Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2017 Feb 13;114(12):E2460–E2465. doi: 10.1073/pnas.1620058114

Modular electron-transport chains from eukaryotic organelles function to support nitrogenase activity

Jianguo Yang a,1, Xiaqing Xie a,1, Mingxuan Yang a, Ray Dixon b,2, Yi-Ping Wang a,2
PMCID: PMC5373397  PMID: 28193863

Significance

Engineering nitrogenase into cereal crops requires detailed understanding of the components required for efficient nitrogen fixation. We have used a synthetic biology modular approach to evaluate components from chloroplast, root plastids, and mitochondria that function as electron donors to both conventional Mo nitrogenase and the alternative Fe nitrogenase systems. The knowledge obtained in this study not only identifies electron-transfer components from plant organelles that can be used to support nitrogenase activity, but also is likely to enable reduction of the number of target genes required to engineer nitrogen fixation in plants.

Keywords: nitrogen fixation, electron transport, plant organelles, nitrogenase engineering

Abstract

A large number of genes are necessary for the biosynthesis and activity of the enzyme nitrogenase to carry out the process of biological nitrogen fixation (BNF), which requires large amounts of ATP and reducing power. The multiplicity of the genes involved, the oxygen sensitivity of nitrogenase, plus the demand for energy and reducing power, are thought to be major obstacles to engineering BNF into cereal crops. Genes required for nitrogen fixation can be considered as three functional modules encoding electron-transport components (ETCs), proteins required for metal cluster biosynthesis, and the “core” nitrogenase apoenzyme, respectively. Among these modules, the ETC is important for the supply of reducing power. In this work, we have used Escherichia coli as a chassis to study the compatibility between molybdenum and the iron-only nitrogenases with ETC modules from target plant organelles, including chloroplasts, root plastids, and mitochondria. We have replaced an ETC module present in diazotrophic bacteria with genes encoding ferredoxin–NADPH oxidoreductases (FNRs) and their cognate ferredoxin counterparts from plant organelles. We observe that the FNR–ferredoxin module from chloroplasts and root plastids can support the activities of both types of nitrogenase. In contrast, an analogous ETC module from mitochondria could not function in electron transfer to nitrogenase. However, this incompatibility could be overcome with hybrid modules comprising mitochondrial NADPH-dependent adrenodoxin oxidoreductase and the Anabaena ferredoxins FdxH or FdxB. We pinpoint endogenous ETCs from plant organelles as power supplies to support nitrogenase for future engineering of diazotrophy in cereal crops.


Nitrogen is one of the primary nutrients limiting plant productivity in agriculture (1). Industrial nitrogen fertilizers are used to circumvent this limitation, but have resulted in environmental pollution and expensive economic costs, especially in developing countries (2, 3). These factors have potentiated a renewed focus toward engineering biological nitrogen fixation (BNF) in cereal crops. BNF, the process that converts gaseous nitrogen to ammonia by nitrogenase enzymes, contributes >60% of the total atmospheric N2 fixed in the biogeochemical nitrogen cycle (4). Nitrogenases are a family of metalloenzymes that consist of two separable components, dinitrogenase reductase (Fe protein) and dinitrogenase (XFe protein, where X is equivalent to Mo, V, or Fe, depending on the heterometal composition of the active site cofactor) (Fig. 1 and refs. 5 and 6). All three nitrogenases catalyze the biological reduction of N2 according to the following equation: N2 + (6 + 2n)H+ + (6 + 2n)e → 2NH3 + nH2 (n ≥ 1) (79). In this process, electrons are first transferred to the Fe protein, which, in turn, donates electrons to the XFe protein with hydrolysis of two ATP molecules per electron (Fig. 1) (1012). Although Fe protein is the obligate electron donor for XFe protein in all characterized nitrogenase systems, the in vivo electron donor for Fe protein is less stringently conserved (9). Direct electron donors to Fe protein are either reduced flavodoxin or reduced ferredoxin, which, in turn, are reduced by a variety of oxidoreductase systems, depending on the physiology of the host diazotroph (1317).

Fig. 1.

Fig. 1.

(Upper) Modular arrangement of genes required for MoFe and the minimal FeFe nitrogenase systems. Letters within the arrows represent the corresponding nif genes or the anfHDGK structural genes encoding FeFe nitrogenase. (Lower) Schematic pathways for electron donation to nitrogenase and electron transfer within nitrogenase. Structures of representative proteins are shown. PFOR (NifJ), pyruvate–ferredoxin (flavodoxin) oxidoreductase [Protein Data Bank (PDB) ID code 1B0P]; Rnf complex, NADH–ferredoxin oxidoreductase [the structure shown in gray is a homology model based on the Nqr complex (Na+-translocating NADH–quinone oxidoreductase from Vibrio alginolyticus; PDB ID code 4P6V) using the online software from https://swissmodel.expasy.org]; FNR (PDB ID code 1QUE); NifF, flavodoxin (PDB ID code 2WC1); FdxN, 2[4Fe–4S]-type ferredoxin (PDB ID code 2OKF); FdxH, [2Fe–2S]-type ferredoxin (PDB ID code 1FRD); Fe protein, dinitrogenase reductase (PDB ID code 1G5P); and XFe protein (where X refers to Mo, V, or Fe), dinitrogenase (PDB ID code 3K1A; MoFe nitrogenase). The cofactors of the Fe and XFe proteins are shown as ball-and-stick models. Atom colors are Fe in rust, S in yellow, C in gray, O in red, and heterometal X in purple.

A number of studies have suggested chloroplasts, root plastids, or mitochondria as suitable locations for expression of nitrogenase in eukaryotes (1820). These energy-conversion organelles can potentially provide reducing power and ATP required for the nitrogen-fixation process. Diverse reduction reactions carried out in these organelles rely on different electron-transport chains (21). Multiple gene copies of ferredoxins have been identified in all plants, including photosynthetic or nonphotosynthetic ferredoxins mainly expressed in chloroplasts or root plastids, respectively; and ferredoxin-like adrenodoxins located in the mitochondria (21, 22). The major function of the photosynthetic ferredoxins is to transfer electrons from photosystem I to NADPH, catalyzed by leaf-type ferredoxin–NADPH oxidoreductase (LFNR) (23). In addition, photosynthetic ferredoxins work to distribute reducing power derived from the photosynthetic process to several ferredoxin-dependent enzymes for nitrogen and sulfur assimilation (24). Electron transfer between root-type ferredoxin–NADPH oxidoreductase (RFNR) and ferredoxin in the root plastid is reversed, with NADPH generated in the oxidative pentose-phosphate pathway being used to reduce RFNR and, in turn, ferredoxin (25). In mitochondria, adrenodoxin serves to transfer electrons from NADPH-dependent adrenodoxin oxidoreductase (MFDR) to the cysteine desulfurase Nfs1 to participate in the biosynthesis of the biotin (26).

Recently, we successfully reassembled the Klebsiella oxytoca (Ko) MoFe (27) and the “minimal” Azotobacter vinelandii (Av) FeFe (28) nitrogenase systems in Escherichia coli (Fig. 1). From the synthetic biology point of view, these two nitrogenase systems can be divided into three functional modules: the electron-transport component (ETC) module, the metal cluster biosynthesis module, and the “core” enzyme module (Fig. 1). In the present study, ETC modules from plastids and mitochondria, representing potential locations for nitrogenase in plants, were used to test their capability to support the activity of either MoFe or FeFe nitrogenase in E. coli. Our results indicate that intact ETC modules from the chloroplast and root plastid, or hybrid modules from mitochondria, can functionally support nitrogenase activity. Therefore, our results unravel the requirements for electron-transport components for engineering diazotrophy in different plant organelles.

Results

Hybrid ETC Modules Consisting of the NifJ Protein and Plastid Ferredoxins Can Functionally Support Nitrogenase Activity.

In many diazotrophs, the direct electron donor to nitrogenase is either reduced ferredoxin or reduced flavodoxin, which have been demonstrated to receive electrons from the pyruvate–flavodoxin (ferredoxin) oxidoreductase, encoded by nifJ, in some cases (1315). Most plants are known to have multiple copies of ferredoxins located in different organelles (21). Through preliminary sequence analysis, we found that chloroplast and root-plastid ferredoxins from plants show high sequence identity with the Anabaena sp. PCC 7120 (As) fdxH gene product (Fig. S1), which is the primary electron donor for nitrogenase in cyanobacteria (29). To investigate whether hybrid ETC modules formed by the NifJ protein and plastid ferredoxins could support nitrogenase activity in E. coli, coding sequences of several representative plastid ferredoxins from Chlamydomonas reinhardtii (Cr; CrPETF), Arabidopsis thaliana (At; AtFD2 and AtFD3), Zea mays (Zm; ZmFDI and ZmFDIII), Oryza sativa (Os; OsFD1 and OsFD4), and Triticum aestivum (Ta; TaFD) were selected for further study. These ferredoxin-encoding genes were codon-optimized for E. coli (Dataset S1), and expressed from the inducible PLteto-1 promoter (Fig. 2A; details are provided in SI Materials and Methods). The fdxH gene from As was also introduced as a control to verify effectiveness of the inducible system.

Fig. S1.

Fig. S1.

Sequence alignment of the AsFdxH protein with ferredoxins from plastids (A) or mitochondria (B). Sequences shaded in green in A or crimson in B are signal peptides for the plant-type ferredoxins. Cysteine residues for liganding the [2Fe–2S] cluster are highlighted in yellow. Abbreviations are the same as in the main text (Fig. 2).

Fig. 2.

Fig. 2.

Influence of hybrid ETC modules consisting of the NifJ protein with ferredoxins (FDs) from different plant organelles on nitrogenase activity in E. coli. (A) Schematic diagram for electron transport between hybrid ETC modules and nitrogenases. (BG) In each case, nifF was replaced by FDs to form hybrid modules consisting of NifJ with chloroplast FDs (B and C), NifJ with root-plastid FDs (D and E), or NifJ with mitochondrial FDs (F and G). In all cases, cultures were assayed for acetylene reduction either in the absence (filled bars) or presence (open bars) of the appropriate inducer required to express heterologous ETCs as described in SI Materials and Methods and Fig. S9. The activity of FeFe or MoFe nitrogenases when expressed in the presence of the NifJ–NifF module from native nif promoters represents 100% activity in each case (in the absence of added inducer). FeFe represents the minimal FeFe nitrogenase system, and MoFe represents the reassembled MoFe nitrogenase system. Assembly of these nitrogenases requires both the “metal cluster biosynthesis module” and the “core enzyme” module, respectively, as shown in Fig. 1. As, Anabaena sp. PCC 7120; At, Arabidopsis thaliana; Cr, Chlamydomonas reinhardtii; Os, Oryza sativa; Ta, T. aestivum; Zm, Zea mays. Error bars indicate the SD observed from at least three independent experiments.

To assay electron transport by plant ferredoxins, the flavodoxin encoded by nifF in the Ko NifJ–NifF module was replaced by coexpression of the respective plant gene and the resultant activity of the reassembled MoFe (27) or the minimal FeFe (28) nitrogenase was analyzed by the method of acetylene reduction. Although significant background activity was observed in the absence of the NifJ–NifF module, all hybrid ETC modules stimulated nitrogenase activity by both the MoFe and FeFe systems, to varying extents (Fig. 2 BE). Stimulation of activity by plastid ferredoxins was dependent on the presence of NifJ, indicating that reducing power is provided by the pyruvate oxidoreductase activity of this electron donor (Fig. S2). Interestingly, values >100% were observed for the FeFe nitrogenase system when NifF was replaced with the ferredoxins from As (FdxH), Cr (PETF), or Os (FD1), respectively (Fig. 2B and Table S1). This phenomenon suggests that the AvAnfH protein in the hybrid minimal FeFe nitrogenase (28) may prefer ferredoxin, rather than flavodoxin, as an electron donor. All of the chloroplast ferredoxins could restore ∼100% activity for the FeFe nitrogenase system, with the exception of the NifJ–AtFD2 hybrid module, which showed ∼76% activity (Fig. 2B). In contrast, only the NifJ–CrPETF and NifJ–TaFD hybrid modules could restore >90% activity to the MoFe nitrogenase system, whereas the NifJ–AtFD2, NifJ–ZmFDI, and NifJ–OsFD1 hybrid modules exhibited <70% activity (Fig. 2C). All root-plastid ferredoxin-derived hybrid modules showed lower nitrogenase activities compared with their chloroplast ferredoxin counterparts derived from the same organism (Fig. 2 BE). It is possible that the different activities observed on substitution of ferredoxins from various plant origins could result from variations in expression levels. However, we were unable to confirm this possibility because we could not detect ferredoxin proteins by Western blotting using an anti-His monoclonal antibody (details are provided in SI Materials and Methods). Together, these results indicated that hybrid ETC modules formed with the NifJ protein, and plastid ferredoxins can direct the transfer of electrons to nitrogenase.

Fig. S2.

Fig. S2.

Plant-type ferredoxins are insufficient to support nitrogenase activity for either the FeFe (A) or MoFe (B) nitrogenase in the absence of NifJ. In all cases, cultures were assayed for acetylene reduction either in the absence (filled bars) or presence (open bars) of the appropriate inducer required to express heterologous ETCs as described in SI Materials and Methods. The activity of FeFe or MoFe nitrogenases when expressed in the presence of the NifJ–NifF module in the absence of added inducer represents 100% activity in each case. FeFe represents the minimal FeFe nitrogenase system; MoFe represents the reassembled MoFe nitrogenase system. Error bars indicate the SD observed from at least three independent experiments.

Table S1.

The redox potential of ferredoxins is not the primary factor limiting electron transfer to nitrogenase from pyruvate–ferredoxin (flavodoxin) oxidoreductase (NifJ)

Organism* Location Genes Redox potential Relative nitrogenase activity, % acetylene reduction
FeFe MoFe
Ko nifF −412 mV (34) 100 ± 6 100 ± 15
ΔnifF 31 ± 6 10 ± 3
As Heterocyst fdxH −351 mV (46) 153 ± 21 100 ± 6
Vegetative petF −384 mV (46) 110 ± 14 81 ± 11
Cr Chloroplast PETF −398 mV (47) 105 ± 11 90 ± 5
Chloroplast FDX2 −321 mV (47) 92 ± 4 87 ± 4
Zm Chloroplast FDI −423 mV (48) 96 ± 9 56 ± 6
Chloroplast FDII −406 mV (48) 75 ± 7 50 ± 7
Root plastid FDIII −321 mV (49) 82 ± 9 51 ± 7
At Chloroplast FD1 −425 mV (50) 59 ± 6 36 ± 6
Chloroplast FD2 −433 mV (50) 76 ± 11 50 ± 11
Root plastid FD3 −337 mV (50) 68 ± 4 34 ± 8
*

Origin of flavodoxin or ferredoxin genes expressed in E. coli.

Redox potentials reported previously with references in parentheses.

Nitrogenase activity of E. coli strains expressing NifJ in the presence of the indicated flavodoxin or ferredoxin components. Activities observed from the minimal FeFe or reassembled MoFe nitrogenase systems in the presence of the NifJ–NifF module represent 100% activity, respectively. Data presented are mean values based on at least three independent experiments.

Because mitochondria represent another potential location for nitrogenase in plants, the capability of mitochondrial ferredoxins to support nitrogen fixation was also investigated in E. coli. The same strategy was used to clone the mitochondrial adrenodoxin coding genes from At (AtMFD1 and AtMFD2) as described above. When mitochondrial ferredoxin-derived hybrid ETC modules (NifJ–AtMFD1 or NifJ–AtMFD2) were introduced into E. coli, no restoration of activity was observed compared with the nifF-deficient MoFe or FeFe nitrogenase systems (Fig. 2 F and G). To exclude the possibility that the GroESL proteins are potentially required for proper and efficient folding of mitochondrial ferredoxins in E. coli (26), a high-copy plasmid carrying the GroESL-encoding genes was cotransformed with MoFe or FeFe nitrogenase systems carrying either the NifJ–AtMFD1 or NifJ–AtMFD2 hybrid modules. Similar negative results were obtained (Fig. S3). Interestingly, phylogenetic analysis showed that the mitochondrial ferredoxins are not clustered with any of the well-defined electron donors for nitrogenase (Fig. S4). Overall, these results suggest that the mitochondria ferredoxins cannot couple with the NifJ protein to form functional ETC modules for nitrogenase systems.

Fig. S3.

Fig. S3.

Acetylene reduction by FeFe (A) or MoFe nitrogenase systems (B) expressed in E. coli with hybrid ETC modules consisting of NifJ and mitochondrial ferredoxins from At (AtMFDs) and cotransformed with a high-copy plasmid carrying the groES operon. In all cases, cultures were assayed for acetylene reduction either in the absence (filled bars) or presence (open bars) of the appropriate inducer required to express heterologous ETCs as described in SI Materials and Methods. The activity of FeFe or MoFe nitrogenases when expressed in the presence of the NifJ–NifF module in the absence of inducer represents 100% activity in each case. FeFe represents the minimal FeFe nitrogenase system; MoFe represents the reassembled MoFe nitrogenase system. Error bars indicate the SD observed from at least three independent experiments.

Fig. S4.

Fig. S4.

Phylogenetic analysis of the ferredoxins and flavodoxins from different organisms using the maximum-likelihood method with MAGA software (Version 6.0). A Poisson model was used to analysis the substitution rate of each site. The scale with the solid line represents 0.5 expected substitutions per site, and the scale with the dashed line represents 2.5 expected substitutions per site. Bootstrap values of 500 replicates are indicated in numbers at each junction. FdxH from As is colored in red; chloroplast FDs are colored in green; root-plastid FDs are shown in gray; and mitochondria FDs are represented in brown. FD or Fld sequence assignments are as follows: (As) AsFdxH, WP_010995602.1; AsPetF, WP_010998287.1; (Cr) CrPETF, XP_001692808.1; CrFDX2, XP_001697912.1; (At) AtFD1, NP_172565.1; AtFD2, NP_176291.1; AtFD3, NP_180320.1; AtMFD1, AAL82812.1; AtMFD2, AAL82813.1; (Zm) ZmFDI, P27787.1; ZmFDII, O80429.1; ZmFDIII, NP_001105346.1; (Os) OsFD1, NP_001060779.1; OsFD4, NP_001055675.1; OsMFD1, XP_015612701.1; OsMFD2, XP_015647181.1; (T. aestivum) TaFD, P00228.2; (Ko) KoNifF, WP_004138775.1; (Av) AvNifF, ACO76434.1; AvFdxN, WP_012703542.1; AvVnfFd, WP_012698954.1; (R. capsulatus) RcNifF, AAC05792.1; RcFdxN, CAA35699.1; (Rhizobium meliloti) SmFdxN, NP_435687.1; (Saccharomyces cerevisiae) ScYAH1, NP_015071.1; and (Mus musculus) MmADX, NP_032022.1.

Influence of Hybrid ETC Modules Consisting of Plant-Type Ferredoxin–NADPH Oxidoreductases and Well-Defined Electron Donors on Nitrogenase Activity.

In plants, three different types of the ferredoxin–NADPH oxidoreductases (FNRs) are identified and exist in different organelles. All of these FNRs function to mediate electron transfer between ferredoxins and NADPH (23, 25, 26). From phylogenetic analysis, we found that the chloroplast FNRs form a subgroup with the FNR proteins from cyanobacteria, whereas the root-plastid FNRs form another subgroup with the FNR proteins from algae (Fig. S5). In contrast, the mitochondrial MFDRs diverged away from these two subgroups during early evolution (Fig. S5).

Fig. S5.

Fig. S5.

Phylogenetic analysis of FNRs from different organisms using the maximum-likelihood method with MAGA software (Version 6.0). A Poisson model was used to analysis the substitution rate of each site. The scale with the solid line represents 0.2 expected substitutions per site, and the scale with the dashed line represents 0.8 expected substitutions per site. Bootstrap values of 500 replicates are indicated in numbers at each junction. FNR from At is shown in red, chloroplast FNRs in green, root-plastid FNRs in gray, and mitochondrial FNRs in brown. FNR sequence assignments are as follows (As) AsFNR, WP_010998260.1; (Synechocystis sp. PCC 6803) SsFNR, WP_010873079.1; (Cr) CrFNR, XP_001697352.1; (Ostreococcus taurii) OtFNR, XP_003084170.1; (At) AtLFNR1, AAF79911.1; AtLFNR2, CAB52472.1; AtRFNR1, CAB81081.1; AtRFNR2, NP_174339.1; AtMFR, NP_194962.2; (Zm) ZmLFNR1, NP_001105568.1; ZmLFNR2, NP_001104851.1; ZmLFNR3, NP_001149023.2; ZmRFNR1, NP_001295409.1; ZmRFNR2, ACG35047.1; (Os) OsLFNR1, XP_015640980.1; OsLFNR2, BAS76542.1; OsRFNR1, XP_015629836.1;OsRFNR2, XP_015646844.1; OsMFDR, XP_015626908.1; (E. coli) EcFpr, WP_000796332.1; (S. cerevisiae) ScARH1, AAC49500.1; and (M. musculus) MmFDXR, NP_032023.1.

To investigate whether hybrid ETC modules consisting of the plant-type FNRs and well-defined electron donors (KoNifF, AsFdxH, and AsFdxB) could direct electron transfer to nitrogenase in E. coli, well-characterized chloroplast or root-plastid FNRs from Cr and Zm, plus the At mitochondrial MFDR, were selected to verify this assumption. These hybrid modules were transformed into the E. coli, and their activities were assayed by the acetylene reduction method. None of the hybrid ETC modules consisting of the plant-type FNRs and NifF could stimulate acetylene reduction by either the MoFe or FeFe nitrogenase systems (Fig. S6). In contrast, all hybrid modules formed with the plant-type FNRs and AsFdxH could partially restore nitrogenase activity to both the MoFe and FeFe nitrogenase systems (Fig. 3 B and C). However, only MFDR from mitochondria could support electron transfer to the nitrogenases when coupled to AsFdxB (Fig. S6).

Fig. S6.

Fig. S6.

Influence of hybrid ETC modules consisting of KoNifF or AsFdxB with FNRs from different plant organelles on nitrogenase activity in E. coli. The NifJ–NifF module was replaced either by hybrid modules consisting of KoNifF with plant-type FNRs (A and B) or AsFdxB with plant-type FNRs (C and D). In all cases, cultures were assayed for acetylene reduction either in the absence (filled bars) or presence (open bars) of the appropriate inducer required to express heterologous ETCs as described in SI Materials and Methods. The activity of FeFe or MoFe nitrogenases when expressed in the presence of the NifJ–NifF module in the absence of added inducer represents 100% activity in each case. MoFe represents the reassembled MoFe nitrogenase system. Error bars indicate the SD observed from at least three independent experiments.

Fig. 3.

Fig. 3.

Influence of hybrid ETC modules consisting of the cyanobacterial ferredoxin AsFdxH combined with FNRs from different plant organelles on nitrogenase activity in E. coli. (A) Schematic diagram for electron transport between hybrid ETC modules and nitrogenases. (B and C) The NifJ–NifF ETC module was replaced by hybrid modules consisting of plant-type FNRs and AsFdxH. Experimental details and abbreviations are the same as in Fig. 2. Error bars indicate the SD observed from at least three independent experiments.

The fact that ETC modules formed with the plant-type FNRs and NifF could not functionally support nitrogenase activity is not surprising, because FNRs known to reduce flavodoxins are limited to a few bacteria, such as Fpr from E. coli and FNRs from some species of cyanobacteria, which are usually related to the biosynthesis of cytochrome P450 (30). Microbial FNRs showed a distant evolutionary relationship with the chloroplast and root-plastid FNRs (Fig. S5). In addition, no flavodoxins have been identified in plants (31). Therefore, the plant-type FNRs do not appear to have coevolved with the flavodoxins.

Intact ETC Modules from Chloroplasts and Root Plastids Can Functionally Support Nitrogenase Activity.

After evaluating the function of the hybrid modules, further experiments were carried out to investigate whether intact ETC modules, consisting of FNRs and their cognate ferredoxin counterparts from plant organelles, could support nitrogenase activity. By combining the Ptac-controlled FNRs with PLtetO-1-controlled ferredoxins (details are provided in SI Materials and Methods), two chloroplast ETC modules, CrFNR–PETF and ZmLFNR–FDI; one root-plastid ETC module, ZmRFNR–FDIII; and one mitochondrial ETC module, AtMFDR–MFD1 were constructed. Because it is known that the AsPetH–FdxH module from cyanobacteria can function to support nitrogen fixation in its original host, this module was also constructed and used as a control.

The ability of the intact ETC modules to support nitrogenase activity as replacement for the NifJ–NifF module was assayed by both acetylene reduction and 15N-assimilation methods. With the exception of the AtMFDR–MFD1 module from mitochondria, all other ETC modules conferred ability to partially restore acetylene reduction and 15N assimilation to both MoFe and FeFe nitrogenases (Fig. 4 BE). In contrast, no obvious stimulation of activity could be observed, when either of the two components from each of the modules was expressed individually, demonstrating that the plant-type ferredoxins do not function in electron transport in the absence of their cognate FNR and, conversely, that the plant FNRs cannot donate electrons to nitrogenase in the absence of a coexpressed ferredoxin component (Fig. S7). Thus, both members of the plant ETC pairs are necessary for functionality.

Fig. 4.

Fig. 4.

Influence of intact ETC modules consisting of FNRs from different plant organelles with their cognate ferredoxins (FDs) on nitrogenase activity in E. coli. (A) Schematic representation for electron transport between intact plant ETC modules and nitrogenases. (BE) Nitrogen fixation by FeFe or MoFe nitrogenases was assayed either by acetylene reduction (B and C; black bars) or 15N assimilation (D and E; striped bars). Experimental details and abbreviations are the same as in Fig. 2. Error bars for the acetylene reduction assay indicate the SD observed from at least three independent experiments. Error bars for 15N assimilation indicate the SD observed from at least two independent experiments.

Fig. S7.

Fig. S7.

Single components of each ETC module are insufficient to support nitrogenase activity for either the FeFe (A) or MoFe (B) nitrogenase. In all cases, cultures were assayed for acetylene reduction either in the absence (filled bars) or presence (open bars) of the appropriate inducer required to express heterologous ETCs as described in SI Materials and Methods. The activity of FeFe or MoFe nitrogenases when expressed in the presence of the NifJ–NifF module in the absence of added inducer represents 100% activity in each case. FeFe represents the minimal FeFe nitrogenase system; MoFe represents the reassembled MoFe nitrogenase system. Error bars indicate the SD observed from at least three independent experiments.

Two of the chloroplast modules, CrFNR–PETF and ZmLFNR–FDI, showed comparable amounts of activity to those observed with the AsPetH–FdxH module from cyanobacteria (∼45% acetylene reduction and ∼30% 15N assimilation) with both FeFe and MoFe nitrogenases (Fig. 4). However, the ZmRFNR–FDIII module from the root plastid was less active than its corresponding chloroplast module from the same plant with respect to both nitrogenases (Fig. 4 BE). Interestingly, weak complementation was observed with the AtMFDR–MFD1 module (11% 15N assimilation activity) when combined with MoFe nitrogenase, compared with the NifJ–NifF-deficient negative control (6% 15N assimilation activity) (Fig. 4E). However, this phenotype was not observed with FeFe nitrogenase. (Fig. 4D). Because multiple copies of ferredoxins exist in E. coli, it is possible that the enhanced activity of MoFe nitrogenase resulted from the contribution of hybrid modules formed between AtMFDR and those from E. coli ferredoxins. This effect may not be observed with the FeFe nitrogenase system because of its higher background activity in the absence of nifF and nifJ (Fig. 4D). Together, these results demonstrate that intact ETC modules from plastids, but not their equivalents from mitochondria, are capable of transferring electrons to nitrogenases.

SI Materials and Methods

Bacterial Strains and Growth Media.

Luria–Bertani broth for E. coli growth contained 10 g/L tryptone, 5 g/L yeast extract, and 10 g/L NaCl. KPM minimal medium used in this study contained 10.4 g/L Na2HPO4, 3.4 g/L KH2PO4, 26 mg/L CaCl2·2H2O, 30 mg/L MgSO4, 0.3 mg/L MnSO4, 36 mg/L ferric citrate, 10 mg/L para-aminobenzoic acid, 5 mg/L biotin, 1 mg/L vitamin B1, and 0.8% (wt/vol) glucose, with 20 mM ammonia (KPM-HN) or 10 mM glutamate (KPM-LN) as the nitrogen source. Casamino acids (BD Biosciences, catalog no. 223050) at a final concentration of 0.05% was also added to the KPM minimal medium to ensure normal growth. Antibiotics were used at the following concentrations: 50 μg/mL for ampicillin and 25 μg/mL for chloramphenicol.

Construction of Recombinant Plasmids.

Plasmid pKU7815 and pKU7017 are pACYC184 derivatives carrying the minimal FeFe (28) or the reassembled MoFe nitrogenase system (27). nifF and nifJ double deletion derivatives of pKU7815 and pKU7017, designated as pKU7830 and pKU7831 in this study, were constructed by direct removal of the nifF and nifJ operons by using the specific restriction sites flanking each transcription unit: SwaI for nifF and ScaI for nifJ, respectively. The complementary plasmid for pKU7830 and pKU7831 is a pBR322M derivative (pKU7832) carrying both the nifF and nifJ operons. The pBR322M-PLtetO-1 plasmid was constructed by direct assembly of the tetR expression cassette and the PLtetO-1 promoter region with the pBR322M as backbone by using the Gibson Assembly Cloning kit (NEB, catalog no. E5520S). The tetR expression cassette consists of a strong constitutive promoter (BBa_J23100; parts.igem.org/Main_Page), a medium strength ribosome-binding site (RBS, BBa_B0032; parts.igem.org/Main_Page), and the thrL terminator from E. coli. Similarly, the pBR322M-Ptac plasmid was constructed by direct assembly of the lacI expression cassette and the Ptac promoter carrying a weak RBS with pBR322M as backbone. To lower the leaky expression of plant-type FNRs, a LacI mutant LacIWF (43) with higher affinity for the lac operator expressed from PlacIq was used for construction of the pBR322M-Ptac plasmid. The pKU7833 plasmid was constructed by replacing the nifF gene in the pKU7832 with the SwaI restricted tetR-PLtetO-1 fragment from plasmid pBR322M-PLtetO-1. To construct plasmids with hybrid modules expressing NifJ and plant-type ferredoxins, original gene sequences or chemically synthesized gene sequences (Dataset S1) were cloned downstream of the PLtetO-1 promoter of the pKU7833 plasmid using NdeI/SpeI restriction sites. To facilitate detection of the expression level of different ferredoxins, a sequence encoding a hexahistidine tag was added to each of the synthesized ferredoxin sequences as shown by the gray shadows in Dataset S1. The pKU7846 plasmid was constructed by replacing the nifJ gene in pKU7832 with the ScaI-restricted lacI-Ptac fragment from plasmid pBR322M-Ptac. For construction of plasmids with hybrid modules expressing plant-type FNRs and NifF, original gene sequences or chemically synthesized gene sequences (Dataset S1) were cloned downstream of the Ptac promoter of the pKU7846 plasmid using NdeI/SpeI restriction sites. The pKU7853 plasmid, carrying Ptac-AspetHori/PLtetO-1-AsfdxHori, was constructed by replacing the nifF gene in plasmid pKU7847 with the ScaI PLtetO-1-AsfdxH fragment from pKU7834. Similar methods were used to construct the pKU7854-pKU7857 and pKU7859-pKU7865 plasmids. The PCR product carrying groES with its original promoter, flanked by XbaI/SpeI restriction sites, was PCR amplified from the genome of E. coli and directly ligated to the pEASY-Blunt vector (TransGene, catalog no. CB101) to generate pKU7845. Plasmids maps are provided in Fig. S8. Each of the above constructs was confirmed by DNA sequencing before any further experiments.

Fig. S8.

Fig. S8.

Plasmid maps of the main vectors used in this study. rrnB T1 is an E. coli terminator (BBa_B0010); TL is the E. coli thrL gene terminator; T0 is the phage lambda t0 terminator; and Ta is an artificial terminator L3S2P21 described by Chen et al. (45).

Acetylene Reduction Assay.

To measure the activity of E. coli JM109 derivatives expressing reassembled nitrogenase systems, cells were initially grown overnight in KPM-HN medium. The cells were then diluted into 2 mL of KPM-LN medium in 20-mL sealed tubes to a final OD600 of ∼0.3. To optimize expression of the plant components, 200 ng/mL anhydrotetracycline (aTc) or 200 μM isopropyl-β-d-thiogalactoside (IPTG) was added to induce expression of the ferredoxins or FNRs, respectively, as indicated by the results in Fig. S9. Air in the tubes was repeatedly evacuated and replaced with argon. After incubation at 30 °C for 6–8 h, 2 mL of C2H2 was injected, and the gas phase was analyzed ∼16 h later with a Shimadzu GC-2014 gas chromatograph. Data presented are mean values based on at least three replicates.

Fig. S9.

Fig. S9.

Induction profiles of CrPETF or CrFNR expression in relation to nitrogenase activity. (A and B) Induction profile of CrPETF expression, controlled by the PLtetO-1 promoter in response to aTc concentration. (C and D) Induction profile of CrFNR expression controlled by the Ptac promoter in response to IPTG concentration. Acetylene reduction activity shown on the y axis is relative to the activity obtained from the FeFe or MoFe nitrogenase system expressed in the presence of the NifJ–NifF module. FeFe represents the minimal FeFe nitrogenase system; MoFe represents the recombined MoFe nitrogenase system. Error bars indicate the SD observed from at least three independent experiments.

15N2 Assimilation Assay.

To detect 15N2 assimilation, E. coli JM109 derivatives expressing reassembled nitrogenase systems were grown as described for the acetylene reduction assay. The tubes were filled with N2 gas, and then 3 mL of gas was removed, and 2 mL of 15N2 gas (99%+; Shanghai Engineering Research Center for Stable Isotopes) was injected. After 48 h of incubation at 30 °C, the cultures were collected, freeze -dried, ground, weighed, and then sealed into tin capsules. Isotope ratios are expressed as δ15N, where values are a linear transform of the isotope ratios 15N/14N, representing the per mille difference between the isotope ratios in a sample and in atmospheric N2 (44). Data presented are mean values based on at least two replicates.

Discussion

It is generally accepted that several factors may limit successful engineering of nitrogenase in cereal crops, including provision of the appropriate physiological environment for the enzyme [e.g., availability of energy (ATP), reducing power, and a low-oxygen environment], in addition to the relative large number of nif genes required for biosynthesis of nitrogenase itself. Attempts have been made to reduce the number of genes required for both MoFe (32) and FeFe nitrogenase biosynthesis (28) by using E. coli as a chassis. However, in the case of nif genes originating from Paenibacillus sp., it was observed that nitrogenase activity decreased drastically when the number of genes were decreased to nine for MoFe nitrogenase (32), but activity could be recovered when additional genes were recruited back into the system (33). These findings have highlighted the difficulties in minimizing nif gene sets for synthetic biology.

One approach to reducing the number of genes required for nitrogenase activity is to make use of host-encoded ETCs. Bioinformatic analysis indicated that ETC modules from plant organelles show close evolutionary relationships with nitrogenase ETCs from nitrogen-fixing cyanobacteria (Figs. S4 and S5). Therefore, it is reasonable to question whether these ETCs can replace the NifJ–NifF module and support nitrogen fixation. We have used the modularity concept to analyze compatibility between ETCs from different plant organelles with nitrogenase enzyme modules. Our results indicate that plastid ferredoxins can efficiently participate in electron transfer to nitrogenase from either the nif-specific pyruvate–flavodoxin (ferredoxin) reductase encoded by nifJ or from its FNR counterparts in plants. Overall, we observed little difference in the response of the MoFe or FeFe nitrogenases to the various ETC modules we analyzed, suggesting that plant ETCs can function with either enzyme.

We found that almost all plant-originated ferredoxins used in this study (except adrenodoxins from mitochondria) could functionally substitute for NifF for both FeFe and MoFe nitrogenases (Fig. 2). This finding implies that the interface between these ferredoxins and NifH/AnfH are competent for electron transfer. In many cases, the previously reported redox potentials of these ferredoxins (Table S1) are higher than those of the Fe protein NifH (−412 mV) (34), which disfavors electron transfer to nitrogenase in vitro (35). Furthermore, the redox potential of the NADH/NAD+ couple (midpoint potential −380 mV) is, in theory, too high to drive electron donation to nitrogenase by FNRs, although NAD(P)H-dependent electron-transfer chains apparently support nitrogenase activity in vivo in aerobic diazotrophs. To account for this conundrum, it has been proposed that either proton motive force or electron bifurcation drives the reversed electron flow required to overcome these bioenergetic constraints in vivo (36, 37). Because multiple factors, including expression levels, interface compatibility, and protein ratios, may affect electron transfer to nitrogenase, it is not possible to determine whether ferredoxin redox potentials relate to the efficiency of electron transfer in our experimental system.

Overall, our results lay a solid foundation for the use of extant ETC modules from plant organelles to engineer BNF in cereal crops. The observation that ETC modules from both chloroplasts and root plastids can functionally support nitrogenase activity implies that engineering diazotrophy in plastids does not require insertion of an ETC module. For instance, for the FeFe nitrogenase system, of the minimal 10 genes required (28), 2 of these function as the bacterial ETC module and presumably would not be essential for engineering nitrogenase activity in plants. To further reduce the number of nif/anf genes required, plant genes that can substitute for components of the “metal cluster biosynthesis” module (Fig. 1), such as those required for iron–sulfur cluster assembly, could also be considered. Two recent studies have shown that the [Fe–S] cluster biosynthesis components from eukaryotic organelles can substitute for NifU and NifS to incorporate the [4Fe–4S] cluster into the Fe protein of nitrogenase (19, 20).

Accumulated mutational studies indicate that many diazotrophs have multiple pathways that mediate electron flow to nitrogenase. These pathways include multiple electron-transfer proteins (ferredoxins and flavodoxins) and multiple oxidoreductases, including pyruvate–flavodoxin (ferredoxin) oxidoreductase, FNR, and the Rnf complex [a membrane-bound ion-translocating NADH–ferredoxin oxidoreductase, first identified in Rhodobacter capsulatus by Schmehl et al. (16)] (Fig. 1). The putative membrane-associated electron-transferring flavoprotein complex encoded by the fixABCX operon has also been proposed to transfer electrons to nitrogenase (38). Such redundancy has endowed nitrogenase with the ability to cope with heterogenous ETC modules. Conversely, the [2Fe–2S]-type ferredoxin is the most extensively used electron shuttle, which is maintained throughout the tree of life and is involved in a plethora of metabolic, regulatory, dissipative, and developmental processes (31). To satisfy the requirements for electron transfer to diverse proteins, ferredoxins have evolved multiple interfaces to cope with different metabolic partners (39). This interface flexibility may provide another explanation for the compatibility of the nitrogenase system with heterogenous ETC modules from plant chloroplasts and root plastids.

Combining the results obtained from this study, a schematic model for electron transfer to nitrogenase in organelles of engineered plants is proposed (Fig. 5). Our results imply that the endogenous ETC module present in mitochondria is not competent to support nitrogenase activity (Fig. 4), reflecting the requirement for an additional ETC component, such as AsFdxH or AsFdxB, that can function as a hybrid module for direct electron transfer to nitrogenase (Fig. 3). In this scenario, NADPH generated either by glycolysis or by isocitrate dehydrogenase can provide reducing equivalents for electron transfer to nitrogenase through the hybrid MFDR–ferredoxin pathway (Fig. 5). In root plastids, the NADPH generated by degradation of glucose through the oxidative pentose phosphate pathway is used by RFNR to catalyze reduction of ferredoxins. We propose that this electron-transfer pathway can be used to provide reducing power for nitrogenase in the absence of the nif-specific NifJ–NifF ETC module (Fig. 5).

Fig. 5.

Fig. 5.

Schematic model illustrating potential routes for electron transfer to nitrogenase in engineered plant organelles. The diagram depicts an artificial plant cell in which a chloroplast and root plastid coexists in the same cell. The main components or processes for generation of reducing power are shown. Components within organelles with solid outlines represent existing proteins present in plants, whereas components with dashed outlines represent those required to engineer nitrogenase activity as suggested by our results. The red arrow in the root plastid represents RFNR–ferredoxin-mediated electron transfer from NADPH to nitrogenase, with reducing equivalents being supplied by the oxidative pentose phosphate pathway (OxPPP). In mitochondria, the red arrow indicates the election transfer pathway from MFDR to nitrogenase, which can function if a heterologous ferredoxin such as AsFdxH or AsFdxB is introduced. NADPH can be supplied either through glycolysis or via the oxidative TCA cycle by isocitrate dehydrogenase (ICDH). In chloroplasts, light-activated photosystem II (PSII) extracts electrons from water and transfers them to plastoquinone (PQ), and then through cytochrome b6f (Cytb6f) to plastocyanin (PC), which then feeds electrons to the light-oxidized photosystem I (PSI). PSI-derived electrons are used to reduce ferredoxin, which then transfers reductant to either LFNR, for NADPH production, or to nitrogenase, for nitrogen fixation. The NADPH generated can also promote reverse electron transfer to nitrogenase via ferredoxin, catalyzed by LFNR. The bottom lighter half of the chloroplast represents the light condition, with the blue arrow representing the photo-coupled “charging” process for accumulating NADPH; the darker half of the chloroplast represents the dark condition, with the yellow arrow representing the LFNR–ferredoxin mediated “discharging” process for transferring electrons from NADPH to nitrogenase (N2ase). As the major site for ammonia assimilation, ammonia produced in the chloroplast can be immediately assimilated by glutamine synthetase (GS) (42). Glc-6P, glucose-6-P; Rib-5P, ribose-5-P.

Although chloroplasts are potentially problematic for nitrogenase engineering in terms of oxygen evolution, it is possible to consider solutions based on mechanisms of oxygen avoidance used by cyanobacteria. The electron-transfer route to LFNR in chloroplasts involves light activation of photosystems II and I and the subsequent reduction of ferredoxin, which feeds electrons into LFNR. The reaction catalyzed by LFNR is reversible; whereas reduction by ferredoxin favors biosynthesis of NADPH (23), the reverse reaction uses NADPH to reduce oxidized ferredoxin (40). In this case, LFNR can be considered to be analogous to a battery, which accumulates reducing equivalents in the light and discharges this power (as NADPH) in the dark to fuel redox reactions (Fig. 5). Because the chloroplast ETC modules studied in this work function with photosystem I and can support nitrogenase activity in E. coli (Fig. 4), it is possible to envisage a scenario for energetic coupling of photosynthesis with nitrogen fixation, with temporal separation of their activities in the light and dark periods. In nonheterocystous cyanobacteria, nature has been able to reconcile the processes of nitrogen fixation and oxygenic photosynthesis through temporal separation, in which photosynthetic CO2 fixation occurs in the light and nitrogen fixation is carried out in the dark (41). According to our model, electron transfer from the ferredoxin to LFNR would occur during the light period, resulting in accumulation of excess NADPH. This reductant could be used in the dark to drive the reverse-catalytic reaction of LFNR, enabling reduction of nitrogenase by reduced ferredoxin (Fig. 5).

Materials and Methods

Bacterial Strains and Growth Medium.

Bacterial strains used in this study are listed in Table S2. Media and antibiotics were used as described (28) and are detailed in SI Materials and Methods.

Table S2.

Strains and plasmids used in this study

Strain or plasmid Relevant feature Source
Ec strains
 Top10 F mcrA Δ(mrr-hsdRMS-mcrBC) φ80lacZΔM15 ΔlacX74 nupG recA1 araD139 Δ(ara-leu)7697 galE15 galK16 rpsL(StrR) endA1 λ Invitrogen
 JM109 recA endA1 gyrA96 hsdR17 supE44 relA1 Δ(lac-proAB)/F′[traD36 proAB+ lacIq lacZΔM15] Takara Bio
Plasmids
 pBR322M pBR322 derivative, AmpR, labeled as ΔnifJ/ΔnifF in the text Ref. 28
 pBR322M-PLtetO-1 pBR322M derivative carrying PLtetO-1 inducible expression cassette This study
 pBR322M-Ptac pBR322M derivative carrying Ptac inducible expression cassette This study
 pKU7815 pACYC184 derivative carrying the minimal FeFe nitrogenase system, CmR Ref. 28
 pKU7017 pACYC184 derivative carrying the reassembled MoFe nitrogenase system, CmR Ref. 27
 pKU7830 pKU7815 derivative, ΔnifJ/ΔnifF This study
 pKU7831 pKU7017 derivative, ΔnifJ/ΔnifF This study
 pKU7832 pBR322M derivative carrying KonifJ/nifF genes, labeled as nifJ/nifF in the text This study
 pKU7833 pKU7830 derivative with nifF gene replaced by PLtetO-1 inducible expression cassette, labeled as nifJ/ΔnifF in the text This study
 pKU7834 pKU7833 derivative carrying nifJ/PLtetO-1-AsfdxHori This study
 pKU7835 pKU7833 derivative carrying nifJ/PLtetO-1-CrPETFsyn This study
 pKU7836 pKU7833 derivative carrying nifJ/PLtetO-1-AtFD2syn This study
 pKU7837 pKU7833 derivative carrying nifJ/PLtetO-1-ZmFDIsyn This study
 pKU7838 pKU7833 derivative carrying nifJ/PLtetO-1-OsFD1syn This study
 pKU7839 pKU7833 derivative carrying nifJ/PLtetO-1-TaFDsyn This study
 pKU7840 pKU7833 derivative carrying nifJ/PLtetO-1-AtFD3syn This study
 pKU7841 pKU7833 derivative carrying nifJ/PLtetO-1-ZmFDIIIsyn This study
 pKU7842 pKU7833 derivative carrying nifJ/PLtetO-1-OsFD4syn This study
 pKU7843 pKU7833 derivative carrying nifJ/PLtetO-1-AtMFD1syn This study
 pKU7844 pKU7833 derivative carrying nifJ/PLtetO-1-AtMFD2syn This study
 pKU7845 pEASY-Blunt derivative carrying EcgroES operon with its original promoter This study
 pKU7846 pKU7832 derivative with nifJ gene replaced by Ptac inducible expression cassette, labeled as ΔnifJ/nifF in the text This study
 pKU7847 pKU7846 derivative carrying Ptac-AspetHori/nifF This study
 pKU7848 pKU7846 derivative carrying Ptac-CrFNRsyn/nifF This study
 pKU7849 pKU7846 derivative carrying Ptac-ZmLFNRsyn/nifF This study
 pKU7850 pKU7846 derivative carrying Ptac-ZmRFNRsyn/nifF This study
 pKU7851 pKU7846 derivative carrying Ptac-AtMFDRsyn/nifF This study
 pKU7852 pKU7834 derivative with nifJ gene replaced by Ptac inducible expression cassette, labeled as ΔnifJ/PLtetO-1-AsfdxHori in the text This study
 pKU7853 pKU7847 derivative carrying Ptac-AspetHori/PLtetO-1-AsfdxHori This study
 pKU7854 pKU7848 derivative carrying Ptac-CrFNRsyn/PLtetO-1-AsfdxHori This study
 pKU7855 pKU7849 derivative carrying Ptac-ZmLFNRsyn/PLtetO-1-AsfdxHori This study
 pKU7856 pKU7850 derivative carrying Ptac-ZmRFNRsyn/PLtetO-1-AsfdxHori This study
 pKU7857 pKU7851 derivative carrying Ptac-AtMFDRsyn/PLtetO-1-AsfdxHori This study
 pKU7858 pKU7833 derivative carrying nifJ/PLtetO-1-AsfdxBori This study
 pKU7859 pKU7847 derivative carrying Ptac-AspetHori/PLtetO-1-AsfdxBori This study
 pKU7858 pKU7848 derivative carrying Ptac-CrFNRsyn/PLtetO-1-AsfdxBori This study
 pKU7859 pKU7849 derivative carrying Ptac-ZmLFNRsyn/PLtetO-1-AsfdxBori This study
 pKU7860 pKU7850 derivative carrying Ptac-ZmRFNRsyn/PLtetO-1-AsfdxBori This study
 pKU7861 pKU7851 derivative carrying Ptac-AtMFDRsyn/PLtetO-1-AsfdxBori This study
 pKU7862 pKU7848 derivative carrying Ptac-CrFNRsyn/PLtetO-1-CrPETFsyn This study
 pKU7863 pKU7849 derivative carrying Ptac-ZmLFNRsyn/PLtetO-1-ZmFDIsyn This study
 pKU7864 pKU7850 derivative carrying Ptac-ZmRFNRsyn/PLtetO-1-ZmFDIIIsyn This study
 pKU7865 pKU7851 derivative carrying Ptac-AtMFDRsyn/PLtetO-1-AtMFDsyn This study
 pKU7866 pKU7833 derivative carrying nifJ/PLtetO-1-AspetFori This study
 pKU7867 pKU7833 derivative carrying nifJ/PLtetO-1-CrFDX2syn This study
 pKU7868 pKU7833 derivative carrying nifJ/PLtetO-1-AtFD1syn This study
 pKU7869 pKU7833 derivative carrying nifJ/PLtetO-1-ZmFDIIsyn This study

Amp, ampicillin; Cm, chloramphenicol; Ec, E. coli; R, resistance; Δ, deletion. syn represents the codon optimized and synthetic gene sequence; ori represents original gene sequence amplified from the genome of its original host.

Construction of Recombinant Plasmids.

Plasmids used in this study are listed in Table S2 and Fig. S8. Plasmids were constructed by using procedures described in SI Materials and Methods. Each of the constructs was confirmed by DNA sequencing before any further experimentation.

Acetylene Reduction Assay.

The C2H2 reduction method was used to assay nitrogenase activity as described (28). Details are provided in SI Materials and Methods. Data presented are mean values based on at least three replicates.

15N2 Assimilation Assay.

To detect 15N2 assimilation, E. coli JM109 derivatives expressing reassembled nitrogenase systems were grown as described (28). More details are provided in SI Materials and Methods. Data presented are mean values based on at least two replicates.

Supplementary Material

Supplementary File
pnas.1620058114.sd01.pdf (56.5KB, pdf)

Acknowledgments

This work was supported by National Science Foundation of China (NSFC) Grant 31530081; 973 National Key Basic Research Program in China Grant 2015CB755700; China Postdoctoral Science Foundation Grant 2015M580014; and State Key Laboratory of Protein and Plant Gene Research Grant B02. Y.-P.W. is recipient of NSFC National Science Fund for Distinguished Young Scholars Grant 39925017. R.D. was supported by U.K. Biotechnology and Biological Sciences Research Council Grant BB/J004553/1. J.Y. was supported in part by the Postdoctoral Fellowship of Peking-Tsinghua Center for Life Sciences.

Footnotes

The authors declare no conflict of interest.

This article is a PNAS Direct Submission.

See Commentary on page 3009.

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1620058114/-/DCSupplemental.

References

  • 1.Xu G, Fan X, Miller AJ. Plant nitrogen assimilation and use efficiency. Annu Rev Plant Biol. 2012;63(1):153–182. doi: 10.1146/annurev-arplant-042811-105532. [DOI] [PubMed] [Google Scholar]
  • 2.Galloway JN, et al. Transformation of the nitrogen cycle: Recent trends, questions, and potential solutions. Science. 2008;320(5878):889–892. doi: 10.1126/science.1136674. [DOI] [PubMed] [Google Scholar]
  • 3.Tilman D, Balzer C, Hill J, Befort BL. Global food demand and the sustainable intensification of agriculture. Proc Natl Acad Sci USA. 2011;108(50):20260–20264. doi: 10.1073/pnas.1116437108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Canfield DE, Glazer AN, Falkowski PG. The evolution and future of Earth’s nitrogen cycle. Science. 2010;330(6001):192–196. doi: 10.1126/science.1186120. [DOI] [PubMed] [Google Scholar]
  • 5.Schindelin H, Kisker C, Schlessman JL, Howard JB, Rees DC. Structure of ADP x AIF4(-)-stabilized nitrogenase complex and its implications for signal transduction. Nature. 1997;387(6631):370–376. doi: 10.1038/387370a0. [DOI] [PubMed] [Google Scholar]
  • 6.Tezcan FA, et al. Nitrogenase complexes: Multiple docking sites for a nucleotide switch protein. Science. 2005;309(5739):1377–1380. doi: 10.1126/science.1115653. [DOI] [PubMed] [Google Scholar]
  • 7.Eady RR, Richardson TH, Miller RW, Hawkins M, Lowe DJ. The vanadium nitrogenase of Azotobacter chroococcum. Purification and properties of the Fe protein. Biochem J. 1988;256(1):189–196. doi: 10.1042/bj2560189. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Schneider K, Gollan U, Dröttboom M, Selsemeier-Voigt S, Müller A. Comparative biochemical characterization of the iron-only nitrogenase and the molybdenum nitrogenase from Rhodobacter capsulatus. Eur J Biochem. 1997;244(3):789–800. doi: 10.1111/j.1432-1033.1997.t01-1-00789.x. [DOI] [PubMed] [Google Scholar]
  • 9.Halbleib CM, Ludden PW. Regulation of biological nitrogen fixation. J Nutr. 2000;130(5):1081–1084. doi: 10.1093/jn/130.5.1081. [DOI] [PubMed] [Google Scholar]
  • 10.Duval S, et al. Electron transfer precedes ATP hydrolysis during nitrogenase catalysis. Proc Natl Acad Sci USA. 2013;110(41):16414–16419. doi: 10.1073/pnas.1311218110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Hageman RV, Burris RH. Nitrogenase and nitrogenase reductase associate and dissociate with each catalytic cycle. Proc Natl Acad Sci USA. 1978;75(6):2699–2702. doi: 10.1073/pnas.75.6.2699. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Hu Y, et al. Nitrogenase Fe protein: A molybdate/homocitrate insertase. Proc Natl Acad Sci USA. 2006;103(46):17125–17130. doi: 10.1073/pnas.0602651103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Shah VK, Stacey G, Brill WJ. Electron transport to nitrogenase. Purification and characterization of pyruvate:flavodoxin oxidoreductase. The nifJ gene product. J Biol Chem. 1983;258(19):12064–12068. [PubMed] [Google Scholar]
  • 14.Schmitz O, Gurke J, Bothe H. Molecular evidence for the aerobic expression of nifJ, encoding pyruvate:ferredoxin oxidoreductase, in cyanobacteria. FEMS Microbiol Lett. 2001;195(1):97–102. doi: 10.1111/j.1574-6968.2001.tb10504.x. [DOI] [PubMed] [Google Scholar]
  • 15.Yakunin AF, Hallenbeck PC. Purification and characterization of pyruvate oxidoreductase from the photosynthetic bacterium Rhodobacter capsulatus. Biochim Biophys Acta Bioenergetics. 1998;1409(1):39–49. doi: 10.1016/s0005-2728(98)00145-5. [DOI] [PubMed] [Google Scholar]
  • 16.Schmehl M, et al. Identification of a new class of nitrogen fixation genes in Rhodobacter capsulatus: A putative membrane complex involved in electron transport to nitrogenase. Mol Gen Genet. 1993;241(5-6):602–615. doi: 10.1007/BF00279903. [DOI] [PubMed] [Google Scholar]
  • 17.Neuer G, Bothe H. Electron donation to nitrogenase in heterocysts of cyanobacteria. Arch Microbiol. 1985;143(2):185–191. [Google Scholar]
  • 18.Cheng Q, Day A, Dowson-Day M, Shen G-F, Dixon R. The Klebsiella pneumoniae nitrogenase Fe protein gene (nifH) functionally substitutes for the chlL gene in Chlamydomonas reinhardtii. Biochem Biophys Res Commun. 2005;329(3):966–975. doi: 10.1016/j.bbrc.2005.02.064. [DOI] [PubMed] [Google Scholar]
  • 19.Ivleva NB, Groat J, Staub JM, Stephens M. Expression of active subunit of nitrogenase via integration into plant organelle genome. PLoS One. 2016;11(8):e0160951. doi: 10.1371/journal.pone.0160951. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.López-Torrejón G, et al. Expression of a functional oxygen-labile nitrogenase component in the mitochondrial matrix of aerobically grown yeast. Nat Commun. 2016;7:11426. doi: 10.1038/ncomms11426. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Hanke G, Mulo P. Plant type ferredoxins and ferredoxin-dependent metabolism. Plant Cell Environ. 2013;36(6):1071–1084. doi: 10.1111/pce.12046. [DOI] [PubMed] [Google Scholar]
  • 22.Fukuyama K. Structure and function of plant-type ferredoxins. Photosynth Res. 2004;81(3):289–301. doi: 10.1023/B:PRES.0000036882.19322.0a. [DOI] [PubMed] [Google Scholar]
  • 23.Kurisu G, et al. Structure of the electron transfer complex between ferredoxin and ferredoxin-NADP(+) reductase. Nat Struct Biol. 2001;8(2):117–121. doi: 10.1038/84097. [DOI] [PubMed] [Google Scholar]
  • 24.Knaff DB, Hirasawa M. Ferredoxin-dependent chloroplast enzymes. Biochim Biophys Acta Bioenergetics. 1991;1056(2):93–125. doi: 10.1016/s0005-2728(05)80277-4. [DOI] [PubMed] [Google Scholar]
  • 25.Green LS, et al. Ferredoxin and ferredoxin-NADP reductase from photosynthetic and nonphotosynthetic tissues of tomato. Plant Physiol. 1991;96(4):1207–1213. doi: 10.1104/pp.96.4.1207. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Picciocchi A, Douce R, Alban C. The plant biotin synthase reaction. Identification and characterization of essential mitochondrial accessory protein components. J Biol Chem. 2003;278(27):24966–24975. doi: 10.1074/jbc.M302154200. [DOI] [PubMed] [Google Scholar]
  • 27.Wang X, et al. Using synthetic biology to distinguish and overcome regulatory and functional barriers related to nitrogen fixation. PLoS One. 2013;8(7):e68677. doi: 10.1371/journal.pone.0068677. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Yang J, Xie X, Wang X, Dixon R, Wang Y-P. Reconstruction and minimal gene requirements for the alternative iron-only nitrogenase in Escherichia coli. Proc Natl Acad Sci USA. 2014;111(35):E3718–E3725. doi: 10.1073/pnas.1411185111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Masepohl B, Schölisch K, Görlitz K, Kutzki C, Böhme H. The heterocyst-specific fdxH gene product of the cyanobacterium Anabaena sp. PCC 7120 is important but not essential for nitrogen fixation. Mol Gen Genet. 1997;253(6):770–776. doi: 10.1007/s004380050383. [DOI] [PubMed] [Google Scholar]
  • 30.Sancho J. Flavodoxins: Sequence, folding, binding, function and beyond. Cell Mol Life Sci. 2006;63(7-8):855–864. doi: 10.1007/s00018-005-5514-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Pierella Karlusich JJ, Lodeyro AF, Carrillo N. The long goodbye: The rise and fall of flavodoxin during plant evolution. J Exp Bot. 2014;65(18):5161–5178. doi: 10.1093/jxb/eru273. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Wang L, et al. A minimal nitrogen fixation gene cluster from Paenibacillus sp. WLY78 enables expression of active nitrogenase in Escherichia coli. PLoS Genet. 2013;9(10):e1003865. doi: 10.1371/journal.pgen.1003865. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Li X-X, Liu Q, Liu X-M, Shi H-W, Chen S-F. Using synthetic biology to increase nitrogenase activity. Microb Cell Fact. 2016;15:43. doi: 10.1186/s12934-016-0442-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Deistung J, Thorneley RN. Electron transfer to nitrogenase. Characterization of flavodoxin from Azotobacter chroococcum and comparison of its redox potentials with those of flavodoxins from Azotobacter vinelandii and Klebsiella pneumoniae (nifF-gene product) Biochem J. 1986;239(1):69–75. doi: 10.1042/bj2390069. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Braaksma A, Haaker H, Grande HJ, Veeger C. The effect of the redox potential on the activity of the nitrogenase and on the Fe-protein of Azotobacter vinelandii. Eur J Biochem. 1982;121(3):483–491. doi: 10.1111/j.1432-1033.1982.tb05813.x. [DOI] [PubMed] [Google Scholar]
  • 36.Haaker H, Veeger C. Involvement of the cytoplasmic membrane in nitrogen fixation by Azotobacter vinelandii. Eur J Biochem. 1977;77(1):1–10. doi: 10.1111/j.1432-1033.1977.tb11634.x. [DOI] [PubMed] [Google Scholar]
  • 37.Buckel W, Thauer RK. Energy conservation via electron bifurcating ferredoxin reduction and proton/Na(+) translocating ferredoxin oxidation. Biochim Biophys Acta. 2013;1827(2):94–113. doi: 10.1016/j.bbabio.2012.07.002. [DOI] [PubMed] [Google Scholar]
  • 38.Edgren T, Nordlund S. The fixABCX genes in Rhodospirillum rubrum encode a putative membrane complex participating in electron transfer to nitrogenase. J Bacteriol. 2004;186(7):2052–2060. doi: 10.1128/JB.186.7.2052-2060.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Kameda H, Hirabayashi K, Wada K, Fukuyama K. Mapping of protein-protein interaction sites in the plant-type [2Fe-2S] ferredoxin. PLoS One. 2011;6(7):e21947. doi: 10.1371/journal.pone.0021947. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Onda Y, et al. Differential interaction of maize root ferredoxin:NADP(+) oxidoreductase with photosynthetic and non-photosynthetic ferredoxin isoproteins. Plant Physiol. 2000;123(3):1037–1045. doi: 10.1104/pp.123.3.1037. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Bothe H, Schmitz O, Yates MG, Newton WE. Nitrogen fixation and hydrogen metabolism in cyanobacteria. Microbiol Mol Biol Rev. 2010;74(4):529–551. doi: 10.1128/MMBR.00033-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Lam HM, Coschigano KT, Oliveira IC, Melo-Oliveira R, Coruzzi GM. The molecular-genetics of nitrogen assimilation into amino acids in higher plants. Annu Rev Plant Physiol Plant Mol Biol. 1996;47(1):569–593. doi: 10.1146/annurev.arplant.47.1.569. [DOI] [PubMed] [Google Scholar]
  • 43.Gatti-Lafranconi P, Dijkman WP, Devenish SR, Hollfelder F. A single mutation in the core domain of the lac repressor reduces leakiness. Microb Cell Fact. 2013;12:67. doi: 10.1186/1475-2859-12-67. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Montoya JP, Voss M, Kahler P, Capone DG. A simple, high-precision, high-sensitivity tracer assay for N(inf2) fixation. Appl Environ Microbiol. 1996;62(3):986–993. doi: 10.1128/aem.62.3.986-993.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Chen Y-J, et al. Characterization of 582 natural and synthetic terminators and quantification of their design constraints. Nat Methods. 2013;10(7):659–664. doi: 10.1038/nmeth.2515. [DOI] [PubMed] [Google Scholar]
  • 46.Hurley JK, et al. Structure-function relationships in Anabaena ferredoxin: Correlations between X-ray crystal structures, reduction potentials, and rate constants of electron transfer to ferredoxin:NADP+ reductase for site-specific ferredoxin mutants. Biochemistry. 1997;36(37):11100–11117. doi: 10.1021/bi9709001. [DOI] [PubMed] [Google Scholar]
  • 47.Terauchi AM, et al. Pattern of expression and substrate specificity of chloroplast ferredoxins from Chlamydomonas reinhardtii. J Biol Chem. 2009;284(38):25867–25878. doi: 10.1074/jbc.M109.023622. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Matsumura T, et al. Complementary DNA cloning and characterization of ferredoxin localized in bundle-sheath cells of maize leaves. Plant Physiol. 1999;119(2):481–488. doi: 10.1104/pp.119.2.481. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Akashi T, et al. Comparison of the electrostatic binding sites on the surface of ferredoxin for two ferredoxin-dependent enzymes, ferredoxin-NADP(+) reductase and sulfite reductase. J Biol Chem. 1999;274(41):29399–29405. doi: 10.1074/jbc.274.41.29399. [DOI] [PubMed] [Google Scholar]
  • 50.Hanke GT, Kimata-Ariga Y, Taniguchi I, Hase T. A post genomic characterization of Arabidopsis ferredoxins. Plant Physiol. 2004;134(1):255–264. doi: 10.1104/pp.103.032755. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary File
pnas.1620058114.sd01.pdf (56.5KB, pdf)

Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES