Enzymatic assembly of carbon–carbon bonds via iron-catalysed sp3 C–H functionalization (2024)

  • Journal List
  • HHS Author Manuscripts
  • PMC6440214

As a library, NLM provides access to scientific literature. Inclusion in an NLM database does not imply endorsem*nt of, or agreement with, the contents by NLM or the National Institutes of Health.
Learn more: PMC Disclaimer | PMC Copyright Notice

Enzymatic assembly of carbon–carbon bonds via iron-catalysed sp3 C–H functionalization (1)

About Author manuscriptsSubmit a manuscriptHHS Public Access; Author Manuscript; Accepted for publication in peer reviewed journal;

Nature. Author manuscript; available in PMC 2019 Jun 19.

Published in final edited form as:

Nature. 2019 Jan; 565(7737): 67–72.

Published online 2018 Dec 19. doi:10.1038/s41586-018-0808-5

PMCID: PMC6440214

NIHMSID: NIHMS1511383

PMID: 30568304

Ruijie K. Zhang,1 Kai Chen,1 Xiongyi Huang,1 Lena Wohlschlager,1,2 Hans Renata,1,3 and Frances H. Arnold1,*

Author information Copyright and License information PMC Disclaimer

The publisher's final edited version of this article is available at Nature

Associated Data

Supplementary Materials

Though abundant in organic molecules, carbon–hydrogen (C–H) bonds are typically considered unreactive and unavailable for chemical manipulation. Recent advances in C–H functionalization technology have begun to transform this logic, while emphasizing the challenge and importance of selective installation of sp3 C–alkyl groups into a hydrocarbon framework1,2. Here we describe the first iron-based catalysts for enantio-, regio-, and chemo-selective intermolecular alkylation of sp3 C–H bonds through carbene C–H insertion. The catalysts, derived from a cytochrome P450 enzyme whose native cysteine axial ligand has been substituted for serine (“cytochrome P411”), are fully genetically encoded and produced in bacteria, where they can be tuned by directed evolution for activity and selectivity. That these proteins activate iron, the most abundant transition metal, to perform this challenging chemistry provides a desirable alternative to noble metal catalysts, which have dominated the field of C–H functionalization1,2. The laboratory-evolved enzymes functionalize diverse substrates containing benzylic, allylic, or α-amino C–H bonds with high turnover and exquisite selectivity. Furthermore, these highly efficient enzymes have enabled the development of concise routes to several natural products. The demonstration that these enzymes mediate sp3 C–H alkylation using their native iron-haem cofactor unlocks a vast natural haem protein diversity for this abiological transformation and will facilitate the development of new enzymatic C–H functionalization reactions for applications in chemistry and synthetic biology.

Biological systems use a limited set of chemical strategies to form carbon–carbon (C–C) bonds during construction of organic molecules3. Whereas many of these approaches rely on the manipulation of functional groups, certain enzymes, including members of the radical S-adenosylmethionine (SAM) family, can perform alkylation of sp3 C–H bonds. This has been an especially versatile strategy for structural diversification, as seen by its essential role in the biosynthesis of structurally varied natural products and cofactors46. Known biological machineries for this transformation, however, are limited to enzymes that transfer a methyl group5,6 or conjugate an activated radical acceptor substrate4,7 to specific molecules, with methylation as a common mode for sp3 C–alkyl installation by radical SAM enzymes (Fig. 1a).

Open in a separate window

Figure 1 |

Enzymatic C–H functionalization systems.

a, Methylation catalysed by cobalamin-dependent radical SAM enzymes, as illustrated by Fom3 in fosfomycin biosynthesis6. b, Oxygenation catalysed by cytochrome P450 monooxygenase (top) and envisioned alkylation reaction achieved under haem protein catalysis (bottom). Structural illustrations are adapted from Protein Data Bank (PDB) ID code 5UL4 (radical SAM enzyme) and PDB 2IJ2 (cytochrome P450BM3). Ad, adenosyl; Cys, cysteine; R, organic group; X, amino acid.

We sought to introduce a new enzymatic strategy for the alkylation of sp3 C–H bonds. For our design, we drew inspiration from the most widely used biological C–H functionalization transformation, C–H oxygenation. Enzymes such as the cytochromes P450 accomplish C–H oxygenation using a haem cofactor; their activities rely on activation of molecular oxygen for the controlled generation of a high-energy iron-oxo intermediate capable of selective insertion into a substrate C–H bond8. Analogously, we anticipated that the combination of a haem protein and a diazo compound would generate a protein-enclosed iron-carbene species and that this carbene could participate in a selective C–H insertion reaction with a second substrate (Fig. 1b). While it has been shown that haem proteins are capable of performing carbene transfer processes such as cyclopropanation and heteroatom–hydrogen bond insertions911, their functionalization of sp3 C–H bonds remained elusive.

Metal-carbene sp3 C–H insertion in small-molecule catalysis, especially intermolecular and stereoselective versions of this reaction, typically relies on transition metal complexes based on rhodium12, iridium13, and others1416. Artificial metalloproteins for carbene C–H insertion have been created by introducing an iridium-porphyrin into variants of apo haem proteins17. Though rare, there are a few examples of iron-carbene sp3 C–H insertion. The iron-catalysed examples employ elevated temperatures (e.g. 80 °C)18, are stoichiometric19, or are restricted to intramolecular reactions20, indicating a high activation energy barrier for C–H insertion with an iron-carbene. However, because the protein framework of an enzyme can impart significant rate enhancements to reactions21 and even confer activity to an otherwise unreactive cofactor22, we surmised that directed evolution could reconfigure a haem protein to overcome the barrier for the iron-carbene C–H insertion reaction and acquire this new function (Fig. 1b).

In initial studies, we tested a panel of seventy-eight haem proteins which included variants of cytochromes P450, cytochromes c, and globin hom*ologs. The haem proteins in whole Escherichia coli (E. coli) cells were combined with p-methoxybenzyl methyl ether (1a) and ethyl diazoacetate (2) at room temperature under anaerobic conditions; the resulting reactions were analysed for formation of C–H alkylation product 3a (Fig. 2a., see Supplementary Information for the complete list of tested haem proteins). We found haem proteins from two superfamilies that showed low levels of this promiscuous activity, establishing the possibility of creating C–H alkylation enzymes with very different protein architectures. An engineered variant of cytochrome P450BM3 from Bacillus megaterium with an axial cysteine-to-serine mutation (cytochrome “P411”), P-4 A82L (ref. 22), provided 3a with 13 total turnovers (TTN). In addition, nitric oxide dioxygenase from Rhodothermus marinus containing the Y32G mutation (Rma NOD Y32G) catalysed the reaction with 7 TTN. A second alkane substrate, 4-ethylanisole (1i), was also accepted by the nascent C–H alkylation enzymes, albeit with lower turnover numbers (Supplementary Table S2). The haem cofactor alone (iron protoporphyrin IX) or in the presence of bovine serum albumin were inactive (Supplementary Tables S1 and S2).

Open in a separate window

Figure 2 |

Haem protein-catalysed sp3 C–H alkylation.

a, Select subset of haem proteins tested for promiscuous C–H alkylation activity. Structural illustrations are of representative superfamily members with the haem cofactor shown as red sticks: cytochrome P450BM3 (PDB 2IJ2), sperm whale myoglobin (PDB 1A6K), and Rma cytochrome c (PDB 3CP5). TTN, total turnover number; n. d., not detected; WT, wild type; Mb, sperm whale myoglobin, HGG, Hell’s Gate globin; cyt c, cytochrome c; Hth, Hydrogenobacter thermophilus.b, Directed evolution of a cytochrome P411 for enantioselective C–H alkylation (reaction shown in (a)). Bars represent average TTNs from reactions performed in quadruplicate; each TTN data point is shown as a grey dot. Enantioselectivity results are represented by green diamonds. Unless otherwise indicated, reaction conditions were haem protein in E. coli whole cells (OD600 = 30, (a)) or in clarified E. coli lysate (b), 10 mM substrate 1a, 10 mM ethyl diazoacetate, 5 vol% EtOH in M9-N buffer at room temperature under anaerobic conditions for 18 hours. Reactions performed with lysate contain 1 mM Na2S2O4. TTN is defined as the amount of indicated product divided by total haem protein as measured by the hemochrome assay. See Supplementary Information for the complete list of haem proteins tested and detailed experimental procedures.

With P411 P-4 A82L as the starting template, sequential rounds of site-saturation mutagenesis and screening in whole E. coli cells were performed to identify increasingly active and enantioselective biocatalysts for C–H alkylation. Amino acid residues chosen for mutagenesis included those which line the active site pocket, reside on loops and other flexible regions of the protein, or possess a nucleophilic side chain23. Improved variants were subsequently evaluated in reactions using clarified E. coli lysate with p-methoxybenzyl methyl ether (1a) and 4-ethylanisole (1i) (Fig. 2b and Supplementary Fig. S1). Five rounds of mutagenesis and screening yielded variant P411-gen6, which furnished product 3a with 60 TTN. Unlike the native monooxygenase activity, the C–H alkylation process does not require reducing equivalents from the FAD and FMN domains of the enzyme. Surmising that these domains may not be needed for the C–H alkylation reaction, we performed systematic truncations of P411-gen6 to determine the minimally sufficient domain(s) for retaining catalytic activity. Curiously, removal of the FAD domain, containing 37% of the amino acids in the full-length protein, created an enzyme with higher C–H alkylation activity: P411ΔFAD-gen6 delivers 3a with 100 TTN, a 1.7-fold increase in TTN compared with P411-gen6 (Supplementary Fig. S2). This indicates that the FAD domain may have (negative) allosteric effects on C–H alkylation activity. Further studies with these truncated enzymes revealed that they could be used in whole E. coli cells, in clarified E. coli cell lysate, and as purified proteins (Supplementary Table S3). Eight additional rounds of mutagenesis and screening yielded P411-CHF (P411ΔFAD CH Functionalization enzyme, full list of changes provided in the Supplementary Information).

P411-CHF displays 140-fold improvement in activity over P-4 A82L and delivers 3a with excellent stereoselectivity (2020 TTN, 96.7 : 3.3 e.r. using clarified E. coli lysate). Subsequent studies showed that the stereoselectivity could be improved by conducting the reaction at lower temperature (e.g. 4 °C) with no significant change to TTN (Supplementary Table S4). Enzymatic C–H alkylation can be performed on millimole scale: using 1.0 mmol substrate 1a, E. coli harbouring P411-CHF at 4 °C furnished 3a in 82% isolated yield, 1060 TTN, and 98.0 : 2.0 e.r. (Fig. 2b). Preliminary mechanistic investigations were pursued to interrogate the nature of the C–H insertion step. Independent initial rates measured for reactions with substrate 1a or deuterated substrate 1a-d2 revealed a normal kinetic isotope effect (KIE) of 5.1 for C–H alkylation catalysed by P411-CHF, suggesting that C–H insertion is rate-determining (Supplementary Fig. S5).

Using E. coli harbouring P411-CHF, we assayed a range of benzylic substrates for coupling with ethyl diazoacetate (Fig. 3). Both electron-rich and electron-deficient functionalities on the aromatic ring are well-tolerated (3a3e, 3h); cyclic substrates are also suitable coupling partners (3f, 3g). Functionalization of alkyl benzenes is successful at secondary benzylic sp3 C–H bonds (3i3l). Notably, in the biotransformation of substrate 1l containing both tertiary and secondary benzylic C–H bonds, P411-CHF preferentially functionalizes the secondary position despite its higher C–H bond dissociation energy (BDE). The carbene intermediate derived from ethyl diazoacetate belongs to the acceptor-only class. In contrast to the more widely-used donor/acceptor carbenes, acceptor-only intermediates are more electrophilic, and as a result selective reactions with this carbene class are still a major challenge for small-molecule catalysts13,16. Our results show that P411-CHF can control this highly reactive intermediate to furnish the desired sp3 C–H alkylation products and do so with high enantioselectivity.

Open in a separate window

Figure 3 |

Substrate scope for benzylic C–H alkylation with P411-CHF.

a, Experiments were performed using E. coli expressing cytochrome P411-CHF (OD600 = 30) with 10 mM substrate 1a1l and 10 mM ethyl diazoacetate at room temperature (RT) under anaerobic conditions for 18 hours; each reported TTN is the average of quadruplicate reactions. See Supplementary Fig. S12 for the full list of alkane substrates. Si–H insertion product 3h’ is also observed (Supplementary Fig. S7). b, Reaction selectivity for carbene C–H insertion or cyclopropanation can be controlled by the protein scaffold. Experiments were performed as in (a) using the indicated P411 variant. ξd.r. is given as cis : trans; e.r. was not determined.

Enzymes can exhibit excellent reaction selectivity arising from their ability to form multiple interactions with substrates and intermediates throughout a reaction cycle. We hypothesized that the protein scaffold could be tuned to create complementary enzymes which access different reaction outcomes available to a substrate. When P411-CHF was challenged with 4-allylanisole (1m), a substrate which can undergo both C–H alkylation and cyclopropanation, we observed that C–H alkylation product 3m dominates, with selectivity > 25:1 (Fig. 3b, Supplementary Fig. S6). In contrast, a related full-length P411 variant P-I263F, containing thirteen mutations in the haem domain relative to P411-CHF, catalysed only the formation of cyclopropane product 3m’. Additionally, despite the established reactivity of silanes with iron-carbene10, P411-CHF delivered C–H alkylation product 3h when substrate 1h was employed in the reaction (Si–H insertion product 3h’ was also observed but its formation may not be catalysed by P411-CHF, Supplementary Fig. S7). Reaction with P-I263F, in contrast, provided only the Si–H insertion product. These examples demonstrate an exceptional feature of macromolecular enzymes: different products can be obtained simply by changing the amino acid sequence of the protein catalyst.

Enzymatic C–H alkylation is not limited to functionalization of benzylic C–H bonds. Structurally dissimilar molecules containing allylic or propargylic C–H bonds are excellent substrates for this chemistry (Fig. 4a). In contrast to 1a1m, which contain a rigid benzene ring, compounds 4a4c and 4e feature flexible linear alkyl chains. Their successful enantioselective alkylation suggests that the enzyme active site can accommodate substrate conformational flexibility while enforcing a favoured substrate orientation relative to the carbene intermediate. To demonstrate the utility of this biotransformation, we applied the methodology to the formal synthesis of lyngbic acid (Fig. 4a). Marine cyanobacteria incorporate this versatile biomolecule into members of the malyngamide family of natural products; likewise, total synthesis approaches to these natural products typically access lyngbic acid as a strategic intermediate en route to the target molecules24. Using E. coli harbouring P411-CHF, intermediate 5a was produced on 2.4 mmol scale in 86% isolated yield, 2810 TTN, and 94.7 : 5.3 e.r.. Subsequent hydrogenation and hydrolysis provided (R)-(+)-6 in quantitative yield, which can be elaborated to (R)-(+)-lyngbic acid by decarboxylative alkenylation25.

Open in a separate window

Figure 4 |

Application of P411 enzymes for sp3 C–H alkylation.

a, Allylic and propargylic C–H alkylation. Unless otherwise indicated, experiments were performed using E. coli expressing cytochrome P411-CHF with 10 mM substrate 4a4e and 10 mM ethyl diazoacetate; each reported TTN is the average of quadruplicate reactions. #TTN was calculated based on isolated yield from a reaction performed at 0.25 mmol scale. Cyclopropene product was also observed (Supplementary Fig. S8). *Hydrogenation, followed by hydrolysis. b, Enzymatic alkylation of substrates containing α-amino C–H bonds. Unless otherwise indicated, experiments were performed at 0.5 mmol scale using E. coli expressing cytochrome P411-CHF with substrates 7a7f and ethyl diazoacetate; TTNs were calculated based on isolated yields of products shown. ξIsolated in 9 : 1 r.r. for 8f : 8f’. eReduction, halogen exchange, and Suzuki-Miyaura cross-coupling. c, Enzymatic C–H alkylation with alternative diazo reagents. Unless otherwise indicated, reactions were performed at 0.5 mmol scale using E. coli expressing cytochrome P411-CHF with coupling partner 1a or 7a and diazo compounds 9a9d; TTNs were calculated based on isolated yields of products shown. Variant P411-IY T327I was used. See Supplementary Information for the complete list substrates (Fig. S12 and Fig. S13), information about enzyme variants, and full experimental details.

As part of our substrate scope studies, we challenged P411-CHF with alkyl amine compounds. Compounds of this type are typically challenging for C–H functionalization methods because the amine functionality may coordinate to and inhibit the catalyst or create the opportunity for undesirable side reactions (e.g. ylide formation and its associated rearrangements)26. Using 7a or 7b, substrates which have both benzylic C–H bonds and α-amino C–H bonds, P411-CHF delivered the corresponding β-amino ester product with high efficiency (8a and 8b, Fig. 4b). Notably, benzylic C–H insertion was not observed (with 7a, Supplementary Fig. S9) or significantly suppressed (with 7b, Supplementary Fig. S10), despite the typically lower BDEs of benzylic C–H bonds compared to α-amino C–H bonds. Additionally, N-aryl pyrrolidines (7c7e) served as excellent substrates and were selectively alkylated at the α-amino sp3 position. Using P411-CHF, the sp3 C–H alkylation of 7c outcompetes a Friedel-Crafts type reaction on the aryl ring, which is a favourable process with other carbene-transfer systems27, 28. Furthermore, alkylation product 8d offers a conceivable strategy for the synthesis of β-hom*oproline, a motif which has been investigated for medicinal chemistry applications29.

Given that P411-CHF alkylates both primary and secondary α-amino C–H bonds, we interrogated whether the enzyme could be selective for one of these positions. Employing N-methyl tetrahydroquinoline 7f as the alkane substrate, P411-CHF afforded β-amino ester products with 1050 TTN and a 9 : 1 ratio of regioisomers (C2 : C1, and 73.0 : 27.0 e.r. for 8f) (Fig. 4b). As the tetrahydroquinoline ring is a privileged structural motif in natural products and bioactive molecules30, its selective functionalization could provide a concise strategy for the synthesis of alkaloids. To improve the selectivity for alkylation of 7f, we tested variants along the evolutionary lineage from P-4 A82L to P411-CHF. We found that P411-gen5 had even better regioselectivity and delivered product with the opposite stereo-preference. In a 3.0 mmol scale reaction, E. coli harbouring P411-gen5 delivered 8f in 85% yield with excellent selectivity (1310 TTN, > 50 : 1 r.r., 8.9 : 91.1 e.r.). In only a few steps, the enzymatic product was successfully transformed to alkaloid (R)-(+)-cuspareine30 (Fig. 4b).

Finally, we probed the introduction of different alkyl groups. Using different diazo reagents, enzymatic C–H alkylation can diversify one alkane substrate, such as 7a, to several products (10a10c in Fig. 4c and Supplementary Fig. S11). The diazo substrate scope extends beyond ester-based reagents: Weinreb amide diazo compound 9c and diazoketone 9d were found to participate in enzymatic C–H alkylation to furnish products 10c and 10d, respectively. Additional substitution at the α-position of the carbene, however, is generally not well-tolerated by P411-CHF and current related enzymes. With the exception of 10b, reactions using disubstituted carbene reagents failed to yield appreciable amounts of desired products (Supplementary Fig. S11).

This study demonstrates that a cytochrome P450 can acquire the ability to construct C–C bonds from sp3 C–H bonds and that activity and selectivity can be greatly enhanced using directed evolution. Nature provides a huge collection of possible alternative starting points for expanding the scope of this reaction even further and for achieving other selectivities. The cytochrome P450 superfamily can access an immense set of organic molecules for its native oxygenation chemistry; we envision that P411-derived enzymes and other natural haem protein diversity can be leveraged to generate families of C–H alkylation enzymes that emulate the scope and selectivity of nature’s C–H oxygenation catalysts.

Supplementary Material

1

Click here to view.(29K, docx)

2

Click here to view.(19M, pdf)

Acknowledgements

This work was supported by the National Science Foundation (NSF), Division of Molecular and Cellular Biosciences (grant MCB-1513007). R.K.Z. acknowledges support from the NSF Graduate Research Fellowship (grant DGE-1144469) and the Donna and Benjamin M. Rosen Bioengineering Center. X.H. is supported by a Ruth L. Kirschstein National Institutes of Health Postdoctoral Fellowship (grant F32GM125231). L.W. received support from the Austrian Marshall Plan Foundation. We thank A. Z. Zhou for experimental assistance; N. W. Goldberg, S. C. Hammer, K. E. Hernandez, Z. Jia, A. M. Knight, G. Kubik, R. D. Lewis, C. K. Prier, D. K. Romney, and J. Zhang for helpful discussions; S. Virgil, M. Shahgholi, and D. VanderVelde for analytical support; B. Stoltz for use of polarimeter and gas chromatography equipment.

Footnotes

Data Availability

Full experimental details, information about enzyme variants, Supplementary Figures S1–S15, and Supplementary Tables S1–S13 are available in the Supplementary Information. Any additional information is available from the corresponding author upon request.

Author Information

A provisional patent application has been filed through the California Institute of Technology based on the results presented here.

Supplementary Information is available in the online version of the paper.

References

1. Hartwig JF & Larsen MAUndirected, hom*ogeneous C–H bond functionalization: Challenges and opportunities. ACS Cent. Sci2, 281–292 (2016). [PMC free article] [PubMed] [Google Scholar]

2. Saint-Denis TG, Zhu R-Y, Chen G, Wu Q-F & Yu J-QEnantioselective C(sp3)–H bond activation by chiral transition metal catalysts. Science359, doi: 10.1126/science.aao4798 (2018). [PMC free article] [PubMed] [CrossRef] [Google Scholar]

3. Frey PA & Hegeman ADEnzymatic Reaction Mechanisms (Oxford University Press, New York, 2007), chap. 14. [Google Scholar]

4. Yokoyama K & Lilla EAC–C bond forming radical SAM enzymes involved in the construction of carbon skeletons of cofactors and natural products. Nat. Prod. Rep35, 660–694 (2018). [PMC free article] [PubMed] [Google Scholar]

5. Bauerle MR, Schwalm EL & Booker SJMechanistic diversity of radical S-adenosylmethionine (SAM)-dependent methylation. J. Biol. Chem290, 3995–4002 (2015). [PMC free article] [PubMed] [Google Scholar]

6. McLaughlin MI & van der Donk WAStereospecific radical-mediated B12-dependent methyl transfer by the fosfomycin biosynthesis enzyme Fom3. Biochemistry57, 4967–4971 (2018). [PMC free article] [PubMed] [Google Scholar]

7. Shisler KA & Broderick JBGlycyl radical activating enzymes: Structure, mechanism, and substrate interactions. Arch. Biochem. Biophys546, 64–71 (2014). [PMC free article] [PubMed] [Google Scholar]

8. Poulos TLHeme enzyme structure and function. Chem. Rev114, 3919–3962 (2014). [PMC free article] [PubMed] [Google Scholar]

9. Coelho PS, Brustad EM, Kannan A & Arnold FHOlefin cyclopropanation via carbene transfer catalyzed by engineered cytochrome P450 enzymes. Science339, 307–310 (2013). [PubMed] [Google Scholar]

10. Kan SBJ, Lewis RD, Chen K & Arnold FHDirected evolution of cytochrome c for carbon–silicon bond formation: Bringing silicon to life. Science354, 1048–1051 (2016). [PMC free article] [PubMed] [Google Scholar]

11. Brandenberg OF, Fasan R & Arnold FHExploiting and engineering hemoproteins for abiological carbene and nitrene transfer reactions. Curr. Opin. Biotechnol47, 102–111 (2017). [PMC free article] [PubMed] [Google Scholar]

12. Liao K, Pickel TC, Boyarskikh V, Bacsa J, Musaev DG & Davies HMLSite-selective and stereoselective functionalization of non-activated tertiary C–H bonds. Nature551, 609–613 (2017). [PubMed] [Google Scholar]

13. Weldy NM, Schafer AG, Owens CP, Herting CJ, Varela-Alvarez A, Chen S, Niemeyer Z, Musaev DG, Sigman MS, Davies HML & Blakey SBIridium(III)-bis(imidazolinyl)phenyl catalysts for enantioselective C–H functionalization with ethyl diazoacetate. Chem. Sci7, 3142–3146 (2016). [PMC free article] [PubMed] [Google Scholar]

14. Wang Y, Wen X, Cui X & Zhang XPEnantioselective radical cyclization for construction of 5-membered ring structures by metalloradical C‒H alkylation. J. Am. Chem. Soc140, 4792–4796 (2018). [PMC free article] [PubMed] [Google Scholar]

15. Caballero A, Despagnet-Ayoub E, Díaz-Requejo MM, Díaz-Rodríguez A, González-Núñez ME, Mello R, Muñoz BK, Ojo W-S, Asensio G, Etienne M & Pérez PJSilver-catalyzed C‒C bond formation between methane and ethyl diazoacetate in supercritical CO2. Science332, 835–838 (2011). [PubMed] [Google Scholar]

16. Wu W-T, Yang Z-P & You S-Lin Asymmetric Functionalization of C–H Bonds, ed. You S-L (Royal Society of Chemistry, Cambridge, UK, 2015), chap. 1. [Google Scholar]

17. Key HM, Dydio P, Clark DS & Hartwig JFAbiological catalysis by artificial haem proteins containing noble metals in place of iron. Nature534, 534–537 (2016). [PubMed] [Google Scholar]

18. Mbuvi HM & Woo LKCatalytic C–H insertions using iron(III) porphyrin complexes. Organometallics27, 637–645 (2008). [Google Scholar]

19. Li Y, Huang J-S, Zhou Z-Y, Che C-M, & You X-ZRemarkably stable iron porphyrins bearing nonheteroatom-stabilized carbene or (alkoxycarbonyl)carbenes: Isolation, X-ray crystal structures, and carbon atom transfer reactions with hydrocarbons. J. Am. Chem. Soc124, 13185–13193 (2002). [PubMed] [Google Scholar]

20. Griffin JR, Wendell CI, Garwin JA & White MCCatalytic C(sp3)–H alkylation via an iron carbene intermediate. J. Am. Chem. Soc139, 13624–13627 (2017). [PMC free article] [PubMed] [Google Scholar]

21. Herschlag D & Natarajan AFundamental challenges in mechanistic enzymology: Progress toward understanding the rate enhancements of enzymes. Biochemistry52, 2050–2067 (2013). [PMC free article] [PubMed] [Google Scholar]

22. Prier CK, Zhang RK, Buller AR, Brinkmann-Chen S & Arnold FHEnantioselective, intermolecular benzylic C‒H amination catalysed by an engineered iron-haem enzyme. Nat. Chem9, 629–634 (2017). [PMC free article] [PubMed] [Google Scholar]

23. Renata H, Lewis RD, Sweredoski MJ, Moradian A, Hess S, Wang ZJ & Arnold FHIdentification of mechanism-based inactivation in P450-catalyzed cyclopropanation facilitates engineering of improved enzymes. J. Am. Chem. Soc138, 12527–12533 (2016). [PMC free article] [PubMed] [Google Scholar]

24. Zhang J-T, Qi X-L, Chen J, Li B-S, Zhou Y-B & Cao X-PTotal synthesis of malyngamides K, L, and 5”-epi-C and absolution configuration of malyngamide L. J. Org. Chem76, 3946–3959 (2011). [PubMed] [Google Scholar]

25. Edwards JT, Merchant RR, McClymont KS, Knouse KW, Qin T, Malins LR, Vokits B, Shaw SA, Bao D-H, Wei F-L, Zhou T, Eastgate MD & Baran PSDecarboxylative alkenylation. Nature545, 213–218 (2017). [PMC free article] [PubMed] [Google Scholar]

26. He J, Hamann LG, Davies HML & Beckwith REJLate-stage C–H functionalization of complex alkaloids and drug molecules via intermolecular rhodium-carbenoid insertion. Nat. Commun6, 5943 (2015). [PubMed] [Google Scholar]

27. Yang J-M, Cai Y, Zhu S-F & Zhou Q-LIron-catalyzed arylation of α-aryl-α-diazoesters. Org. Biomol. Chem14, 5516–5519 (2016). [PubMed] [Google Scholar]

28. Xu B, Li M-L, Zuo X-D, Zhu S-F & Zhou Q-LCatalytic asymmetric arylation of α‐aryl-α-diazoacetates with aniline derivatives. J. Am. Chem. Soc137, 8700–8703 (2015). [PubMed] [Google Scholar]

29. Cardillo G, Gentilucci L, Qasem AR, Sgarzi F & Spampinato SEndomorphin-1 analogues containing β-Proline are μ-opioid receptor agonists and display enhanced enzymatic hydrolysis resistance. J. Med. Chem45, 2571–2578 (2002). [PubMed] [Google Scholar]

30. Sridharan V, Suryavanshi PA & Menéndez JCAdvances in the chemistry of tetrahydroquinolines. Chem. Rev111, 7157–7259 (2011). [PubMed] [Google Scholar]

Enzymatic assembly of carbon–carbon bonds via iron-catalysed sp3 C–H functionalization (2024)

FAQs

Which enzyme catalyzes the formation of a carbon carbon bond? ›

Enzymatic carbon–carbon bond forming reactions [1] catalysed by aldolases, transketolases [2–5], hydroxynitrile lyases [1,6] and thiamin diphosphate (ThDP)-depending α-hydroxy ketone forming enzymes [1,7] are well established for synthetic purposes.

What is the enzyme for co2 fixation? ›

Ribulose-1,5-bisphosphate carboxylase/oxygenase, better known by the name Rubisco, is the key enzyme responsible for photosynthetic and chemoautotrophic carbon fixation and oxygen metabolism.

What is the role of enzymes as catalysts in organic reactions? ›

Enzymes (and other catalysts) act by reducing the activation energy, thereby increasing the rate of reaction. The increased rate is the same in both the forward and reverse directions, since both must pass through the same transition state.

What is the name of the enzyme that catalyzes carbon fixation in the c3 cycle? ›

The enzyme that catalyzes this specific reaction is ribulose bisphosphate carboxylase (RuBisCO). RuBisCO is identified as the most abundant enzyme on earth, to date. RuBisCO is the first enzyme utilized in the process of carbon fixation and its enzymatic activity is highly regulated.

Which enzymes catalyze bond formation? ›

Three specific enzymes are discussed: pyruvate-formate lyase (PFL), spore photoproduct lyase (SPL), and benzylsuccinate synthase (BSS). What these enzymes have in common is that they use radical chemistry to catalyze C–C bond formation or cleavage reactions.

Where are the enzymes required for carbon fixation located? ›

The enzymes required for carbon fixation are located only in the grana of chloroplasts. Photosynthesis is a redox process, in which water is oxidised and carbon dioxide is reduced. In green plants, both PS- I and PS- II are required for the formation of NADPH + H+.

Where are the enzymes for the carbon fixation reactions located? ›

The enzymes that catalyse the dark reaction of carbon fixation are present outside the thylakoids. These enzymes are present in the stromal matrix of chloroplasts.

What is formed as a result of enzyme RuBisCO? ›

The main function of RuBisCO is in photosynthesis and photorespiration. It catalyses the first step of carbon fixation in the C3 pathway or Calvin cycle, i.e. carboxylation of RuBP. It results in the formation of 2 molecules of 3-PGA.

Why do all enzymatic reactions need activation? ›

Activation energy is needed so reactants can move together, overcome forces of repulsion, and start breaking bonds.

What is the mechanism of enzyme-catalyzed reaction? ›

The mechanism of enzymatic action. An enzyme attracts substrates to its active site, catalyzes the chemical reaction by which products are formed, and then allows the products to dissociate (separate from the enzyme surface). The combination formed by an enzyme and its substrates is called the enzyme–substrate complex.

What is the mechanism of enzyme catalysed reaction? ›

Ans : Enzymes can catalyse reactions through a variety of mechanisms. Some of these include Bond strain: enzymes can destabilise bonds within the substrate. Proximity and orientation: conformational changes in the enzyme upon substrate binding can bring reactive groups closer together or orient them so they can react.

What is the formation of carbon carbon bond? ›

Carbon–Carbon σ-Bond Formation

Traditionally, carbon–carbon bond forming processes have required that both organic fragments be specifically activated. This approach is exemplified by the alkylation of an enolate anion (1) with an alkyl halide (2; equation 1).

Which enzyme will catalyze the cleavage of a carbon carbon bond? ›

Enzymes that catalyze a carbon-carbon bond cleavage are known as lyases.

What is the carbon carbon bond formation reaction? ›

Some examples of reactions which form carbon–carbon bonds are aldol reactions, Diels–Alder reaction, the addition of a Grignard reagent to a carbonyl group, a Heck reaction, a Michael reaction and a Wittig reaction.

Top Articles
Latest Posts
Article information

Author: Terrell Hackett

Last Updated:

Views: 6216

Rating: 4.1 / 5 (52 voted)

Reviews: 83% of readers found this page helpful

Author information

Name: Terrell Hackett

Birthday: 1992-03-17

Address: Suite 453 459 Gibson Squares, East Adriane, AK 71925-5692

Phone: +21811810803470

Job: Chief Representative

Hobby: Board games, Rock climbing, Ghost hunting, Origami, Kabaddi, Mushroom hunting, Gaming

Introduction: My name is Terrell Hackett, I am a gleaming, brainy, courageous, helpful, healthy, cooperative, graceful person who loves writing and wants to share my knowledge and understanding with you.