Annabel
Serpico§
a,
Silvia
De Cesare
a,
Jon
Marles-Wright
b,
M. Kalim
Akhtar¶
c,
Gary J.
Loake
c and
Dominic J.
Campopiano
*a
aSchool of Chemistry, University of Edinburgh, Joseph Black Building, David Brewster Road, Edinburgh, EH9 3FJ, UK. E-mail: Dominic.Campopiano@ed.ac.uk
bSchool of Natural and Environmental Sciences, Newcastle University, Newcastle upon Tyne, NE1 7RX, UK
cInstitute for Molecular Plant Sciences, School of Biological Sciences, University of Edinburgh, King's Buildings, Edinburgh EH9 3BF, UK
First published on 28th July 2020
Enantiopure amines are key building blocks in the synthesis of many pharmaceuticals, so a route to their production is a current goal for biocatalysis. The stereo-inverting D-phenylglycine aminotransferase (D-PhgAT), isolated from Pseudomonas stutzeri ST-201, catalyses the reversible transamination from L-glutamic acid to benzoylformate, yielding α-ketoglutarate and D-phenylglycine (D-Phg). Detailed kinetic analysis revealed a range of amine donor and acceptor substrates that allowed the synthesis of enantiopure aromatic D-amino acids at a preparative scale. We also determined the first X-ray crystal structure of D-PhgAT with its bound pyridoxal 5′-phosphate (PLP) cofactor at 2.25 Å resolution. A combination of structural analysis and site-directed mutagenesis of this class III aminotransferase revealed key residues that are potentially involved in the dual substrate recognition, as well as controlling the stereo-inverting behaviour of D-PhgAT. Two arginine residues (Arg34 and Arg407) are involved in substrate recognition within P and O binding pockets respectively. These studies lay the foundation for further enzyme engineering and promote D-PhgAT as a useful biocatalyst for the sustainable production of high value, aromatic D-amino acids.
Scheme 1 Simplified, two-step mechanism of the D-PhgAT-catalysed reaction. In step 1, L-Glu donates the amino group to the pyridoxal 5′-phosphate (PLP) cofactor generating pyridoxal amine (PMP) and α-ketoglutarate (AKG). In step 2, benzoyl formate (BZF, R = H) or its 4′-hydroxy derivative (HBF, R = OH) accepts the amino group from PMP yielding D-Phg or D-Hpg respectively.11 |
The 3D structures of many PLP enzymes have been determined and they display seven different folds with TAs typically found in classes I and IV.17–22 To determine the substrate scope of so many AT-catalysed reactions, powerful high-throughput assays have been developed that can be used to study the activities of wild-type ATs or identify new substrates for engineered enzymes.23–27
The X-ray structures and screening methods for these biocatalysts have been comprehensively reviewed6 but there are many ATs that display “unusual” characteristics which require further study.
The D-phenylglycine aminotransferase (D-PhgAT), isolated from the soil bacterium Pseudomonas stutzeri ST-20128,29 catalyses the reversible transamination of L-glutamic acid (L-Glu) with the amino acceptors benzoylformate (BZF) and 4-hydroxy benzoylformate (HBF) yielding the R-enantiomers D-Phg and D-Hpg respectively. The keto product derived from L-Glu is α-ketoglutarate (AKG) (Scheme 1). What is unique to this AT is its so-called “stereo-inverting” characteristic, where the amino donors (L-Glu and D-Phg) in this reversible reaction exhibit inverse absolute configurations. Since it uses an inexpensive L- (or S-) amino acid donor, the P. stutzeriD-PhgAT represents an attractive biocatalyst for the synthesis of enantiomerically pure D-Phg (or R-Phg) derivatives in a single step.5D-PhgAT has also been used in qualitative and quantitative analysis of L-Glu concentrations in food when coupled with L-glutamate dehydrogenase30 and for quality control of amoxicillin in pharmaceuticals when coupled with a penicillin acylase.31 The homologous D-PhgAT from P. putida32–34 (82% amino acid sequence identity) has been successfully combined with hydroxymandelate synthase (HmaS) and hydroxymandelate oxidase (Hmo) to engineer an E. coli strain that is able to produce D-Phg.34
To expand its potential as a valuable biocatalyst, comprehensive substrate and structural analyses of the P. stutzeriD-PhgAT are required. An understanding of the mechanism of its unique stereo-inverting activity will also aid in increasing its substrate scope by enzyme engineering.35 In this study we describe the isolation of recombinant P. stutzeriD-PhgAT and, by a combination of spectrophotometric and chiral product analysis, we explore the amino substrate range and enantioselectivity of this enzyme. This revealed a broader substrate range than previously anticipated, which led us to use this biocatalyst to prepare a range of aromatic D-amino acids in high conversion and e.e. Furthermore, we determined the first X-ray structure of the PLP-bound form of the P. stutzeriD-PhgAT at 2.25 Å resolution, which confirms its classification as a member of the class III TA family. To explore the dual substrate binding mechanism, we then compared the D-PhgAT structure with the PLP-bound, external aldimine complexes of similar TAs. Furthermore, a combined sequence and structural analysis led to site-directed mutagenesis studies which identified potential residues required for catalytic activity and substrate recognition. This work lays the foundation for future enzyme engineering, and strengthens the utility of D-PhgAT as an important biocatalyst for the preparation of enantiopure D-aromatic amino acids.
A convenient, high-throughput, coupled assay used the α-ketoglutarate dehydrogenase (AKGDH) enzyme to monitor transamination of L-Glu through NADH production (Fig. S3 and S4‡). This assay allowed the determination of kinetic parameters (apparent KM for L-Glu and BZF are 9.85 ± 0.62 mM, 1.81 ± 0.57 mM respectively) in the “forward” D-amino acid-producing direction (Fig. S5, Table S1‡). We confirmed the chirality of the product using a modified Chirobiotic-T chiral HPLC method and demonstrated the enzyme enantioselectivity with production of D-Phg in >98% e.e. (Fig. S6 and S7‡).36 Furthermore, we used this method to determine the D-PhgAT kinetics, which match well with the coupled assay (Fig. S8 and Table S1‡). With BZF as the amino acceptor the chiral HPLC assay was used to screen D-Phg production with various commonly used amino donors. These included L- and D-amino acids, R- and S-α-methylbenzylamine (MBA), isopropylamine (iPrA), as well as o-xylene diamine (OXD) (Fig. S9 and S10‡). The production of D-Phg is observed with several amino donors (apart from iPrA) and the activity is ranked relative to L-Glu as the best amino donor. This broad utility is in contrast with a previous report which observed a very limited substrate scope.28 Regardless of the chirality of the amino donor, the enantiopurity of the final D-Phg product is not affected. Since L-Glu was the best amino donor it was used to scale up the reaction with 1 g BZF substrate. Using D-PhgAT (1 mg mL−1) 93% conversion to D-Phg (99% e.e.) was observed after ∼3 h (Fig. S11). We purified the D-Phg product by preparative HPLC (Fig. S12‡) and characterized it using 1H and 13C NMR and ESI-MS (Fig. S13‡) which are identical to an authentic standard.
Since D-Phe, D-Trp and D-Tyr are used as chiral building blocks for many clinically useful drugs, we tested the promiscuity of the P. stutzeriD-PhgAT towards various aromatic acceptors.37–39 We used D-PhgAT with L-Glu and the five amino acceptors (HBF, BZF, indole pyruvic acid (IPA), phenylpyruvic acid (PPA), and 4-hydroxyphenylpyruvic acid (HPPA)) that would yield the corresponding aromatic amino acids D-Hpg, D-Phg, D-Trp, D-Phe, and D-Tyr (Scheme 2). The enzyme shows good affinity towards all acceptor substrates tested (apparent KM values from 0.79 mM for IPA to 9.24 mM for HPPA, Table 1).
Scheme 2 D-PhgAT reactions for the synthesis of aromatic D-amino acids from L-Glu and aromatic amino acceptors. |
Substrate | Conversion % | ee % | K M (mM) | k cat/KM (M−1 s−1) |
---|---|---|---|---|
HBF | 50 | >98 | 1.01 ± 0.07 | 899.15 ± 0.27 |
BZF | 93 | >98 | 3.16 ± 0.46 | 302.80 ± 0.10 |
IPA | 30 | >98 | 0.39 ± 0.19 | 262.12 ± 0.13 |
PPA | 57 | >98 | 6.91 ± 0.99 | 14.49 ± 0.82 |
HPPA | 15 | >98 | 9.24 ± 2.40 | 0.06 ± 0.01 |
Furthermore, we found again that the conversion was highly stereoselective with only the D-enantiomer produced (Fig. 1). Kinetic parameters and % conversions were calculated using the chiral HPLC method and are summarized in Table 1. Using this data larger scale biotransformations were carried out at up to 100 mg scale for each amino acceptor. The D-PhgAT displayed modest to excellent conversion for these aromatic acids, with the highest observed for the BZF to D-Phg conversion (93%). This analysis emphasizes the broad utility of the enzyme.
Fig. 1 Chiral HPLC traces of the synthesis of enantiopure D-Tyr, D-Phe, D-Trp, D-Phg and D-Hpg. Reactions described in Scheme 2 were analysed for product formation by chiral HPLC using a Chirobiotic T column with monitoring at λ = 205 nm. |
To understand the molecular basis of the substrate promiscuity and the origin of the D-enantioselectivity, as well as revealing the residues involved in catalysis, we determined the crystal structure of the D-PhgAT in complex with the PLP-cofactor bound as the internal aldimine form at 2.25 Å resolution (Fig. 2, S14 and S15‡). Here we used the incomplete structure of the P. stutzeri enzyme lacking the bound PLP cofactor (PDB code: 2CY840) as a molecular replacement model (Table S2,‡ PDB code: 6G1F). The overall D-PhgAT structure confirmed its classification as a member of group III of the aspartate aminotransferase family, which falls within the type I PLP-dependent superfamily fold.41
Fig. 2 Crystal structure of P. stutzeriD-PhgAT. (A) Functional dimer of the D-PhgAT protein, shown with one chain as a surface representation and the other as a cartoon showing the position of the PLP cofactor in stick representation. (B) Active site pockets of the D-PhgAT protein showing the PLP internal aldimine and key residues in the substrate binding pockets. The O- and P-pockets are labelled following the convention in Wybenga et al.43 the protein backbone is shown as a cartoon, with key residues shown as sticks. The two chains in the structure are coloured green and blue. The final experimental 2mFo-DFc electron density for the PLP is shown as blue mesh. (C) Cut-away view of the substrate binding tunnel showing the depth and breadth of the active site region of the protein. Amino acids and PLP internal aldimine are shown as stick representations. |
The final refined model contains three homo-dimers of D-PhgAT with the PLP cofactor present as an internal aldimine bound to residue Lys269 in each chain (Fig. 2, S14 and S15‡). Each monomer can be subdivided into two distinct domains: a small discontinuous domain comprising the residues 1–72 and 336–453; and a large domain formed by residues 73–335 (Fig. S14‡). The N-terminal part of the small domain comprizes a kinked α-helix followed by a three-stranded antiparallel β-sheet, while the C- terminal part consists of an α-helix followed by two antiparallel β-strands, an extended loop interspersed with a short α-helix and continues into a longer α-helix. This is followed by a β-strand that extends the N-terminal β-sheet and an additional β-strand, which extends the C-terminal β-sheet, the structure finishes with an α-helix opposed between the two other major helices in this domain. The large domain consists of a central seven-stranded β-sheet, with five parallel strands and two in an anti-parallel orientation. The β-sheet is connected by α-helices, which harbor the PLP cofactor binding site and main dimerization interface (Fig. S14‡).
The essential Lys269 is located in the loop that connects strands β9 and β10, with the aromatic ring of the PLP sandwiched between Val243 and Tyr149 (Fig. S15‡). Water molecules (red spheres) form bridges between the PLP phosphate oxygens and the side chains of Glu124, Thr303 and the backbone of Phe304 from the opposite monomer. The phosphate is also coordinated by H-bonds to Ser121, Gly122 and Thr123 from the same monomer.42
Since the ATs bind two amino-acid substrates, it has been proposed that this dual-substrate recognition is accommodated in two binding sites. These were named the O-pocket (or O-site, defined by the proximity of residues to the 3′-O of the PLP) and P-pockets (or P-site, defined by the proximity to the PLP-phosphate) by Dijkstra and colleagues in their study of the S-selective TA (MesAT).43 They determined the structures of this useful enzyme with the PLP-external aldimine forms of S-β-Phe and R-3-methylhexanoic acid, as well as with the amino acceptor AKG. These complexes allowed them to identify residues in the binding pockets as well as a pair of key arginine residues (Arg54 and Arg412) in this type I enzyme. The O- and P-pocket hypothesis was also explored by a comprehensive structural and sequence analysis of the AT superfamily recently carried out by Bornscheuer and colleagues to identify 13 key residues (including specific arginines) that they propose as crucial in controlling substrate binding and reaction specificity.11 Bornscheuer and others have also put forward an elegant model known as the ‘arginine flip or switch’ that is thought to control the dual specificity of these enzymes.44–47 By modelling into the incomplete apo D-PhgAT structure (2CY8) they suggested that a key Arg residue, equivalent to Arg407 in our structure, binds to the α-carboxylate of the L-Glu donor and may switch in and out of the active site. A recent paper by Walton et al. also determined the crystal structure of D-PhgAT (PDB code: 6DBS) and they used this to propose a two pocket substrate binding model and residues involved in catalysis.48 Unfortunately, like 2CY8, the D-PhgAT had no PLP bound (but instead had phosphate) and lacked defined electron density for important residues 28–36 and 292–302.42,48 These incomplete apo- and phosphate-bound structures are missing important features such as the P-pocket. In contrast, with the PLP bound in our structure we can now better define both pockets for the first time (Fig. 2B). The O-pocket is formed by the side chains of Phe63, His66, His213, Phe304, and Arg407. On the other side of the PLP cofactor the P-pocket is formed by residues Arg34, Gln301, and Thr303.
The enzyme is unusual in that it can bind both L- and D- forms of the amino acid substrates (e.g.L-Glu and D-Phg) depending on which reaction (forward or reverse) it is catalyzing (Schemes 1 and 2). This inherent substrate promiscuity of the enzyme and the product enantioselectivity appears to be due to a combination of these two pockets (Fig. 2B) and a cavity (Fig. 2C) that can accommodate large hydrophobic amino acceptors. This cavity extends from the surface of the protein and has a wide mouth with a constriction near the active site formed by the side chains of His66 and His213. This constriction accommodates hydrophobic substrates through pi-stacking interactions, while orienting the keto substrate to accept the amine group from the PMP intermediate. This results in the formation of the key PLP:substrate external aldimine intermediate that undergoes the enantiomeric H+ transfer at C4′ to form the D-product (Fig. S16‡). We were also struck by the pair of arginine residues provided by one of the subunits, Arg34 (in the P-pocket) and Arg407 (in the O-pocket) (Fig. 2B). In the PLP bound structure, in the absence of an L-amino acid donor or a D-amino acid product, both of the side chains are orientated away from their respective pockets but are free to move upon ligand binding.
To understand how D-PhgAT bound its substrates Walton et al. superimposed the active sites of their PLP-free structure with MesAT with PLP and AKG (PDB: 2YKX), MesAT with PLP:(S)-3-phenyl-β-alanine (PDB: 2YKY) and a GSAM AT with PLP:(4S)-4,5-diaminopentanoate (PDB: 2HP2).48 The Arg54 and Arg412 side chains of MesAT were proposed to be equivalent residues to the Arg34 and Arg407 respectively in D-PhgAT. This allowed the “missing” Arg34 in the P-pocket to be proposed as the residue that engages the sidechain of the various substrates e.g. the γ-carboxylate of L-Glu. Similarly, in the O-pocket Arg407 is thought to interact with the carboxylate at the C-α of amino acids such as L-Glu.
Now with the Arg34-containing loop observed, we carried out similar structural comparisons of our PLP-bound structure with other TA enzymes to gain further insight into the residues involved in ligand binding in both pockets. We superimposed our PLP-bound internal aldimine structure with TAs that have structures with useful bound ligands. Here we used an interesting pair of S-selective ω-TAs, Bacillus megaterium (BM-ωTA) and Arthrobacter Ars-ωTA studied by Dijkstra and colleagues.49 These enzymes share 95% sequence identity but surprisingly display somewhat different substrate profiles and structural studies revealed insights into their mechanism and specificity. The X-ray structure of the PLP:R-MBA external aldimine bound complex of the BM-ωTA (PDB code: 5G09,)49 showed the phenyl group of R-MBA binding in the O-pocket and methyl group in the P-pocket. Dijkstra and colleagues found that the O-pocket is large enough to accommodate the phenyl group, in contrast to the smaller P-pocket. The structure also revealed why this PLP:R-MBA external intermediate was captured in the unproductive state - the amine of the lysine 298 side chain is unable to act as a base since it is on the wrong face to remove the proton from C-α. By simply inverting the stereochemistry at C-α to the productive S-MBA configuration the enzyme could achieve the optimal orientation for catalysis. The D-PhgAT also prefers S-MBA to R-MBA (Fig. S9‡) suggesting a similar discrimination mechanism.
A comparison of the two structures (Fig. 3A) shows how this substrate could be accommodated into the O-pocket; in the BM-ωTA:PLP:R-MBA complex the Arg442 is swung out to make way for the phenyl ring and Arg407, the equivalent residue in D-PhgAT, is also swung out. Since D-PhgAT prefers L-Ala, L-Glu and L-Asp over their opposite enantiomers we also carried out a comparison (Fig. 3B) of the structure of the Arthrobacter Ars-ωTA in complex with the PLP:L-Ala external aldimine (5G2Q). In Ars-ωTA the side chain of Arg442 is proposed as the arginine switch; swung out in the PLP-bound complex and swung in to engage with the C-α carboxylate when L-Ala binds. Our overlay of suggests the D-PhgAT Arg407 residue plays the equivalent flipping role to Arg442 in Ars-ωTA. Our models also suggests that residues His66, His213 from one monomer and Thr303 from the partner monomer play a role in the enantioselective mechanism of the enzyme. These residues orientate the ligand to be protonated by Lys269 from the Si-face of the PLP-bound intermediate, generating the D-Phg enantiomer (Fig. 2B, S15 and S16‡).
Fig. 3 Structural comparison of D-PhgAT with other aminotransferases. (A) Overlay of the P. stutzeri DPhgAT apo (6DVS, pink sticks) and PLP internal aldimine (6G1F, green and cyans cartoons and sticks) structures with the PLP:R-α-methylbenzylamine (R-α-MBA) external aldimine complex of the B. megaterium S-selective aminotransferase (5G09, orange). In the B. megaterium complex Arg442 swings away from the position of the ring of the R-α-MBA to accommodate this bulky intermediate. The equivalent residue in the P. stutzeri structure is Arg407, which is found in different conformations in the apo- and PLP-bound structures. (B) Overlay of the D-PhgAT:PLP internal aldimine with the PLP-external aldimine of Ars ω-TA PLP:L-Ala external aldimine complex (5G2Q). In Ars ω-TA PLP:L-Ala structure, the side chain of Arg442 swings in to interact with the C-α carboxylate of the PLP:L-Ala, this is a 180° flip around the CG of the arginine side chain in comparison to the B. megaterium structure. (C) Schematic representation of the dual substrate recognition mechanisms of D-PhgAT and the role of Arg34 and Arg407 in the arginine flip/switch model. In the forward direction the ‘flipping’ Arg407 in the O-pocket is involved in the recognition of the α-carboxylic group of amino donor L-Glu and can move out of the active site to accommodate the benzyl ring of the product D-Phg. In the P-pocket, the Arg34 side chain is involved in recognition of the α-carboxylate of the L-Glu (and L-Asp) amino donor. |
We also used this structural analysis to rationalize the L- to D-stereoinversion selectivity of D-PhgAT. Proton transfer of the C-4′ of the resulting PLP:quinonoid intermediate is the key to the enantioselectivity and Jomrit et al., determined that this occurred from the Si face of the intermediate.50 The L-Glu and D-Phg must bind in inverted orientations with respect to each other. The aromatic side chain of D-Phg (and the other aromatic amino acids we generated) must therefore exchange for the C-α carboxylate of L-Glu (O-pocket) and, similarly, the carboxylate at C-α of D-Phg can swap into the site (P-pocket) that binds the γ-carboxylate of the L-Glu side chain. This conformational flexibility requires mobile side chains in the two pockets and the two arginine residues (R34 and R407) have the ability to provide such dynamic movement. They are ideally placed to be able to swing in to engage with carboxylate residues directly in electrostatic interactions. Clearly, the active site must be highly dynamic during catalytic cycles that bind and release amino acceptors and products, depending on which direction it is operating (Schemes 1 and 2). Moreover, we also noted the presence of Q301 in the P-pocket in our structure, and this adds further weight to its important role in the catalytic mechanism. The importance of Q301 was identified in a recent cell-free extract, high-throughput, saturation mutagenesis screen of D-PhgAT catalysing the reverse reaction of D-Phg conversion to L-Glu.48 This contrasts with our assay which monitors aromatic amino acid production and AKG release. In their screen Walton et al. focused on three residues in the O-pocket (H66, H213 and R407) and two, in what they predicted, in the absence of defined electron density, to be the P-pocket (R34 and Q301). It is interesting to note that, like us, they also found D-PhgAT displayed some substrate promiscuity – D-Phg was the best substrate but also showed ∼8.6% relatively activity with D-Trp and ∼82% relative activity with (S-)-4-phenyl-4-aminobutyrate.
We gained further insight into the properties of D-PhgAT by carrying out a sequence alignment of the two D-PhgATs from P. stutzeri and P. putida with well characterized ATs from the same group III fold, but with opposite product enantioselectivity (Fig. S17‡). This revealed that the key residues (Arg34, His66, His213, Gln301 and Arg407) identified in our structural analysis are found only in the stereo-inverting enzymes (also referred to as R-selective) that produce D-products from L-amino acid donors. This suggests that this stereo-inverting property has arisen from using the group III TA fold and incorporating these four key amino acids. Since we found these arginine residues in the active site, we carried out a targeted mutagenesis study and mutants of these two residues, R34A and R407A, as well as the two histidine residues, H66A and H213A, were obtained. We also included the Q301A mutant to allow comparison with the recent results presented by Walton et al.48 The results of this mutagenesis study, summarized in Table 2, show the effect of these changes on the affinity and catalytic efficiency towards both L-Glu and BZF in the forward, D-Phg synthesis direction.
D-PhgAT | L-Glu | BZF | ||||
---|---|---|---|---|---|---|
K M (mM) | k cat (s−1) | k cat/KM (M−1 s−1) | K M (mM) | k cat (s−1) | k cat/KM (M−1 s−1) | |
a The KM could not be accurately determined as the highest achievable amount of L-Glu in the reaction was 500 mM. | ||||||
WT | 26.17 ± 3.63 | 1.65 ± 0.12 | 63.21 ± 0.78 | 3.16 ± 0.46 | 0.95 ± 0.056 | 302.80 ± 0.10 |
R34Aa | 368.33 ± 18.78 | 0.111 ± 0.0088 | 0.304 ± 0.004 | 0.842 ± 0.15 | 0.079 ± 0.016 | 93.33 ± 0.037 |
R407A | 133.50 ± 30.5 | 0.42 ± 0.012 | 3.22 ± 0.09 | 1.51 ± 0.45 | 0.169 ± 0.012 | 112.98 ± 0.12 |
H66A | 9.65 ± 0.32 | 0.092 ± 0.006 | 9.76 ± 0.009 | 11.66 ± 5.95 | 0.078 ± 0.01 | 9.57 ± 0.09 |
H213A | — | — | — | — | — | — |
Q301A | 43.23 ± 7.90 | 0.38 ± 0.015 | 20.59 ± 0.020 | 9.02 ± 0.95 | 0.15 ± 0.012 | 16.43 ± 0.05 |
As we predicted, the two arginine mutants R34A and R407A displayed a much lower affinity towards L-Glu with a KM of 368.33 (∼14 fold) and 133.50 mM (∼5 fold) respectively vs. the wild-type KM of 26.17 mM. The kcat and the catalytic efficiency is substantially disrupted by the mutations for both substrates by approximately 200- and 20-fold respectively. The proposed role of Arg34 as a key residue in binding L-Glu is strengthened since we found it was not possible to saturate the enzyme with the highest achievable concentration of L-Glu in the assay (500 mM). Like all AT enzymes, the catalytic cycle involves many PLP-derived intermediates including the true amine donor PMP which delivers the –NH2 group to AKG to give D-Phg in that direction. In the absence of PLP-bound intermediate structures we used our combined sequence, structural and mutagenesis results to envisage how the substrates and products are recognized (Fig. 3C). We suggest that Arg407 is the key “flipping” residue that recognizes the C-α carboxylate of the L-Glu amino donor as well as being able to make way for the side chains of the aromatic products. Its partner Arg34 is involved in the recognition of the amino donor side chain carboxylic acid group, L-Glu is preferred but the shorter L-Asp can also be accepted. This arginine ‘switch’ and substrate promiscuity requires a flexible side chain and the mobility of Arg407 was revealed by comparing our PLP-bound structure with the two apo-D-PhgAT structures (PDBID:2CY8 and PDBID:6DVS).40,50 We found that this sidechain adopts a different orientation in each structure, suggesting that it is able to move to accommodate substrate binding during catalysis (Fig. 4).
Fig. 4 Comparison of apo- and PLP-bound forms of D-PhgAT highlights the mobility and movement of Arg407 and loops around PLP-binding site. Our PLP-bound structure shown in salmon with two apo-forms in orange (PDBID: 2CY8) and blue (PDBID: 6DVS). The disordered, flexible loops, defined only in the PLP-bound structure, are shown as red ribbons with terminal ordered residues labelled and shown as sticks. For clarity the inset shows a zoom in of the positions of the Arg407 residue in each structure. |
The H213A mutant was totally inactive in all the tested conditions, suggesting that this residue is crucial to positioning the substrate in the O-pocket. This is in contrast to the H213A mutant described by Walton et al. which displays ∼45% activity in the reverse direction. Interestingly, H213N was 150% more active than the wild type but a rationale for this dramatic increase is not clear. In our hands both the H66A and Q301A mutants retained some activity and by analysis of the structure we suggest that these residues are involved in amino-acceptor binding and a hydrogen bonding network surrounding the PLP cofactor respectively (Fig. S15‡). Similarly, in their screen Walton et al. found that mutations at H66 and Q301 resulted in diminished activity.
A recent report described the engineering of three recombinant E. coli strains that co-expressed four, seven and nine enzymes, including the P. stutzeriD-PhgAT as the final step. These whole cell biotransformations produced a range of D-Phg derivatives in one pot from racemic mandelic acid, styrene and L-Phe starting materials.53 Since we have shown the D-PhgAT to be a versatile biocatalyst in being able to use a range of amino acceptors it suggests that similar cascades could be constructed to allow conversion of simple building blocks to produce a variety of enantiopure aromatic D-amino acids.53
The determination of the first crystal structure of the D-PhgAT with its bound PLP cofactor has shed light on the unique stereo-inverting and enantioselective properties of the enzyme. The broad substrate scope is explained by a large active site cavity and two pockets each containing an essential Arg residue that is crucial for the catalytic activity. Our study highlights key active site residues that are involved in the catalytic mechanism and potentially control the exquisite R-selectivity of the enzyme. These features suggest that the O-pocket of D-PhgAT could be engineered to accept even bulkier pro-chiral substrates and convert them to enantiopure R-products. Our study paves the way for D-PhgAT engineering in order to further expand the substrate scope of this enzyme.54,55 In future, structures of the enzyme with substrates and products, combined with a directed evolution campaign, will allow the product scope of D-PhgAT to be further expanded to take advantage of the unique properties of this versatile biocatalyst.
Reactions containing 0.51 mg mL−1D-PhgAT, 10 mM L/D amino donor, 10 mM BZF in 0.1 M CAPS pH 9.5, 150 mM NaCl, 50 μM PLP buffer were incubated at 37 °C before being terminated at 18 h by diluting 40-fold in the chiral mobile phase. Reactions were then analysed by chiral HPLC using a Chirobiotic-T column as stated before. Percentage conversions were normalized relative to L-Glu as the best amino donor.
Crystal structures were overlayed using the Pymol Align feature within Pymol (The PyMOL Molecular Graphics System, Version 1.2r3pre, Schrödinger, LLC). The RMSD values for each overlay are: 6G1F with 5G2Q – 2.18 Å over 1610 atoms; 6G1F with 5G09 – 2.12 Å over 1546 atoms; 6G1F with 6DVS – 0.351 Å over 2205 Å.
Footnotes |
† The X-ray crystal structure of the D-PhgAT PLP-bound internal aldimine reported here is available in the Protein Data Bank (https://www.rcsb.org/) with the accession code 6G1F. |
‡ Electronic supplementary information (ESI) available. See DOI: 10.1039/d0cy01391a |
§ Current address: LEITAT Technological Center, Terrassa (Barcelona), Spain. |
¶ Current address: Department of Chemistry, United Arab Emirates University, Al Ain (UAE). |
This journal is © The Royal Society of Chemistry 2020 |