You are on page 1of 4

Angewandte

Communications Chemie

International Edition: DOI: 10.1002/anie.201707918


Biocatalytic Amide-Bond Formation German Edition: DOI: 10.1002/ange.201707918

Adenylation Activity of Carboxylic Acid Reductases Enables the


Synthesis of Amides
Alexander J. L. Wood, Nicholas J. Weise, Joseph D. Frampton, Mark S. Dunstan,
Michael A. Hollas, Sasha R. Derrington, Richard C. Lloyd, Daniela Quaglia,
Fabio Parmeggiani, David Leys, Nicholas J. Turner, and Sabine L. Flitsch*
Abstract: Carboxylic acid reductases (CARs) catalyze the
reduction of a broad range of carboxylic acids to aldehydes
using the cofactors adenosine triphosphate and nicotinamide
adenine dinucleotide phosphate, and have become attractive
biocatalysts for organic synthesis. Mechanistic understanding
of CARs was used to expand reaction scope, generating
biocatalysts for amide bond formation from carboxylic acid
and amine. CARs demonstrated amidation activity for various
acids and amines. Optimization of reaction conditions, with
respect to pH and temperature, allowed for the synthesis of the
anticonvulsant ilepcimide with up to 96 % conversion. Mech-
anistic studies using site-directed mutagenesis suggest that,
following initial enzymatic adenylation of substrates, amida-
tion of the carboxylic acid proceeds by direct reaction of the
acyl adenylate with amine nucleophiles.
Figure 1. CARs are composed of three domains: an adenylation

Carboxylic acids are common reagents in organic synthesis, domain, which is divided into a core-domain (A(core)) and a subdo-
main (A(sub)), a peptidyl carrier protein (PCP) and a reduction
and their interconversion to other functional groups generally domain (R). CARs catalyze reduction of the carboxylic acid at the
requires activation to reactive acyl intermediates as the key expense of ATP and NADPH by production of acyl adenylate and
step.[1] In nature, there are many enzymes that catalyze enzyme–thioester intermediates. Both intermediates on the reaction
transformation of carboxylic acids using activated esters as pathway might potentially be intercepted by nucleophiles (Nu) such as
intermediates.[2] An interesting group are carboxylic acid amines (image adapted from previous work, see Ref. [7]). Key: P2O74@
(PPi), adenosine monophosphate (AMP).
reductases (CAR),[3, 4] which catalyze the reduction of car-
boxylic acids to aldehydes via acyl adenylates and thioesters
(Figure 1). Given that both activated esters are shared with exploring the activity of CARs towards other nucleophiles
other enzyme pathways, such as those of nonribosomal (Figure 1). Of specific interest were amine nucleophiles that
peptide synthetases (NRPS),[5] we were interested in would lead to amide bond formation from acid and amine, an
important challenge in organic chemistry.[6]
Our recent structural work[7] has shown that CARs are
[*] A. J. L. Wood, Dr. N. J. Weise, J. D. Frampton, M. A. Hollas, particularly good candidates for this approach. CARs are
Dr. S. R. Derrington, Dr. D. Quaglia, Dr. F. Parmeggiani, Prof. D. Leys, composed of three distinct protein domains: the adenylation
Prof. N. J. Turner, Prof. S. L. Flitsch
domain, which itself is divided into a core-domain and
School of Chemistry & Manchester Institute of Biotechnology
The University of Manchester
subdomain,[7] the peptidyl carrier protein with the bound
131 Princess Street, M1 7DN, Manchester (UK) 4’-phosphopantetheine (PPant) prosthetic group, and a reduc-
E-mail: sabine.flitsch@manchester.ac.uk tase domain. The adenylation domain activates carboxylic
Dr. M. S. Dunstan acids by catalyzing the formation of an acyl adenylate
Manchester Centre for Synthetic Biology of Fine and Speciality intermediate, at the expense of adenosine triphosphate
Chemicals (SYNBIOCHEM), Manchester Institute of Biotechnology (ATP).[8] The activated substrate is then attacked by the
The University of Manchester thiol of the enzyme-bound PPant group, generating a thioester
Manchester M1 7DN (UK)
intermediate that is transferred to the reduction domain.
Dr. R. C. Lloyd nicotinamide adenine dinucleotide phosphate (NADPH) is
Dr. Reddy’s Laboratories (EU) Ltd.
then expended to reduce the intermediate to an aldehyde
410 Cambridge Science Park, Milton Road, Cambridge CB4 0PE (UK)
(Figure 1).[3] CARs have a broad substrate range and
Dr. D. Quaglia
Chemistry Department, Universit8 de Montr8al a number of examples from different organisms are now
2900, Edouard-Montpetit, H3C 3J7, Montr8al (Canada) available, making them attractive for applications in single-
Supporting information and the ORCID identification number(s) for step biotransformations,[9, 10] as well as in enzymatic
the author(s) of this article can be found under: cascades.[11–13]
https://doi.org/10.1002/anie.201707918.

14498 T 2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2017, 56, 14498 –14501
Angewandte
Communications Chemie

The three-domain topology of the enzyme suggests that, Table 1: CARs generate a range of primary amides 7–12 from acids 1–6
in the presence of ATP but the absence of NADPH, the by reaction with ATP and ammonia.[a]
substrate carboxylic acid might still be activated, yielding the
acyl adenylate or thioester. Both intermediates could poten-
tially react with an external nucleophile, such as an amine, to
generate amides (Figure 1). Amidation would be of particular Conversion [%][b]
interest because many natural amide synthetases, such as Acid Amide CARmm CARni CARtp CARse
NRPS[14] and ATP-grasp enzymes[15, 16] , suffer from very
narrow substrate specificity, which limits their applications 11 3 1 11
in biocatalysis. Alternatively, hydrolases (such as lipases[17]
and proteases[18]), which are commonly used for amide
formation,[2] require systems in which very little water is
22 12 4 8
present, such as organic solvents, to drive the reaction towards
amide bond formation.[19–22]
Four CAR candidates (CARmm from Mycobacterium
marinum,[3] CARni from Nocardia iowensis,[4] CARtp from 25 13 6 15
Tsukamurella paurometabola,[10] and CARse from Segnilipa-
rus rotundus,[23]) were produced recombinantly in Escherichia
coli, with the coexpressed gene for the Bacillus subtilis
8 2 1 <1
phosphopantetheinyl transferase (PPTase; Sfp),[24] which is
required for post-translational addition of the PPant group.[4]
These CARs were purified (Supporting Information, Fig-
10 3 2 1
ure S1) and the amidation of carboxylic acid substrates 1–6
was tested. Instead of NADPH, ammonia was added to the
biotransformations in excess, with a ratio of 100:1 amine to
acid, at neutral pH. However, only trace amounts of amide
13 3 1 8
were produced under these conditions (Supporting Informa-
tion, Table S1).
Given that amidation might have very different require-
ments to the native reduction reaction, a wider range of [a] 1–6 (1 mm), ammonia (100 mm), purified CAR (100 mg mL@1), ATP
conditions was investigated. Firstly, in preliminary work, a pH (5 mm), MgCl2 (10 mm), sodium carbonate buffer (100 mm), pH 9.0,
37 8C (30 8C for CARse), 24 h. [b] Conversion determined by HPLC on
range of pH 7.5–10.5 was investigated for the production of
a non-chiral phase at 24 h.
primary amides, with pH 9.0 found to be optimal (Supporting
Information, Table S1). Using pH 9.0, formation of primary
amides 7–12 was clearly detected by HPLC and products were
identified by comparison with authentic synthetic standards.
Comparatively, controls in which either ATP or enzyme was Table 2: CARs can generate secondary and tertiary amides 16–18 by
reaction of 1 with ATP and amines 13–15.[a]
omitted showed no amide production. Table 1 shows that
conversions of up to 25 % were observed, demonstrating clear
evidence of CAR-dependent amide production. Higher
conversions were observed for the cinnamic acid derivatives
(2 and 3) compared to benzoic acid derivatives (1 and 4–6;
Table 1). The four CARs demonstrated different reaction Conversion [%][b]
Amine Amide CARmm CARni CARtp CARse
profiles depending on the temperature (Supporting Informa-
tion, Table S2).
Subsequently, amines other than ammonia were inves- 25 12 3 14
tigated for amidation of 1; primary amines 13 and 14, and
secondary amine 15 were explored (Table 2). Comparison of
conversions for all enzymes showed that amines 13 and 15
4 –[c] –[c] –[c]
were better reactants than ammonia, but propargylamine 14
only afforded amide 17 with a conversion of 4 % by CARmm.
After establishing that various carboxylic acids and
amines could be used for amide production with this
8 7 2 8
method, the CARs were tested as biocatalysts for the
formation of a commercially relevant target amide, ilepcimide
20, which possesses anticonvulsant properties.[25] Preparative-
[a] 1 (1 mm), 13–15 (100 mm), purified CAR (100 mg mL@1), ATP (5 mm),
scale CAR-mediated production of ilepcimide 20 was carried MgCl2 (10 mm), sodium carbonate buffer (100 mm), pH 9.0, 30 8C (37 8C
out at 37 8C, using cell lysate containing CARmm, with 15 for CARmm, 22 8C for CARni), 24 h. [b] Conversion determined by HPLC
on a non-chiral phase at 24 h. [c] No activity detected.

Angew. Chem. Int. Ed. 2017, 56, 14498 –14501 T 2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org 14499
Angewandte
Communications Chemie

(100 mm, 19 equiv), ATP (15 mm), MgCl2 (20 mm), and 19
(100 mg), affording 25 mg of 20 (19 % yield).
Further optimization using purified CARs at pH 9.0 at
22 8C, 30 8C, and 37 8C, resulted in conversions of 71 % by
CARmm at 30 8C, 68 % by CARni at 22 8C, 32 % by CARtp
and 7 % by CARse at 37 8C (Table 3). Lastly, conversion of
96 % to ilepcimide 20 by CARmm was achieved at 30 8C,
pH 9.0, when the reaction was left for up to 72 h (Table 3). A

Table 3: Synthesis of ilepcimide 20 by CAR amidation.[a]

Conversion [%][b]
T [88C] CARmm CARni CARtp CARse
22 52 68 5 1
30 71 (96)[c] 50 27 4
37 17 48 32 7
[a] 19 (1 mm), 15 (100 mm), purified CAR (100 mg mL@1), ATP (5 mm),
MgCl2 (10 mm), sodium carbonate buffer (100 mm), pH 9.0, 22–37 8C.
[b] Conversion determined by HPLC at the 24 h mark. [c] Value
determined at 72 h.
Figure 2. A) CARs are bound to a PPant group at serine (Ser689 on
CARni or Ser685 on CARmm), which is replaced with an alanine group
in mutant CARni Ser689Ala. CARmmD729-1175 lacks the reduction
pH profile was also conducted using CARmm for the domain (images adapted from previous work, see Ref. [7]). B) Mecha-
production of 20 (Supporting Information, Figure S2), which nism proposed for the amidation reaction.
confirmed that pH 9.0 was optimal for this reaction with 71 %
conversion. However, conversion of 21 % could still be
observed at neutral pH, while at pH values higher than 9.0, was constructed, as described previously.[7] This construct
the conversion was lower, with 40 % conversion at pH 10.0. lacks the terminal reduction domain but retains the
These results suggest that both amine deprotonation and adenylation domain and carrier protein (CARmmD729-
decreased enzyme stability at higher pH may have an effect 1175; Figure 2 A). In the amidation assay, this variant gave
on conversion. The specific activity of CARmm on the a conversion of 69 % to 20.
production of 20 was shown to be 12.22 : 0.15 mU mg@1 Notably, both the CARni mutant and the CARmm
(Supporting Information, Figure S3). As with the other truncation still retain amido synthetase activity, but have
amide products, control reactions without enzyme or ATP lost reductase activity (even in the presence of NADPH), and
showed no conversion to 20. therefore present a new class of amido synthetase biocatalysts
Finally, the mechanism of amide bond formation was with no residual CAR activity; that is, a full switch of activity.
explored, in particular addressing the important question as These variants might have application as amidation catalysts
to whether the acyl phosphate or the thioester was inter- in more complex enzyme cascades, both in cell-free and
cepted by the amine (Figure 1). To probe this question, the whole-cell systems, where the presence of NADPH would
Ser689Ala variant of CARni[4] (Figure 2 A) was generated. normally suppress CAR-mediated amidation in favor of
This variant lacks the Ser689 attachment site to the PPant reduction.[12]
group that is involved in thioester formation and therefore In conclusion, we have shown that in the absence of
would only allow the formation of the acyl adenylate and not NADPH, the adenylation activity of CARs can be exploited
the thioester. Conversion to ilepcimide 20 (79 %) was still for amide bond formation. The reaction was tolerant to
observed with this variant, suggesting that adenylation a range of carboxylic acids and amines and could be
activity alone is sufficient for amidation, presumably by performed in an aqueous medium using ATP as the driving
nucleophilic attack onto the acyl adenylate (Figure 2 B), force. The method was applied to a target drug molecule,
a reaction which has been observed with other adenylating ilepcimide 20, with up to 96 % conversion and preparative
and phosphorylating enzymes, including NRPS and glutamine scale-up. Given that all four CARs tested exhibited this
synthetase.[26–30] Our current results are not able to establish if activity, the range of enzymes that could be used for this
this reaction is enzyme catalyzed. amidation reaction might be rapidly expanded by tapping into
To demonstrate that the reduction domain was the collection of known CAR enzymes with differing and
not required for amidation, and to remove the wild-type broad substrate specificities.[9–10] While our method is limited
reduction activity of the enzyme, a truncated form of CARmm by the use of the expensive cofactor ATP, this may be

14500 www.angewandte.org T 2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2017, 56, 14498 –14501
Angewandte
Communications Chemie

overcome in future by established ATP regenerating sys- [10] W. Finnigan, A. Thomas, H. Cromar, B. Gough, R. Snajdrova,
tems.[31] Two CAR variants were generated by the rational J. P. Adams, J. A. Littlechild, N. J. Harmer, ChemCatChem 2017,
redesign of the CARs, which possess no reduction activity but 9, 1005 – 1017.
[11] S. P. France, S. Hussain, A. M. Hill, L. J. Hepworth, R. M.
retain amido synthetase activity. The success with these
Howard, K. R. Mulholland, S. L. Flitsch, N. J. Turner, ACS Catal.
variants suggests that the adenylation activity alone is 2016, 6, 3753 – 3759.
sufficient for activation of carboxylic acids for subsequent [12] L. J. Hepworth, S. P. France, S. Hussain, P. Both, N. J. Turner,
amidation, and also lays the foundations for future work using S. L. Flitsch, ACS Catal. 2017, 7, 2920 – 2925.
this method in more complex enzyme systems, such as in [13] S. P. France, L. J. Hepworth, N. J. Turner, S. L. Flitsch, ACS
enzyme cascades or whole-cell biotransformations. Catal. 2017, 7, 710 – 724.
[14] T. Stachelhaus, H. D. Mootz, M. A. Marahiel, Chem. Biol. 1999,
6, 493 – 505.
[15] M. Y. Galperin, E. V. Koonin, Protein Sci. 1997, 6, 2639 – 2643.
Acknowledgements [16] J. Pitzer, K. Steiner, J. Biotechnol. 2016, 235, 32 – 46.
[17] T. Nuijens, E. Piva, J. A. W. Kruijtzer, D. T. S. Rijkers, R. M. J.
This work was funded by EPSRC, BBSRC, SYNBIOCHEM, Liskamp, P. J. L. M. Quaedflieg, Tetrahedron Lett. 2012, 53,
and the Centre of Excellence for Biocatalysis, Biotransfor- 3777 – 3779.
mations and Biocatalytic Manufacture (CoEBio3) and its [18] G. A. Homandberg, J. A. Mattis, M. Laskowski, Jr., Biochem-
istry 1978, 17, 5220 – 5227.
affiliates, in particular Dr. ReddyQs Laboratories.
[19] V. Gotor, Bioorg. Med. Chem. 1999, 7, 2189 – 2197.
[20] C. Shui-Tein, H. Shu-Chyong, W. Kung-Tsung, Bioorg. Med.
Chem. Lett. 1991, 1, 445 – 450.
Conflict of interest [21] R. V. Ulijn, B. Baragana, P. J. Halling, S. L. Flitsch, J. Am. Chem.
Soc. 2002, 124, 10988 – 10989.
The authors declare no conflict of interest. [22] R. V. Ulijn, N. Bisek, P. J. Halling, S. L. Flitsch, Org. Biomol.
Chem. 2003, 1, 1277 – 1281.
[23] Y. Duan, P. Yao, X. Chen, X. Liu, R. Zhang, J. Feng, Q. Wu, D.
Keywords: amidation · amides · amido synthetase ·
Zhu, J. Mol. Catal. B 2015, 115, 1 – 7.
biocatalysis · carboxylic acid reductase [24] R. H. Lambalot, A. M. Gehring, R. S. Flugel, P. Zuber, M.
LaCelle, M. A. Marahiel, R. Reid, C. Khosla, C. T. Walsh, Chem.
How to cite: Angew. Chem. Int. Ed. 2017, 56, 14498 – 14501 Biol. 1996, 3, 923 – 936.
Angew. Chem. 2017, 129, 14690 – 14693 [25] Q. S. Yan, P. K. Mishra, R. L. Burger, A. F. Bettendorf, P. C.
Jobe, J. W. Dailey, J. Pharmacol. Exp. Ther. 1992, 261, 652 – 659.
[1] C. A. G. N. Montalbetti, V. Falque, Tetrahedron 2005, 61, 10827 – [26] R. Dieckmann, T. Neuhof, M. Pavela-Vrancic, H. von Dçhren,
10852. FEBS Lett. 2001, 498, 42 – 45.
[2] A. Goswami, S. G. Van Lanen, Mol. Biosyst. 2015, 11, 338 – 353. [27] T. Abe, Y. Hashimoto, S. Sugimoto, K. Kobayashi, T. Kumano,
[3] M. K. Akhtar, N. J. Turner, P. R. Jones, Proc. Natl. Acad. Sci. M. Kobayashi, J. Antibiot. 2017, 70, 435 – 442.
USA 2013, 110, 87 – 92. [28] C. Maruyama, J. Toyoda, Y. Kato, M. Izumikawa, M. Takagi, K.
[4] P. Venkitasubramanian, L. Daniels, J. P. N. Rosazza, J. Biol. Shin-ya, H. Katano, T. Utagawa, Y. Hamano, Nat. Chem. Biol.
Chem. 2007, 282, 478 – 485. 2012, 8, 791 – 797.
[5] M. A. Marahiel, T. Stachelhaus, H. D. Mootz, Chem. Rev. 1997, [29] C. Ji, Q. Chen, Q. Li, H. Huang, Y. Song, J. Ma, J. Ju, Tetrahedron
97, 2651 – 2674. Lett. 2014, 55, 4901 – 4904.
[6] V. R. Pattabiraman, J. W. Bode, Nature 2011, 480, 471 – 479. [30] S. H. Liaw, D. Eisenberg, Biochemistry 1994, 33, 675 – 681.
[7] D. Gahloth, M. S. Dunstan, D. Quaglia, E. Klumbys, M. P. [31] J. N. Andexer, M. Richter, ChemBioChem 2015, 16, 380 – 386.
Lockhart-Cairns, A. M. Hill, S. R. Derrington, N. S. Scrutton,
N. J. Turner, D. Leys, Nat. Chem. Biol. 2017, DOI: https://doi.
org/10.1038/nchembio.2434.
[8] M. Moura, D. Pertusi, S. Lenzini, N. Bhan, L. J. Broadbelt, Manuscript received: August 2, 2017
K. E. J. Tyo, Biotechnol. Bioeng. 2016, 113, 944 – 952. Revised manuscript received: September 5, 2017
[9] K. Napora-Wijata, G. A. Strohmeier, M. Winkler, Biotechnol. J. Accepted manuscript online: September 22, 2017
2014, 9, 822 – 843. Version of record online: October 13, 2017

Angew. Chem. Int. Ed. 2017, 56, 14498 –14501 T 2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org 14501

You might also like