Next Article in Journal
The Potential Effects of Light Irradiance in Glaucoma and Photobiomodulation Therapy
Next Article in Special Issue
A Single Active-Site Mutagenesis Confers Enhanced Activity and/or Changed Product Distribution to a Pentalenene Synthase from Streptomyces sp. PSKA01
Previous Article in Journal
Engineering In Situ Weldable Vascular Devices
Previous Article in Special Issue
Development of a Transcriptional Factor PuuR-Based Putrescine-Specific Biosensor in Corynebacterium glutamicum
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

N-Amidation of Nitrogen-Containing Heterocyclic Compounds: Can We Apply Enzymatic Tools?

1
State Key Laboratory Base of Eco-Chemical Engineering, College of Chemical Engineering, Qingdao University of Science and Technology, Qingdao 266042, China
2
CAS Key Laboratory of Bio-Based Materials, Qingdao Institute of Bioenergy and Bioprocess Technology, Chinese Academy of Sciences, Qingdao 266101, China
3
Institute of Corrosion Science and Technology, Guangzhou 510535, China
*
Authors to whom correspondence should be addressed.
Bioengineering 2023, 10(2), 222; https://doi.org/10.3390/bioengineering10020222
Submission received: 7 December 2022 / Revised: 20 January 2023 / Accepted: 31 January 2023 / Published: 7 February 2023

Abstract

:
Amide bond is often seen in value-added nitrogen-containing heterocyclic compounds, which can present promising chemical, biological, and pharmaceutical significance. However, current synthesis methods in the preparation of amide-containing N-heterocyclic compounds have low specificity (large amount of by-products) and efficiency. In this study, we focused on reviewing the feasible enzymes (nitrogen acetyltransferase, carboxylic acid reductase, lipase, and cutinase) for the amidation of N-heterocyclic compounds; summarizing their advantages and weakness in the specific applications; and further predicting candidate enzymes through in silico structure-functional analysis. For future prospects, current enzymes demand further engineering and improving for practical industrial applications and more enzymatic tools need to be explored and developed for a broader range of N-heterocyclic substrates.

1. Introduction

Nitrogen-containing heterocycles contain at least one nitrogen in their ring framework, most of which exhibits important biological and pharmacological properties [1]. N-heterocycle structure presents diversified inter-molecular interactions, such as hydrogen bonding, dipole–dipole interaction, π-stacking interaction, hydrophobic interaction, etc.; thus, many small molecule drugs containing N-heterocycle structure to interact with targeting enzymes or receptors [2]. According to the registration information by the Food and Drug Administration (FDA), N-heterocycles drugs accounted for more than 59% of all approved drugs in 2014, and more than 80% of drugs containing one or more N-heterocyclic nucleus among the 200 best-selling small molecule drugs in 2018 [3,4].
Another advantage of N-heterocyclic structure is that it can provide an active ligation site to prepare functional derivatives, and N-amidation is quite often seen in the synthesis of N-heterocyclic drugs [5]. According to the data from the FDA, around two-thirds of registered small-molecule drugs present amide bond groups [6]. For example, N-amidation is found in the well-sold N-heterocycles drugs of Lisinopril, Ramipril, Enalapril, Ketoconazole, Loratadine, Sitagliptin, and Doxazosin (Figure 1). Aiming at high conversion rates, low levels of epimerization and limited by-products, N-amidation of heterocyclic compounds still relied on the well-developed chemical methods [7,8], such as acid-catalyzed Schmidt reaction [9,10], Ritter reaction [11], Schotten–Baumann reaction, Passerini reaction [12], Ugi reaction [13,14], and Beckmann rearrangement [15]. More green methods were also developed, which utilized more efficient catalysts [16]. Take ruthenium catalyst as an example; this catalyst can promote the coupling of amines and alcohols in the absence of any acids, bases, and stoichiometric oxidant, producing hydrogen as the only by-product [17]. However, in the preparation of structurally complex chemicals, the majority of these chemical methods have vast amounts of byproducts to present overall low atomic economy [18,19,20]. For example, the preparation of 1 kg of the anti-HIV drug Fuzeon (containing amide bond) through chemical process requires 45 kg of raw materials, excluding the solvents used in the synthesis and purification steps [17]. Thus, recent enzymatic methods with higher selectivity and efficiency were fast developed [20]. This study reviewed the current chemical and enzymatic methods for the N-amidation of nitrogen-containing heterocyclic compounds and predicted potential enzymatic tools, which can be applied in this area through in silico analysis.

2. Chemical Methods

2.1. N-Heterocycles Production by Metal Catalysts

Metal-catalyzed cross-coupling reactions are quite often utilized in the preparation of complex N-heterocycles [21]. For example, a study synthesized carbazoles from anilides and boronic acids with NNO pincer-type Pd(II) complex B catalyst under aerobic condition in aqueous solution. This green method highlights the advantage of a recyclable catalyst, wide substrate range, one-pot process, high yield, and low catalyst stacking. The minimum loading of the catalytic rate can reach 0.0001 mol%, the catalyst can still be recycled six times under this circumstance [22]. The detailed information is summarized in Figure 2A.
Another study utilized the cooperative Cu(II)/Fe(II) catalyst, which could catalyze the Ritter-type C−H activation followed by amination to produce a series of amide-containing N-heterocycles [23]. This method is applicable to a wide range of substrates that enable the rapid synthesis of amide ligation on nitrogen-containing heterocyclic compounds, including co-synthetic precursors for two anxiolytic drugs: Pagoclone and Pazinaclone. The detailed information is summarized in Figure 2B.
Enantioselective metal catalyst was also developed. For example, chiral catalysts were applied in the enantioselective addition reaction for the synthesis of hemiaminal intermediates, and the intramolecular cyclization can process under mildly basic conditions with excellent enantioselectivities and high yields [24]. The detailed information is summarized in Figure 2C.

2.2. Transition Metal Catalysis

Transition metal catalysts were applied in amides forming from alkenes, amines and CO, with no stoichiometric by-products. Based on the pioneering studies [25,26], a recent study established a cooperative catalytic system with transition metal catalysts [27], and the method would synthesis target products in mild conditions. The detailed information is summarized in Figure 2D.

2.3. Cycloaddition Methods

Cycloaddition is another popular strategy in the preparation of N-heterocycles. Cycloaddition prefers Lewis acid/base or chiral catalysts other than metal catalysts. For example, a diarylprolinol silyl ether catalyst was applied as chiral catalyst in the aza-Diels–Alder reaction to produce bicyclic aza-heterocyclic products with high enantioselectivity, [28]. However, one of the weaknesses of the aza-Diels–Alder cycloaddition is that it usually demands chiral precursors to promote the cycloaddition [29]. The detailed information is summarized in Figure 2E.

2.4. Photocatalytic Methods

More recent photocatalytic protocols were applied in N-amidation reaction through radical aminocarbonylation. For example, alkyl amides and amide-containing N-heterocycles could be prepared by metal photoredox catalysts, and the reaction can occur under visible light irradiation [30,31]. Non-metal organic photoredox catalysts were also developed. For example, organophotoredox catalyst eosin Y was applied in the synthesis of alkyl N-amides from thioamides [32]. The detailed information is summarized in Figure 2F,G, respectively.

3. Enzymatic Tools

3.1. Nitrogen Acetyltransferase

In the area of synthetic chemistry, alternative enzymatic tools are in fast development, with the advantages of high selectivity and enantioselectivity under mild conditions [33,34,35]. Some enzymes were found to catalyze the N-amidation reaction for N-heterocycle substrates. Nitrogen acetyltransferase (N-acetyltransferase) is one of the enzymatic tools known to catalyze the amidation reaction towards a broader range of substrates, from small molecules, such as aminoglycoside, to macromolecules of various proteins [36]. Earlier study found that one N-acetyltransferase (pyrB) can catalyze the N-acetylation on small N-heterocyclic compounds (imidazole) in which acetyl coenzyme A was utilized as acetyl donor [37]. A recent study reported that the N-acetyltransferase FbsK can catalyze the N-acetylation in the biosynthesis of fimsbactin (having N-amide heterocyclic structure) [38].
However, N-acetyltransferase has limitation in the in vitro system, as the acyl donor (acetyl coenzyme A or succinyl coenzyme A) is not an economic raw material for the in vitro production. The second limitation of N-acetyltransferase is that it can only catalyze acetylation (acetyl coenzyme A as acyl donor) on or succinylation (succinyl coenzyme A as acyl donor) during amide bond formation.

3.2. Carboxylic Acid Reductase

Carboxylic acid reductase (CAR) is another well-known biocatalytic tool, especially in the biosynthesis of aldehydes, alcohols, and alkanes [39]. Recent studies found that CAR can be applied in the formation of amide bonds, including the N-amidation, towards a broader range of substrates including N-heterocycles (piperidine, piperazine, homopiperazine) [40,41,42]. Different with N-acetyltransferase, which utilizes acetyl-coenzyme A in formation of amide structures, CAR utilizes carboxylic acid to start the bio-conversion. In detail, CARs consist of three parts: core-domain (A-domain), peptidyl carrier protein (PCP), and reduction domain (R-domain). In A-domain, carboxylic acid is activated by ATP to form acyl-AMP complex and then is further converted to acyl-thioester by phosphopantetheine thiol (Figure 3). If acyl-thioester is not transferred to R-domain for the biosynthesis of aldehyde (after reduction by NADH), it can react with the nucleophilic reagent, such as amine, to generate target amides. Thus, engineered CAR (only contains A-domain) is often utilized in amides formation [39,40]. For example, engineered CARmm-A (from Mycobacterium maritimum) can be utilized in vitro to form amides (including N-heterocycles) from a broader range of substrates. As ATP is demanded to activate CAR, class III polyphosphate (CHU) is included in the in vitro system to regenerate ATP [43,44]. At optimized conditions (175 mM piperazine or homopiperazine, 1 mM trans-cinnamic acid, 3 mg/mL CARmm-A lysate, 100 mM MgCl2, and CHU as ATP circulatory system), the yield of corresponding amides can reach 99%. When the amount of acyl donor (carboxylic acid) increased, the yield reduced to 74% [43,44].
The catalytic mechanism of CARs was explored in recent studies. In general, it is believed that acyl-AMP is important for nucleophilic attack during the catalytic process [45]; however, the exact mechanism of CARs in amide forming or lactamization needs to be further investigated.
Comparing with N-acetyltransferase, CAR has several advantages: (1) can be utilized in the in vitro system; and (2) CAR can recognize a broader range of substrates, including long chain fatty acids [46]. However, the practical studies for the amidation reaction by CAR is limited, and further research is demanded in the area [43].

3.3. Lipase and Cutinase

Lipase and cutinase all belong to the α/β-hydrolase superfamily, their typical catalytic function is to form or hydrolyze ester bond [33,47,48]. Lipase is widely applied in the industry due to its temperature, solvent tolerance, and broad catalytic properties, and amide bond lysis/formation is one of its untypical properties [33]. Similarly, cutinase can also be applied for the synthesis or hydrolysis of amides [40,48]. Comparing with lipase, cutinase has the same Ser-His-Asp catalytic triad [49], but does not have the hydrophobicity lip [50], thus the active serine residue may be directly exposed to environments for broader substrates.
Thermo and solvent stability is one of advantages of lipases and cutinases as amide-forming catalysts. The stability may benefit from the non-polar hydrophobic interactions between protein side chains with resistance to the breaking of hydrogen bonds in their three dimensional structures [51]. The zinc binding sites, which are also responsible for the enhanced thermal stability [52]. For example, Bacillus lipase shows excellent stability in hydrophobic organic solvents, even in 10–50% (v/v) of short chain alkanes, benzene and toluene [53]. Fusarium solani pisi cutinase can tolerate high temperature (70 °C) and organic solvent [54]. In addition, structures and mechanisms of lipase and cutinase were intensively studied [55,56] as referable factors for further modification and engineering. The commercial accessibility, broad substrate specificity, broad reaction type, and highly selective and specific absence of any cofactor dependence, could work in mild conditions [57,58,59].
Cutinase from different sources present high diversity. For example, fungal cutinase has Mw of approximately 22–26 kDa, with an optimum pH of 10 and optimum temperature of 30–40 °C. While bacteria cutinase has larger Mw (about 30 kDa), with an optimum pH of 8.5 to 10.5 and a higher thermal stability than fungal cutinase [60], it was indicated that bacterial cutinase may be preferable in amide forming application.
In the amide forming process, different with N-acetyltransferase and CAR, which demand CoA or ATP to activate the acyl donor during amide bond formation, lipase and cutinase can catalyze the amidation through the acyl-enzyme intermediate. During biocatalytic amidation by lipase or cutinase (Figure 4), acyl donor (carboxylic acid, ester, and vinyl acetate) firstly interacts with the Ser-His-Asp catalytic triad to form tetrahedral intermediate; then acyl group ligates the serine residue to form acyl-enzyme intermediate; at last, the acyl-enzyme intermediate can be attacked by nucleophile amine to form an amide bond [61,62].
In practical application, Candida antarctica lipase A (CAL-A) could catalyze the N- acetylation on methyl 2-piperidinecarboxylate in the persistent of PrCO2CH2CF3 and TBME [63]. CAL-A (Mw 45 kDa, pI 7.5) has a large catalytic pocket with the catalytic triad of Ser184-Asp334-His366, the oxyanion hole (Gly185 and Asp95), and hydrophobic lid (residues 217–308). CAL-A exhibits excellent thermal stability (up to 90 °C) and the lid may be important for the acylation behavior of CAL-A [64,65,66].
Although the N-amidation activities of lipase and cutinase were not reported for nitrogen-containing heterocyclic substrates, this study speculated a lipase/cutinase catalyzed amidation process towards piperazine (Figure 4). The testing experiments are in process by our lab and will be published later.
As the formation of acyl-enzyme intermediate is important for the amidation reaction by lipase and cutinase, this study further in silico analyzed the known lipases and cutinases by using the AutoDock tool [67] to screen candidate lipases or cutinases, which can form acyl-enzyme intermediates with the active acyl donor of vinyl acetate (Figure 5). Some positive results are summarized in Figure 5 and Table 1.

4. Concluding Remarks and Future Prospects

In summary, this study demonstrates the feasibility of enzymatic tools (nitrogen acetyltransferase, carboxylic acid reductase, lipase, and cutinase) for selective amide bond formation towards N-heterocyclic compounds under mild conditions, which may present an advantage towards currently known chemical methods (Table 2).
The enzymatic method provides a highly selective alternative to current synthesis methods of amide-containing N-heterocycles from variable acyl donors and amines without side reactions observed in chemical methods. The enzymatic method provides a valuable starting point, although candidate enzymes have limitations in current applications in organic synthesis: (1) nitrogen acetyltransferase and CAR demand co-enzyme or co-factor (CoA and ATP) in amidation; (2) lipase and cutinase may have side-reactions (ester-bond formation), and their amide forming activities towards N-heterocyclic substrates demand further testing.
To improve the current known enzymatic tools, the candidate enzymes demand further engineering and modification for practical applications in organic synthesis and other relevant fields [79,80]. For example, although lipase is one of the industrial enzymatic tools in organic synthesis with the advantages of stereoselectivity and resistance to extreme reaction conditions [65], the binding pocket of natural lipase demands further engineering for the target N-heterocyclic substrates for efficient amidation reaction [56]. One candidate strategy is to replace the residues of long chains with the residues of shorter chains engineering the substrate binding pocket. As seen in the engineering of lipase (PDB: 1CEX), which expands the workable substrates [81]. On the contrary, if residues with larger side chains in binding pockets are engineered, the substrate range will be decreased [82]. Considering the properties of N-heterocyclic substrate, another candidate engineering strategy is to include aromatic residues in the binding pocket, thus increasing the π–π stacking interaction with N-heterocyclic substrate [83]. Fusion strategy can also be utilized in engineering robust amidation bio-tools. For example, the ICCG cutinase was confused with α-synuclein (αSP) in its C-terminal to facilitate its adhesion to PET [84]. A similar strategy can be applied in engineering evolved cutinase to reduce steric hindrance to N-heterocyclic substrate. Other than the well-known lipase CAL-B, CAL-A is another promising starting lipase for further evolution [85,86]. For further improvement, the acyl donor-binding pocket demands further engineering to increase its acyltransferase activity [87], which is important for the amidation reaction by lipase (Figure 4).
As coenzyme or cofactor are demanded for some enzymatic tools (N-acetyltransferase and CAR) in amide forming process, robust coenzyme or cofactor regeneration systems demand to be developed and applied for these enzymatic tools. Similar with the ATP regeneration system (CHU) in the CAR catalyzing process [43,44], acyl coenzyme A regeneration system is demanded for future application of N-acetyltransferase in the N-amidation of N-heterocyclic compounds. A flexible polyphosphate-driven regeneration system may be suitable to perform this work [88]. The system has the advantage of recognizing a broader range of acyl donors to regenerate corresponding acyl coenzyme A, thus both acetyl coenzyme A and succinyl coenzyme A can be regenerated for amidation reaction catalyzed by N-acetyltransferase.
For future prospects, more amide ligating enzymes need to be screened for their potential application in amidation of a broader range of N-heterocyclic substrates. Similar with CAR, major amide ligases are ATP-dependent [62,89], they work in aqueous condition and demand ATP to form acyl-adenylate intermediates to activate amide bond formation. Amino acid ligases and non-ribosomal peptide synthetases are representatives for this group and majorities of them were approved in the biosynthesis of natural metabolites, such as nikkomycin [90], bacilysin [91], tabtoxin [92], rhizocticin [93], and shinorine [94]. Another group of amide-forming biocatalysts are ATP-independent serine proteases (including transpeptidases) and lipases [61,62,95,96], which do not demand costly cofactors and can perform a broader range of amide-forming reactions. For example, transpeptidases can incorporate exogenous building blockings into peptidoglycan via forming fresh amide ligation [95,96]. The strategy may be applicable once we develop novel enzymatic tools and design novel synthetic routes towards amide-containing N-heterocyclic compounds.
However, majorities of these amide-forming enzymes (amino acid ligases, peptide synthetases, and proteases) have a limited substrate scope, and their catalytic feasibilities towards N-heterocyclic substrates demand further research, judgment, and improvement. In view of industrial and scale-up amidation techniques, the enzymatic tools and methods must compare with corresponding chemical tools and methods [97], of their cost, efficiency, safety, and toxicity. It is generally accepted that enzymatic tools may have advantages when amidation steps are required in the in vitro production of structurally complex natural products.
In conclusion, limited enzymatic tools were reported in the preparation of amide-containing N-heterocyclic compounds. With the fast development of enzyme engineering and biotechnology, more robust enzymes will be discovered and engineered to improve their performance and application, which in turn will reduce the barrier and cost in design and produce value-added N-heterocyclic compounds.

Author Contributions

A.Y., X.M. and L.Y., writing—original draft preparation; C.X., M.X., W.L. and H.Z.; writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Talent Foundation funded by the Province and Ministry Co-construction Collaborative Innovation Center of Eco-chemical Engineering [STHGYX2221]; National Key R&D Program of China [2022YFC2104703]; and Shandong Province Natural Science Foundation [ZR2022MB014].

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, F.; Yao, Y.; Zhu, H.; Zhang, Y. Nitrogen-containing Heterocycle: A Privileged Scaffold for Marketed Drugs. Curr. Top. Med. Chem. 2021, 21, 439–441. [Google Scholar] [CrossRef]
  2. Kerru, N.; Gummidi, L.; Maddila, S.; Gangu, K.K.; Jonnalagadda, S. A Review on Recent Advances in Nitrogen-Containing Molecules and Their Biological Applications. Molecules 2020, 25, 1909. [Google Scholar] [CrossRef] [PubMed]
  3. Jagadeesh, R.V.; Murugesan, K.; Alshammari, A.S.; Neumann, H.; Pohl, M.-M.; Radnik, J.; Beller, M. MOF-derived cobalt nanoparticles catalyze a general synthesis of amines. Science 2017, 358, 326–332. [Google Scholar] [CrossRef] [PubMed]
  4. Vitaku, E.; Smith, D.T.; Njardarson, J.T. Analysis of the Structural Diversity, Substitution Patterns, and Frequency of Nitrogen Heterocycles among U.S. FDA Approved Pharmaceuticals. J. Med. Chem. 2014, 57, 10257–10274. [Google Scholar] [CrossRef]
  5. Li, G.; Szostak, M. Highly selective transition-metal-free transamidation of amides and amidation of esters at room temperature. Nat. Commun 2018, 9, 4165. [Google Scholar] [CrossRef]
  6. Roughley, S.D.; Jordan, A.M. The Medicinal Chemist's Toolbox: An Analysis of Reactions Used in the Pursuit of Drug Candidates. J. Med. Chem. 2011, 42, 3451–3479. [Google Scholar] [CrossRef]
  7. Valeur, E.; Bradley, M. Amide bond formation: Beyond the myth of coupling reagents. Chem. Soc. Rev. 2009, 38, 606–631. [Google Scholar] [CrossRef]
  8. Carey, J.S.; Laffan, D.; Thomson, C.; Williams, M.T. Analysis of the reactions used for the preparation of drug candidate molecules. Org. Biomol. Chem. 2006, 4, 2337–2347. [Google Scholar] [CrossRef]
  9. Schuerch, C.; Huntress, E.H. The Schmidt Reaction. I. Conditions and Reaction Mechanism with Primary, Secondary and Tertiary Aliphatic Acids. J. Am. Chem. Soc. 1949, 71, 2233–2237. [Google Scholar] [CrossRef]
  10. Smith, P.A.S. The Schmidt Reaction: Experimental Conditions and Mechanism. J. Am. Chem. Soc. 1948, 70, 320–323. [Google Scholar] [CrossRef]
  11. Ritter, J.J.; Minieri, P.P. A new reaction of nitriles; amides from alkenes and mononitriles. J. Am. Chem. Soc. 1948, 70, 4045–4048. [Google Scholar] [CrossRef]
  12. Soeta, T.; Kojima, Y.; Ukaji, Y.; Inomata, K. O-Silylative Passerini Reaction: A New One-Pot Synthesis of alpha-Siloxyamides. Org. Lett. 2010, 12, 4341–4343. [Google Scholar] [CrossRef]
  13. Ugi, I. The α-Addition of Immonium Ions and Anions to Isonitriles Accompanied by Secondary Reactions. Angew. Chem. Int. Ed. 1962, 1, 8–21. [Google Scholar] [CrossRef]
  14. Xu, Z.G.; Ding, Y.; Meng, J.P.; Tang, D.Y.; Li, Y.; Lei, J.; Xu, C.; Chen, Z.Z. Facile Construction of Hydantoin Scaffolds via a Post-Ugi Cascade Reaction. Synlett 2018, 29, 2199–2202. [Google Scholar] [CrossRef]
  15. Eaton, P.E.; Carlson, G.R.; Lee, J.T. Phosphorus pentoxide-methanesulfonic acid. Convenient alternative to polyphosphoric acid. J. Org. Chem. 1973, 38, 4071–4073. [Google Scholar] [CrossRef]
  16. Todorovic, M.; Perrin, D.M. Recent developments in catalytic amide bond formation. Peptide Science 2020, 112, e24210. [Google Scholar] [CrossRef]
  17. Pattabiraman, V.R.; Bode, J.W. Rethinking amide bond synthesis. Nature 2011, 480, 471–479. [Google Scholar] [CrossRef]
  18. de Figueiredo, R.M.; Suppo, J.-S.; Campagne, J.-M. Nonclassical Routes for Amide Bond Formation. Chem. Rev. 2016, 116, 12029–12122. [Google Scholar] [CrossRef]
  19. Santos, A.S.; Silva, A.M.S.; Marques, M.M.B. Sustainable Amidation Reactions – Recent Advances. Eur. J. Org. Chem. 2020, 2020, 2501–2516. [Google Scholar] [CrossRef]
  20. Petchey, M.R.; Grogan, G. Enzyme-Catalysed Synthesis of Secondary and Tertiary Amides. Adv. Synth. Catal. 2019, 361, 3895–3914. [Google Scholar] [CrossRef]
  21. Taylor, A.P.; Robinson, R.P.; Fobian, Y.M.; Blakemore, D.C.; Jones, L.H.; Fadeyi, O. Modern advances in heterocyclic chemistry in drug discovery. Org. Biomol. Chem. 2016, 14, 6611–6637. [Google Scholar] [CrossRef] [PubMed]
  22. Arumugam, V.; Kaminsky, W.; Nallasamy, D. Pd(ii) pincer type complex catalyzed tandem C–H and N–H activation of acetanilide in aqueous media: A concise access to functionalized carbazoles in a single step. Green Chem. 2016, 18, 3295–3301. [Google Scholar] [CrossRef]
  23. Liu, C.; Zhang, Q.; Li, H.; Guo, S.; Xiao, B.; Deng, W.; Liu, L.; He, W. Cu/Fe Catalyzed Intermolecular Oxidative Amination of Benzylic C−H Bonds. Chem. Eur. J. 2016, 22, 6208–6212. [Google Scholar] [CrossRef] [PubMed]
  24. Nimmagadda, S.K.; Zhang, Z.; Antilla, J.C. Asymmetric One-Pot Synthesis of 1,3-Oxazolidines and 1,3-Oxazinanes via Hemiaminal Intermediates. Org. Lett. 2014, 16, 4098–4101. [Google Scholar] [CrossRef] [PubMed]
  25. Cheng, J.; Qi, X.; Li, M.; Chen, P.; Liu, G. Palladium-Catalyzed Intermolecular Aminocarbonylation of Alkenes: Efficient Access of β-Amino Acid Derivatives. J. Am. Chem. Soc. 2015, 137, 2480–2483. [Google Scholar] [CrossRef]
  26. Fang, X.; Li, H.; Jackstell, R.; Beller, M. Selective Palladium-Catalyzed Aminocarbonylation of 1,3-Dienes: Atom-Efficient Synthesis of β,γ-Unsaturated Amides. J. Am. Chem. Soc. 2014, 136, 16039–16043. [Google Scholar] [CrossRef]
  27. Zhang, G.; Gao, B.; Huang, H. Palladium-catalyzed hydroaminocarbonylation of alkenes with amines: A strategy to overcome the basicity barrier imparted by aliphatic amines. Angew. Chem. Int. Ed 2015, 54, 7657–7661. [Google Scholar] [CrossRef]
  28. Li, Y.; Barløse, C.; Jørgensen, J.; Carlsen, B.D.; Jørgensen, K.A. Asymmetric Catalytic Aza-Diels–Alder/Ring-Closing Cascade Reaction Forming Bicyclic Azaheterocycles by Trienamine Catalysis. Chem. Eur. J. 2017, 23, 38–41. [Google Scholar] [CrossRef]
  29. Vinogradov, M.G.; Turova, O.V.; Zlotin, S.G. Catalytic Asymmetric Aza-Diels-Alder Reaction: Pivotal Milestones and Recent Applications to Synthesis of Nitrogen-Containing Heterocycles. Adv. Synth. Catal. 2021, 363, 1466–1526. [Google Scholar] [CrossRef]
  30. Chow, S.Y.; Stevens, M.Y.; Åkerbladh, L.; Bergman, S.; Odell, L.R. Mild and Low-Pressure fac-Ir(ppy)3-Mediated Radical Aminocarbonylation of Unactivated Alkyl Iodides through Visible-Light Photoredox Catalysis. Chem. Eur. J. 2016, 22, 9155–9161. [Google Scholar] [CrossRef]
  31. Veatch, A.M.; Alexanian, E.J. Cobalt-catalyzed aminocarbonylation of (hetero)aryl halides promoted by visible light. Chem. Sci 2020, 11, 7210–7213. [Google Scholar] [CrossRef]
  32. Yadav, A.K.; Srivastava, V.P.; Yadav, L.D.S. Visible-light-mediated eosin Y catalyzed aerobic desulfurization of thioamides into amides. New J. Chem. 2013, 37, 4119–4124. [Google Scholar] [CrossRef]
  33. Ismail, A.R.; Kashtoh, H.; Baek, K.H. Temperature-resistant and solvent-tolerant lipases as industrial biocatalysts: Biotechnological approaches and applications. Int. J. Biol. Macromol. 2021, 187, 127–142. [Google Scholar] [CrossRef]
  34. Pellis, A.; Cantone, S.; Ebert, C.; Gardossi, L. Evolving biocatalysis to meet bioeconomy challenges and opportunities. N. Biotechnol. 2018, 40, 154–169. [Google Scholar] [CrossRef]
  35. Sharma, S.; Das, J.; Braje, W.M.; Dash, A.K.; Handa, S. A Glimpse into Green Chemistry Practices in the Pharmaceutical Industry. ChemSusChem 2020, 13, 2859–2875. [Google Scholar] [CrossRef]
  36. Zhang, P.; Liu, P.; Xu, Y.; Liang, Y.; Wang, P.G.; Cheng, J. N-acetyltransferases from three different organisms displaying distinct selectivity toward hexosamines and N-terminal amine of peptides. Carbohydr. Res. 2019, 472, 72–75. [Google Scholar] [CrossRef]
  37. Kinsky, S.C. Assay, Purification, and Properties of Imidazole Acetylase. J. Biol. Chem. 1960, 235, 94–98. [Google Scholar] [CrossRef]
  38. Yang, J.; Wencewicz, T.A. In Vitro Reconstitution of Fimsbactin Biosynthesis from Acinetobacter baumannii. ACS Chem. Biol. 2022, 17, 2923–2935. [Google Scholar] [CrossRef]
  39. Derrington, S.R.; Turner, N.J.; France, S.P. Carboxylic acid reductases (CARs): An industrial perspective. J. Biotechnol. 2019, 304, 78–88. [Google Scholar] [CrossRef]
  40. Wood, A.J.L.; Weise, N.J.; Frampton, J.D.; Dunstan, M.S.; Hollas, M.A.; Derrington, S.R.; Lloyd, R.C.; Quaglia, D.; Parmeggiani, F.; Leys, D.; et al. Adenylation Activity of Carboxylic Acid Reductases Enables the Synthesis of Amides. Angew. Chem. Int. Ed. 2017, 56, 14498–14501. [Google Scholar] [CrossRef]
  41. Qu, G.; Guo, J.; Yang, D.; Sun, Z. Biocatalysis of carboxylic acid reductases: Phylogenesis, catalytic mechanism and potential applications. Green Chem. 2018, 20, 777–792. [Google Scholar] [CrossRef]
  42. Gahloth, D.; Aleku, G.A.; Leys, D. Carboxylic acid reductase: Structure and mechanism. J. Biotechnol. 2020, 307, 107–113. [Google Scholar] [CrossRef] [PubMed]
  43. Lubberink, M.; Finnigan, W.; Schnepel, C.; Baldwin, C.R.; Turner, N.J.; Flitsch, S.L. One-Step Biocatalytic Synthesis of Sustainable Surfactants by Selective Amide Bond Formation. Angew. Chem. Int. Ed 2022, 61, e202205054. [Google Scholar] [CrossRef] [PubMed]
  44. Lubberink, M.; Schnepel, C.; Citoler, J.; Derrington, S.R.; Finnigan, W.; Hayes, M.A.; Turner, N.J.; Flitsch, S.L. Biocatalytic Monoacylation of Symmetrical Diamines and Its Application to the Synthesis of Pharmaceutically Relevant Amides. ACS Catalysis 2020, 10, 10005–10009. [Google Scholar] [CrossRef]
  45. Qin, Z.; Zhang, X.; Sang, X.; Zhang, W.; Qu, G.; Sun, Z. Carboxylic acid reductases enable intramolecular lactamization reactions. Green Synth. Catal. 2022, 3, 294–297. [Google Scholar] [CrossRef]
  46. Akhtar, M.K.; Turner, N.J.; Jones, P.R. Carboxylic acid reductase is a versatile enzyme for the conversion of fatty acids into fuels and chemical commodities. Proc. Natl. Acad. Sci. USA 2013, 110, 87–92. [Google Scholar] [CrossRef]
  47. Roussel, A.; Amara, S.; Nyyssölä, A.; Mateos-Diaz, E.; Blangy, S.; Kontkanen, H.; Westerholm-Parvinen, A.; Carrière, F.; Cambillau, C. A Cutinase from Trichoderma reesei with a Lid-Covered Active Site and Kinetic Properties of True Lipases. J. Mol. Biol. 2014, 426, 3757–3772. [Google Scholar] [CrossRef]
  48. Ferrario, V.; Pellis, A.; Cespugli, M.; Guebitz, G.M.; Gardossi, L. Nature Inspired Solutions for Polymers: Will Cutinase Enzymes Make Polyesters and Polyamides Greener? Catalysts 2016, 6, 205. [Google Scholar] [CrossRef]
  49. Chen, S.; Su, L.; Chen, J.; Wu, J. Cutinase: Characteristics, preparation, and application. Biotechnol. Adv. 2013, 31, 1754–1767. [Google Scholar] [CrossRef]
  50. Martinez, C.; Degeus, P.; Lauwereys, M.; Matthyssens, G.; Cambillau, C. Fusarium solani cutinase is a lipolytic enzyme with a catalytic serine accessible to solvent. Nature 1992, 356, 615–618. [Google Scholar] [CrossRef]
  51. Wang, W.; Nema, S.; Teagarden, D. Protein aggregation--pathways and influencing factors. Int. J. Pharm. 2010, 390, 89–99. [Google Scholar] [CrossRef]
  52. Tyndall, J.; Sinchaikul, S.; Fothergill-Gilmore, L.; Taylor, P.; Walkinshaw, M. Crystal Structure of a Thermostable Lipase from Bacillus stearothermophilus P1. J. Mol. Biol. 2002, 323, 859–869. [Google Scholar] [CrossRef]
  53. Kumar, A.; Dhar, K.; Kanwar, S.S.; Arora, P.K. Lipase catalysis in organic solvents: Advantages and applications. Biol. Proced. Online 2016, 18, 2. [Google Scholar] [CrossRef]
  54. Stavila, E.; Arsyi, R.Z.; Petrovic, D.M.; Loos, K. Fusarium solani pisi cutinase-catalyzed synthesis of polyamides. Eur. Polym. J. 2013, 49, 834–842. [Google Scholar] [CrossRef]
  55. Nardini Marco, D.B.W. α/β Hydrolase fold enzymes: The family keeps growing. Curr. Opin. Struct. Biol. 1999, 9, 732–737. [Google Scholar] [CrossRef]
  56. Pleiss, J.; Fischer, M.; Schmid, R.D. Anatomy of lipase binding sites: The scissile fatty acid binding site. Chem. Phys. Lipids 1998, 93, 67–80. [Google Scholar] [CrossRef]
  57. Alfonso, I.; Gotor, V. Biocatalytic and biomimetic aminolysis reactions: Useful tools for selective transformations on polyfunctional substrates. Chem. Soc. Rev. 2004, 33, 201–209. [Google Scholar] [CrossRef]
  58. Al-Ghanayem, A.A.; Joseph, B.; Alhussaini, M.S.; Ramteke, P.W. Microbial Extremozymes; Academic Press: Cambridge, MA, USA, 2022; pp. 223–230. ISBN 978-0-12-822945-3. [Google Scholar]
  59. Salgado, C.A.; dos Santos, C.I.A.; Vanetti, M.C.D. Microbial lipases: Propitious biocatalysts for the food industry. Food Bioscience 2022, 45, 101509. [Google Scholar] [CrossRef]
  60. Liang, X.; Zou, H. Biotechnological Application of Cutinase: A Powerful Tool in Synthetic Biology. SynBio 2023, 1, 54–64. [Google Scholar] [CrossRef]
  61. Castillo, E.; Casas-Godoy, L.; Sandoval, G. Medium-engineering: A useful tool for modulating lipase activity and selectivity. Biocatalysis 2016, 1, 178–188. [Google Scholar] [CrossRef]
  62. Goswami, A.; Van Lanen, S.G. Enzymatic strategies and biocatalysts for amide bond formation: Tricks of the trade outside of the ribosome. Mol. Biosyst. 2015, 11, 338–353. [Google Scholar] [CrossRef] [PubMed]
  63. Liljeblad, A.; Kallio, P.; Vainio, M.; Niemi, J.; Kanerva, L.T. Formation and hydrolysis of amide bonds by lipase A from Candida antarctica; exceptional features. Org. Biomol. Chem. 2010, 8, 886–895. [Google Scholar] [CrossRef] [PubMed]
  64. Brundiek, H.; Padhi, S.K.; Kourist, R.; Evitt, A.; Bornscheuer, U.T. Altering the scissile fatty acid binding site of Candida antarctica lipase A by protein engineering for the selective hydrolysis of medium chain fatty acids. Eur. J. Lipid Sci. Technol. 2012, 114, 1148–1153. [Google Scholar] [CrossRef]
  65. Lima, R.N.; dos Anjos, C.S.; Orozco, E.V.M.; Porto, A.L.M. Versatility of Candida antarctica lipase in the amide bond formation applied in organic synthesis and biotechnological processes. Mol. Catal. 2019, 466, 75–105. [Google Scholar] [CrossRef]
  66. Monteiro, R.R.C.; Virgen-Ortiz, J.J.; Berenguer-Murcia, Á.; da Rocha, T.N.; dos Santos, J.C.S.; Alcántara, A.R.; Fernandez-Lafuente, R. Biotechnological relevance of the lipase A from Candida antarctica. Catal. Today 2021, 362, 141–154. [Google Scholar] [CrossRef]
  67. Morris, G.M.; Huey, R.; Lindstrom, W.; Sanner, M.F.; Belew, R.K.; Goodsell, D.S.; Olson, A.J. AutoDock4 and AutoDockTools4: Automated Docking with Selective Receptor Flexibility. J. Comput. Chem. 2009, 30, 2785–2791. [Google Scholar] [CrossRef]
  68. Guex, N.; Peitsch, M.C.; Schwede, T. Automated comparative protein structure modeling with SWISS-MODEL and Swiss-PdbViewer: A historical perspective. Electrophoresis 2009, 30, S162–S173. [Google Scholar] [CrossRef]
  69. Waterhouse, A.; Bertoni, M.; Bienert, S.; Studer, G.; Tauriello, G.; Gumienny, R.; Heer, F.T.; de Beer, T.A.P.; Rempfer, C.; Bordoli, L.; et al. SWISS-MODEL: Homology modelling of protein structures and complexes. Nucleic Acids Res. 2018, 46, W296–W303. [Google Scholar] [CrossRef]
  70. Kitadokoro, K.; Kakara, M.; Matsui, S.; Osokoshi, R.; Thumarat, U.; Kawai, F.; Kamitani, S. Structural insights into the unique polylactate-degrading mechanism of Thermobifida alba cutinase. FEBS J. 2019, 286, 2087–2098. [Google Scholar] [CrossRef]
  71. Nicolas, A.; Egmond, M.; Verrips, C.T.; de Vlieg, J.; Longhi, S.; Cambillau, C.; Martinez, C. Contribution of cutinase serine 42 side chain to the stabilization of the oxyanion transition state. Biochemistry 1996, 35, 398–410. [Google Scholar] [CrossRef]
  72. Flores-Castañón, N.; Sarkar, S.; Banerjee, A. Structural, functional, and molecular docking analyses of microbial cutinase enzymes against polyurethane monomers. J. Hazard. Mater. Lett. 2022, 3, 100063. [Google Scholar] [CrossRef]
  73. Pellis, A.; Ferrario, V.; Hufnagel, B.; Brandauer, M.; Gamerith, C.; Herrero Acero, E.; Ebert, C.; Gardossi, L.; Guebitz, G. Enlarging the tools for efficient enzymatic polycondensation: Structural and catalytic features of cutinase 1 from: Thermobifida cellulosilytica. Catal. Sci. Technol. 2015, 6. [Google Scholar] [CrossRef]
  74. Shirke, A.N.; Basore, D.; Holton, S.; Su, A.; Baugh, E.; Butterfoss, G.L.; Makhatadze, G.; Bystroff, C.; Gross, R.A. Influence of surface charge, binding site residues and glycosylation on Thielavia terrestris cutinase biochemical characteristics. Appl. Microbiol. Biotechnol. 2016, 100, 4435–4446. [Google Scholar] [CrossRef]
  75. Kitadokoro, K.; Thumarat, U.; Nakamura, R.; Nishimura, K.; Karatani, H.; Suzuki, H.; Kawai, F. Crystal structure of cutinase Est119 from Thermobifida alba AHK119 that can degrade modified polyethylene terephthalate at 1.76Å resolution. Polym. Degrad. Stab. 2012, 97, 771–775. [Google Scholar] [CrossRef]
  76. Chen, Z. Selective Mono-Acylation of Piperazine Derivatives with Pseudomonas Stutzeri lipase (PSL). Master’s Thesis, The University of Manchester, Manchester, UK, 2016. [Google Scholar]
  77. Nikolaivits, E.; Kanelli, M.; Dimarogona, M.; Topakas, E. A Middle-Aged Enzyme Still in Its Prime: Recent Advances in the Field of Cutinases. Catalysts 2018, 8, 612. [Google Scholar] [CrossRef]
  78. Austin, H.P.; Allen, M.D.; Donohoe, B.S.; Rorrer, N.A.; Kearns, F.L.; Silveira, R.L.; Pollard, B.C.; Dominick, G.; Duman, R.; El Omari, K.; et al. Characterization and engineering of a plastic-degrading aromatic polyesterase. Proc. Natl. Acad. Sci. USA 2018, 115, E4350–E4357. [Google Scholar] [CrossRef]
  79. Sinha, R.; Shukla, P. Current Trends in Protein Engineering: Updates and Progress. Curr. Protein Peptide Sci. 2019, 20, 398–407. [Google Scholar] [CrossRef]
  80. Chen, H.; Ma, L.; Dai, H.; Fu, Y.; Wang, H.; Zhang, Y. Advances in Rational Protein Engineering toward Functional Architectures and Their Applications in Food Science. J. Agric. Food Chem. 2022, 70, 4522–4533. [Google Scholar] [CrossRef]
  81. Araújo, R.; Silva, C.; O'Neill, A.; Micaelo, N.; Guebitz, G.; Soares, C.M.; Casal, M.; Cavaco-Paulo, A. Tailoring cutinase activity towards polyethylene terephthalate and polyamide 6,6 fibers. J. Biotechnol. 2007, 128, 849–857. [Google Scholar] [CrossRef]
  82. Lan, D.; Xu, H.; Xu, J.; Dubin, G.; Liu, J.; Iqbal Khan, F.; Wang, Y. Malassezia globosa MgMDL2 lipase: Crystal structure and rational modification of substrate specificity. Biochem. Biophys. Res. Commun. 2017, 488, 259–265. [Google Scholar] [CrossRef]
  83. Ding, X.; Tang, X.; Zheng, R.; Zheng, Y. Identification and engineering of the key residues at the crevice-like binding site of lipases responsible for activity and substrate specificity. Biotechnol. Lett. 2019, 41, 137–146. [Google Scholar] [CrossRef] [PubMed]
  84. Su, L.; Chen, K.; Bai, S.; Yu, L.; Sun, Y. Cutinase fused with C-terminal residues of α-synuclein improves polyethylene terephthalate degradation by enhancing the substrate binding. Biochem. Eng. J. 2022, 188, 108709. [Google Scholar] [CrossRef]
  85. Kirk, O.; Christensen, M.W. Lipases from Candida antarctica: Unique Biocatalysts from a Unique Origin. Org. Process Res. Dev. 2002, 6, 446–451. [Google Scholar] [CrossRef]
  86. Krishna, S.H.; Persson, M.; Bornscheuer, U.T. Enantioselective transesterification of a tertiary alcohol by lipase A from Candida antarctica. Tetrahedron Asymmetry 2002, 13, 2693–2696. [Google Scholar] [CrossRef]
  87. Muller, J.; Sowa, M.A.; Fredrich, B.; Brundiek, H.; Bornscheuer, U.T. Enhancing the Acyltransferase Activity of Candida antarctica Lipase A by Rational Design. ChemBioChem 2015, 16, 1791–1796. [Google Scholar] [CrossRef]
  88. Mordhorst, S.; Maurer, A.; Popadić, D.; Brech, J.; Andexer, J.N. A Flexible Polyphosphate-Driven Regeneration System for Coenzyme A Dependent Catalysis. ChemCatChem 2017, 9, 4164–4168. [Google Scholar] [CrossRef]
  89. Winn, M.; Richardson, S.M.; Campopiano, D.J.; Micklefield, J. Harnessing and engineering amide bond forming ligases for the synthesis of amides. Curr. Opin. Chem. Biol. 2020, 55, 77–85. [Google Scholar] [CrossRef]
  90. Bruntner, C.; Lauer, B.; Schwarz, W.; Möhrle, V.; Bormann, C. Molecular characterization of co-transcribed genes from Streptomyces tendae Tü901 involved in the biosynthesis of the peptidyl moiety of the peptidyl nucleoside antibiotic nikkomycin. Mol. Gen. Genet. MGG 1999, 262, 102–114. [Google Scholar] [CrossRef]
  91. Tsuda, T.; Asami, M.; Koguchi, Y.; Kojima, S. Single mutation alters the substrate specificity of L-amino acid ligase. Biochemistry 2014, 53 16, 2650–2660. [Google Scholar] [CrossRef]
  92. Kinscherf, T.G.; Willis, D.K. The biosynthetic gene cluster for the beta-lactam antibiotic tabtoxin in Pseudomonas syringae. J. Antibiot. (Tokyo) 2005, 58, 817–821. [Google Scholar] [CrossRef] [Green Version]
  93. Kino, K.; Kotanaka, Y.; Arai, T.; Yagasaki, M. A Novel l-Amino Acid Ligase from Bacillus subtilis NBRC3134, a Microorganism Producing Peptide-Antibiotic Rhizocticin. Biosci. Biotechnol. Biochem. 2009, 73, 901–907. [Google Scholar] [CrossRef]
  94. Gao, Q.; Garcia-Pichel, F. An ATP-grasp ligase involved in the last biosynthetic step of the iminomycosporine shinorine in Nostoc punctiforme ATCC 29133. J. Bacteriol. 2011, 193, 5923–5928. [Google Scholar] [CrossRef]
  95. Lebar, M.D.; May, J.M.; Meeske, A.J.; Leiman, S.A.; Lupoli, T.J.; Tsukamoto, H.; Losick, R.; Rudner, D.Z.; Walker, S.; Kahne, D. Reconstitution of peptidoglycan cross-linking leads to improved fluorescent probes of cell wall synthesis. J. Am. Chem. Soc. 2014, 136, 10874–10877. [Google Scholar] [CrossRef]
  96. Qiao, Y.; Lebar, M.D.; Schirner, K.; Schaefer, K.; Tsukamoto, H.; Kahne, D.; Walker, S. Detection of lipid-linked peptidoglycan precursors by exploiting an unexpected transpeptidase reaction. J. Am. Chem. Soc. 2014, 136, 14678–14681. [Google Scholar] [CrossRef]
  97. Magano, J. Large-Scale Amidations in Process Chemistry: Practical Considerations for Reagent Selection and Reaction Execution. Org Process Res Dev 2022, 26, 1562–1689. [Google Scholar] [CrossRef]
Figure 1. The representative marketing drugs that have amide ligation on the nitrogen-containing heterocycle structure. The blue color indicates the featured structure of each molecule.
Figure 1. The representative marketing drugs that have amide ligation on the nitrogen-containing heterocycle structure. The blue color indicates the featured structure of each molecule.
Bioengineering 10 00222 g001
Figure 2. Details of representative chemical amide forming methods towards N-heterocyclic compounds. (A) acetanilide (1a, 2 mmol), phenylboronic acid (1b, 2 mmol), Pd(II) complexes catalyst (1d, 0.01 mol%), K2CO3 (4 mmol), solvent (water), 60 °C, 8 h, yield of the target product (1c) can reach 93 %; (B) 2-methylbenzoic acid (2a) (27.2 mg, 0.2 mmol, 1 equiv), selectfulor (141.7 mg, 0.4 mmol, 2 equiv), catalyst (20% Fe(acac)3, 1% Cu(NO3)2), Schlenk tube, 80 °C, 8 h, the yield of product (2a) can reach 85%. (C) 3a (1.0 equiv), 3b (2.0 equiv), catalyst (2.5 mol%), solvent (ethyl acetate). After step 1, ethyl acetate is removed; DMF and Cs2CO3 (2.0 equiv) are added. The yield of product (3c) can reach 92%, 95%ee. (D) 4a (0.8 mmol), CO (10 atm), 4b (0.4 mmol), Pd(TFA)2 (0.01 mmol), DPPPen (0.012 mmol), NH2CH2CO2Me·HCl (0.04 mmol), (HCHO)n (0.04 mmol), solvent (anisole, 1.0 mL), 120 °C, 21 h. (E) 5a (0.2 mmol), 5b (0.1 mmol), 5c (0.02 mmol), DABCO (0.02 mmol), solvent (toluene, 0.5 mL), 40 °C. (F) 5 mol % Co2(CO)8, 2 atm CO, 2 equiv TMP, 390 nm LEDs. (G) 7a (1.0 mmol), eosin Y catalyst (2 mol%), solvent (DMF, 3 mL), and green LEDs (2.6 W, 161 lm) irradiation.
Figure 2. Details of representative chemical amide forming methods towards N-heterocyclic compounds. (A) acetanilide (1a, 2 mmol), phenylboronic acid (1b, 2 mmol), Pd(II) complexes catalyst (1d, 0.01 mol%), K2CO3 (4 mmol), solvent (water), 60 °C, 8 h, yield of the target product (1c) can reach 93 %; (B) 2-methylbenzoic acid (2a) (27.2 mg, 0.2 mmol, 1 equiv), selectfulor (141.7 mg, 0.4 mmol, 2 equiv), catalyst (20% Fe(acac)3, 1% Cu(NO3)2), Schlenk tube, 80 °C, 8 h, the yield of product (2a) can reach 85%. (C) 3a (1.0 equiv), 3b (2.0 equiv), catalyst (2.5 mol%), solvent (ethyl acetate). After step 1, ethyl acetate is removed; DMF and Cs2CO3 (2.0 equiv) are added. The yield of product (3c) can reach 92%, 95%ee. (D) 4a (0.8 mmol), CO (10 atm), 4b (0.4 mmol), Pd(TFA)2 (0.01 mmol), DPPPen (0.012 mmol), NH2CH2CO2Me·HCl (0.04 mmol), (HCHO)n (0.04 mmol), solvent (anisole, 1.0 mL), 120 °C, 21 h. (E) 5a (0.2 mmol), 5b (0.1 mmol), 5c (0.02 mmol), DABCO (0.02 mmol), solvent (toluene, 0.5 mL), 40 °C. (F) 5 mol % Co2(CO)8, 2 atm CO, 2 equiv TMP, 390 nm LEDs. (G) 7a (1.0 mmol), eosin Y catalyst (2 mol%), solvent (DMF, 3 mL), and green LEDs (2.6 W, 161 lm) irradiation.
Bioengineering 10 00222 g002
Figure 3. The A-domain of carboxylic acid reductase (CAR) presents amidation activity towards N-heterocycles. The A, R domains and peptidyl carrier protein (PCP) are shown in red, blue, and yellow. In the absence of NADPH, acyl-thioester intermediates react with variable amines (including piperazine) for amide formation in the A domain. On the other hand, the acyl-thioester intermediate is eventually reduced (by NADPH) into aldehyde in the R domain.
Figure 3. The A-domain of carboxylic acid reductase (CAR) presents amidation activity towards N-heterocycles. The A, R domains and peptidyl carrier protein (PCP) are shown in red, blue, and yellow. In the absence of NADPH, acyl-thioester intermediates react with variable amines (including piperazine) for amide formation in the A domain. On the other hand, the acyl-thioester intermediate is eventually reduced (by NADPH) into aldehyde in the R domain.
Bioengineering 10 00222 g003
Figure 4. Mechanism of a speculated lipase/cutinase catalyzed amidation reaction towards piperazine. The Ser-His-Asp catalytic triad center can activate the acyl donor to form a tetrahedral intermediate, which is stabilized by the oxyanion hole. Then, this high-energy transition state product decomposes into acyl-enzyme intermediate. The acyl-enzyme intermediate is further attacked by nucleophilic piperazine to obtain N-amidated piperazine.
Figure 4. Mechanism of a speculated lipase/cutinase catalyzed amidation reaction towards piperazine. The Ser-His-Asp catalytic triad center can activate the acyl donor to form a tetrahedral intermediate, which is stabilized by the oxyanion hole. Then, this high-energy transition state product decomposes into acyl-enzyme intermediate. The acyl-enzyme intermediate is further attacked by nucleophilic piperazine to obtain N-amidated piperazine.
Bioengineering 10 00222 g004
Figure 5. Autodock analysis of representative cutinases, which can form acyl-enzyme intermediates with the acyl donor of vinyl acetate. (A) Docking of vinyl acetate in the structure of cutinase (PDB: 6AID). (B) The primary sequence of one Aspergillus oryzae cutinase (NCBI: OOO09588.1) was modelled in its structure by SWISS-Model [68,69] before docking analysis with vinyl acetate. (C) Docking of vinyl acetate in the structure of cutinase (PDB: 1FFB).
Figure 5. Autodock analysis of representative cutinases, which can form acyl-enzyme intermediates with the acyl donor of vinyl acetate. (A) Docking of vinyl acetate in the structure of cutinase (PDB: 6AID). (B) The primary sequence of one Aspergillus oryzae cutinase (NCBI: OOO09588.1) was modelled in its structure by SWISS-Model [68,69] before docking analysis with vinyl acetate. (C) Docking of vinyl acetate in the structure of cutinase (PDB: 1FFB).
Bioengineering 10 00222 g005
Table 1. Candidate cutinases and lipases for the bio-amidation towards N-heterocyclic substrates.
Table 1. Candidate cutinases and lipases for the bio-amidation towards N-heterocyclic substrates.
EnzymeSourceDescriptionReference
Cutinase, PDB: 6AIDThermobifida albaIndicated by docking analysis (Figure 4a), this enzyme has potential ability to form acyl-enzyme intermediate for further amidation towards N-heterocyclic substrates.[70]
Cutinase, NCBI: OOO09588.1Aspergillus oryzaeIndicated by docking analysis (Figure 4b), this enzyme has potential ability to form acyl-enzyme intermediate for further amidation towards N-heterocyclic substrates.Not published yet
Cutinase, PDB: 1FFBFusarium solani pisiIndicated by docking analysis (Figure 4c), this enzyme has potential ability to form acyl-enzyme intermediate for further amidation towards N-heterocyclic substrates.[71]
Cutinase, PDB: 4OYY and 4OYLHumicola insolensThis enzyme can form acyl-enzyme intermediate with acyl donor. In addition, the aromatic ring in the catalytic pocket can help to form π–π stacking interactions with N-heterocyclic rings.[48,72]
Cutinase, NCBI: HQ147785Thermobifida cellulosilyticaThe catalytic triad is located near to the surface of this enzyme, which helps to accommodate acyl donors and N-heterocyclic rings.[73]
Cutinase, PDB: 3GBSThielavia terrestrisBoth hydrophilic and hydrophobic residues are found in the catalytic pocket, which contribute to recognize a broader range of acyl donors and amine substrates (including N-heterocyclic substrates).[74]
Cutinase, NCBI: AB445476.2Thermobifida albaIn the substrate binding pocket, the arrangement of alternating hydrophobic and hydrophilic amino acids can facilitate its binding to acyl donors and hydrophilic amine substrates.[75]
Lipase, NCBI: AEA86017Pseudomonas stutzeriThe enzyme can catalyze the monoacetylation reaction towards the N-heterocyclic substrate of piperazine.[76]
PETase, NCBI: GAP38373.1Ideonella sakaiensisThis PETase has a more open and wider substrate binding pocket than cutinase and lipase. The S238F mutation can further provide π–π stacking interactions towards N-heterocyclic substrates.[77,78]
Table 2. Comparison of chemical and enzymatic methods in amides forming of N-heterocyclic substrates.
Table 2. Comparison of chemical and enzymatic methods in amides forming of N-heterocyclic substrates.
AdvantageDisadvantageReaction Condition
Chemical methodsMatured techniques, high yield and efficiency.Require extreme conditions (high temperature, strong acid/base), and large amounts of by products and solvents.Detailed information is summarized in Figure 2.
Nitrogen acetyltransferaseMild condition, workable in aqueous solvent, and high selectivity. Demands high-cost acyl coenzyme A (acetyl coenzyme A or succinyl coenzyme A) supplementation.In aqueous solution, 0.05 M potassium phosphate buffer, pH 7.4, mild temperature, the concentrate of acyl CoA up to 5 × 10−5.
Carboxylic acid reductaseMild condition, workable in aqueous solvent, and recognizes a broader range of substrates for amide bond forming. Demands ATP circulatory system.In aqueous solution, carboxylic acid (1–10 mM), amine (175 mM), MgCl2 (100 mM), CARmm-A (3 mg/mL), Tris-buffer (100 mM), polyphosphate (4 mg/mL), pH 8.5, 30 °C, 16 h.
Lipase and CutinaseThermo and organic solvent tolerance, industrial resources, do not demand expensive supplementation such as acyl CoA and ATP, and can recognize broad range of acyl donors.N-amidation activities demand further research for nitrogen-containing heterocyclic substrates.In organic solvents, broad temperature range (40–70 °C), broad acyl donors (acid, esters, and vinyl acetate).
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yang, A.; Miao, X.; Yang, L.; Xu, C.; Liu, W.; Xian, M.; Zou, H. N-Amidation of Nitrogen-Containing Heterocyclic Compounds: Can We Apply Enzymatic Tools? Bioengineering 2023, 10, 222. https://doi.org/10.3390/bioengineering10020222

AMA Style

Yang A, Miao X, Yang L, Xu C, Liu W, Xian M, Zou H. N-Amidation of Nitrogen-Containing Heterocyclic Compounds: Can We Apply Enzymatic Tools? Bioengineering. 2023; 10(2):222. https://doi.org/10.3390/bioengineering10020222

Chicago/Turabian Style

Yang, Anran, Xue Miao, Liu Yang, Chao Xu, Wei Liu, Mo Xian, and Huibin Zou. 2023. "N-Amidation of Nitrogen-Containing Heterocyclic Compounds: Can We Apply Enzymatic Tools?" Bioengineering 10, no. 2: 222. https://doi.org/10.3390/bioengineering10020222

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop