ACS Publications. Most Trusted. Most Cited. Most Read
Metabolic Pathway Rerouting in Paraburkholderia rhizoxinica Evolved Long-Overlooked Derivatives of Coenzyme F420
My Activity

Figure 1Loading Img
  • Open Access
Articles

Metabolic Pathway Rerouting in Paraburkholderia rhizoxinica Evolved Long-Overlooked Derivatives of Coenzyme F420
Click to copy article linkArticle link copied!

  • Daniel Braga
    Daniel Braga
    Junior Research Group Synthetic Microbiology, Leibniz Institute for Natural Product Research and Infection Biology − Hans Knöll Institute, Beutenbergstr. 11a, 07745 Jena, Germany
    Friedrich Schiller University, Jena, Germany
    More by Daniel Braga
  • Daniel Last
    Daniel Last
    Junior Research Group Synthetic Microbiology, Leibniz Institute for Natural Product Research and Infection Biology − Hans Knöll Institute, Beutenbergstr. 11a, 07745 Jena, Germany
    More by Daniel Last
  • Mahmudul Hasan
    Mahmudul Hasan
    Junior Research Group Synthetic Microbiology, Leibniz Institute for Natural Product Research and Infection Biology − Hans Knöll Institute, Beutenbergstr. 11a, 07745 Jena, Germany
    Friedrich Schiller University, Jena, Germany
  • Huijuan Guo
    Huijuan Guo
    Junior Research Group, Chemical Biology of Microbe−Host Interactions, Leibniz Institute for Natural Product Research and Infection Biology − Hans Knöll Institute, Beutenbergstr. 11a, 07745 Jena, Germany
    More by Huijuan Guo
  • Daniel Leichnitz
    Daniel Leichnitz
    Junior Research Group, Chemical Biology of Microbe−Host Interactions, Leibniz Institute for Natural Product Research and Infection Biology − Hans Knöll Institute, Beutenbergstr. 11a, 07745 Jena, Germany
  • Zerrin Uzum
    Zerrin Uzum
    Department of Biomolecular Chemistry, Leibniz Institute for Natural Product Research and Infection Biology − Hans Knöll Institute, Beutenbergstr. 11a, 07745 Jena, Germany
    More by Zerrin Uzum
  • Ingrid Richter
    Ingrid Richter
    Department of Biomolecular Chemistry, Leibniz Institute for Natural Product Research and Infection Biology − Hans Knöll Institute, Beutenbergstr. 11a, 07745 Jena, Germany
  • Felix Schalk
    Felix Schalk
    Junior Research Group, Chemical Biology of Microbe−Host Interactions, Leibniz Institute for Natural Product Research and Infection Biology − Hans Knöll Institute, Beutenbergstr. 11a, 07745 Jena, Germany
    More by Felix Schalk
  • Christine Beemelmanns
    Christine Beemelmanns
    Junior Research Group, Chemical Biology of Microbe−Host Interactions, Leibniz Institute for Natural Product Research and Infection Biology − Hans Knöll Institute, Beutenbergstr. 11a, 07745 Jena, Germany
  • Christian Hertweck
    Christian Hertweck
    Department of Biomolecular Chemistry, Leibniz Institute for Natural Product Research and Infection Biology − Hans Knöll Institute, Beutenbergstr. 11a, 07745 Jena, Germany
    Friedrich Schiller University, Jena, Germany
  • Gerald Lackner*
    Gerald Lackner
    Junior Research Group Synthetic Microbiology, Leibniz Institute for Natural Product Research and Infection Biology − Hans Knöll Institute, Beutenbergstr. 11a, 07745 Jena, Germany
    Friedrich Schiller University, Jena, Germany
    *Phone: +49 3641 532 1104. E-mail: [email protected]
Open PDFSupporting Information (1)

ACS Chemical Biology

Cite this: ACS Chem. Biol. 2019, 14, 9, 2088–2094
Click to copy citationCitation copied!
https://doi.org/10.1021/acschembio.9b00605
Published August 30, 2019

Copyright © 2019 American Chemical Society. This publication is licensed under CC-BY-NC-ND.

Abstract

Click to copy section linkSection link copied!

Coenzyme F420 is a specialized redox cofactor with a negative redox potential. It supports biochemical processes like methanogenesis, degradation of xenobiotics, and the biosynthesis of antibiotics. Although well-studied in methanogenic archaea and actinobacteria, not much is known about F420 in Gram-negative bacteria. Genome sequencing revealed F420 biosynthetic genes in the Gram-negative, endofungal bacterium Paraburkholderia rhizoxinica, a symbiont of phytopathogenic fungi. Fluorescence microscopy, high-resolution LC-MS, and structure elucidation by NMR demonstrated that the encoded pathway is active and yields unexpected derivatives of F420 (3PG-F420). Further analyses of a biogas-producing microbial community showed that these derivatives are more widespread in nature. Genetic and biochemical studies of their biosynthesis established that a specificity switch in the guanylyltransferase CofC reprogrammed the pathway to start from 3-phospho-d-glycerate, suggesting a rerouting event during the evolution of F420 biosynthesis. Furthermore, the cofactor activity of 3PG-F420 was validated, thus opening up perspectives for its use in biocatalysis. The 3PG-F420 biosynthetic gene cluster is fully functional in Escherichia coli, enabling convenient production of the cofactor by fermentation.

Copyright © 2019 American Chemical Society
Cofactors are essential for the catalytic power of many enzymes and thus play a key role in virtually all metabolic pathways. Knowledge of their catalytic functions and biosynthesis is highly important for the understanding of biochemical reactions as well as their application in biocatalysis and biotechnology. An important subclass comprises redox cofactors that mediate electron transfer between molecules. The deazaflavin coenzyme F420 (Figure 1) is a specialized redox cofactor with a lower redox potential (ca. −340 mV) than NAD. (1,2) F420 is a prominent electron carrier in methanogenic archaea. Additionally, F420 has attracted considerable interest as a cofactor of the human pathogen Mycobacterium tuberculosis, being involved in respiration, nitrosative stress response, (3) or prodrug-activation. (4) In filamentous actinomycetes, coenzyme F420 is involved in secondary metabolism, e.g., the biosynthesis of antibiotics like oxytetracycline, (5) pyrrolobenzodiazepines, (6) or thiopeptins. (7) Its low redox potential renders coenzyme F420 a strong reducing agent for challenging reactions in biocatalysis. (2) Enzymatic processes involving F420 facilitate the degradation of pollutants like aromatic nitro compounds (8) or the carcinogen aflatoxin. (9) Furthermore, F420-dependent enzymes are important for asymmetric ene reductions. (10,11)

Figure 1

Figure 1. Biosynthesis of coenzyme F420. (A) FO synthase FbiC (in archaea: CofG/H) catalyzes formation of the deazaflavin ring from tyrosine and 5-amino-6-(ribitylamino)-uracil, an intermediate of riboflavin biosynthesis. (B) Biosynthetic scheme of F420-n starting from 2-PL: CofC and CofD catalyze the activation of 2-PL and transfer of the 2-PL moiety, respectively. CofE performs (oligo-)γ-glutamylation. The number of glutamate residues (n) varies depending on the organism. Enzymes producing 2-PL are elusive, and it has been questioned that 2-PL is an intermediate of F420 biosynthesis. (C) Biosynthesis of F420-n starting from PEP: CofC and CofD activate PEP, resulting in DF420 formation. The C-terminal domain of FbiB reduces DF420 to F420. A pathway starting from 3-phosphoglycerate was established in this study (Figure 4B). EPPG: enolpyruvyl-diphosphoguanosine. LPPG: lactyl-diphosphoguanosine. 2-PL: 2-phospho-l-lactate.

A key step during the biosynthesis of F420 is the formation of the deazaflavin fluorophore FO (Figure 1A), a stable metabolic precursor of F420 originating from tyrosine and an intermediate of riboflavin biosynthesis. This chemically challenging step is catalyzed by the radical SAM enzyme complex CofG/H in archaea or the homologous dual-domain protein FbiC in actinobacteria. (12) FO is then further processed by CofC (EC 2.7.7.68) and CofD (EC 2.7.8.28). This pair of enzymes is responsible for the biosynthesis of the formal phospholactate moiety of the F420 molecule (Figure 1B). Previous studies have shown that CofC from Methanocaldococcus jannaschii directly activates 2-phospho-l-lactate (2-PL) by guanylylation, (13) resulting in the formation of the short-lived metabolite lactyl-diphosphoguanosine (LPPG). CofD then forms F420-0 by transfer of the 2-PL moiety from LPPG to FO. (14) The enzymes producing 2-PL, however, have remained elusive in all F420 producers so far. Recently, Bashiri et al. proposed a revised biosynthetic pathway by demonstrating that phosphoenolpyruvate (PEP) instead of 2-PL can serve as a substrate of CofC in mycobacteria and presented evidence that CofC from Methanosarcina mazei accepts PEP as well. (15) The resulting dehydro-F420(DF420) is reduced to F420 by a flavin-dependent reductase domain present in the FbiB protein (Figure 1C). In mycobacteria, FbiB is a dual-domain protein consisting of a γ-glutamyl ligase domain (CofE-like, EC 6.3.2.31) and the C-terminal DF420 reductase domain. (16) The γ-glutamyl ligase CofE finally decorates F420-0 with a varying number of (oligo-)γ-glutamate residues. (17)
F420 is not ubiquitous in prokaryotes, but it is associated with certain phyla. (18,19) First discovered in methanogenic archaea, (20,21) it was extensively studied as a potential drug target of pathogenic mycobacteria (3) or as a cofactor enabling antibiotics biosynthesis in streptomycetes. (2) Genome sequencing as well as fluorescence microscopy and analytical chemistry revealed that some Gram-negative bacteria have acquired F420 genes by horizontal transfer. (19,22) However, virtually nothing is known about the biosynthesis and role of F420 in these organisms. By genome mining, we found a biosynthetic gene cluster (BGC) homologous to those previously implicated in the biosynthesis of F420 in the endofungal bacterium Paraburkholderia rhizoxinica HKI 454 (Figure 2A and Supporting Information Table S3). This organism is an intracellular endosymbiont of the phytopathogenic fungus Rhizopus microsporus supplying its host with antimitotic toxins that act as virulence factors during infection of rice plants. (23−25) We hypothesized that genes from Gram-negative bacteria related to F420 biosynthesis could facilitate F420 production in E. coli or could reveal novel biosynthetic routes toward this molecule. Therefore, we set out to investigate if the BGC is active and if it can be refactored to produce F420 in E. coli.

Figure 2

Figure 2. Deazaflavin biosynthesis in P. rhizoxinica. (A) BGC of 3PG-F420. Core genes are shown in dark gray. (B–I) Microscopy photographs depict fluorescence characteristic of deazaflavins in blue. (B–D) Axenic M. smegmatis (B), P. rhizoxinica (C), and E. coli/pDB045 (D). In R. microsporus ATCC 62417, deazaflavins are correlated to the presence of P. rhizoxinica symbionts (green, Syto9 staining; E). No fluorescence was detected in cured ATCC 62417 mycelium (F) or in the naturally symbiont-free strain CBS 344.29 (G). The same pattern was observed in spores of either wild-type ATCC 62417 (H) or CBS 344.29 (I). Scale bars represent 10 μm. (J) Refactored versions of the BGC and corresponding plasmids for heterologous expression in E. coli. Asterisks mark genes from M. jannaschii.

Here, we show that P. rhizoxinica produces unexpected F420 derivatives (3PG-F420) both in symbiosis as well as in axenic culture. Heterologous expression and large-scale production in E. coli allowed for elucidation of their chemical structure. By comparative analyses, we discovered related metabolites in a biogas-producing microbial community, thus indicating their broader abundance and relevance. Enzyme assays showed that a switch in substrate specificity of CofC is responsible for the biosynthesis of 3PG-F420 and proved that it can serve as a substitute for F420 in biochemical reactions.

Results and Discussion

Click to copy section linkSection link copied!

Discovery and Structure of 3PG-F420

To test whether P. rhizoxinica is capable of producing deazaflavins, we investigated axenic cultures of symbiont (P. rhizoxinica) and host (R. microsporus) as well as symbiotic cultures by fluorescence microscopy (Figure 2B–I and Supporting Information Figures S1–S4). Indeed, symbiotic P. rhizoxinica and the cytosol of colonized mycelia emitted strong fluorescence characteristic of deazaflavins (excitation at approximately 420 nm, emission at 470 nm). Notably, even fluorescence of bacteria present inside of fungal spores was observed. Axenic fungi, however, showed no fluorescence, whereas only low signals were measured from axenic bacteria under the same conditions. To corroborate the results obtained by microscopy, we extracted metabolites from axenic P. rhizoxinica and axenic R. microsporus as well as a fungal host containing endosymbionts and analyzed the extracts by LC-MS/MS. To our surprise, only FO could be detected in axenic P. rhizoxinica and in the fungal host containing endosymbionts, but none of the expected F420-n species.
To further test the biosynthetic capacity of the full BGC, we refactored it to obtain a single operon under the control of a T7 promoter for heterologous expression yielding E. coli/pDB045 (Figure 2J). Examination of transformed bacteria by fluorescence microscopy revealed strong fluorescence as in the endofungal bacteria (Figure 2D), but LC-MS analyses again yielded only mass traces of FO (mainly found in the culture supernatant), but not of F420-n. Since the impeded production could be attributed to nonfunctional proteins, we analyzed proteins by SDS-PAGE and found that CofD was poorly soluble in E. coli. Replacement of the cofD gene with the corresponding M. jannaschii homologue (14) provided soluble protein (E. coli/pDB060); however, again no trace of F420 could be detected. Therefore, we re-examined the metabolome of E. coli/pDB045 for characteristic MS/MS fragments (Figure 3B) derived from the FO moiety (m/z 230.06, 346.10, 364.11). Surprisingly, the analysis revealed spectra with a similar fragmentation pattern, yet derived from precursor ions that had a mass shift of +15.995 compared to F420, indicating the presence of an additional oxygen atom (Figure 3A,B and Supporting Information Figures S5–S8). Further analyses revealed an (oligo-)γ-glutamate series of the oxygenated compound suggesting these species are congeners. According to MS/MS fragmentation, the additional oxygen was present in the “phospholactyl” moiety of F420-n, thus forming a “phosphoglyceryl” moiety. Extensive 1D- and 2D-NMR experiments (Figure 3C and Supporting Information Section 2.3) and comparison to classical F420 (20) corroborated that this moiety corresponds to 3-phosphoglycerate (3-PG). Therefore, we named the molecules 3PG-F420. This finding was unexpected because 2-phosphoglycerate is structurally more similar to PEP and 2-PL than 3-PG. Chemical degradation followed by chiral UHPLC-MS finally substantiated that the additional stereocenter of 3PG-F420 is R-configured (Figure 3D and Supporting Information Figure S51). Large-scale cultivation of E. coli/pDB045 also revealed traces of dehydro-F420 (DF420), but the yields were too low for NMR studies. The structure and occurrence of 3PG-F420 have not been reported before. To date, the only known derivatives of F420-n are factor F390-A and F390-G, 8-OH-AMP and 8-OH-GMP esters of F420, respectively. (26) In methanogens, they are formed reversibly, e.g., during oxygen exposure, acting as a reporter compound for hydrogen starvation. (27) In contrast, the modifications seen in 3PG-F420 are not temporary. Rather, 3PG-F420 seems to replace F420 as a natural deazaflavin-cofactor in P. rhizoxinica. At least in this organism, it does not coexist with classical F420. This situation is reminiscent of mycothiol, a specialized thiol cofactor that replaced glutathione in actinobacteria. (28)

Figure 3

Figure 3. Chemical analysis of 3PG-F420. (A) Extracted ion chromatograms of 3PG-F420-2 produced in E. coli. I: E. coli/pDB045. II: cofD exchanged by M. jannaschii homologue (pDB060). III: cofD and cofC exchanged by M. jannaschii homologues (pDB070). IV: empty vector (pETDuet). (B) Excerpt of the MS/MS spectrum of 3PG-F420-2. Gray bars highlight m/z used for fragment ion search of F420 derivatives. (C) 1H NMR comparison of F420-n (D2O), 3PG-F420-n (0.1% ND3 in D2O), and 3PG-F420-0 (0.1% ND3 in D2O) indicated the replacement of the lactyl moiety in F420 with a glyceryl moiety in 3PG-F420. (D) Proposed structures of 3PG-F420-0, 3PG-F420-n, and DF420-n.

Occurrence of 3PG-F420 in Nature

To investigate if 3PG-F420 is produced by wild-type P. rhizoxinica, we reanalyzed LC-MS data for the presence of corresponding mass signals. Indeed, 3PG-F420 species were found in samples containing bacteria (axenic culture, symbiosis) but not in symbiont-free host mycelia (Supporting Information Figures S9–S14). In extracts of P. rhizoxinica, no traces of F420 and DF420 were detected. The presence of 3PG-F420 was restricted to the cell pellet, whereas FO was abundant in culture supernatants. We thus conclude that the fluorescence observed in bacterial cells of P. rhizoxinica is derived from 3PG-F420 and FO. So far, there is evidence for the occurrence of deazaflavin cofactors in a few Gram-negative bacteria, e.g., Oligotropha carboxidovorans and Paracoccus denitrificans, (19) as well as in the uncultured, but biosynthetically highly prolific “Candidatus Entotheonella factor.” (22) The exact structure and function of F420 in most Gram-negative bacteria that harbor corresponding biosynthetic genes, however, is unknown. The fact that P. rhizoxinica produced 3PG-F420 under symbiotic cultivation conditions allows for the conclusion that it provides a fitness benefit in its natural habitat. Notably, none of the well-characterized F420-dependent enzyme families (9,18) are encoded in the P. rhizoxinica genome according to BLAST and conserved domains searches. Not even any of the widespread regeneration systems like Fno or F420-dependent glucose-6-phosphate dehydrogenase could be identified. Therefore, future investigations of 3PG-F420-producing organisms are likely to reveal novel enzymes, regeneration systems, and cellular pathways depending on this cofactor.
To assess if 3PG-F420 is restricted to fungal endosymbionts or if it might be more widespread in the environment, we examined M. jannaschii and M. smegmatis for the presence of any F420 congeners. Only F420-n was found from extracts of M. jannaschii, while F420-n and DF420-n were detected in extracts of M. smegmatis (Supporting Information Figures S15, S16). None of the 3PG-F420 derivatives were found in the reference organisms. Since methanogens are a common source of F420 in nature, we analyzed (two independent) sludge samples from a local biogas production plant. To our surprise, extraction of the microbial community present in the biogas-producing sludge followed by LC-MS/MS eluted, besides classical F420, a compound with an identical retention time, the exact mass, and the same MS/MS fragmentation pattern to that of 3PG-F420 (Supporting Information Figure S17). As fluorescence and UV-based detection do usually not resolve classical F420 and 3PG-F420, these derivatives might have been misidentified as F420 in the past. Since neither P. rhizoxinica nor its host R. microsporus are able to grow under anaerobic and thermophilic (temperatures >42 °C) conditions, they can be excluded as the source of these cofactors. The high complexity of biogas-producing microbiomes (29) does not allow for an educated guess of the producer.

Biosynthesis of 3PG-F420

In order to rationalize how the biosynthetic pathway was redirected to form 3PG-F420 instead of F420, we examined key steps of the biosynthesis more closely. We observed that production of 3PG-F420 was not abolished by the exchange of cofD from P. rhizoxinica by cofD from M. jannaschii (plasmid pDB060). Hence, the phospholactyl transferase CofD could not be held accountable for the switch toward 3PG-F420. According to the existing biosynthetic model, the most plausible scenario was that CofC incorporated 3-phospho-d-glycerate (3-PG), an intermediate of glycolysis, instead of 2-PL to form 3PG-F420-0 and, to a minor extent, PEP to form DF420-0. To test this hypothesis, we exchanged cofC and cofD for the corresponding M. jannaschii homologues. The resulting strain E. coli/pDB070 (Figure 2J and Supporting Information Figure S18) produced neither 3PG-F420 nor F420 but traces of DF420. To further investigate the substrate specificity of CofC, we performed an in vitro assay using CofC and CofD. (13) Genes cofC of P. rhizoxinica as well as cofC and cofD from M. jannaschii were cloned, corresponding proteins produced as hexahistidine fusions in E. coli and purified by metal affinity chromatography for in vitro assays. The physiologically relevant isomers 2-phospho-d-glycerate (2-PG) and 3-phospho-d-glycerate (3-PG), as well as PEP and 2-PL, served as substrates. Reaction products were monitored by LC-MS. Indeed, when CofC from P. rhizoxinica was tested, the mass of 3PG-F420-0 appeared after reaction with d-3-PG eluting at the same retention time as the in vivo product (Supporting Information Section 2.4). In addition, the formation of DF420-0 and F420-0 was detected, when the enzymes were incubated with PEP and 2-PL, respectively. In contrast, reaction with 2-PG yielded mass signals close to the noise level. Controls lacking CofC did not generate any of these products. In a direct substrate competition assay (Figure 4A), 3-PG was found to be the preferred substrate with a relative turnover of ca. 73% (2-PL, 23%; PEP, 4%). This finding is in agreement with the structure of 3PG-F420 and the occurrence of DF420 as a minor biosynthetic product in E. coli. Note that CofC from M. jannaschii displayed a strong turnover of 2-PL (96.5%), weak turnover of PEP (3.5%), and no turnover of 3-PG. This finding supports the notion that the CofC of P. rhizoxinica has undergone a substrate specificity switch during evolution.

Figure 4

Figure 4. Combined CofC/D in vitro assay. (A) Relative turnover of substrates estimated from a substrate competition assay (d-3-PG, 2-PL, and PEP). CofC from P. rhizoxinica accepted 3-PG (72.7%), 2-PL (23.4%), and PEP (3.9%). CofC from M. jannaschii preferred 2-PL (96.5%) and PEP (3.5%). 3-PG was not turned over. CofD from M. jannaschii was used in all assays. Error bars represent the standard deviation (SD) of three independent biological replicates (N = 3). (B) Proposed model of 3PG-F420 biosynthesis. 3-GPPG: 3-(guanosine-5′-disphospho)-d-glycerate.

Recently, Bashiri et al. claimed that PEP is the substrate of CofC in prokaryotes. (15) Our results confirm the hypothesis that PEP is the physiological substrate in mycobacteria, since we observed turnover of PEP by all CofC homologues tested. However, in contrast to Bashiri et al., 2-PL was the best substrate of M. jannaschii CofC in our assay. Since Graupner and White detected significant amounts of 2-PL in methanogenic archaea and observed the conversion of lactate into 2-PL by isotope labeling, (30) we conclude that 2-PL might still be a relevant substrate in archaea. Further investigations into archaeal F420 biosynthesis are warranted to account for these discrepancies. From a phylogenetic perspective, our results suggest that multiple metabolic rewiring events occurred in the evolution of F420 biosynthesis. While actinobacteria evolved the DF420 reductase (C-terminal domain of FbiB), archaea accomplished production of the (unusual) metabolite 2-PL. Other organisms, as exemplified by P. rhizoxinca, rerouted the biosynthesis to the ubiquitous metabolite 3-PG.
To address the question if the γ-glutamyl ligase CofE adapted its substrate specificity to 3PG-F420, we individually coexpressed cofE genes from P. rhizoxinica, M. jannaschii, and M. smegmatis (fbiB) together with a minimal BGC consisting of fbiC, cofC, and cofD in a two-plasmid system. Extraction of metabolites and LC-MS/MS revealed that all three CofE homologues elongated 3PG-F420-0 to oligo-glutamate chain lengths up to n = 6 (Supporting Information Figures S63–S65). Thus, we conclude that CofE does not act as an additional specificity filter during chain elongation of 3PG-F420.

Cofactor Role of 3PG-F420

The successful isolation of 3PG-F420 and reconstitution of its biosynthesis in E. coli motivated us to address the question of whether 3PG-F420 could substitute F420 in biocatalysis. To this end, we cloned a gene encoding Fno (F420:NADPH oxidoreductase), an enzyme that serves as a regeneration system for F420H2 using NADPH/H+ as an electron donor. (31) We first examined if Fno can accept 3PG-F420 as a substrate. Indeed, we observed an efficient reduction of 3PG-F420 by recombinant Fno as mirrored by a rapid decrease of characteristic UV absorption. An examination of kinetic parameters (Figure 5A) revealed that the apparent KM of Fno for F420 was 3.6 ± 0.7 μM. This value is similar to the reported KM of 10 μM. (32) Under identical assay conditions, the KM for 3PG-F420 was only slightly higher (5.1 ± 1.0 μM). The vmax values were in a similar range as well (F420, 1.3 ± 0.2 μM min–1; 3PG-F420, 0.88 ± 0.07 μM min–1) pointing toward only a minor reduction of maximal turnover. These findings are in line with recent studies on classical F420, demonstrating that the side chain modulates cofactor binding and in turn substrate turnover. (33,34) Encouraged by the finding that 3PG-F420 can substitute F420, we aimed at an in vivo application of the cofactor for malachite green reduction as a proof of principle. To this end, we combined the fno gene with a minimal BGC producing 3PG-F420-0 (Figure 5C and Supporting Information Figure S62) on a single vector (pDB071). Additionally, the F420-dependent malachite green reductase gene MSMEG_5998 (35) from M. smegmatis was cloned and expressed from a compatible vector backbone (pDB061). Finally, coexpression of all components in E. coli yielded a strain (pDB061/pDB071) that was able to decolorize malachite green significantly faster than control strains expressing the reductase or the cofactor alone (Figure 5B). Thus, we conclude that 3PG-F420 can substitute F420 as a redox cofactor in this case. The production of classical F420 and its use for biotransformations in E. coli has just recently been achieved in moderate yields using Mycobacterium genes including the DF420 reductase domain. (15)

Figure 5

Figure 5. Cofactor function of 3PG-F420. (A) Michaelis–Menten kinetics of Fno for F420 (left) and 3PG-F420 as substrates (right). Three biological replicates were used to determine parameters. KM for F420 was 3.6 ± 0.7 μM (standard error). KM for 3PG-F420 was 5.1 ± 1.0 μM. Error bars indicate standard deviation of replicates (N = 3). (B) In vivo reduction of malachite green (absorbance: 618 nm) by the F420-dependent reductase MSMEG_5998. Fno was used to regenerate 3PG-F420H2. Left panel: Time course of the malachite green depletion assay. Right panel: Bar chart of residual malachite green after 20 h; wt, E. coli BL21(DE3); pDB061, E. coli producing MSMEG_5998; pDB071, E. coli producing 3PG-F420-0 + Fno. Exact means ± SD of biological triplicates were 0.234 ± 0.017 (wt), 0.124 ± 0.003 (pDB061), 0.169 ± 0.011 (pDB071), and 0.082 ± 0.0139 (pDB061/pDB071). An asterisk indicates statistical significance (one-way ANOVA, p < 0.05, N = 3). (C) Engineered E. coli combining 3PG-F420, Fno, and reductase MSMEG_5998 (red) for reduction of malachite green.

In summary, we discovered a derivative of the redox cofactor F420 that is produced by the Gram-negative endofungal bacterium P. rhizoxinica. We fully elucidated its chemical structure and show its potential cofactor function. Thus, our work is a solid basis to unveil unknown enzyme families and bioprocesses depending on 3PG-F420. Intriguingly, its presence in a biogas-producing digester suggests that the cofactor is more widespread in nature than expected. Furthermore, we could demonstrate that the guanylyltransferase CofC is responsible for the biosynthetic switch leading to the production of 3PG-F420. Our results thus significantly refine and extend the biosynthetic pathway models to deazaflavin cofactors in several phyla. Notably, the pathway discovered here offered an alternative route to heterologous production and reconstitution of F420-dependent bioprocesses in E. coli. In recent years, there has been increasing interest in F420-dependent enzymes for biocatalysis. (10,11,36,37) Future applications will comprise, for instance, enantioselective biotransformations or the creation of a universal expression host for the production of antibiotics and other high-value compounds.

Methods

Click to copy section linkSection link copied!

Materials and methods are summarized in the Supporting Information (section 1).

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acschembio.9b00605.

  • Methods section, supporting figures, data, and discussion (fluorescence microscopy, mass spectrometry, NMR, enzyme assays) (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Author
    • Gerald Lackner - Junior Research Group Synthetic Microbiology, Leibniz Institute for Natural Product Research and Infection Biology − Hans Knöll Institute, Beutenbergstr. 11a, 07745 Jena, GermanyFriedrich Schiller University, Jena, GermanyOrcidhttp://orcid.org/0000-0002-0307-8319 Email: [email protected]
  • Authors
    • Daniel Braga - Junior Research Group Synthetic Microbiology, Leibniz Institute for Natural Product Research and Infection Biology − Hans Knöll Institute, Beutenbergstr. 11a, 07745 Jena, GermanyFriedrich Schiller University, Jena, Germany
    • Daniel Last - Junior Research Group Synthetic Microbiology, Leibniz Institute for Natural Product Research and Infection Biology − Hans Knöll Institute, Beutenbergstr. 11a, 07745 Jena, Germany
    • Mahmudul Hasan - Junior Research Group Synthetic Microbiology, Leibniz Institute for Natural Product Research and Infection Biology − Hans Knöll Institute, Beutenbergstr. 11a, 07745 Jena, GermanyFriedrich Schiller University, Jena, Germany
    • Huijuan Guo - Junior Research Group, Chemical Biology of Microbe−Host Interactions, Leibniz Institute for Natural Product Research and Infection Biology − Hans Knöll Institute, Beutenbergstr. 11a, 07745 Jena, Germany
    • Daniel Leichnitz - Junior Research Group, Chemical Biology of Microbe−Host Interactions, Leibniz Institute for Natural Product Research and Infection Biology − Hans Knöll Institute, Beutenbergstr. 11a, 07745 Jena, Germany
    • Zerrin Uzum - Department of Biomolecular Chemistry, Leibniz Institute for Natural Product Research and Infection Biology − Hans Knöll Institute, Beutenbergstr. 11a, 07745 Jena, Germany
    • Ingrid Richter - Department of Biomolecular Chemistry, Leibniz Institute for Natural Product Research and Infection Biology − Hans Knöll Institute, Beutenbergstr. 11a, 07745 Jena, Germany
    • Felix Schalk - Junior Research Group, Chemical Biology of Microbe−Host Interactions, Leibniz Institute for Natural Product Research and Infection Biology − Hans Knöll Institute, Beutenbergstr. 11a, 07745 Jena, Germany
    • Christine Beemelmanns - Junior Research Group, Chemical Biology of Microbe−Host Interactions, Leibniz Institute for Natural Product Research and Infection Biology − Hans Knöll Institute, Beutenbergstr. 11a, 07745 Jena, Germany
    • Christian Hertweck - Department of Biomolecular Chemistry, Leibniz Institute for Natural Product Research and Infection Biology − Hans Knöll Institute, Beutenbergstr. 11a, 07745 Jena, GermanyFriedrich Schiller University, Jena, GermanyOrcidhttp://orcid.org/0000-0002-0367-337X
  • Author Contributions

    D.B. performed research, analyzed data (molecular biology, malachite green assay, mass spectrometry), and contributed to writing the manuscript. D.L. performed research and analyzed data (structure elucidation, Fno assay, biogas plant studies). M.H. performed research (CofC/D enzyme assays). H.G. performed research and analyzed data (structure elucidation). D.L. performed research (chemical synthesis). Z.U. performed research (microscopy). I.R. performed research (microscopy). F.S. performed research (cofE constructs). C.B. designed research, acquired funding, analyzed data (structure elucidation, synthesis), and edited the manuscript. C.H. designed research, acquired funding, and edited the manuscript. G.L. designed the study, acquired funding, and wrote the original manuscript.

  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

We thank D. Grohmann and G. Bashiri for kindly providing M. jannaschii and M. smegmatis/fbiABC, respectively. We thank Biogas Jena GmbH and Co. KG for the kind donation of biogas plant samples. We thank H. Heinecke for conducting NMR experiments. G.L. thanks the Deutsche Forschungsgemeinschaft (DFG Grant LA 4424/1-1) and the Carl Zeiss Foundation for funding. Financial support from the DFG (CRC 1127 ChemBioSys) to C.H. and C.B. and BE-4799/2-1 to C.B., a Leibniz Award to C.H., support by the ERC (MSCA-IF-EF-RI Project 794343, to I.R.), and support from the JSMC to Z.U. are gratefully acknowledged.

References

Click to copy section linkSection link copied!

This article references 37 other publications.

  1. 1
    Jacobson, F. and Walsh, C. (1984) Properties of 7,8-didemethyl-8-hydroxy-5-deazaflavins relevant to redox coenzyme function in methanogen metabolism. Biochemistry 23, 979988,  DOI: 10.1021/bi00300a028
  2. 2
    Greening, C., Ahmed, F. H., Mohamed, A. E., Lee, B. M., Pandey, G., Warden, A. C., Scott, C., Oakeshott, J. G., Taylor, M. C., and Jackson, C. J. (2016) Physiology, biochemistry, and applications of F420- and Fo-dependent redox reactions. Microbiol. Mol. Biol. Rev. 80, 451493,  DOI: 10.1128/MMBR.00070-15
  3. 3
    Purwantini, E. and Mukhopadhyay, B. (2009) Conversion of NO2 to NO by reduced coenzyme F420 protects mycobacteria from nitrosative damage. Proc. Natl. Acad. Sci. U. S. A. 106, 63336338,  DOI: 10.1073/pnas.0812883106
  4. 4
    Singh, R., Manjunatha, U., Boshoff, H. I., Ha, Y. H., Niyomrattanakit, P., Ledwidge, R., Dowd, C. S., Lee, I. Y., Kim, P., Zhang, L., Kang, S., Keller, T. H., Jiricek, J., and Barry, C. E., 3rd (2008) PA-824 kills nonreplicating Mycobacterium tuberculosis by intracellular NO release. Science 322, 13921395,  DOI: 10.1126/science.1164571
  5. 5
    Wang, P., Bashiri, G., Gao, X., Sawaya, M. R., and Tang, Y. (2013) Uncovering the enzymes that catalyze the final steps in oxytetracycline biosynthesis. J. Am. Chem. Soc. 135, 71387141,  DOI: 10.1021/ja403516u
  6. 6
    Li, W., Khullar, A., Chou, S., Sacramo, A., and Gerratana, B. (2009) Biosynthesis of sibiromycin, a potent antitumor antibiotic. Appl. Environ. Microbiol. 75, 28692878,  DOI: 10.1128/AEM.02326-08
  7. 7
    Ichikawa, H., Bashiri, G., and Kelly, W. L. (2018) Biosynthesis of the thiopeptins and identification of an F420H2-dependent dehydropiperidine reductase. J. Am. Chem. Soc. 140, 1074910756,  DOI: 10.1021/jacs.8b04238
  8. 8
    Heiss, G., Hofmann, K. W., Trachtmann, N., Walters, D. M., Rouviere, P., and Knackmuss, H. J. (2002) npd gene functions of Rhodococcus (opacus) erythropolis HL PM-1 in the initial steps of 2,4,6-trinitrophenol degradation. Microbiology 148, 799806,  DOI: 10.1099/00221287-148-3-799
  9. 9
    Taylor, M. C., Jackson, C. J., Tattersall, D. B., French, N., Peat, T. S., Newman, J., Briggs, L. J., Lapalikar, G. V., Campbell, P. M., Scott, C., Russell, R. J., and Oakeshott, J. G. (2010) Identification and characterization of two families of F420 H2-dependent reductases from Mycobacteria that catalyse aflatoxin degradation. Mol. Microbiol. 78, 561575,  DOI: 10.1111/j.1365-2958.2010.07356.x
  10. 10
    Taylor, M., Scott, C., and Grogan, G. (2013) F420-dependent enzymes - potential for applications in biotechnology. Trends Biotechnol. 31, 6364,  DOI: 10.1016/j.tibtech.2012.09.003
  11. 11
    Mathew, S., Trajkovic, M., Kumar, H., Nguyen, Q. T., and Fraaije, M. W. (2018) Enantio- and regioselective ene-reductions using F420H2-dependent enzymes. Chem. Commun. 54, 1120811211,  DOI: 10.1039/C8CC04449J
  12. 12
    Decamps, L., Philmus, B., Benjdia, A., White, R., Begley, T. P., and Berteau, O. (2012) Biosynthesis of F0, precursor of the F420 cofactor, requires a unique two radical-SAM domain enzyme and tyrosine as substrate. J. Am. Chem. Soc. 134, 1817318176,  DOI: 10.1021/ja307762b
  13. 13
    Grochowski, L. L., Xu, H., and White, R. H. (2008) Identification and characterization of the 2-phospho-L-lactate guanylyltransferase involved in coenzyme F420 biosynthesis. Biochemistry 47, 30333037,  DOI: 10.1021/bi702475t
  14. 14
    Graupner, M., Xu, H., and White, R. H. (2002) Characterization of the 2-phospho-L-lactate transferase enzyme involved in coenzyme F(420) biosynthesis in Methanococcus jannaschii. Biochemistry 41, 37543761,  DOI: 10.1021/bi011937v
  15. 15
    Bashiri, G., Antoney, J., Jirgis, E. N. M., Shah, M. V., Ney, B., Copp, J., Stuteley, S. M., Sreebhavan, S., Palmer, B., Middleditch, M., Tokuriki, N., Greening, C., Scott, C., Baker, E. N., and Jackson, C. J. (2019) A revised biosynthetic pathway for the cofactor F420 in prokaryotes. Nat. Commun. 10, 1558,  DOI: 10.1038/s41467-019-09534-x
  16. 16
    Bashiri, G., Rehan, A. M., Sreebhavan, S., Baker, H. M., Baker, E. N., and Squire, C. J. (2016) Elongation of the poly-gamma-glutamate tail of F420 requires both domains of the F420:gamma-glutamyl ligase (FbiB) of Mycobacterium tuberculosis. J. Biol. Chem. 291, 68826894,  DOI: 10.1074/jbc.M115.689026
  17. 17
    Li, H., Graupner, M., Xu, H., and White, R. H. (2003) CofE catalyzes the addition of two glutamates to F420-0 in F420 coenzyme biosynthesis in Methanococcus jannaschii. Biochemistry 42, 97719778,  DOI: 10.1021/bi034779b
  18. 18
    Selengut, J. D. and Haft, D. H. (2010) Unexpected abundance of coenzyme F(420)-dependent enzymes in Mycobacterium tuberculosis and other actinobacteria. J. Bacteriol. 192, 57885798,  DOI: 10.1128/JB.00425-10
  19. 19
    Ney, B., Ahmed, F. H., Carere, C. R., Biswas, A., Warden, A. C., Morales, S. E., Pandey, G., Watt, S. J., Oakeshott, J. G., Taylor, M. C., Stott, M. B., Jackson, C. J., and Greening, C. (2017) The methanogenic redox cofactor F420 is widely synthesized by aerobic soil bacteria. ISME J. 11, 125137,  DOI: 10.1038/ismej.2016.100
  20. 20
    Eirich, L. D., Vogels, G. D., and Wolfe, R. S. (1978) Proposed structure for coenzyme F420 from Methanobacterium. Biochemistry 17, 45834593,  DOI: 10.1021/bi00615a002
  21. 21
    Cheeseman, P., Toms-Wood, A., and Wolfe, R. S. (1972) Isolation and properties of a fluorescent compound, factor 420, from Methanobacterium strain M.o.H. J. Bacteriol. 112, 527531
  22. 22
    Lackner, G., Peters, E. E., Helfrich, E. J., and Piel, J. (2017) Insights into the lifestyle of uncultured bacterial natural product factories associated with marine sponges. Proc. Natl. Acad. Sci. U. S. A. 114, E347E356,  DOI: 10.1073/pnas.1616234114
  23. 23
    Lackner, G., Moebius, N., Partida-Martinez, L. P., Boland, S., and Hertweck, C. (2011) Evolution of an endofungal lifestyle: deductions from the Burkholderia rhizoxinica genome. BMC Genomics 12, 210,  DOI: 10.1186/1471-2164-12-210
  24. 24
    Lackner, G. and Hertweck, C. (2011) Impact of endofungal bacteria on infection biology, food safety, and drug development. PLoS Pathog. 7, e1002096  DOI: 10.1371/journal.ppat.1002096
  25. 25
    Scherlach, K., Busch, B., Lackner, G., Paszkowski, U., and Hertweck, C. (2012) Symbiotic cooperation in the biosynthesis of a phytotoxin. Angew. Chem., Int. Ed. 51, 96159618,  DOI: 10.1002/anie.201204540
  26. 26
    Kiener, A., Orme-Johnson, W. H., and Walsh, C. T. (1988) Reversible conversion of coenzyme F420 to the 8-OH-AMP and 8-OH-GMP esters, F390-A and F390-G, on oxygen exposure and reestablishment of anaerobiosis in Methanobacterium thermoautotrophicum. Arch. Microbiol. 150, 249253,  DOI: 10.1007/BF00407788
  27. 27
    Vermeij, P., Pennings, J. L., Maassen, S. M., Keltjens, J. T., and Vogels, G. D. (1997) Cellular levels of factor 390 and methanogenic enzymes during growth of Methanobacterium thermoautotrophicum deltaH. J. Bacteriol. 179, 66406648,  DOI: 10.1128/jb.179.21.6640-6648.1997
  28. 28
    Newton, G. L., Buchmeier, N., and Fahey, R. C. (2008) Biosynthesis and functions of mycothiol, the unique protective thiol of Actinobacteria. Microbiol Mol. Biol. Rev. 72, 471494,  DOI: 10.1128/MMBR.00008-08
  29. 29
    Zhang, L., Loh, K. C., Lim, J. W., and Zhang, J. X. (2019) Bioinformatics analysis of metagenomics data of biogas-producing microbial communities in anaerobic digesters: A review. Renewable Sustainable Energy Rev. 100, 110126,  DOI: 10.1016/j.rser.2018.10.021
  30. 30
    Graupner, M. and White, R. H. (2001) Biosynthesis of the phosphodiester bond in coenzyme F(420) in the methanoarchaea. Biochemistry 40, 1085910872,  DOI: 10.1021/bi0107703
  31. 31
    Warkentin, E., Mamat, B., Sordel-Klippert, M., Wicke, M., Thauer, R. K., Iwata, M., Iwata, S., Ermler, U., and Shima, S. (2001) Structures of F420H2: NADP+ oxidoreductase with and without its substrates bound. EMBO J. 20, 65616569,  DOI: 10.1093/emboj/20.23.6561
  32. 32
    Kunow, J., Schwörer, B., Stetter, K. O., and Thauer, R. K. (1993) A F(420)-dependent NADP reductase in the extremely thermophilic sulfate-reducing Archaeoglobus fulgidus. Arch. Microbiol. 160, 199205
  33. 33
    Drenth, J., Trajkovic, M., and Fraaije, M. W. (2019) Chemoenzymatic synthesis of an unnatural deazaflavin cofactor that can fuel F-420-dependent enzymes. ACS Catal. 9, 64356443,  DOI: 10.1021/acscatal.9b01506
  34. 34
    Ney, B., Carere, C. R., Sparling, R., Jirapanjawat, T., Stott, M. B., Jackson, C. J., Oakeshott, J. G., Warden, A. C., and Greening, C. (2017) Cofactor tail length modulates catalysis of bacterial F-420-dependent oxidoreductases. Front. Microbiol. 8, 1902,  DOI: 10.3389/fmicb.2017.01902
  35. 35
    Jirapanjawat, T., Ney, B., Taylor, M. C., Warden, A. C., Afroze, S., Russell, R. J., Lee, B. M., Jackson, C. J., Oakeshott, J. G., Pandey, G., and Greening, C. (2016) The redox cofactor F420 protects mycobacteria from diverse antimicrobial compounds and mediates a reductive detoxification system. Appl. Environ. Microbiol. 82, 68106818,  DOI: 10.1128/AEM.02500-16
  36. 36
    Mascotti, M. L., Kumar, H., Nguyen, Q. T., Ayub, M. J., and Fraaije, M. W. (2018) Reconstructing the evolutionary history of F-420-dependent dehydrogenases. Sci. Rep. 8, 17571,  DOI: 10.1038/s41598-018-35590-2
  37. 37
    Kumar, H., Nguyen, Q. T., Binda, C., Mattevi, A., and Fraaije, M. W. (2017) Isolation and characterization of a thermostable F420:NADPH oxidoreductase from Thermobifida fusca. J. Biol. Chem. 292, 1012310130,  DOI: 10.1074/jbc.M117.787754

Cited By

Click to copy section linkSection link copied!
Citation Statements
Explore this article's citation statements on scite.ai

This article is cited by 25 publications.

  1. Misun Lee, Marco W. Fraaije. Equipping Saccharomyces cerevisiae with an Additional Redox Cofactor Allows F420-Dependent Bioconversions in Yeast. ACS Synthetic Biology 2024, 13 (3) , 921-929. https://doi.org/10.1021/acssynbio.3c00718
  2. Suk Woo Kang, James Antoney, Rebecca L. Frkic, David W. Lupton, Robert Speight, Colin Scott, Colin J. Jackson. Asymmetric Ene-Reduction of α,β-Unsaturated Compounds by F420-Dependent Oxidoreductases A Enzymes from Mycobacterium smegmatis. Biochemistry 2023, 62 (3) , 873-891. https://doi.org/10.1021/acs.biochem.2c00557
  3. Misun Lee, Jeroen Drenth, Milos Trajkovic, René M. de Jong, Marco W. Fraaije. Introducing an Artificial Deazaflavin Cofactor in Escherichia coli and Saccharomyces cerevisiae. ACS Synthetic Biology 2022, 11 (2) , 938-952. https://doi.org/10.1021/acssynbio.1c00552
  4. Fulin Sun, Hongqiang Yang, Fei Tan, Xiyang Zhang, Mingzhuang Wang, Guan Wang, Qi Shi. Impact of coral bleaching on the archaeal community and carbon metabolic functions. Coral Reefs 2025, 44 (2) , 721-730. https://doi.org/10.1007/s00338-025-02641-w
  5. Annika Lenić, Bettina Bardl, Florian Kloss, Gundela Peschel, Ivan Schlembach, Gerald Lackner, Lars Regestein, Miriam A. Rosenbaum. Pilot Scale Production of a F 420 Precursor Under Microaerobic Conditions. Biotechnology Journal 2025, 20 (3) https://doi.org/10.1002/biot.70002
  6. Natalie Mladenov, Scott Sanfilippo, Laura Panduro, Chelsi Pascua, Armando Arteaga, Bjoern Pietruschka. Tracking performance and disturbance in decentralized wastewater treatment systems with fluorescence spectroscopy. Environmental Science: Water Research & Technology 2024, 10 (6) , 1506-1516. https://doi.org/10.1039/D3EW00671A
  7. Ingrid Richter, Mahmudul Hasan, Johannes W Kramer, Philipp Wein, Jana Krabbe, K Philip Wojtas, Timothy P Stinear, Sacha J Pidot, Florian Kloss, Christian Hertweck, Gerald Lackner. Deazaflavin metabolite produced by endosymbiotic bacteria controls fungal host reproduction. The ISME Journal 2024, 18 (1) https://doi.org/10.1093/ismejo/wrae074
  8. Yan‐Hua Fu, Liguo Yang, Zhongyuan Zhou, Fang Wang, Guang‐Bin Shen, Xiao‐Qing Zhu. Structure–activity relationship of alkanes and alkane derivatives for the abilities of C(sp 3 )H bonds toward their H‐atom transfer reactions. Journal of Physical Organic Chemistry 2023, 36 (10) https://doi.org/10.1002/poc.4550
  9. Guang Yang, Hein J. Wijma, Henriëtte J. Rozeboom, Maria Laura Mascotti, Marco W. Fraaije. Identification and characterization of archaeal and bacterial F 420 ‐dependent thioredoxin reductases. The FEBS Journal 2023, 290 (19) , 4777-4791. https://doi.org/10.1111/febs.16896
  10. Hao-Qiang Liu, Ze-long Zhao, Hong-Jun Li, Shi-Jiang Yu, Lin Cong, Li-Li Ding, Chun Ran, Xue-Feng Wang. Accurate prediction of huanglongbing occurrence in citrus plants by machine learning-based analysis of symbiotic bacteria. Frontiers in Plant Science 2023, 14 https://doi.org/10.3389/fpls.2023.1129508
  11. Suk Woo Kang, James Antoney, David W. Lupton, Robert Speight, Colin Scott, Colin J. Jackson. Asymmetric Ene‐Reduction by F 420 ‐Dependent Oxidoreductases B (FDOR‐B) from Mycobacterium smegmatis. ChemBioChem 2023, 24 (8) https://doi.org/10.1002/cbic.202200797
  12. Yan‐Hua Fu, Fang Wang, Ling Zhao, Yanwei Zhang, Guang‐Bin Shen, Xiao‐Qing Zhu. Thermodynamic and Kinetic Studies of the Activities of Aldehydic C−H Bonds toward Their H‐Atom Transfer Reactions. ChemistrySelect 2023, 8 (9) https://doi.org/10.1002/slct.202204789
  13. Yan‐Hua Fu, Li‐Guo Yang, Zhong‐Yuan Zhou, Taixuan Jia, Guang‐Bin Shen, Xiao‐Qing Zhu. Comparison of Thermodynamic Energies for Elementary Steps of Anionic Hydrides to Release Hydride Ions in Acetonitrile. ChemistrySelect 2022, 7 (43) https://doi.org/10.1002/slct.202203626
  14. Yan-Hua Fu, Zhen Wang, Kai Wang, Guang-Bin Shen, Xiao-Qing Zhu. Evaluation and comparison of antioxidant abilities of five bioactive molecules with C–H and O–H bonds in thermodynamics and kinetics. RSC Advances 2022, 12 (42) , 27389-27395. https://doi.org/10.1039/D2RA04839F
  15. Daniel Last, Mahmudul Hasan, Linda Rothenburger, Daniel Braga, Gerald Lackner. High-yield production of coenzyme F420 in Escherichia coli by fluorescence-based screening of multi-dimensional gene expression space. Metabolic Engineering 2022, 73 , 158-167. https://doi.org/10.1016/j.ymben.2022.07.006
  16. Mahmudul Hasan, Sabrina Schulze, Leona Berndt, Gottfried J. Palm, Daniel Braga, Ingrid Richter, Daniel Last, Michael Lammers, Gerald Lackner, . Diversification by CofC and Control by CofD Govern Biosynthesis and Evolution of Coenzyme F 420 and Its Derivative 3PG-F 420. mBio 2022, 13 (1) https://doi.org/10.1128/mbio.03501-21
  17. Mihir V. Shah, Hadi Nazem-Bokaee, James Antoney, Suk Woo Kang, Colin J. Jackson, Colin Scott. Improved production of the non-native cofactor F420 in Escherichia coli. Scientific Reports 2021, 11 (1) https://doi.org/10.1038/s41598-021-01224-3
  18. Rhys Grinter, Chris Greening. Cofactor F420: an expanded view of its distribution, biosynthesis and roles in bacteria and archaea. FEMS Microbiology Reviews 2021, 45 (5) https://doi.org/10.1093/femsre/fuab021
  19. Jeroen Drenth, Marco W. Fraaije. Natural Flavins: Occurrence, Role, and Noncanonical Chemistry. 2021, 29-65. https://doi.org/10.1002/9783527830138.ch2
  20. Friedhelm Pfeiffer, Mike Dyall-Smith. Open Issues for Protein Function Assignment in Haloferax volcanii and Other Halophilic Archaea. Genes 2021, 12 (7) , 963. https://doi.org/10.3390/genes12070963
  21. Ghader Bashiri, Edward N Baker. Convergent pathways to biosynthesis of the versatile cofactor F420. Current Opinion in Structural Biology 2020, 65 , 9-16. https://doi.org/10.1016/j.sbi.2020.05.002
  22. Kirstin Scherlach, Christian Hertweck. Chemical Mediators at the Bacterial-Fungal Interface. Annual Review of Microbiology 2020, 74 (1) , 267-290. https://doi.org/10.1146/annurev-micro-012420-081224
  23. Rhys Grinter, Blair Ney, Rajini Brammananth, Christopher K. Barlow, Paul R. F. Cordero, David L. Gillett, Thierry Izoré, Max J. Cryle, Liam K. Harold, Gregory M. Cook, George Taiaroa, Deborah A. Williamson, Andrew C. Warden, John G. Oakeshott, Matthew C. Taylor, Paul K. Crellin, Colin J. Jackson, Ralf B. Schittenhelm, Ross L. Coppel, Chris Greening, . Cellular and Structural Basis of Synthesis of the Unique Intermediate Dehydro-F 420 -0 in Mycobacteria. mSystems 2020, 5 (3) https://doi.org/10.1128/msystems.00389-20
  24. Daniel Braga, Mahmudul Hasan, Tabea Kröber, Daniel Last, Gerald Lackner, . Redox Coenzyme F 420 Biosynthesis in Thermomicrobia Involves Reduction by Stand-Alone Nitroreductase Superfamily Enzymes. Applied and Environmental Microbiology 2020, 86 (12) https://doi.org/10.1128/AEM.00457-20
  25. Mihir V. Shah, James Antoney, Suk Woo Kang, Andrew C. Warden, Carol J. Hartley, Hadi Nazem-Bokaee, Colin J. Jackson, Colin Scott. Cofactor F420-Dependent Enzymes: An Under-Explored Resource for Asymmetric Redox Biocatalysis. Catalysts 2019, 9 (10) , 868. https://doi.org/10.3390/catal9100868

ACS Chemical Biology

Cite this: ACS Chem. Biol. 2019, 14, 9, 2088–2094
Click to copy citationCitation copied!
https://doi.org/10.1021/acschembio.9b00605
Published August 30, 2019

Copyright © 2019 American Chemical Society. This publication is licensed under CC-BY-NC-ND.

Article Views

2053

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. Biosynthesis of coenzyme F420. (A) FO synthase FbiC (in archaea: CofG/H) catalyzes formation of the deazaflavin ring from tyrosine and 5-amino-6-(ribitylamino)-uracil, an intermediate of riboflavin biosynthesis. (B) Biosynthetic scheme of F420-n starting from 2-PL: CofC and CofD catalyze the activation of 2-PL and transfer of the 2-PL moiety, respectively. CofE performs (oligo-)γ-glutamylation. The number of glutamate residues (n) varies depending on the organism. Enzymes producing 2-PL are elusive, and it has been questioned that 2-PL is an intermediate of F420 biosynthesis. (C) Biosynthesis of F420-n starting from PEP: CofC and CofD activate PEP, resulting in DF420 formation. The C-terminal domain of FbiB reduces DF420 to F420. A pathway starting from 3-phosphoglycerate was established in this study (Figure 4B). EPPG: enolpyruvyl-diphosphoguanosine. LPPG: lactyl-diphosphoguanosine. 2-PL: 2-phospho-l-lactate.

    Figure 2

    Figure 2. Deazaflavin biosynthesis in P. rhizoxinica. (A) BGC of 3PG-F420. Core genes are shown in dark gray. (B–I) Microscopy photographs depict fluorescence characteristic of deazaflavins in blue. (B–D) Axenic M. smegmatis (B), P. rhizoxinica (C), and E. coli/pDB045 (D). In R. microsporus ATCC 62417, deazaflavins are correlated to the presence of P. rhizoxinica symbionts (green, Syto9 staining; E). No fluorescence was detected in cured ATCC 62417 mycelium (F) or in the naturally symbiont-free strain CBS 344.29 (G). The same pattern was observed in spores of either wild-type ATCC 62417 (H) or CBS 344.29 (I). Scale bars represent 10 μm. (J) Refactored versions of the BGC and corresponding plasmids for heterologous expression in E. coli. Asterisks mark genes from M. jannaschii.

    Figure 3

    Figure 3. Chemical analysis of 3PG-F420. (A) Extracted ion chromatograms of 3PG-F420-2 produced in E. coli. I: E. coli/pDB045. II: cofD exchanged by M. jannaschii homologue (pDB060). III: cofD and cofC exchanged by M. jannaschii homologues (pDB070). IV: empty vector (pETDuet). (B) Excerpt of the MS/MS spectrum of 3PG-F420-2. Gray bars highlight m/z used for fragment ion search of F420 derivatives. (C) 1H NMR comparison of F420-n (D2O), 3PG-F420-n (0.1% ND3 in D2O), and 3PG-F420-0 (0.1% ND3 in D2O) indicated the replacement of the lactyl moiety in F420 with a glyceryl moiety in 3PG-F420. (D) Proposed structures of 3PG-F420-0, 3PG-F420-n, and DF420-n.

    Figure 4

    Figure 4. Combined CofC/D in vitro assay. (A) Relative turnover of substrates estimated from a substrate competition assay (d-3-PG, 2-PL, and PEP). CofC from P. rhizoxinica accepted 3-PG (72.7%), 2-PL (23.4%), and PEP (3.9%). CofC from M. jannaschii preferred 2-PL (96.5%) and PEP (3.5%). 3-PG was not turned over. CofD from M. jannaschii was used in all assays. Error bars represent the standard deviation (SD) of three independent biological replicates (N = 3). (B) Proposed model of 3PG-F420 biosynthesis. 3-GPPG: 3-(guanosine-5′-disphospho)-d-glycerate.

    Figure 5

    Figure 5. Cofactor function of 3PG-F420. (A) Michaelis–Menten kinetics of Fno for F420 (left) and 3PG-F420 as substrates (right). Three biological replicates were used to determine parameters. KM for F420 was 3.6 ± 0.7 μM (standard error). KM for 3PG-F420 was 5.1 ± 1.0 μM. Error bars indicate standard deviation of replicates (N = 3). (B) In vivo reduction of malachite green (absorbance: 618 nm) by the F420-dependent reductase MSMEG_5998. Fno was used to regenerate 3PG-F420H2. Left panel: Time course of the malachite green depletion assay. Right panel: Bar chart of residual malachite green after 20 h; wt, E. coli BL21(DE3); pDB061, E. coli producing MSMEG_5998; pDB071, E. coli producing 3PG-F420-0 + Fno. Exact means ± SD of biological triplicates were 0.234 ± 0.017 (wt), 0.124 ± 0.003 (pDB061), 0.169 ± 0.011 (pDB071), and 0.082 ± 0.0139 (pDB061/pDB071). An asterisk indicates statistical significance (one-way ANOVA, p < 0.05, N = 3). (C) Engineered E. coli combining 3PG-F420, Fno, and reductase MSMEG_5998 (red) for reduction of malachite green.

  • References


    This article references 37 other publications.

    1. 1
      Jacobson, F. and Walsh, C. (1984) Properties of 7,8-didemethyl-8-hydroxy-5-deazaflavins relevant to redox coenzyme function in methanogen metabolism. Biochemistry 23, 979988,  DOI: 10.1021/bi00300a028
    2. 2
      Greening, C., Ahmed, F. H., Mohamed, A. E., Lee, B. M., Pandey, G., Warden, A. C., Scott, C., Oakeshott, J. G., Taylor, M. C., and Jackson, C. J. (2016) Physiology, biochemistry, and applications of F420- and Fo-dependent redox reactions. Microbiol. Mol. Biol. Rev. 80, 451493,  DOI: 10.1128/MMBR.00070-15
    3. 3
      Purwantini, E. and Mukhopadhyay, B. (2009) Conversion of NO2 to NO by reduced coenzyme F420 protects mycobacteria from nitrosative damage. Proc. Natl. Acad. Sci. U. S. A. 106, 63336338,  DOI: 10.1073/pnas.0812883106
    4. 4
      Singh, R., Manjunatha, U., Boshoff, H. I., Ha, Y. H., Niyomrattanakit, P., Ledwidge, R., Dowd, C. S., Lee, I. Y., Kim, P., Zhang, L., Kang, S., Keller, T. H., Jiricek, J., and Barry, C. E., 3rd (2008) PA-824 kills nonreplicating Mycobacterium tuberculosis by intracellular NO release. Science 322, 13921395,  DOI: 10.1126/science.1164571
    5. 5
      Wang, P., Bashiri, G., Gao, X., Sawaya, M. R., and Tang, Y. (2013) Uncovering the enzymes that catalyze the final steps in oxytetracycline biosynthesis. J. Am. Chem. Soc. 135, 71387141,  DOI: 10.1021/ja403516u
    6. 6
      Li, W., Khullar, A., Chou, S., Sacramo, A., and Gerratana, B. (2009) Biosynthesis of sibiromycin, a potent antitumor antibiotic. Appl. Environ. Microbiol. 75, 28692878,  DOI: 10.1128/AEM.02326-08
    7. 7
      Ichikawa, H., Bashiri, G., and Kelly, W. L. (2018) Biosynthesis of the thiopeptins and identification of an F420H2-dependent dehydropiperidine reductase. J. Am. Chem. Soc. 140, 1074910756,  DOI: 10.1021/jacs.8b04238
    8. 8
      Heiss, G., Hofmann, K. W., Trachtmann, N., Walters, D. M., Rouviere, P., and Knackmuss, H. J. (2002) npd gene functions of Rhodococcus (opacus) erythropolis HL PM-1 in the initial steps of 2,4,6-trinitrophenol degradation. Microbiology 148, 799806,  DOI: 10.1099/00221287-148-3-799
    9. 9
      Taylor, M. C., Jackson, C. J., Tattersall, D. B., French, N., Peat, T. S., Newman, J., Briggs, L. J., Lapalikar, G. V., Campbell, P. M., Scott, C., Russell, R. J., and Oakeshott, J. G. (2010) Identification and characterization of two families of F420 H2-dependent reductases from Mycobacteria that catalyse aflatoxin degradation. Mol. Microbiol. 78, 561575,  DOI: 10.1111/j.1365-2958.2010.07356.x
    10. 10
      Taylor, M., Scott, C., and Grogan, G. (2013) F420-dependent enzymes - potential for applications in biotechnology. Trends Biotechnol. 31, 6364,  DOI: 10.1016/j.tibtech.2012.09.003
    11. 11
      Mathew, S., Trajkovic, M., Kumar, H., Nguyen, Q. T., and Fraaije, M. W. (2018) Enantio- and regioselective ene-reductions using F420H2-dependent enzymes. Chem. Commun. 54, 1120811211,  DOI: 10.1039/C8CC04449J
    12. 12
      Decamps, L., Philmus, B., Benjdia, A., White, R., Begley, T. P., and Berteau, O. (2012) Biosynthesis of F0, precursor of the F420 cofactor, requires a unique two radical-SAM domain enzyme and tyrosine as substrate. J. Am. Chem. Soc. 134, 1817318176,  DOI: 10.1021/ja307762b
    13. 13
      Grochowski, L. L., Xu, H., and White, R. H. (2008) Identification and characterization of the 2-phospho-L-lactate guanylyltransferase involved in coenzyme F420 biosynthesis. Biochemistry 47, 30333037,  DOI: 10.1021/bi702475t
    14. 14
      Graupner, M., Xu, H., and White, R. H. (2002) Characterization of the 2-phospho-L-lactate transferase enzyme involved in coenzyme F(420) biosynthesis in Methanococcus jannaschii. Biochemistry 41, 37543761,  DOI: 10.1021/bi011937v
    15. 15
      Bashiri, G., Antoney, J., Jirgis, E. N. M., Shah, M. V., Ney, B., Copp, J., Stuteley, S. M., Sreebhavan, S., Palmer, B., Middleditch, M., Tokuriki, N., Greening, C., Scott, C., Baker, E. N., and Jackson, C. J. (2019) A revised biosynthetic pathway for the cofactor F420 in prokaryotes. Nat. Commun. 10, 1558,  DOI: 10.1038/s41467-019-09534-x
    16. 16
      Bashiri, G., Rehan, A. M., Sreebhavan, S., Baker, H. M., Baker, E. N., and Squire, C. J. (2016) Elongation of the poly-gamma-glutamate tail of F420 requires both domains of the F420:gamma-glutamyl ligase (FbiB) of Mycobacterium tuberculosis. J. Biol. Chem. 291, 68826894,  DOI: 10.1074/jbc.M115.689026
    17. 17
      Li, H., Graupner, M., Xu, H., and White, R. H. (2003) CofE catalyzes the addition of two glutamates to F420-0 in F420 coenzyme biosynthesis in Methanococcus jannaschii. Biochemistry 42, 97719778,  DOI: 10.1021/bi034779b
    18. 18
      Selengut, J. D. and Haft, D. H. (2010) Unexpected abundance of coenzyme F(420)-dependent enzymes in Mycobacterium tuberculosis and other actinobacteria. J. Bacteriol. 192, 57885798,  DOI: 10.1128/JB.00425-10
    19. 19
      Ney, B., Ahmed, F. H., Carere, C. R., Biswas, A., Warden, A. C., Morales, S. E., Pandey, G., Watt, S. J., Oakeshott, J. G., Taylor, M. C., Stott, M. B., Jackson, C. J., and Greening, C. (2017) The methanogenic redox cofactor F420 is widely synthesized by aerobic soil bacteria. ISME J. 11, 125137,  DOI: 10.1038/ismej.2016.100
    20. 20
      Eirich, L. D., Vogels, G. D., and Wolfe, R. S. (1978) Proposed structure for coenzyme F420 from Methanobacterium. Biochemistry 17, 45834593,  DOI: 10.1021/bi00615a002
    21. 21
      Cheeseman, P., Toms-Wood, A., and Wolfe, R. S. (1972) Isolation and properties of a fluorescent compound, factor 420, from Methanobacterium strain M.o.H. J. Bacteriol. 112, 527531
    22. 22
      Lackner, G., Peters, E. E., Helfrich, E. J., and Piel, J. (2017) Insights into the lifestyle of uncultured bacterial natural product factories associated with marine sponges. Proc. Natl. Acad. Sci. U. S. A. 114, E347E356,  DOI: 10.1073/pnas.1616234114
    23. 23
      Lackner, G., Moebius, N., Partida-Martinez, L. P., Boland, S., and Hertweck, C. (2011) Evolution of an endofungal lifestyle: deductions from the Burkholderia rhizoxinica genome. BMC Genomics 12, 210,  DOI: 10.1186/1471-2164-12-210
    24. 24
      Lackner, G. and Hertweck, C. (2011) Impact of endofungal bacteria on infection biology, food safety, and drug development. PLoS Pathog. 7, e1002096  DOI: 10.1371/journal.ppat.1002096
    25. 25
      Scherlach, K., Busch, B., Lackner, G., Paszkowski, U., and Hertweck, C. (2012) Symbiotic cooperation in the biosynthesis of a phytotoxin. Angew. Chem., Int. Ed. 51, 96159618,  DOI: 10.1002/anie.201204540
    26. 26
      Kiener, A., Orme-Johnson, W. H., and Walsh, C. T. (1988) Reversible conversion of coenzyme F420 to the 8-OH-AMP and 8-OH-GMP esters, F390-A and F390-G, on oxygen exposure and reestablishment of anaerobiosis in Methanobacterium thermoautotrophicum. Arch. Microbiol. 150, 249253,  DOI: 10.1007/BF00407788
    27. 27
      Vermeij, P., Pennings, J. L., Maassen, S. M., Keltjens, J. T., and Vogels, G. D. (1997) Cellular levels of factor 390 and methanogenic enzymes during growth of Methanobacterium thermoautotrophicum deltaH. J. Bacteriol. 179, 66406648,  DOI: 10.1128/jb.179.21.6640-6648.1997
    28. 28
      Newton, G. L., Buchmeier, N., and Fahey, R. C. (2008) Biosynthesis and functions of mycothiol, the unique protective thiol of Actinobacteria. Microbiol Mol. Biol. Rev. 72, 471494,  DOI: 10.1128/MMBR.00008-08
    29. 29
      Zhang, L., Loh, K. C., Lim, J. W., and Zhang, J. X. (2019) Bioinformatics analysis of metagenomics data of biogas-producing microbial communities in anaerobic digesters: A review. Renewable Sustainable Energy Rev. 100, 110126,  DOI: 10.1016/j.rser.2018.10.021
    30. 30
      Graupner, M. and White, R. H. (2001) Biosynthesis of the phosphodiester bond in coenzyme F(420) in the methanoarchaea. Biochemistry 40, 1085910872,  DOI: 10.1021/bi0107703
    31. 31
      Warkentin, E., Mamat, B., Sordel-Klippert, M., Wicke, M., Thauer, R. K., Iwata, M., Iwata, S., Ermler, U., and Shima, S. (2001) Structures of F420H2: NADP+ oxidoreductase with and without its substrates bound. EMBO J. 20, 65616569,  DOI: 10.1093/emboj/20.23.6561
    32. 32
      Kunow, J., Schwörer, B., Stetter, K. O., and Thauer, R. K. (1993) A F(420)-dependent NADP reductase in the extremely thermophilic sulfate-reducing Archaeoglobus fulgidus. Arch. Microbiol. 160, 199205
    33. 33
      Drenth, J., Trajkovic, M., and Fraaije, M. W. (2019) Chemoenzymatic synthesis of an unnatural deazaflavin cofactor that can fuel F-420-dependent enzymes. ACS Catal. 9, 64356443,  DOI: 10.1021/acscatal.9b01506
    34. 34
      Ney, B., Carere, C. R., Sparling, R., Jirapanjawat, T., Stott, M. B., Jackson, C. J., Oakeshott, J. G., Warden, A. C., and Greening, C. (2017) Cofactor tail length modulates catalysis of bacterial F-420-dependent oxidoreductases. Front. Microbiol. 8, 1902,  DOI: 10.3389/fmicb.2017.01902
    35. 35
      Jirapanjawat, T., Ney, B., Taylor, M. C., Warden, A. C., Afroze, S., Russell, R. J., Lee, B. M., Jackson, C. J., Oakeshott, J. G., Pandey, G., and Greening, C. (2016) The redox cofactor F420 protects mycobacteria from diverse antimicrobial compounds and mediates a reductive detoxification system. Appl. Environ. Microbiol. 82, 68106818,  DOI: 10.1128/AEM.02500-16
    36. 36
      Mascotti, M. L., Kumar, H., Nguyen, Q. T., Ayub, M. J., and Fraaije, M. W. (2018) Reconstructing the evolutionary history of F-420-dependent dehydrogenases. Sci. Rep. 8, 17571,  DOI: 10.1038/s41598-018-35590-2
    37. 37
      Kumar, H., Nguyen, Q. T., Binda, C., Mattevi, A., and Fraaije, M. W. (2017) Isolation and characterization of a thermostable F420:NADPH oxidoreductase from Thermobifida fusca. J. Biol. Chem. 292, 1012310130,  DOI: 10.1074/jbc.M117.787754
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acschembio.9b00605.

    • Methods section, supporting figures, data, and discussion (fluorescence microscopy, mass spectrometry, NMR, enzyme assays) (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.