (2018) described a crystal structure for Arabidopsis Rubisco that was homogenous for the low abundance isoform RbcS1B, which represents only 3–5% of the total RbcS pool (Khumsupan et al., 2020). Structural characterization of Rubisco using alternative approaches (e.g. cryo-EM) should help to elucidate if such findings are an artefact of the crystallization process, or biologically relevant. If the Rubisco assembly process does favour one RbcS isoform per L8S8 complex, this raises interesting questions about the potential mechanisms involved. One possibility is that binding of the first monomeric RbcS protomer could modify the structural conformation of the (L2)4 assembly to strongly favour the subsequent addition of the same isoform type. Although hypothetical, such a scenario could offer an efficient means of regulating the catalytic properties of the Rubisco pool through the expression of different RbcS isoforms.
Influence of the RbcS on the assembly and catalytic properties of Rubisco
The efficiency of CO2 assimilation by Rubisco appears to be constrained by several catalytic trade-offs. Mechanistic models have hypothesized an apparent inverse relationship between kcatc and Sc/o, or that any improvement in carboxylation efficiency may also improve oxygenation efficiency (for a more detailed, review see Flamholz et al., 2019). Several authors have argued that Form I Rubiscos may now be optimized to their existing environments, having reached a ‘Pareto optimality’ of activity and specificity in which neither parameter can be further improved without negatively affecting overall fitness (Tcherkez et al., 2006; Savir et al., 2010; Erb and Zarzycki, 2018). Nevertheless, new lines of evidence now suggest that Rubisco may have more room to manoeuvre than earlier thought. Firstly, previous arguments have been based on data from a relatively small pool of organisms. More expanded analyses have indicated that Rubisco has a greater degree of catalytic variability than expected and that the apparent catalytic trade-offs observed for Form I Rubiscos are weaker when larger groups are considered (Young et al., 2016; Flamholz et al., 2019). Bouvier et al. (2021) have also argued that the evolution of Rubisco in plants has been limited more by phylogenetic constraints than potential catalytic trade-offs. Such constraints may include the necessity for Rubisco to exhibit high levels of expression and protein stability, and a requirement to maintain complementarity with chaperones involved in assembly and regulation (i.e. the native Rubiscosome). Secondly, engineering efforts have screened predicted ancestral Rubiscos and produced Rubisco variants that deviate from the ‘canonical’ catalytic trade-offs between kcatc, Kc, and Sc/o (Wilson et al., 2018; Zhou and Whitney, 2019; Martin-Avila et al., 2020; Lin et al., 2022). Together, these studies suggest a wider scope for engineering improvements in the catalytic parameters of Rubisco, and that furthering our understanding of the underlying structural basis that confers the catalytic properties of Rubisco is still critically important (Valegård et al., 2018).
In the context of the RbcS, it has been clear for decades that it influences assembly and the catalytic performance of the Form I Rubisco complex, and earlier work has been well reviewed (e.g. Spreitzer, 2003). For example, even in rare cases in cyanobacteria where RbcL can assemble in the absence of RbcS, the L8 complex exhibited only 0.15% of the carboxylase activity of the L8S8 complex (Lee and Tabita, 1990), while kcatc was reduced to 0.6–1% of wild-type values (Andrews, 1988; Gutteridge, 1991). More recent studies can be divided into those that have examined the impact of native RbcS mutations on Rubisco performance, and those that have attempted to assemble heterologous RbcS with native RbcL to produce a hybrid Rubisco complex. Several of these modifications have resulted in significant changes to the catalytic parameters and/or content of Rubisco (Table 2).
Table 2.
Recent engineering studies involving modifications to the small subunit of Rubisco
Background | RbcL | RbcS | Rubisco source and modification | Rubisco characteristics | Reference |
---|---|---|---|---|---|
E. coli | AtRbcL | AtRbcS3B | Arabidopsis | Similar catalytic properties of leaf Rubisco | Aigner et al. (2017) |
NtRbcL | NtRbcSA | Tobacco NtRbcS family | All M-type RbcS: similar to leaf RubiscoRbcS-T: kcatc↑28%; KC↑108% | Lin et al. (2020) | |
TeRbcL | TeRbcS | Thermosynechococcus elongatus BP1 with PMs in RbcL and RbcS | TeRbcL (P415A)/TeRbcS (V98M):kcatc↑28%; Kc↓13%; kcatc/Kcair↑42%; SC/O↑6% | Wilson et al. (2018) | |
Chlamydomonas reinhardtii (RbcS KO mutant rbcSΔ-T60-3) | CrRbcL | CrRbcS1 | Native RbcS with vascular plant or Synechococcus elongatus PCC 7942 βA–βB loop | Spinach: Vc↓51%; Kc↓26%; Ko↓14%; SC/O≈Synechococcus: Vc↓46%; Kc≈; Ko↓30%; SC/O↓11% | Karkehabadi et al. (2005) |
CrRbcL | CrRbcS1 | Native RbcS with plantβA–βB loop and five RbcL PMs | k cat c↓44%; Vc↓45%; Kc↑26%; Vo↓42%; Ko↓12%; Sc/o↑12% | Spreitzer et al. (2005) | |
CrRbcL | CrRbcS1 | Native RbcS with various PMs | V c↓63-92%; Kc↑0-207%; Ko↑0-208%; SC/O↓0-9% | Genkov and Spreitzer (2009) | |
CrRbcL | SoRbcSAtRbcS1BHaRbcS | Hybrid Rubiscos with plant RbcS | Spinach: Vc↓13%; Kc≈; Ko↑8%; SC/O↑7%Arabidopsis: Vc↓6%; Kc≈; Ko↑10%; SC/O↑11%Sunflower: Vc≈; Kc≈; Ko↑12%; SC/O↑3% | Genkov et al. (2010) | |
CrRbcL | SoRbcS | Spinach RbcS with Chlamydomonas α-helices | V c↓88%; Kc↑17%; Ko↓43%; Sc/o↓11% | Meyer et al. (2012) | |
CrRbcL | CrRbcS1 | Chlamydomonas RbcS with PM: I58W3 | V c↓73%; Kc↑43%; Ko↑133%; SC/O↓15% | Esquivel et al. (2013) | |
CrRbcL | NtRbcS-S1 | Hybrid Rubisco with tobacco RbcS | V c↓22%; Kc↓21%; Ko↓14%; SC/O↑8% | Wachter et al. (2013) | |
CrRbcL | NtRbcS-S1 NtRbcS-Tc | Tobacco M-type or T-type RbcS | NtRbcS-S1 (M-Type): Vc↓13%; Kc↓24%NtRbcS-Tc (T-type): Vc↑7%; Kc↑9% | Laterre et al. (2017) | |
Arabidopsis thaliana (Col-0) | AtRbcL | PsRbcS3A | Wild type with pea RbcS | Pea RbcS=15–18% of total RbcS pool; carboxylase activity↓11% | Getzoff et al. (1998) |
AtRbcL | AtRbcS | 1a3b mutant | RbcS transcript pool↓80%; Rubisco content↓63%; FW↓81%; Aamb↓65% | Izumi et al. (2012) | |
AtRbcL | CrRbcS2 | 1a3b mutant with Chlamydomonas RbcS | k cat c↓12%;Kc≈; Kcair≈; kcatc/Kcair≈; Sc/o↓5%; Rubisco content↑117%; FW↑463%; Vcmax↑52% | Atkinson et al. (2017) | |
AtRbcL | AtRbcS | 1a2b3b mutant | Rubisco content↓97%; FW↓99%; Vcmax↓90% | Khumsupan et al. (2020) | |
Nicotiana tabacum(cv. Samsung, NC89 or Petite Havana) | NtRbcL | NtRbcS | OE of native RbcS in plastid | Plastid RbcS assembled into Rubisco, but only represented ~1% of total RbcS pool | Whitney and Andrews (2001) |
NtRbcL | NtRbcS | OE of native RbcS in plastid. KD of nuclear RbcS | RbcS protein pool≈; Aamb↓22% | Dhingra et al. (2004) | |
NtRbcL | NtRbcS | Plastid-expressed linked RbcL–RbcS |
k cat c≈%; Kc↓15%; Ko↓45; Sc/o≈; Rubisco content↓20% | (Whitney et al., 2009) | |
LeRbcL | NtRbcS | Tomato RbcL andnative RbcS family | k cat c↑27; Rubisco content↓; DW↓23%; Vcmax↓45% | Zhang et al. (2011) | |
NtRbcL | NtRbcS | CRISPR repression of native RbcS | Rubisco content↓93%; FW↓92%; Vcmax↓58%; Aamb↓57% | Donovan et al. (2020) | |
SeRbcL | NtRbcS | S. elongatus PCC 7942 RbcL, CcmM35 and native RbcS | V c↑154%; Kc↑1444%; Vo↓64%; Ko↑47%; Sc/o↓33%Rubisco content↓91%; Growth↓; Aamb↓ | Orr et al. (2020) | |
StRbcL | NtRbcS1StRbcS1StRbcS2StRbcS3StRbcS-T | Potato RbcL with native RbcS (nuclear), or potato RbcS1, RbcS2, RbcS,3 or RbcS-T expressed in plastid | StRbcL/NtRbcS: kcatc≈;Kc≈; kcatc/Kcair≈; Sc/o≈; Rubisco content≈StRbcL/StRbcS1: kcatc≈; Kc≈; kcatc/Kcair≈; Sc/o↑5%; Rubisco content↓55%StRbcL/StRbcS2: kcatc≈; Kc≈; kcatc/Kcair≈; Sc/o≈; Rubisco content↓65%StRbcL/StRbcS3: kcatc↑13%; Kc≈; kcatc/Kcair↑7%; Sc/o↑8%; Rubisco content↓73%StRbcL/StRbcS-T: kcatc↑42%; Kc↑174%; kcatc/Kcair↓49%; Sc/o↓12%; Rubisco content↓63% | Martin-Avila et al. (2020) | |
Oryza sativa(cv. Notohikari, cv. Nipponbare, or subsp. indica cv. IR64) | OsRbcL | OsRbcS | OE of native RbcS2 | RbcS transcript pool↑166%; Rubisco content↑52%; Activity↑60%; Aamb≈ | Suzuki et al. (2007) |
OsRbcL | SbRbcS | Sorghum RbcS | k cat c↑48%; Kc↑39%; Sc/o↓12%; Rubisco content↑24%; Activation↓11%; Growth≈; Vcmax↑31% | Ishikawa et al. (2011) | |
OsRbcL | OsRbcS | OE of nativeT-type RbcS | k cat c↑46%; Kc↑200%; Sc/o↓9%; Rubisco content↑21%; Activation↓32%; DW↓23% Vcmax↑117% | Morita et al. (2014) | |
OsRbcL | SbRbcS | Sorghum RbcS andKO of native RbcS family | k cat c↑79%; Kc↑91%; Kcair↑107%; kcatc/Kcair↑15%; Sc/o↓14%; Rubisco content↓56%; Activation↑42%; DW↓57%; Vcmax≈ | Matsumura et al. (2020) | |
OsRbcL | OsRbcS | KD of native RbcS2 and RbcS4 | Rubisco content↓40%; DW↓28% | Maheshwari et al. (2021) | |
OsRbcL | OsRbcS | OE of native RbcS2 | RbcS transcript pool↑130%; Rubisco content↑28%; DW↑23% | Yoon et al. (2020) | |
OsRbcL | OsRbcS2 | OE of native RbcS2 and RCA | Rubisco content↑44%; Activation≈; Vcmax≈; Aamb↑21% at 36°C | Suganami et al. (2021) | |
Zea mays (cv. Hi-II) | ZmRbcL | ZmRbcS | OE of nativeRbcS, RbcL, and RAF1 | RbcS-RAF1: Rubisco content↑40%; Activation↓25%; Activity↑37; FW↑30%; Vcmax≈RbcSL-RAF1: Rubisco content↑35%; Activation↓16%; Activity↑47% FW↑27%; Vcmax≈ | Salesse-Smith et al. (2018) |
Impacts on the catalytic parameters of Rubisco, Rubisco content, activity, and activation state, and subsequent changes in growth and physiological phenotype are shown where data were available. For plant studies, arrows indicate change relative to the host background. For C. reinhardtii, changes are shown in comparison with the wild-type strain. For E. coli, changes are shown in comparison with measurements of Rubisco from the native host. Rubisco catalytic parameters are defined in Box 1. See Spreitzer (2003) for earlier work. Abbreviations/symbols: ≈, no significant difference; Aamb, CO2 assimilation rate at ambient CO2 concentrations; Vcmax, maximum achievable rate of Rubisco carboxylation; PM, point mutation; OE, overexpression; KD, knockdown; KO, knockout; At, Arabidopsis; Nt, tobacco; Te, Thermosynechococcus elongatus BP1; Cr, Chlamydomonas; So, spinach; Ha, sunflower; Ps, pea; Le, tomato; Se, Synechococus elongatus PCC 7942; St, potato; Os, rice; Sb, Sorghum bicolor; Zm, maize
Chlamydomonas has been used extensively to examine the importance of different residues and features of the RbcS, in particular the variable βA–βB loop region. Spreitzer et al. (2001) observed that single residue substitutions in the βA–βB loop resulted in significant changes in the catalytic efficiency and specificity of Rubisco (Khrebtukova and Spreitzer, 1996; Spreitzer et al., 2001). Furthermore, replacing the native βA–βB loop with a shorter variant from spinach or the model cyanobacterium Synechococcus elongatus PCC 7942 showed that the length of the βA–βB loop is not critical for Form IB Rubisco assembly (Karkehabadi et al., 2005). Both heterologous βA–βB loops caused similar reductions in the carboxylation rate of Rubisco, while the cyanobacterial sequence resulted in a reduction in Sc/o values. These and further studies have demonstrated that changes to the βA–βB loop can impact the performance of the active sites of Rubisco even though the loop is structurally remote (i.e. >16 Å in distance), possibly through distant interactions with RbcL in the solvent pore (Spreitzer et al., 2005; Esquivel et al., 2013). Apart from the βA–βB loop, several highly conserved residues on the RbcS that make contact with RbcL near the β/α-barrel through van der Waals or salt bridge interactions have been shown to be important (Genkov and Spreitzer, 2009; Meyer et al., 2012). For example, the substitution L18A in the RbcS N-terminal region can influence Rubisco stability, while Y32A in αA and E43A in βA impact catalytic performance.
More recent vascular plant-based studies have focused on engineering entire heterologous small subunits into host organisms to generate hybrid Rubisco complexes (Table 2). Introduction of an RbcS from C4Sorghum bicolor into rice (Oryza sativa) resulted in a hybrid Rubisco with more C4-like characteristics, including an increased kcatc and decreased Sc/o (Ishikawa et al., 2011; Matsumura et al., 2020). Furthermore, transformation of three potato (Solanum tuberosum) RbcS isoforms into the chloroplast genome in tobacco (Nicotiana tabacum) produced hybrid Rubiscos with the native NtRbcL that had significant differences in catalysis, including a 13% and 8% increase in kcatc and Sc/o, respectively, for StRbcS3 (Martin-Avila et al., 2020). Notably, StRbcS3 differs from two other StRbcS isoforms tested by two amino acid resides at the apex of the βA–βB loop. These results highlight the potential of modified or heterologous RbcS isoforms to enhance Rubisco performance in different crop species.
Introducing different plant RbcS variants into Chlamydomonas not only resulted in increased Sc/o values, more similar to that of higher plants, but also revealed the critical role of the algal RbcS α-helices in pyrenoid assembly (Genkov et al., 2010; Meyer et al., 2012). In comparison, expression of Chlamydomonas CrRbcS2 in the Arabidopsis RbcS-deficient double mutant 1a3b (i.e. lacking expression of AtRbcS1A and AtRbcS3B) produced a hybrid Rubisco pool with reduced Sc/o values (Atkinson et al., 2017). Given the evolutionary distance between plants and algae, this raises interesting questions of how broadly compatible different RbcS isoforms are for Rubisco assembly and stability in non-native species.
RbcS regulates the content and can influence the catalytic characteristics of Rubisco in eukaryotes
In eukaryotes that possess Form IB Rubisco, RbcS is encoded by a family of nuclear-encoded genes that varies in number between species. Some species have a small RbcS gene family, such as Chlamydomonas with two isoforms, while others have much larger gene families, such as tobacco and wheat (Triticum aestivum) with 13 and 25, respectively (Table 3) (Donovan et al., 2020; Caruana et al., 2022). As the RbcS is no longer located with the RbcL on a chloroplastic operon, further regulatory processes have evolved to coordinate the efficient production of both subunits in a 1:1 stoichiometric ratio. Several studies in algae, C3 plants, and C4 plants have shown that RbcL synthesis in the chloroplast is dependent on the presence of RbcS, and that knockdown of RbcS expression reduces the rate of RbcL synthesis (Khrebtukova and Spreitzer, 1996; Rodermel et al., 1996; Wostrikoff and Stern, 2007; Wostrikoff et al., 2012; Khumsupan et al., 2020). RbcL synthesis appears to be regulated by a process known as control by epistasis of synthesis, where in the absence of RbcS, the L8–RAF1 assembly intermediate (or possibly L8–BSD2 in species with BSD2) acts as a suppressor of RbcL translation (Wietrzynski et al., 2021). Thus, RbcS expression plays a key role in regulating the size of the Rubisco pool.
Table 3.
Examples of Rubisco small subunit gene families and their relative transcript abundances
Species | RbcS isoforms | Gene name | Gene accession | Relative abundance | RbcS Type-T expression | References |
---|---|---|---|---|---|---|
Chlamydomonas | 2 | CrRbcS1 | Cre02.g120100 | 44.7 | Fang et al. (2012) | |
reinhardtii | CrRbcS2 | Cre02.g120150 | 53.3 | |||
Arabidopsis | 4 | AtRbcS1A | AT1G67090 | 78.7 | Mergner et al. (2020) | |
thaliana | AtRbcS1B | AT5G38430 | 1.4 | |||
AtRbcS2B | AT5G38420 | 8.2 | ||||
AtRbcS3B | AT5G38410 | 11.7 | ||||
Glycine | 3 | GmRbcS1 | GLYMA_19G046600 | 42.5 | PhytoMine dataset | |
max | GmRbcS2 | GLYMA_19G046800 | 57.4 | |||
GmRbcS-T | GLYMA_13G097100 | 0.1 | Nodules, root | |||
Nicotiana | 13 | NtrbcS-S1a | KM025316.1 | 22.9 | Sierro et al. (2014) | |
tabacum | NtrbcS-S1b | KM025317.1 | 19.9 | |||
NtrbcS-S2 | KM025319.1 | 10.6 | ||||
NtrbcS-S3 | KM025321.1 | 1.6 | ||||
NtrbcS-S4 | KM025323.1 | 0.1 | ||||
NtrbcS-S5 | KM025325.1 | 3.1 | ||||
NtrbcS-T1 | KM025327.1 | 40.5 | ||||
NtrbcS-T2 | KM025329.1 | 0.7 | ||||
NtrbcS-T3a | KM025331.1 | 0.0 | ||||
NtrbcS-T4a | KM025334.1 | 0.7 | ||||
NtrbcS-T5 | KM025337.1 | 0.7 | ||||
NtrbcS-T | DV157962.1 | 0.0 | Trichome | |||
Oryza | 5 | OsRbcS1 | Os02g0152400 | 0.0 | Embryo, seed, pistil | Davidson et al. (2012) |
sativa | OsRbcS2 | Os12g0274700 | 50.3 | |||
OsRbcS3 | Os12g0291100 | 41.1 | ||||
OsRbcS4 | Os12g0292400 | 2.8 | ||||
OsRbcS5 | Os12g0291400 | 5.8 | ||||
Solanum | 6 | SlRbcS1 | Solyc02g063150.3 | 19.7 | Tomato Genome Consortium (2012) | |
lycopersicum | SlRbcS2 | Solyc03g034220.3 | 55.5 | |||
SlRbcS3A-C | Solyc02g085950.3 | 24.8 | ||||
Sly-t | Solyc07g017950.3 | 0.0 | Epidermis, fruit, bud | |||
Solanum | 5 | StRbcS1/pS1 | PGSC0003DMG400019584 | 56.4 | Potato Genome Sequencing Consortium (2011) | |
tuberosum | StRbcS2B/pS2 | PGSC0003DMG400026409 | 6.7 | |||
StRbcS2C | PGSC0003DMG400012666 | 5.8 | ||||
StRbcSC/ps3 | PGSC0003DMG400024182 | 31.1 | ||||
Stu-t/pST | PGSC0003DMG400004097 | 0.0 | Stamen, flower | |||
Triticum | 25 | TaRbcS1A | TraesCS2A02G066700 | 2.9 | Wheat Expression Browser | |
aestivum | TaRbcS2A | TraesCS2A02G066800 | 5.0 | |||
TaRbcS3A | TraesCS2A02G066900 | 1.5 | ||||
TaRbcS4A | TraesCS2A02G067000 | 0.4 | ||||
TaRbcS5A | TraesCS2A02G067100 | 0.8 | ||||
TaRbcS6A | TraesCS2A02G067200 | 1.3 | ||||
TaRbcS7A | TraesCS2A02G067300 | 1.6 | ||||
TaRbcS8A | TraesCS5A02G165400 | 10.3 | ||||
TaRbcS9A | TraesCS5A02G165700 | 11.9 | ||||
TaRbcS1B | TraesCS2B02G078900 | 0.9 | ||||
TaRbcS2B | TraesCS2B02G079100 | 1.1 | ||||
TaRbcS3B | TraesCS2B02G079200 | 2.5 | ||||
TaRbcS4B | TraesCS2B02G079300 | 2.2 | ||||
TaRbcS5B | TraesCS2B02G079400 | 1.1 | ||||
TaRbcS6B | TraesCS2B02G079500 | 2.6 | ||||
TaRbcS7B | TraesCS5B02G162600 | 6.4 | ||||
TaRbcS8B | TraesCS5B02G162800 | 9.3 | ||||
TaRbcS1D | TraesCS2D02G065100 | 3.2 | ||||
TaRbcS2D | TraesCS2D02G065200 | 3.6 | ||||
TaRbcS3D | TraesCS2D02G065300 | 2.2 | ||||
TaRbcS4D | TraesCS2D02G065400 | 1.2 | ||||
TaRbcS5D | TraesCS2D02G065500 | 0.0 | ||||
TaRbcS6D | TraesCS2D02G065600 | 2.4 | ||||
TaRbcS7D | TraesCS5D02G169600 | 10.7 | ||||
TaRbcS8D | TraesCS5D02G169900 | 14.9 | ||||
Sorghum bicolor | 1 | SbRbcS1 | SORBI_3005G042000 | 100 | Davidson et al. (2012) | |
Zea | 2 | RbcS1 | GRMZM2G098520 | 76.9 | NAM Consortium | |
mays | RbcS2 | GRMZM2G113033 | 23.1 |
Only RbcS isoforms with verified gene expression are shown. Italicized gene accessions indicate cluster T-type RbcS isoforms (Laterre et al., 2017). For Solanum lycopersicum, RbcS3A, RbcS3B, and RbcS3C are represented by the same accession number. Gene names for Glycine max, Triticum aestivum, and Sorghum bicolor have been assigned based on genomic location. RNA-seq data were obtained from Expression Atlas (https://www.ebi.ac.uk/gxa/home). For Chlamydomonas, G. max, N. tabacum, and T. aestivum RNA-seq data were obtained from Fang et al. (2012), PhytoMine (https://phytozome-next.jgi.doe.gov/phytomine/begin.do), Sierro et al. (2014), and Wheat Expression Browser (http://www.wheat-expression.com), respectively. Relative abundance refers to the percentage contribution of each RbcS gene to the total RbcS transcript pool for each species
Recent work has shown that increasing native RbcS expression levels can lead to improvements in growth and yield. In rice, overexpression of OsRbcS2 resulted in a 23% and 28% increase in dry weight and rice yield, respectively, compared with wild-type plants (Table 2) (Yoon et al., 2020). In maize, overexpression of ZmRbcS alone had no impact on Rubisco levels (Salesse-Smith et al., 2018). However, co-expression of ZmRbcS with RAF1 resulted in a 30% increase in both Rubisco content and dry weight, which suggests that RAF1 expression may co-regulate Rubisco levels and activity, at least in C4 plants.
The role(s) of RbcS gene families
The expansion of gene families is driven by various gene duplication processes, at both the single-gene and whole-genome level. For example, gene duplication can occur through tandem duplications, which results in gene copies being close together on the same chromosome in a tandem array (Freeling, 2009). Tandem arrays of RbcS genes are present in many organisms, including CrRbcS1 and CrRbcS2 in Chlamydomonas (Goldschmidt-Clermont and Rahire, 1986), AtRbcS1B–AtRbcS 3B in Arabidopsis (Krebbers et al., 1988), OsRbcS2–OsRbcS5 in rice (Morita et al., 2014), five of the six RbcS genes in the facultative Crassulacean acid metabolism (CAM) plant Mesembryanthemum crystallinum (Derocher et al., 1993), six of the eight RbcS genes in petunia (Petunia sp. hybrid Mitchell strain) (Dean et al., 1985), and the entire RbcS gene family in pea (Pisum sativum) (Polans et al., 1985). A further mechanism for gene duplication is transposon activity, which can transfer gene copies between different chromosomes. Transposon activity may account for the chromosomally isolated RbcS isoforms in Arabidopsis (i.e AtRbcS1A) and petunia. At the whole-genome level, duplication can occur through polyploidization (Van De Peer et al., 2017). For example, in wheat, two allopolyploidy events have resulted in the presence of three distinct diploid subgenomes named A, B, and D, which each contributed an RbcS gene family of nine, eight, and eight genes, respectively (Table 3) (Caruana et al., 2022).
Despite these duplication events, RbcS gene families generally show little sequence divergence. For example, four of the five rice RbcS genes, OsRbcS2–OsRbc5, encode the same mature RbcS peptide (Suzuki et al., 2007). Similarly, tomato (Solanum lycopersicum) and Arabidopsis each encode three RbcS genes located in a tandem array that are nearly identical (Wanner and Gruissem, 1991). Given the apparent lack of sequence divergence, the evolutionary benefit of retaining individual members within RbcS gene families is somewhat unclear, and still an active area of ongoing research. One potential explanation is that multiple RbcS gene copies ensure a sufficient supply of RbcS peptides to meet the demands of Rubisco production (Khrebtukova and Spreitzer, 1996). Retaining an RbcS gene family could also help plants to respond more dynamically to a wider range of environmental and developmental stimuli. In line with the latter argument, the promoter elements regulating the expression of different RbcS genes within a family appear diverse and indicate a high degree of subfunctionalization of expression (Sugita et al., 1987; Dedonder et al., 1993; Cheng et al., 1998; Yoon et al., 2001; Sawchuk et al., 2008). Furthermore, evidence for differential expression of RbcS genes in different organs and under different environmental conditions is available in several plant species including Arabidopsis, tomato, tobacco, rice, and maize (Ewing et al., 1998; Narsai et al., 2009; Zhang et al., 2012; Laterre et al., 2017; Buti et al., 2018).
An alternative hypothesis is that RbcS isoforms have evolved for specific functions that impact on the catalytic properties of Rubisco. The apparent lack of sequence diversity within RbcS families argues against this hypothesis, as it is unlikely that the small differences in peptide sequences observed from current data would lead to a significant functional impact on the catalytic performance of Rubisco in photosynthetic tissues (Yamada et al., 2019). Nevertheless, a broader analysis of RbcS families may reveal further complexity. For example, Sharwood et al. (2016) noted that the C4 Paniceae species Megathyrsus maximus and Panicum monticola share identical RbcL peptide sequences but showed significant differences in several catalytic parameters, suggesting that RbcS may influence Rubisco catalysis. Furthermore, E. coli-based characterizations of predicted ancestral Rubiscos within the Solanaceae family (i.e. tobacco and potato) have also shown that relatively small differences in the RbcS sequence can modify catalysis (Lin et al., 2022). Notably, the increased abundance of mutations found in ancestral RbcS sequences indicated that the recent evolution of C3 Rubiscos may have been driven more by changes in the RbcS than in the RbcL.
RbcS in the T-type cluster produce Rubisco with different catalytic properties
A new cluster of ‘T-type’ RbcS isoforms has recently been discovered that are phylogenetically distinct from RbcS isoforms expressed in leaves (i.e. M-type RbcS) (Laterre et al., 2017). Initially identified in tobacco trichomes (hence ‘T’), T-type RbcS appear to have a similar tertiary structure to M-type RbcS (Fig. 2B), but differ significantly in terms of sequence homology and expression (Table 3). T-type RbcS are present in many different plant species, including pteridophytes and bryophytes, but are absent in several dicots and monocots, indicating that independent losses of T-type RbcS genes have occurred during evolution (Pottier et al., 2018). Current evidence suggests that T-type RbcS facilitates Rubisco expression for non-photosynthetic processes, such as CO2 recycling during secondary metabolic processes (e.g. oil or fatty acid biosynthesis). Notably, Rubisco complexes assembled with T-type RbcS can have significantly different catalytic properties compared with those with M-type RbcS (Table 2). For example, Chlamydomonas Rubisco had increased Vc and Kc values when assembled with the tobacco T-type RbcS compared with an M-type NtRbcS (Laterre et al., 2017). Furthermore, Rubisco carrying the T-type RbcS showed higher carboxylation rates at pH values below 8. Laterre et al. (2017) hypothesized that these changes may be a functional adaptation to allow Rubisco to operate in a more acidic, CO2-rich environment, which might be the case in specialized secretory cells compared with the more alkaline environment of the chloroplast stroma in mesophyll cells. More recently, assembly of the tobacco RbcL with different members of the RbcS family in E. coli has confirmed that the T-type RbcS increases the kcatc and Kc values for tobacco Rubisco (Lin et al., 2020). Similar changes in Rubisco catalytic performance were observed in transgenic rice plants overexpressing the rice T-type RbcS, and in transgenic tobacco plants expressing the potato RbcL and potato T-type RbcS (Morita et al., 2014; Martin-Avila et al., 2020).
In flowering plants, expression of T-type RbcS is limited to non-photosynthetic tissues (Voo et al., 2012; Laterre et al., 2017). However, in pteridophytes and bryophytes, T-type RbcS expression appears less restricted, and T-type isoforms are generally more abundant than M-type RbcS. Pottier et al. (2018) suggested this may be due to species in early land plant lineages having emerged before the increase of atmospheric O2 to modern levels (Lenton et al., 2016), when Rubisco would have been adapted to a higher CO2 to O2 ratio (i.e. a higher kcatc and lower Sc/o). Thus, M-type RbcS could have evolved to improve the specificity of Rubisco as atmospheric O2 levels increased. Further measurements of the catalytic characteristics of Rubiscos with different T-type RbcS may help to uncover the past and ongoing ecological roles of the T-type cluster, and if T-type RbcS isoforms could be used to inform strategies to improve the catalytic properties of Rubiscos in crop plants.
RbcS as an engineering target for enhancing plant photosynthesis
Despite the relatively small number of studies to date focused on overexpressing native RbcS or engineering modified or heterologous RbcS isoforms in plants, the observed wide-ranging impacts on the catalytic performance of Rubisco are encouraging (Table 2). In support of the relative ease of nuclear transformation of RbcS, recent advances in tissue culture and protoplast regeneration highlight the growing feasibility and generally robust efficiencies of nuclear genome engineering in a wide selection of crop and non-crop species (Woo et al., 2015; Lin et al., 2018). Research going forward should continue to focus on establishing the extent of non-native RbcS compatibility in terms of sequence and structure for expression and assembly in plants (and algae of biotechnological interest), and the range of catalytic changes achievable by the RbcS alone.
RbcS-deficient mutants of model plant species, such as tobacco or Arabidopsis (Khumsupan et al., 2020; Martin-Avila et al., 2020), are useful proxies to investigate the capacity of heterologous RbcS to assemble with plant RbcL, affect the catalysis of Rubisco, and impact photosynthesis and growth. In planta screens could also be important to confirm that an appropriate transit peptide is being used for efficient chloroplast targeting of heterologous RbcS and that appropriate promoter and terminator combinations are being employed to drive sufficient expression. This could be achieved relatively quickly through transient expression in Nicotiana benthamiana or by protoplast transformation of the target plant (Xu et al., 2022), or an amenable close relative. More high-throughput screening approaches will probably rely on E. coli, which is rapidly developing as a powerful platform for expressing all Form I Rubiscos, screening RbcS and RbcL from different species, and directed evolution (Duraõ et al., 2015; Aigner et al., 2017; Wilson and Whitney, 2017; Lin et al., 2020, 2022). Nevertheless, producing Rubiscos from plants in E. coli still remains challenging due to inefficient RbcL processing and the accumulation of assembly intermediates (Ng et al., 2020), low functional Rubisco yields, and limited current knowledge of the chaperones required for Rubiscos from different plant species. Rubisco biogenesis in E. coli is also currently not a reliable predictor for assembly in chloroplasts (Wilson et al., 2018), while the catalytic parameters of plant Rubiscos assembled in E. coli are similar but not identical to those observed in planta (Lin et al., 2020, 2022). However, significant improvements have already been made (Zhou and Whitney, 2019), and the increasing availability of genomic data could assist with identifying homologues or additional ancillary chaperones required for different plant species (Kress et al., 2022). Further progress in the state of the art of in silico Rubisco models should also help to complement RbcS screening approaches (Van Lun et al., 2011, 2014).
A key challenge to successfully exploiting heterologous RbcS in planta appears to be achieving sufficiently high levels of expression in transformed plants, as production of heterologous RbcS using single expression cassettes driven by high strength promoters has not yet yielded sufficient quantities to achieve expression commensurate with native RbcS levels (Atkinson et al., 2017; Matsumura et al., 2020). Generating a synthetic RbcS family using a multigene expression cassette approach may help to solve this issue (Marillonnet and Grützner, 2020). Furthermore, removal or suppression of the native RbcS gene family may be required. Both CRISPR/Cas and RNAi strategies have been successfully used to remove or significantly suppress native RbcS expression levels in Arabidopsis, rice, and tobacco (Donovan et al., 2020; Khumsupan et al., 2020; Martin-Avila et al., 2020; Matsumura et al., 2020).
If improvements in Rubisco performance are achieved, other factors could become limiting to photosynthetic efficiency. For example, leaf CO2 diffusion is enhanced in Limonium species with faster Rubiscos (i.e. a higher kcatc) (Galmés et al., 2017). Increases in CO2 diffusion could potentially be accomplished through the overexpression of CO2-permeable membrane channels, such as specific aquaporins (Kaldenhoff, 2012; Kromdijk et al., 2020; Clarke et al., 2022), or CO2/HCO3− transporters found in algal and cyanobacterial CCMs (Rae et al., 2017; Rottet et al., 2021). Overexpression of chaperones could also be required, for example RAF1 or Rubisco activase in maize and rice, respectively (Salesse-Smith et al., 2018; Suganami et al., 2021). Furthermore, any increases in Rubisco CO2 assimilation may benefit from additional modifications in downstream fluxes, for example the transport of carbohydrates to sink tissues (Wang et al., 2020). Modified canopy-scale models that account for Rubisco parameters may help further to predict the physiological consequences of replacing native Rubiscos with heterologous enzymes in plants (Iqbal et al., 2021).
Despite these potential challenges, it is feasible that modifications to the RbcS will yield the first generation of crop plants with improved Rubiscos. RbcS changes may ultimately require concurrent modifications in the RbcL to maximize catalytic improvements or compatibility between the subunits, and this will probably become achievable as chloroplast engineering technologies improve (Yu et al., 2020). Nevertheless, it is fascinating that over the past two decades the RbcS has ascended from relative obscurity to becoming a potential game changer to enhance the photosynthetic performance of plants.
Contributor Information
Yuwei Mao, SynthSys & Institute of Molecular Plant Sciences, School of Biological Sciences, King’s Buildings, University of Edinburgh, Edingburgh EH9 3BF, UK.
Ella Catherall, SynthSys & Institute of Molecular Plant Sciences, School of Biological Sciences, King’s Buildings, University of Edinburgh, Edingburgh EH9 3BF, UK.
Aranzazú Díaz-Ramos, SynthSys & Institute of Molecular Plant Sciences, School of Biological Sciences, King’s Buildings, University of Edinburgh, Edingburgh EH9 3BF, UK.
George R L Greiff, School of Biological Sciences, University of Bristol, 24 Tyndall Avenue, Bristol BS8 1TQ, UK.
Stavros Azinas, Department of Cell and Molecular Biology, Uppsala University, S-751 24 Uppsala, Sweden.
Laura Gunn, Department of Cell and Molecular Biology, Uppsala University, S-751 24 Uppsala, Sweden; Plant Biology Section, School of Integrative Plant Science, Cornell University, Ithaca, NY, USA.
Alistair J McCormick, SynthSys & Institute of Molecular Plant Sciences, School of Biological Sciences, King’s Buildings, University of Edinburgh, Edingburgh EH9 3BF, UK.
Robert Sharwood, Western Sydney University, Australia.
Conflict of interest
The authors have no conflicts to declare.
Funding
AJM acknowledges funding from the UKRI Biotechnology and Biological Sciences Research Council (grant no. BB/S015531/1) and Leverhulme Trust (grant no. RPG-2017-402). EC was funded by EASTBIO-BBSRC doctoral training programme. GRLG was funded by the SWBio-BBSRC doctoral training programme.
References
- Aigner H, Wilson RH, Bracher A, Calisse L, Bhat JY, Hartl FU, Hayer-Hartl M.. 2017. Plant RuBisCo assembly in E. coli with five chloroplast chaperones including BSD2. Science 358, 1272–1278. [DOI] [PubMed] [Google Scholar]
- Andersson I, Backlund A.. 2008. Structure and function of Rubisco. Plant Physiology and Biochemistry 46, 275–291. [DOI] [PubMed] [Google Scholar]
- Andrews TJ. 1988. Catalysis by cyanobacterial ribulose-bisphosphate carboxylase large subunits in the complete absence of small subunits. Journal of Biological Chemistry 263, 12213–12219. [PubMed] [Google Scholar]
- Andrews TJ, Ballment B.. 1983. The function of the small subunits of ribulose bisphosphate carboxylase-oxygenase. Journal of Biological Chemistry 258, 7514–7518. [PubMed] [Google Scholar]
- Ashida H, Saito Y, Kojima C, Kobayashi K, Ogasawara N, Yokota A.. 2003. A functional link between RuBisCO-like protein of Bacillus and photosynthetic RuBisCO. Science 302, 286–290. [DOI] [PubMed] [Google Scholar]
- Atkinson N, Leitão N, Orr DJ, Meyer MT, Carmo-Silva E, Griffiths H, Smith AM, McCormick AJ.. 2017. Rubisco small subunits from the unicellular green alga Chlamydomonas complement Rubisco-deficient mutants of Arabidopsis. New Phytologist 214, 655–667. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Atkinson N, Mao Y, Chan KX, McCormick AJ.. 2020. Condensation of Rubisco into a proto-pyrenoid in higher plant chloroplasts. Nature Communications 11, 6303. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Atkinson N, Velanis CN, Wunder T, Clarke DJ, Mueller-Cajar O, McCormick AJ.. 2019. The pyrenoidal linker protein EPYC1 phase separates with hybrid Arabidopsis–Chlamydomonas Rubisco through interactions with the algal Rubisco small subunit. Journal of Experimental Botany 70, 5271–5285. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Badger MR, Bek EJ.. 2008. Multiple Rubisco forms in proteobacteria: their functional significance in relation to CO2 acquisition by the CBB cycle. Journal of Experimental Botany 59, 1525–1541. [DOI] [PubMed] [Google Scholar]
- Badger MR, Hanson D, Price GD.. 2002. Evolution and diversity of CO2-concentrating mechanisms in cyanobacteria. Functional Plant Biology 29, 161–173. [DOI] [PubMed] [Google Scholar]
- Banda DM, Pereira JH, Liu AK, et al. 2020. Novel bacterial clade reveals origin of form I Rubisco. Nature Plants 6, 1158–1166. [DOI] [PubMed] [Google Scholar]
- Bar-On YM, Milo R.. 2019. The global mass and average rate of Rubisco. Proceedings of the National Academy of Sciences, USA 116, 4738–4743. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Barrett J, Girr P, Mackinder LCM.. 2021. Pyrenoids: CO2-fixing phase separated liquid organelles. Biochimica et Biophysica Acta 1868, 118949. [DOI] [PubMed] [Google Scholar]
- Borden JS, Savage DF.. 2021. New discoveries expand possibilities for carboxysome engineering. Current Opinion in Microbiology 61, 58–66. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bouvier JW, Emms DM, Rhodes T, Bolton JS, Brasnett A, Eddershaw A, Nielsen JR, Unitt A, Whitney SM, Kelly S.. 2021. Rubisco adaptation is more limited by phylogenetic constraint than by catalytic trade-off. Molecular Biology and Evolution 38, 2880–2896. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bracher A, Whitney SM, Hartl FU, Hayer-Hartl M.. 2017. Biogenesis and metabolic maintenance of Rubisco. Annual Review of Plant Biology 68, 29–60. [DOI] [PubMed] [Google Scholar]
- Busch FA. 2020. Photorespiration in the context of Rubisco biochemistry, CO2 diffusion and metabolism. The Plant Journal 101, 919–939. [DOI] [PubMed] [Google Scholar]
- Busch FA, Sage RF, Farquhar GD.. 2018. Plants increase CO2 uptake by assimilating nitrogen via the photorespiratory pathway. Nature Plants 4, 46–54. [DOI] [PubMed] [Google Scholar]
- Buti M, Pasquariello M, Ronga D, Milc JA, Pecchioni N, Ho VT, Pucciariello C, Perata P, Francia E.. 2018. Transcriptome profiling of short-term response to chilling stress in tolerant and sensitive Oryza sativa ssp. Japonica seedlings. Functional and Integrative Genomics 18, 627–644. [DOI] [PubMed] [Google Scholar]
- Carmo-Silva E, Scales JC, Madgwick PJ, Parry MAJ.. 2015. Optimizing Rubisco and its regulation for greater resource use efficiency. Plant, Cell & Environment 38, 1817–1832. [DOI] [PubMed] [Google Scholar]
- Caruana L, Orr DJ, Carmo-Silva E.. 2022. Rubiscosome gene expression is balanced across the hexaploid wheat genome. Photosynthesis Research 152, 1–11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cheng SH, Moore BD, Seemann JR.. 1998. Effects of short- and long-term elevated CO2 on the expression of ribulose-1,5-bisphosphate carboxylase/oxygenase genes and carbohydrate accumulation in leaves of Arabidopsis thaliana (L.) Heynh. Plant Physiology 116, 715–723. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Clarke VC, De Rosa A, Massey B, George AM, Evans JR, von Caemmerer S, Groszmann M.. 2022. Mesophyll conductance is unaffected by expression of Arabidopsis PIP1 aquaporins in the plasmalemma of Nicotiana. Journal of Experimental Botany 73, 3625–3636. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Davidson RM, Gowda M, Moghe G, Lin H, Vaillancourt B, Shiu SH, Jiang N, Buell CR.. 2012. Comparative transcriptomics of three Poaceae species reveals patterns of gene expression evolution. The Plant Journal 71, 492–502. [DOI] [PubMed] [Google Scholar]
- Dean C, Van den Elzen P, Tamaki S, Dunsmuir P, Bedbrook J.. 1985. Linkage and homology analysis divides the eight genes for the small subunit of petunia ribulose 1,5-bisphosphate carboxylase into three gene families. Proceedings of the National Academy of Sciences, USA 82, 4964–4968. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dedonder A, Rethy R, Fredericq H, Van Montagu M, Krebbers E.. 1993. Arabidopsis rbcS genes are differentially regulated by light. Plant Physiology 101, 801–808. [DOI] [PMC free article] [PubMed] [Google Scholar]
- DeRocher EJ, Quigley F, Mache R, Bohnert HJ.. 1993. The six genes of the Rubisco small subunit multigene family from Mesembryanthemum crystallinum, a facultative CAM plant. Molecular and General Genetics 239, 450–462. [DOI] [PubMed] [Google Scholar]
- Dhingra A, Portis AR Jr, Daniell H.. 2004. Enhanced translation of a chloroplast-expressed RbcS gene restores small subunit levels and photosynthesis in nuclear RbcS antisense plants. Proceedings of the National Academy of Sciences, USA 101, 6315–6320. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Donovan S, Mao Y, Orr DJ, Carmo-Silva E, McCormick AJ.. 2020. CRISPR–Cas9-mediated mutagenesis of the Rubisco small subunit family in Nicotiana tabacum. Frontiers in Genome Editing 2, 605614. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Duraõ P, Aigner H, Nagy P, Mueller-Cajar O, Hartl FU, Hayer-Hartl M.. 2015. Opposing effects of folding and assembly chaperones on evolvability of Rubisco. Nature Chemical Biology 11, 148–155. [DOI] [PubMed] [Google Scholar]
- Ellis RJ. 1979. The most abundant protein in the world. Trends in Biochemical Sciences 4, 241–244. [Google Scholar]
- Erb TJ, Evans BS, Cho K, Warlick BP, Sriram J, Wood BM, Imker HJ, Sweedler JV, Tabita FR, Gerlt JA.. 2012. A RubisCO-like protein links SAM metabolism with isoprenoid biosynthesis. Nature Chemical Biology 8, 926–932. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Erb TJ, Zarzycki J.. 2018. A short history of RubisCO: the rise and fall (?) of Nature’s predominant CO2 fixing enzyme. Current Opinion in Biotechnology 49, 100–107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Esquivel MG, Genkov T, Nogueira AS, Salvucci ME, Spreitzer RJ.. 2013. Substitutions at the opening of the Rubisco central solvent channel affect holoenzyme stability and CO2/O2 specificity but not activation by Rubisco activase. Photosynthesis Research 118, 209–218. [DOI] [PubMed] [Google Scholar]
- Ewing RM, Jenkins GI, Langdale JA.. 1998. Transcripts of maize RbcS genes accumulate differentially in C3 and C4 tissues. Plant Molecular Biology 36, 593–599. [DOI] [PubMed] [Google Scholar]
- Fang W, Si Y, Douglass S, Casero D, Merchant SS, Pellegrini M, Ladunga I, Liu P, Spalding MH.. 2012. Transcriptome-wide changes in Chlamydomonas reinhardtii gene expression regulated by carbon dioxide and the CO2-concentrating mechanism regulator CIA5/CCM1. The Plant Cell 24, 1876–1893. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Feiz L, Williams-Carrier R, Belcher S, Montano M, Barkan A, Stern DB.. 2014. A protein with an inactive pterin-4a-carbinolamine dehydratase domain is required for Rubisco biogenesis in plants. The Plant Journal 80, 862–869. [DOI] [PubMed] [Google Scholar]
- Fernie AR, Bauwe H.. 2020. Wasteful, essential, evolutionary stepping stone? The multiple personalities of the photorespiratory pathway. The Plant Journal 102, 666–677. [DOI] [PubMed] [Google Scholar]
- Flamholz AI, Prywes N, Moran U, Davidi D, Bar-On YM, Oltrogge LM, Alves R, Savage D, Milo R.. 2019. Revisiting trade-offs between Rubisco kinetic parameters. Biochemistry 58, 3365–3376. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Flamholz A, Shih PM.. 2020. Cell biology of photosynthesis over geologic time. Current Biology 30, R490R451–R490R494. [DOI] [PubMed] [Google Scholar]
- Flügel F, Timm S, Arrivault S, Florian A, Stitt M, Fernie AR, Bauwe H.. 2017. The photorespiratory metabolite 2-phosphoglycolate regulates photosynthesis and starch accumulation in Arabidopsis. The Plant Cell 29, 2537–2551. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Freeling M. 2009. Bias in plant gene content following different sorts of duplication: tandem, whole-genome, segmental, or by transposition. Annual Review of Plant Biology 60, 433–453. [DOI] [PubMed] [Google Scholar]
- Fristedt R, Hu C, Wheatley N, et al. 2018. RAF2 is a RuBisCO assembly factor in Arabidopsis thaliana. The Plant Journal 94, 146–156. [DOI] [PubMed] [Google Scholar]
- Frolov EN, Kublanov IV, Toshchakov SV, Lunev EA, Pimenov NV, Bonch-Osmolovskaya EA, Lebedinsky AV, Chernyh NA.. 2019. Form III RubisCO-mediated transaldolase variant of the Calvin cycle in a chemolithoautotrophic bacterium. Proceedings of the National Academy of Sciences, USA 116, 18638–18646. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Galmés J, Molins A, Flexas J, Conesa MA.. 2017. Coordination between leaf CO2 diffusion and Rubisco properties allows maximizing photosynthetic efficiency in Limonium species. Plant, Cell & Environment 40, 2081–2094. [DOI] [PubMed] [Google Scholar]
- Genkov T, Meyer M, Griffiths H, Spreitzer RJ.. 2010. Functional hybrid Rubisco enzymes with plant small subunits and algal large subunits. Journal of Biological Chemistry 285, 19833–19841. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Genkov T, Spreitzer RJ.. 2009. Highly conserved small subunit residues influence rubisco large subunit catalysis. Journal of Biological Chemistry 284, 30105–30112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Getzoff TP, Zhu G, Bohnert HJ, Jensen RG.. 1998. Chimeric Arabidopsis thaliana ribulose-1,5-bisphosphate carboxylase/oxygenase containing a pea small subunit protein is compromised in carbamylation. Plant Physiology 116, 695–702. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Goldschmidt-Clermont M, Rahire M.. 1986. Sequence, evolution and differential expression of the two genes encoding variant small subunits of ribulose bisphosphate carboxylase/oxygenase in Chlamydomonas reinhardtii. Journal of Molecular Biology 191, 421–432. [DOI] [PubMed] [Google Scholar]
- Goudet MMM, Orr DJ, Melkonian M, Müller KH, Meyer MT, Carmo-Silva E, Griffiths H.. 2020. Rubisco and carbon-concentrating mechanism co-evolution across chlorophyte and streptophyte green algae. New Phytologist 227, 810–823. [DOI] [PubMed] [Google Scholar]
- Gunn LH, Valegӓrd K, Andersson I.. 2017. A unique structural domain in Methanococcoides burtonii ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco) acts as a small subunit mimic. Journal of Biological Chemistry 292, 6838–6850. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gutteridge S. 1991. The relative catalytic specificities of the large subunit core of Synechococcus ribulose bisphosphate carboxylase/oxygenase. Journal of Biological Chemistry 266, 7359–7362. [PubMed] [Google Scholar]
- Hanson TE, Tabita FR.. 2001. A ribulose-1,5-bisphosphate carboxylase/oxygenase (RubisCO)-like protein from Chlorobium tepidum that is involved with sulfur metabolism and the response to oxidative stress. Proceedings of the National Academy of Sciences, USA 98, 4397–4402. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hayer-Hartl M, Hartl FU.. 2020. Chaperone machineries of Rubisco—the most abundant enzyme. Trends in Biochemical Sciences 45, 748–763. [DOI] [PubMed] [Google Scholar]
- He S, Chou HT, Matthies D, et al. 2020. The structural basis of Rubisco phase separation in the pyrenoid. Nature Plants 6, 1480–1490. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hennacy JH, Jonikas MC.. 2020. Prospects for engineering biophysical CO2-concentrating mechanisms into land plants to enhance yields. Annual Review of Plant Biology 71, 461–485. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Iñiguez C, Capó-Bauçà S, Niinemets U, Stoll H, Aguiló-Nicolau P, Galmés J.. 2020. Evolutionary trends in RuBisCO kinetics and their co-evolution with CO2 concentrating mechanisms. The Plant Journal 101, 897–918. [DOI] [PubMed] [Google Scholar]
- Iqbal WA, Miller IG, Moore RL, Hope IJ, Cowan-Turner D, Kapralov MV.. 2021. Rubisco substitutions predicted to enhance crop performance through carbon uptake modelling. Journal of Experimental Botany 72, 6066–6075. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ishikawa C, Hatanaka T, Misoo S, Miyake C, Fukayama H.. 2011. Functional incorporation of sorghum small subunit increases the catalytic turnover rate of Rubisco in transgenic rice. Plant Physiology 156, 1603–1611. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Izumi M, Tsunoda H, Suzuki Y, Makino A, Ishida H.. 2012. RBCS1A and RBCS3B, two major members within the Arabidopsis RBCS multigene family, function to yield sufficient Rubisco content for leaf photosynthetic capacity. Journal of Experimental Botany 63, 2159–2170. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Joshi J, Mueller-Cajar O, Tsai YCC, Hartl FU, Hayer-Hartl M.. 2015. Role of small subunit in mediating assembly of red-type Form I Rubisco. Journal of Biological Chemistry 290, 1066–1074. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kacar B, Hanson-Smith V, Adam ZR, Boekelheide N.. 2017. Constraining the timing of the Great Oxidation Event within the Rubisco phylogenetic tree. Geobiology 15, 628–640. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kaldenhoff R. 2012. Mechanisms underlying CO2 diffusion in leaves. Current Opinion in Plant Biology 15, 276–281. [DOI] [PubMed] [Google Scholar]
- Karkehabadi S, Peddi SR, Anwaruzzaman M, Taylor TC, Cederlund A, Genkov T, Andersson I, Spreitzer RJ.. 2005. Chimeric small subunits influence catalysis without causing global conformational changes in the crystal structure of ribulose-1,5-bisphosphate carboxylase/oxygenase. Biochemistry 44, 9851–9861. [DOI] [PubMed] [Google Scholar]
- Khrebtukova I, Spreitzer RJ.. 1996. Elimination of the Chlamydomonas gene family that encodes the small subunit of ribulose-1,5-bisphosphate carboxylase/oxygenase. Proceedings of the National Academy of Sciences, USA 93, 13689–13693. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Khumsupan P, Kozlowska MA, Orr DJ, Andreou AI, Nakayama N, Patron N, Carmo-Silva E, McCormick AJ.. 2020. Generating and characterizing single- and multigene mutants of the Rubisco small subunit family in Arabidopsis. Journal of Experimental Botany 71, 5963–5975. [DOI] [PubMed] [Google Scholar]
- Kitano K, Maeda N, Fukui T, Atomi H, Imanaka T, Miki K.. 2001. Crystal structure of a novel-type archaeal Rubisco with pentagonal symmetry. Structure 9, 473–481. [DOI] [PubMed] [Google Scholar]
- Knight S, Andersson I, Brändén CI.. 1990. Crystallographic analysis of ribulose 1,5-bisphosphate carboxylase from spinach at 2.4 Å resolution: subunit interactions and active site. Journal of Molecular Biology 215, 113–160. [DOI] [PubMed] [Google Scholar]
- Kono T, Mehrotra S, Endo C, et al. 2017. A RuBisCO-mediated carbon metabolic pathway in methanogenic archaea. Nature Communications 8, 14007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kostrzewa M, Valentin K, Maid U, Radetzky R, Zetsche K.. 1990. Structure of the rubisco operon from the multicellular red alga Antithamnion spec. Current Genetics 18, 465–469. [DOI] [PubMed] [Google Scholar]
- Krebbers E, Seurinck J, Herdies L, Cashmore AR, Timko MP.. 1988. Four genes in two diverged subfamilies encode the ribulose-1,5-bisphosphate carboxylase small subunit polypeptides of Arabidopsis thaliana. Plant Molecular Biology 11, 745–759. [DOI] [PubMed] [Google Scholar]
- Kress WJ, Soltis DE, Kersey PJ, Wegrzyn JL, Leebens-Mack JH, Gostel MR, Liu X, Soltis PS.. 2022. Green plant genomes: what we know in an era of rapidly expanding opportunities. Proceedings of the National Academy of Sciences, USA 119, e2115640118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kromdijk J, Głowacka K, Long SP.. 2020. Photosynthetic efficiency and mesophyll conductance are unaffected in Arabidopsis thaliana aquaporin knock-out lines. Journal of Experimental Botany 71, 318–329. [DOI] [PubMed] [Google Scholar]
- Laterre R, Pottier M, Remacle C, Boutry M.. 2017. Photosynthetic trichomes contain a specific Rubisco with a modified pH-dependent activity. Plant Physiology 173, 2110–2120. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee B, Tabita FR.. 1990. Purification of recombinant ribulose-l,5-bisphosphate carboxylase/oxygenase large subunits suitable for reconstitution and assembly of active L8S8 enzyme. Biochemistry 29, 9352–9357. [DOI] [PubMed] [Google Scholar]
- Lenton TM, Dahl TW, Daines SJ, Mills BJW, Ozaki K, Saltzman MR, Porada P.. 2016. Earliest land plants created modern levels of atmospheric oxygen. Proceedings of the National Academy of Sciences, USA 113, 9704–9709. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lin CS, Hsu CT, Yang LH, et al. 2018. Application of protoplast technology to CRISPR/Cas9 mutagenesis: from single-cell mutation detection to mutant plant regeneration. Plant Biotechnology Journal 16, 1295–1310. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lin MT, Salihovic H, Clark FK, Hanson MR.. 2022. Improving the efficiency of Rubisco by resurrecting its ancestors in the family Solanaceae. Science Advances 8, eabm6871. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lin MT, Stone WD, Chaudhari V, Hanson MR.. 2020. Small subunits can determine enzyme kinetics of tobacco Rubisco expressed in Escherichia coli. Nature Plants 6, 1289–1299. [DOI] [PubMed] [Google Scholar]
- Liu D, Ramya RCS, Mueller-Cajar O.. 2017. Surveying the expanding prokaryotic Rubisco multiverse. FEMS Microbiology Letters 364, fnx156. [DOI] [PubMed] [Google Scholar]
- Loewen PC, Didychuk AL, Switala J, Perez-Luque R, Fita I, Loewen MC.. 2013. Structure of Pisum sativum Rubisco with bound ribulose 1,5-bisphosphate. Acta Crystallographica Section F: Structural Biology and Crystallization Communications 69, 10–14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Long SP, Burgess S, Causton I.. 2019. Redesigning crop photosynthesis. In: Zeigler R, ed. Sustaining global food security: the nexus of science and policy. Clayton, Australia: CSIRO Publishing, 128–147. [Google Scholar]
- Long BM, Hee WY, Sharwood RE, et al. 2018. Carboxysome encapsulation of the CO2-fixing enzyme Rubisco in tobacco chloroplasts. Nature Communications 9, 3570. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mackinder LCM, Meyer MT, Mettler-Altmann T, et al. 2016. A repeat protein links Rubisco to form the eukaryotic carbon-concentrating organelle. Proceedings of the National Academy of Sciences, USA 113, 5958–5963. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Maheshwari C, Coe RA, Karki S, Covshoff S, Tapia R, Tyagi A, Hibberd JM, Furbank RT, Quick WP, Lin HC.. 2021. Targeted knockdown of ribulose-1,5-bisphosphate carboxylase-oxygenase in rice mesophyll cells. Journal of Plant Physiology 260, 153395. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Marillonnet S, Grützner R.. 2020. Synthetic DNA assembly using Golden Gate cloning and the hierarchical modular cloning pipeline. Current Protocols in Molecular Biology 130, e115. [DOI] [PubMed] [Google Scholar]
- Martin-Avila E, Lim YL, Birch R, Dirk LMA, Buck S, Rhodes T, Sharwood RE, Kapralov MV, Whitney SM.. 2020. Modifying plant photosynthesis and growth via simultaneous chloroplast transformation of Rubisco large and small subunits. The Plant Cell 32, 2898–2916. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Matsumura H, Shiomi K, Yamamoto A, Taketani Y, Kobayashi N, Yoshizawa T, Tanaka SI, Yoshikawa H, Endo M, Fukayama H.. 2020. Hybrid Rubisco with complete replacement of rice Rubisco small subunits by sorghum counterparts confers C4 plant-like high catalytic activity. Molecular Plant 13, 1570–1581. [DOI] [PubMed] [Google Scholar]
- Mergner J, Frejno M, List M, et al. 2020. Mass-spectrometry-based draft of the Arabidopsis proteome. Nature 579, 409–414. [DOI] [PubMed] [Google Scholar]
- Meyer MT, Genkov T, Skepper JN, Jouhet J, Mitchell MC, Spreitzer RJ, Griffiths H.. 2012. Rubisco small-subunit α-helices control pyrenoid formation in Chlamydomonas. Proceedings of the National Academy of Sciences, USA 109, 19474–19479. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Meyer MT, Itakura AK, Patena W, Wang L, He S, Emrich-Mills T, Lau CS, Yates G, Mackinder LCM, Jonikas MC.. 2020. Assembly of the algal CO2-fixing organelle, the pyrenoid, is guided by a Rubisco-binding motif. Science Advances 6, eabd2408. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Morell MK, Wilkin JM, Kane HJ, Andrews TJ.. 1997. Side reactions catalyzed by ribulose-bisphosphate carboxylase in the presence and absence of small subunits. Journal of Biological Chemistry 272, 5445–5451. [DOI] [PubMed] [Google Scholar]
- Morita K, Hatanaka T, Misoo S, Fukayama H.. 2014. Unusual small subunit that is not expressed in photosynthetic cells alters the catalytic properties of Rubisco in rice. Plant Physiology 164, 69–79. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Narsai R, Howell KA, Carroll A, Ivanova A, Millar AH, Whelan J.. 2009. Defining core metabolic and transcriptomic responses to oxygen availability in rice embryos and young seedlings. Plant Physiology 151, 306–322. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ng J, Guo Z, Mueller-Cajar O.. 2020. Rubisco activase requires residues in the large subunit N terminus to remodel inhibited plant Rubisco. Journal of Biological Chemistry 295, 16427–16435. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nisbet EG, Grassineau NV, Howe CJ, Abell PI, Regelous M, Nisbet RER.. 2007. The age of Rubisco: the evolution of oxygenic photosynthesis. Geobiology 5, 311–335. [Google Scholar]
- Oltrogge LM, Chaijarasphong T, Chen AW, Bolin ER, Marqusee S, Savage DF.. 2020. Multivalent interactions between CsoS2 and Rubisco mediate α-carboxysome formation. Nature Structural & Molecular Biology 27, 281–287. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Orr DJ, Alcântara A, Kapralov MV, Andralojc PJ, Carmo-Silva E, Parry MAJ.. 2016. Surveying rubisco diversity and temperature response to improve crop photosynthetic efficiency. Plant Physiology 172, 707–717. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Orr DJ, Parry MAJ.. 2020. Overcoming the limitations of Rubisco: fantasy or realistic prospect? Journal of Plant Physiology 254, 153285. [DOI] [PubMed] [Google Scholar]
- Orr DJ, Worrall D, Lin MT, Carmo-Silva E, Hanson MR, Parry MAJ.. 2020. Hybrid cyanobacterial–tobacco Rubisco supports autotrophic growth and procarboxysomal aggregation. Plant Physiology 182, 807–818. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Parry MAJ, Andralojc PJ, Scales JC, Salvucci ME, Carmo-Silva AE, Alonso H, Whitney SM.. 2013. Rubisco activity and regulation as targets for crop improvement. Journal of Experimental Botany 64, 717–730. [DOI] [PubMed] [Google Scholar]
- Parry MAJ, Keys AJ, Gutteridge S.. 1989. Variation in the specificity factor of C3 higher plant Rubiscos determined by the total consumption of ribulose-P2. Journal of Experimental Botany 40, 317–320. [Google Scholar]
- Polans NO, Weeden NF, Thompson WF.. 1985. Inheritance, organization, and mapping of rbcS and cab multigene families in pea. Proceedings of the National Academy of Sciences, USA 82, 5083–5087. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pottier M, Gilis D, Boutry M.. 2018. The hidden face of Rubisco. Trends in Plant Science 23, 382–392. [DOI] [PubMed] [Google Scholar]
- Price DC, Goodenough UW, Roth R, et al. 2019. Analysis of an improved Cyanophora paradoxa genome assembly. DNA Research 26, 287–299. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rae BD, Long BM, Förster B, Nguyen ND, Velanis CN, Atkinson N, Hee WY, Mukherjee B, Price GD, McCormick AJ.. 2017. Progress and challenges of engineering a biophysical CO2-concentrating mechanism into higher plants. Journal of Experimental Botany 68, 3717–3737. [DOI] [PubMed] [Google Scholar]
- Raven JA. 2009. Contributions of anoxygenic and oxygenic phototrophy and chemolithotrophy to carbon and oxygen fluxes in aquatic environments. Aquatic Microbial Ecology 56, 177–192. [Google Scholar]
- Rodermel S. 1999. Subunit control of Rubisco biosynthesis—a relic of an endosymbiotic past? Photosynthesis Research 59, 105–123. [Google Scholar]
- Rodermel S, Haley J, Jiang CZ, Tsai CH, Bogorad L.. 1996. A mechanism for intergenomic integration: abundance of ribulose bisphosphate carboxylase small-subunit protein influences the translation of the large-subunit mRNA. Proceedings of the National Academy of Sciences, USA 93, 3881–3885. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rottet S, Förster B, Hee WY, Rourke LM, Price GD, Long BM.. 2021. Engineered accumulation of bicarbonate in plant chloroplasts: known knowns and known unknowns. Frontiers in Plant Science 12, 727118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ruf S, Forner J, Hasse C, Kroop X, Seeger S, Schollbach L, Schadach A, Bock R.. 2019. High-efficiency generation of fertile transplastomic Arabidopsis plants. Nature Plants 5, 282–289. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ryan P, Forrester TJB, Wroblewski C, Kenney TMG, Kitova EN, Klassen JS, Kimber MS.. 2019. The small RbcS-like domains of the β-carboxysome structural protein CcmM bind RubisCO at a site distinct from that binding the RbcS subunit. Joursnal of Biological Chemistry 294, 2593–2603. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Salesse-Smith CE, Sharwood RE, Busch FA, Kromdijk J, Bardal V, Stern DB.. 2018. Overexpression of Rubisco subunits with RAF1 increases Rubisco content in maize. Nature Plants 4, 802–810. [DOI] [PubMed] [Google Scholar]
- Savir Y, Noor E, Milo R, Tlusty T.. 2010. Cross-species analysis traces adaptation of Rubisco toward optimality in a low-dimensional landscape. Proceedings of the National Academy of Sciences, USA 107, 3475–3480. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sawchuk MG, Donner TJ, Head P, Scarpella E.. 2008. Unique and overlapping expression patterns among members of photosynthesis-associated nuclear gene families in Arabidopsis. Plant Physiology 148, 1908–1924. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sharwood RE. 2017. Engineering chloroplasts to improve Rubisco catalysis: prospects for translating improvements into food and fiber crops. New Phytologist 213, 494–510. [DOI] [PubMed] [Google Scholar]
- Sharwood RE, Ghannoum O, Kapralov MV, Gunn LH, Whitney SM.. 2016. Temperature responses of Rubisco from Paniceae grasses provide opportunities for improving C3 photosynthesis. Nature Plants 2, 16186. [DOI] [PubMed] [Google Scholar]
- Shi X, Bloom A.. 2021. Photorespiration: the futile cycle? Plants 10, 908. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shibata N, Inoue T, Fukuhara K, et al. 1996. Orderly disposition of heterogeneous small subunits in d-ribulose-1,5-bisphosphate carboxylase/oxygenase from spinach. Journal of Biological Chemistry 271, 26449–26452. [DOI] [PubMed] [Google Scholar]
- Shih PM, Occhialini A, Cameron JC, Andralojc PJ, Parry MAJ, Kerfeld CA.. 2016. Biochemical characterization of predicted Precambrian RuBisCO. Nature Communications 7, 10382. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sierro N, Battey JND, Ouadi S, Bakaher N, Bovet L, Willig A, Goepfert S, Peitsch MC, Ivanov NV.. 2014. The tobacco genome sequence and its comparison with those of tomato and potato. Nature Communications 5, 3833. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Spreitzer RJ. 2003. Role of the small subunit in ribulose-1,5-bisphosphate carboxylase/oxygenase. Archives of Biochemistry and Biophysics 414, 141–149. [DOI] [PubMed] [Google Scholar]
- Spreitzer RJ, Esquivel MG, Du YC, McLaughlin PD.. 2001. Alanine-scanning mutagenesis of the small-subunit βA–βB loop of chloroplast ribulose-1,5-bisphosphate carboxylase/oxygenase: substitution at Arg-71 affects thermal stability and CO2/O2 specificity. Biochemistry 40, 5615–5621. [DOI] [PubMed] [Google Scholar]
- Spreitzer RJ, Peddi SR, Satagopan S.. 2005. Phylogenetic engineering at an interface between large and small subunits imparts land-plant kinetic properties to algal Rubisco. Proceedings of the National Academy of Sciences, USA 102, 17225–17230. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Suganami M, Suzuki Y, Tazoe Y, Yamori W, Makino A.. 2021. Co-overproducing Rubisco and Rubisco activase enhances photosynthesis in the optimal temperature range in rice. Plant Physiology 185, 108–119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sugita M, Manzara T, Pichersky E, Cashmore A, Gruissem W.. 1987. Genomic organization, sequence analysis and expression of all five genes encoding the small subunit of ribulose-1,5-bisphosphate carboxylase/oxygenase from tomato. Molecular & General Genetics 209, 247–256. [DOI] [PubMed] [Google Scholar]
- Suzuki Y, Ohkubo M, Hatakeyama H, Ohashi K, Yoshizawa R, Kojima S, Hayakawa T, Yamaya T, Mae T, Makino A.. 2007. Increased Rubisco content in transgenic rice transformed with the ‘sense’ rbcS gene. Plant & Cell Physiology 48, 626–637. [DOI] [PubMed] [Google Scholar]
- Tabita FR, Hanson TE, Li H, Satagopan S, Singh J, Chan S.. 2007. Function, structure, and evolution of the RubisCO-like proteins and their RubisCO homologs. Microbiology and Molecular Biology Reviews 71, 576–599. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tabita FR, Satagopan S, Hanson TE, Kreel NE, Scott SS.. 2008. Distinct form I, II, III, and IV Rubisco proteins from the three kingdoms of life provide clues about Rubisco evolution and structure/function relationships. Journal of Experimental Botany 59, 1515–1524. [DOI] [PubMed] [Google Scholar]
- Taylor SH, Gonzalez-Escobar E, Page R, Parry MAJ, Long SP, Carmo-Silva E.. 2022. Faster than expected Rubisco deactivation in shade reduces cowpea photosynthetic potential in variable light conditions. Nature Plants 8, 118–124. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tcherkez GGB, Farquhar GD, Andrews TJ.. 2006. Despite slow catalysis and confused substrate specificity, all ribulose bisphosphate carboxylases may be nearly perfectly optimized. Proceedings of the National Academy of Sciences, USA 103, 7246–7251. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Timm S. 2020. The impact of photorespiration on plant primary metabolism through metabolic and redox regulation. Biochemical Society Transactions 48, 2495–2504. [DOI] [PubMed] [Google Scholar]
- Valegård K, Hasse D, Andersson I, Gunn LH.. 2018. Structure of Rubisco from Arabidopsis thaliana in complex with 2-carboxyarabinitol-1,5-bisphosphate. Acta Crystallographica Section D: Structural Biology 74, 1–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Valentin K, Zetsche K.. 1989. The genes of both subunits of ribulose-1,5-bisphosphate carboxylase constitute an operon on the plastome of a red alga. Current Genetics 16, 203–209. [DOI] [PubMed] [Google Scholar]
- Van De Peer Y, Mizrachi E, Marchal K.. 2017. The evolutionary significance of polyploidy. Nature Reviews. Genetics 18, 411–424. [DOI] [PubMed] [Google Scholar]
- Van Lun M, Hub JS, Van Der Spoel D, Andersson I.. 2014. CO2 and O2 distribution in Rubisco suggests the small subunit functions as a CO2 reservoir. Journal of the American Chemical Society 136, 3165–3171. [DOI] [PubMed] [Google Scholar]
- Van Lun M, Van Der Spoel D, Andersson I.. 2011. Subunit interface dynamics in hexadecameric Rubisco. Journal of Molecular Biology 411, 1083–1098. [DOI] [PubMed] [Google Scholar]
- Villarreal JC, Renner SS.. 2012. Hornwort pyrenoids, carbon-concentrating structures, evolved and were lost at least five times during the last 100 million years. Proceedings of the National Academy of Sciences, USA 109, 18873–18878. [DOI] [PMC free article] [PubMed] [Google Scholar]
- von Caemmerer S. 2020. Rubisco carboxylase/oxygenase: from the enzyme to the globe: a gas exchange perspective. Journal of Plant Physiology 252, 153240. [DOI] [PubMed] [Google Scholar]
- Voo SS, Grimes HD, Lange BM.. 2012. Assessing the biosynthetic capabilities of secretory glands in Citrus peel. Plant Physiology 159, 81–94. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wachter RM, Salvucci ME, Carmo-Silva AE, Barta C, Genkov T, Spreitzer RJ.. 2013. Activation of interspecies-hybrid Rubisco enzymes to assess different models for the Rubisco–Rubisco activase interaction. Photosynthesis Research 117, 557–566. [DOI] [PubMed] [Google Scholar]
- Wang H, Yan X, Aigner H, Bracher A, Nguyen ND, Hee WY, Long BM, Price GD, Hartl FU, Hayer-Hartl M.. 2019. Rubisco condensate formation by CcmM in β-carboxysome biogenesis. Nature 566, 131–135. [DOI] [PubMed] [Google Scholar]
- Wang LM, Shen BR, Li BD, Zhang CL, Lin M, Tong PP, Cui LL, Zhang ZS, Peng XX.. 2020. A synthetic photorespiratory shortcut enhances photosynthesis to boost biomass and grain yield in rice. Molecular Plant 13, 1802–1815. [DOI] [PubMed] [Google Scholar]
- Wanner LA, Gruissem W.. 1991. Expression dynamics of the tomato rbcS gene family during development. The Plant Cell 3, 1289–1303. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Whitney SM, Andrews TJ.. 2001. The gene for the ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco) small subunit relocated to the plastid genome of tobacco directs the synthesis of small subunits that assemble into Rubisco. The Plant Cell 13, 193–205. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Whitney SM, Houtz RL, Alonso H.. 2011. Advancing our understanding and capacity to engineer nature’s CO2-sequestering enzyme, Rubisco. Plant Physiology 155, 27–35. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Whitney SM, Kane HJ, Houtz RL, Sharwood RE.. 2009. Rubisco oligomers composed of linked small and large subunits assemble in tobacco plastids and have higher affinities for CO2 and O2. Plant Physiology 149, 1887–1895. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wietrzynski W, Traverso E, Wollman FA, Wostrikoff K.. 2021. The state of oligomerization of Rubisco controls the rate of synthesis of the Rubisco large subunit in Chlamydomonas reinhardtii. The Plant Cell 33, 1706–1727. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wilson RH, Martin-Avila E, Conlan C, Whitney SM.. 2018. An improved Escherichia coli screen for Rubisco identifies a protein–protein interface that can enhance CO2-fixation kinetics. Journal of Biological Chemistry 293, 18–27. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wilson RH, Thieulin-Pardo G, Hartl FU, Hayer-Hartl M.. 2019. Improved recombinant expression and purification of functional plant Rubisco. FEBS Letters 593, 611–621. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wilson RH, Whitney SM.. 2017. Improving CO2 fixation by enhancing Rubisco performance. In: Alcalde M, ed. Directed enzyme evolution: advances and applications. Cham: Springer, 101–124. [Google Scholar]
- Woo JW, Kim J, Kwon SI, Corvalán C, Cho SW, Kim H, Kim SG, Kim ST, Choe S, Kim JS.. 2015. DNA-free genome editing in plants with preassembled CRISPR–Cas9 ribonucleoproteins. Nature Biotechnology 33, 1162–1164. [DOI] [PubMed] [Google Scholar]
- Wostrikoff K, Clark A, Sato S, Clemente T, Stern D.. 2012. Ectopic expression of Rubisco subunits in maize mesophyll cells does not overcome barriers to cell type-specific accumulation. Plant Physiology 160, 419–432. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wostrikoff K, Stern D.. 2007. Rubisco large-subunit translation is autoregulated in response to its assembly state in tobacco chloroplasts. Proceedings of the National Academy of Sciences, USA 104, 6466–6471. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wrighton KC, Castelle CJ, Varaljay VA, et al. 2016. RubisCO of a nucleoside pathway known from Archaea is found in diverse uncultivated phyla in bacteria. The ISME Journal 10, 2702–2714. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xia LY, Jiang YL, Kong WW, Sun H, Li WF, Chen Y, Zhou CZ.. 2020. Molecular basis for the assembly of RuBisCO assisted by the chaperone Raf1. Nature Plants 6, 708–717. [DOI] [PubMed] [Google Scholar]
- Xu Y, Li R, Luo H, Wang Z, Li MW, Lam HM, Huang C.. 2022. Protoplasts: small cells with big roles in plant biology. Trends in Plant Science 27, 828–829. [DOI] [PubMed] [Google Scholar]
- Yamada K, Davydov II, Besnard G, Salamin N.. 2019. Duplication history and molecular evolution of the rbcS multigene family in angiosperms. Journal of Experimental Botany 70, 6127–6139. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ying Z, Mulligan RM, Janney N, Houtz RL.. 1999. Rubisco small and large subunit N-methyltransferases: bi- and mono-functional methyltransferases that methylate the small and large subunits of Rubisco. Journal of Biological Chemistry 274, 36750–36756. [DOI] [PubMed] [Google Scholar]
- Yoon DK, Ishiyama K, Suganami M, et al. 2020. Transgenic rice overproducing Rubisco exhibits increased yields with improved nitrogen-use efficiency in an experimental paddy field. Nature Food 1, 134–139. [DOI] [PubMed] [Google Scholar]
- Yoon M, Putterill JJ, Ross GS, Laing WA.. 2001. Determination of the relative expression levels of Rubisco small subunit genes in Arabidopsis by rapid amplification of cDNA ends. Analytical Biochemistry 291, 237–244. [DOI] [PubMed] [Google Scholar]
- Young JN, Heureux AMC, Sharwood RE, Rickaby REM, Morel FMM, Whitney SM.. 2016. Large variation in the Rubisco kinetics of diatoms reveals diversity among their carbon-concentrating mechanisms. Journal of Experimental Botany 67, 3445–3456. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yu Y, Yu PC, Chang WJ, Yu K, Lin CS.. 2020. Plastid transformation: how does it work? Can it be applied to crops? What can it offer? International Journal of Molecular Sciences 21, 4854. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zang K, Wang H, Hartl FU, Hayer-Hartl M.. 2021. Scaffolding protein CcmM directs multiprotein phase separation in β-carboxysome biogenesis. Nature Structural & Molecular Biology 28, 909–922. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang F, Huang L, Wang W, Zhao X, Zhu L, Fu B, Li Z.. 2012. Genome-wide gene expression profiling of introgressed indica rice alleles associated with seedling cold tolerance improvement in a japonica rice background. BMC Genomics 13, 461. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang X, Carter MS, Vetting MW, et al. 2016. Assignment of function to a domain of unknown function: DUF1537 is a new kinase family in catabolic pathways for acid sugars. Proceedings of the National Academy of Sciences, USA 113, E4161–E4169. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang XH, Webb J, Huang YH, Lin L, Tang RS, Liu A.. 2011. Hybrid Rubisco of tomato large subunits and tobacco small subunits is functional in tobacco plants. Plant Science 180, 480–488. [DOI] [PubMed] [Google Scholar]
- Zhou Y, Whitney S.. 2019. Directed evolution of an improved Rubisco; in vitro analyses to decipher fact from fiction. International Journal of Molecular Sciences 20, 5019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhu XG, Long SP, Ort DR.. 2010. Improving photosynthetic efficiency for greater yield. Annual Review of Plant Biology 61, 235–261. [DOI] [PubMed] [Google Scholar]