Research Article
18 July 2019

Crystal Structure and Biophysical Analysis of Furfural-Detoxifying Aldehyde Reductase from Clostridium beijerinckii

ABSTRACT

Many aldehydes, such as furfural, are present in high quantities in lignocellulose lysates and are fermentation inhibitors, which makes biofuel production from this abundant carbon source extremely challenging. Cbei_3974 has recently been identified as an aldo-keto reductase responsible for partial furfural resistance in Clostridium beijerinckii. Rational engineering of this enzyme could enhance the furfural tolerance of this organism, thereby improving biofuel yields. We report an extensive characterization of Cbei_3974 and a single-crystal X-ray structure of Cbei_3974 in complex with NADPH at a resolution of 1.75 Å. Docking studies identified residues involved in substrate binding, and an activity screen revealed the substrate tolerance of the enzyme. Hydride transfer, which is partially rate limiting under physiological conditions, occurs from the pro-R hydrogen of NADPH. Enzyme isotope labeling revealed a temperature-independent enzyme isotope effect of unity, indicating that the enzyme does not use dynamic coupling for catalysis and suggesting that the active site of the enzyme is optimally configured for catalysis with the substrate tested.
IMPORTANCE Here we report the crystal structure and biophysical properties of an aldehyde reductase that can detoxify furfural, a common inhibitor of biofuel fermentation found in lignocellulose lysates. The data contained here will serve as a guide for protein engineers to develop improved enzyme variants that would impart furfural resistance to the microorganisms used in biofuel production and thus lead to enhanced biofuel yields from this sustainable resource.

INTRODUCTION

An ideal source of carbohydrates for biofuel fermentation is lignocellulose, an abundant waste product which is available at low cost and does not affect food security (1). Fermentable sugars are most commonly released from lignocellulose using an acid pretreatment (2). One of the major drawbacks of this method is the release of aldehydes, organic acids, and phenols, which severely inhibit growth and limit the final yield of biofuel (3). While it has been observed that Clostridium spp. have a higher tolerance against the aldehyde inhibitors furfural and hydroxymethylfurfural than other organisms, the high levels of inhibitors found in lignocellulose lysates are nevertheless hugely problematic (46).
A recent investigation has identified two genes from Clostridium beijerinckii that encode enzymes that reduce aldehydes to less-toxic alcohols (7, 8). These enzymes showed activity against furfural, hydroxymethyl furfural, and benzaldehyde, which are all common fermentation inhibitors (7). Furthermore, the genes encoding these enzymes are upregulated during furfural stress, suggesting the physiological relevance of these enzymes to protect C. beijerinckii (8). One of these enzymes, Cbei_3904, belongs to the short-chain dehydrogenase (SDR) family and the other, Cbei_3974, to the aldo-keto reductase (AKR) family (7). It is highly desirable to engineer greater catalytic efficiency into these enzymes to more rapidly eliminate toxic aldehydes, thereby enhancing resistance to aldehyde inhibition.
A prerequisite for rational engineering of an enzyme is a thorough understanding of its mechanism. In this study, the furfural-transforming AKR Cbei_3974 was characterized to provide valuable information for protein engineers. The substrate specificity, steady-state kinetic parameters, and crystal structure of Cbei_3974 have been determined. In addition, the rate-limiting step of the reaction was identified and the coupling of dynamic motions to the active site explored.

RESULTS

Substrate profile.

It has previously been suggested that Cbei_3974 may be useful to alleviate the toxicity of furfural during the fermentation of acid-treated lignocellulose lysates (7). NADPH-dependent activity toward furfural was previously reported for this enzyme, but no characterization of the reaction product was shown (7). To confirm that the enzyme indeed generates the less-toxic alcohol from the aldehyde, the reaction product from an enzyme-substrate-NADPH incubation was analyzed by gas chromatography-mass spectrometry (GC-MS) in parallel with controls containing no NADPH or no enzyme. After 5 h of incubation, a new compound could be detected on the GC trace (Fig. 1). The retention time and fragmentation pattern were identical to those of a commercial standard of furfuryl alcohol. This compound was not detected in either of the controls, showing that its formation was enzyme catalyzed and NADPH dependent.
FIG 1
FIG 1 GC-MS analysis of the Cbei_3974 reaction product. (A) Mixture of enzyme, NADPH, and furfural at time zero; (B) incubation of furfural and NADPH without enzyme after 5 h; (C) incubation of furfural and enzyme without NADPH for 5 h; (D) incubation of furfural, NADPH, and enzyme after 5 h; (E) same as for panel D but doped with furfuryl alcohol; (F) standard of furfuryl alcohol; (G) mass spectrum of reaction product from incubation of furfural, NADPH, and enzyme after 5 h; (H) mass spectrum of furfuryl alcohol standard. The full-length GC traces can be found in Fig. S1 in the supplemental material.
Cbei_3974 has previously been shown to also exhibit activity with hydroxymethyl furfural, benzaldehyde, and butyraldehyde (7). To more fully explore the substrate scope of the enzyme, a selection of aldehydes, ketones, and alcohols was chosen. These putative substrates were incubated at 2 mM with the enzyme and an excess of NADPH, NADH, or NADP+. The change in cofactor concentration was measured continuously by UV spectroscopy to give the reaction rate (Table 1). Surprisingly, l-glyceraldehyde-3-phosphate gave only negligible activity despite the enzyme sharing 57.7% identity with Escherichia coli YghZ, which converts l-glyceraldehyde-3-phosphate to l-glycerol-3-phosphate as part of a novel triose phosphate isomerase (TIM) bypass that allows the formation of dihydroxyacetone phosphate under gluconeogenic conditions, when TIM is genetically inactivated (9). Purified YghZ was shown to reduce l-glyceraldehyde-3-phosphate to l-glycerol-3-phosphate, which can be converted to dihydroxyacetone by l-glycerol-3-phosphate dehydrogenase, thus complementing TIM deficiency. Rather unexpectedly, YghZ is stereospecific for the l-enantiomer of the substrate, whereas the TIM substrate is d-glyceraldehyde-3-phosphate. It was therefore proposed that a spontaneous reaction may interconvert the two enantiomers (9). On this basis, YghZ and enzymes with similar sequences, including Cbei_3974, are annotated in the KEGG database (https://www.kegg.jp/) as l-glyceraldehyde-3-phosphate reductases (10, 11). Our results demonstrate that this annotation is incorrect for Cbei_3974.
TABLE 1
TABLE 1 Substrate screen of activity of Cbei_3974 with putative substrates (2 mM) and 0.4 mM NADPH
SubstrateSp act (μmol/min/mg)a
Furfural416.86 ± 4.81
BenzaldehydeND
4-Pyrridine carboxaldehyde4,652.70 ± 671.05
3-Pyrridine carboxaldehyde1,519.52 ± 61.48
4-Nitrobenzaldehyde4,390.44 ± 61.61
IsatinND
Methylglyoxal2,852.59 ± 335.72
FormaldehydeND
AcetaldehydeND
PropionaldehydeND
Butyraldehyde307.37 ± 23.65
ValeraldehydeND
Furfural alcoholND
Benzyl alcoholND
4-Pyridine methanolND
MethanolND
EthanolND
Isopropyl alcoholND
ButanolND
2,4-PentanedioneND
4-AcetylpyridineND
2-ButanoneND
l-ArabinoseND
d-GlucoseND
β-LactoseND
l-Glyceraldehyde-3-phosphate269.71 ± 34.05
a
Errors indicate standard deviation from three repeats. ND, no detectable activity.
Only minimal activity was found for the previously identified substrates, furfural and butyraldehyde (7), while the enzyme had no measurable activity with benzaldehyde at the concentration tested (2 mM). In contrast, 4-pyridinecarboxaldehyde, which differs from benzaldehyde only by the presence of a nitrogen atom in the aromatic ring, was the most kinetically efficient among all of the substrates examined. Turnover was 17 times faster than for l-glyceraldehyde-3-phosphate and 11 times faster than for furfural. Similarly, propionaldehyde did not show activity at 2 mM, while the more polar methylglyoxal, a dialdehyde of the same chain length, gave strong activity 6.8 times faster than furfural. No activity could be detected with ketones or alcohols at the concentrations tested. The enzyme was specific for NADPH, with no activity detectable with 0.4 mM NADH. To determine whether the differences between substrates were caused by differences in kcat or Km, steady-state kinetics were measured for the 5 fastest substrates, not including 4-nitrobenzaldehyde, which was not soluble enough to achieve saturation. All the aldehyde substrates resulted in Michaelis constants in the millimolar range, but NADPH had higher affinity as indicated by a lower Km of 32 μM (Table 2). The best substrate was 4-pyridine carboxaldehyde, with a kcat of 10 s−1 and a Km of 3.87 mM. This was closely followed by methylglyoxal, with a similar kcat of 8.52 s−1 but a higher Km of 12.9 mM. Furfural has previously been shown to have an extraordinarily high kcat of 1.4 × 105 s−1 at 40°C (7). In our experiments the kcat was measured at 19°C and was 2.72 s−1. This was lower than expected, even considering the lower temperature, but is more realistic. The Km of 34.9 mM measured here is in agreement with the literature value (7).
TABLE 2
TABLE 2 Kinetic measurements for Cbei_3974a
Substratekcat (s−1)Km (mM)kcat/Km (s−1 M−1)
Furfural2.72 ± 0.2734.9 ± 7.678
Butyraldehyde5.16 ± 1.0440.79 ± 4.5127
3-Pyridine carboxaldehyde4.97 ± 1.3815.7 ± 2.6317
Methylglyoxal8.52 ± 0.2612.9 ± 1.5660
4-Pyridine carboxaldehyde10.60 ± 0.23.87 ± 0.342,739
NADPH with:   
    4-Pyridine carboxaldehyde10.0 ± 1.020.032 ± 0.02312,500
    Furfural2.59 ± 0.1450.015 ± 0.003172,666
a
All measurements were made at 19°C. Errors show the standard error from fitting the data to the Michaelis-Menten equation in GraphPad Prism.

Stereochemistry of NADPH transfer.

The hydride transfer in aldehyde reductases occurs from either the pro-R or pro-S hydrogen on C-4 of the nicotinamide ring of NADPH. Typically, short-chain reductases (SDRs) transfer the pro-S hydrogen, while aldo-keto reductases (AKRs) transfer the pro-R hydrogen (12). To determine the stereospecificity of Cbei_3974, the enzyme was incubated with NADPH or (4R)-[4-2H]NADPD with an excess of the substrate 4-pyridinecarboxaldehyde. The reaction products were analyzed to determine whether the deuterium had been incorporated into the alcohol product or remained on the nicotinamide cofactor. The alcohol reaction product was extracted with chloroform and analyzed by GC-MS. The incubation with (4R)-[4-2H]NADPD gave a product which was 1 atomic mass unit larger than that from the incubation with NADPH, consistent with the incorporation of deuterium. In a duplicate reaction, the nucleotide product, NADP+, was purified by ion-exchange chromatography, freeze-dried, and dissolved in D2O. Nuclear magnetic resonance (NMR) analysis of NADP+ from the reaction compared with commercial NADP+ showed identical spectra, confirming that the deuterium at the pro-R position had been transferred from (4R)-[4-2H]NADPD (Fig. 2). Cbei_3974 therefore transfers the pro-R hydrogen from NADPH, in accordance with other members of the AKR superfamily.
FIG 2
FIG 2 (A and B) MS analysis of the alcohol product from an incubation of 4-pyridine carboxaldehyde, enzyme, and either NADPH (A) or (4R)-[4-2H]NADPD (B). The corresponding GC traces can be seen in Fig. S2 in the supplemental material. (C) NMR spectra showing the purified nucleotide reaction product obtained from incubations of Cbei_3974 with (4R)-[4-2H]NADPD and 4-pyridine carboxaldehyde (red) against a standard of NADP+ (blue).

Substrate KIE.

To ascertain if hydride transfer is the rate-limiting step, the substrate kinetic isotope effect (KIE) for NADPH versus (4R)-[4-2H]NADPD was determined for a range of substrates. A KIE on kcat of 2.13 to 2.58 was observed across the 5 substrates tested (Fig. 2), suggesting that the catalytic step is at least partially rate limiting for all substrates tested.

Heavy-enzyme KIE.

The effect of protein dynamics on catalysis was investigated by heavy-enzyme production, where the nonexchangeable carbon and nitrogen atoms were replaced with their heavy counterparts (13C, 15N) to slow protein motions without affecting the electrostatics. A reactivity difference between the “heavy” (labeled) and “light” (natural-abundance) enzymes indicates that protein motions affect the catalysis (13). As the substrate KIE measurements indicated that hydride transfer is partially rate limiting, steady-state measurements were used to determine any effect that slower protein motions in the heavy enzyme may have on the catalytic step.
Heavy enzyme (15N, 13C) was produced in M9 medium with labeled feedstocks and purified to homogeneity (see Fig. S3 in the supplemental material). The incorporation of the heavy isotopes was confirmed by mass spectrometry on the purified enzyme, which revealed a 5.5% mass increase (see Fig. S4 in the supplemental material). To determine if the protein was correctly folded, the circular dichroism (CD) spectrum and melting temperature were recorded and compared for both the “heavy” and “light” enzymes (see Fig. S5 in the supplemental material). The two enzymes gave identical spectra and had nearly identical melting temperatures of 62.4°C ± 0.1 and 63.8°C ± 0.2, respectively, indicating that isotopic labeling does not significantly alter protein folding.
Steady-state kinetics were used to determine the kcat values for the “heavy” and “light” enzymes with a range of substrates at 19°C (3). All substrates gave an enzyme KIE of near unity, implying that there were no mass-dependent effects and that dynamic coupling was minimal (Fig. 3). Although some authors have proposed that enzymes use “promoting motions” to drive catalysis (1417), this result is consistent with a growing body of literature that shows that dynamic effects become significant only outside physiological conditions and only when poorly tolerated substrates that necessitate rearrangement of the active site are utilized (1820). The enzyme does not, therefore, use dynamic motions as a part of its catalytic mechanism.
FIG 3
FIG 3 Kinetic isotope effects. Gray circles, kcat values for natural-abundance enzyme with NADPH; blue circles, enzyme KIEs; red circles, substrate KIEs (NADPH). All measurements were made at 19°C. Errors are standard deviations from three repeats.
A recent study on the thermophilic Geobacillus stearothermophilus alcohol dehydrogenase (BsADH) showed that significant heavy-enzyme KIEs manifest only below its physiological temperature (40°C) and only with poor substrates (19). The temperature dependency of heavy-enzyme KIEs has been suggested to be an indicator of whether an enzyme is optimized for utilization of a particular substrate (19). The temperature dependence of the KIE for Cbei_3974-catalyzed reduction of 3-pyridine carboxaldehyde was constant over the temperature range from 11 to 44°C (Fig. 4), suggesting that the active-site architecture of the enzyme is optimized for this substrate.
FIG 4
FIG 4 Temperature dependency of the enzyme kinetic isotope effect (kcat for light enzyme/kcat for heavy enzyme) with 3-pyridine carboxaldehyde as a substrate. Error bars show the standard deviation from three repeats.

Single-crystal X-ray structure.

The protein was cocrystallized with NADPH and the structure solved by molecular replacement using PDB entry 5T79, which is the crystal structure for STM2406, an AKR from Salmonella enterica serovar Typhimurium of unknown physiological function but with a substrate profile similar to that of Cbei_3974 (21). The two proteins have 60.91% sequence identity and a root mean square deviation of 0.89. The structure was refined at 1.75 Å to an Rfactor of 16.5% (Rfree of 19.3%). The structure consists of alternating α-helices and β-strands forming an 8-stranded TIM barrel with some extra helices (Fig. 5).
FIG 5
FIG 5 Crystal structure of Cbei_3974. (A) Cartoon representation showing the TIM barrel fold. NADPH can be seen in gold, with the nicotinamide ring in the central cavity. (B) Surface representation showing the exposed nature of the active site. NADPH is shown in gold. (C) Residues involved in binding NADPH. Red semicircles identify hydrophobic interactions, and residues involved in hydrogen bonds are shown in blue. The green numbers indicate hydrogen bond distance. The figure was prepared with LigPlot+ (42).
This motif is conserved across the AKR superfamily (22). Both this structure and STM2406 have an unusual N terminus consisting of a long loop and a β-hairpin. Most AKR structures, including the structure of Coptotermes gestroi AKR1 (another AKR known to reduce furfural), have a shorter N-terminal tail consisting of only the β-hairpin or, in the case of the AKR7 family, have no N-terminal tail (21, 23, 24). The function of this extra sequence is unclear. Conversely, the C terminus is truncated and is lacking a loop that is present in many AKRs, leaving the active site exposed to solvent (24). AKRs which omit this loop have low catalytic efficiency consistent with the measured millimolar Michaelis constants (21, 25, 26). Deletion of the C-terminal loop from human aldose reductase AKR1B1 (26), Bacillus subtilis YhdN and YvgN, and Pseudomonas aeruginosa PA1127 (21) resulted in a dramatic loss of catalytic efficiency.
NADPH sits in a mostly open cleft with a hydrophobic center and polar residues at the ends where the adenine base and nicotinamide ring bind. The adenine base is held in place by hydrogen bonds to Glu-307 and Asn-308. The nucleotide 2′-phosphate that distinguishes NADPH from NADH is hydrogen bonded to Gln-304 and Ser-300. The diphosphate makes hydrogen bond contact with the backbone oxygen of Leu-225.
There is an area of missing electron density between residues Ile-238 and Leu-256. In Coptotermes gestroi AKR1 and human aldose reductase, this region forms a mobile loop that would strap the cofactor in place across the diphosphate bridge (23, 27). The lack of density in Cbei_3974 indicates that the region is disordered and suggests that the loop is not trapping the cofactor.
The canonical mechanism of AKRs involves hydride transfer from NADPH to the carbonyl acceptor (12). This is followed by protonation from an active-site tyrosine as part of a proton relay from histidine and bulk water (12). Neighboring aspartate and lysine residues lower the pKa of tyrosine to enable it to function as an acid (12). In the close homologue STM2406, the catalytic tetrad consists of Tyr-66, Asp61, Lys-97, and His-138 (21). All these residues are conserved in Cbei_3974 (identical numbering). It was not possible to obtain crystal structures of protein-product complexes due to the low affinity of the ligands. Therefore, docking was used to predict the possible binding of substrates. AutoDock Vina (28) was used to dock furfural and the best substrate, 4-pyridine carboxaldehyde, into the active site. The best pose was selected on the basis of proximity to NADPH and the proposed catalytic residues. These poses are illustrated in Fig. 6. Both substrates are orientated toward the pro-R hydrogen of NADPH, consistent with the experimentally determined stereochemistry. The carbonyl oxygen of 4-pyridine carboxaldehyde is within hydrogen bonding distance of the exocyclic amide of NADPH and makes hydrophobic contacts with residues Asn-65, Trp-33, and Tyr-100. These residues are conserved in STM2406 and have been shown to be important for binding in that enzyme (21). Furfural docked into the active site in a similar location but with a different orientation, possibly due to its smaller size. The active site has many polar residues, which may explain why the more hydrophobic aldehydes such as benzaldehyde are less favored. Asn-65 and Tyr-100 contribute to the polar surface of the active site, and therefore alteration of these residues to more hydrophobic ones may help improve activity for hydrophobic substrates. In STM2406, which has a very similar active site, the variant Asn-65-Met (both enzymes have the same residue numbering) gave a 341% increase in activity toward 3-pyridinecarboxaldehyde compared with that of the wild type and a 2-fold decrease in Km (21). Alterations of Tyr-100 to aspartate, leucine, isoleucine, or valine mostly resulted in insoluble proteins, while Tyr-100-Ala showed decreased activity, but this may have been due to a loss of steric bulk by replacing a phenyl group with a hydrogen atom (21).
FIG 6
FIG 6 Docking of 4-pyridine carboxaldehyde (A and B) and furfural (C and D) into the active site. Substrates are shown in orange. (A and C) Hydrophobic surface rendering; (B and D) identification of residues involved in binding (gray) and catalysis (green).

DISCUSSION

Cbei_3974, an enzyme that could putatively address the problem of aldehyde toxicity in biofuel fermentation from lignocellulose (7), has been extensively characterized and its crystal structure solved. Though annotated as an l-glyceraldehyde-3-phosphate reductase, it shows only low activity toward this substrate. Instead, it preferentially catalyzes reactions with 4-pyridine carboxaldehyde and methylglyoxal, of which only the latter is likely to be naturally occurring inside the cell. Methylglyoxal is a toxic product formed from dihydroxy-acetone-phosphate by methylglyoxal synthase to release phosphate (29). Detoxification of methylglyoxal by Clostridium results in the formation of 1,2-propanediol in a pathway that requires aldo-keto reductase activity (30). The physiological relevance of methylglyoxal reductase activity of Cbei_3974 is, however, questionable, given the low affinity of the substrate, with a Km in the region of 12 mM. The low affinity may be part of a mechanism to conserve NADPH, as depletion of NADPH can be just as lethal as aldehyde accumulation (31), but this is unlikely given that the 50% lethal dose (LD50) of methylglyoxal is likely to be significantly less than the Km, rendering the enzyme useless for detoxification. It is also possible that the enzyme requires an interaction partner or a posttranslational modification for activity or that the experimental conditions were not optimal. An examination of the C. beijerinckii genome in the KEGG database (10, 11) shows that the cbei_3974 gene is not part of a biosynthetic gene cluster, but it is adjacent to a gene for a putative MerR transcription regulator. These typically respond to environmental stimuli to upregulate stress response proteins (32). In conclusion, the natural substrate of Cbei_3974 remains unclear, but it is likely to be involved in a stress response.
Studies with (4R)-[4-2H]NADPD demonstrated that hydride transfer occurs from the pro-R hydrogen of NADPH and is partially rate limiting. Isotopically labeled heavy enzyme (13C, 15N) gave kcat constants identical to those obtained with the natural-abundance enzyme, indicating that the slower protein motions in the heavy enzyme did not impact the catalytic step. This shows that the enzyme does not use “promoting motions” to drive the chemical transformation.
The enzyme kinetic isotope effect was independent of temperature, suggesting that the enzyme’s active site is optimally configured for the use of the tested substrates; the physiological substrate is therefore likely to be structurally similar (19).
The crystal structure of Cbei_3974 revealed a typical AKR structure based around a TIM barrel fold and is essentially the same as that of STM2406 (21). Docking of 4-pyridinecarboxaldehyde and furfural revealed residues that may be involved in substrate binding. Although no enzyme variants were generated in this study, these residues represent targets for future work to generate an improved enzyme for more efficient detoxification of furfural.

MATERIALS AND METHODS

Materials.

A pET-14b vector harboring a codon-optimized gene encoding Cbei_3974 was purchased from GenScript (the sequence is given in Fig. S6 in the supplemental material). This also encodes a 6× His tag and thrombin cleavage site upstream of Cbei_3974.
NADPH was obtained from Fisher or Apollo Scientific. (4R)-[4-2H]NADPD was prepared from NADP+ (Melford) and isopropanol-d8 (Acros) according to the published protocol (33). 15NH4Cl2 and [13C]glucose were obtained from Goss Scientific, Cheshire, UK. Furfural was obtained from VWR, methylglyoxal from Apollo Scientific, and 3-pyridine carboxaldehyde from Acros. All other chemicals were obtained from Sigma-Aldrich.

Crystallography.

Cbei_3974 was overproduced in BL21(DE3) cells and purified by Ni affinity chromatography as previously described (7). The protein was dialyzed against 10 mM HEPES-NaOH (pH 7.5)–300 mM NaCl and concentrated to 10 mg/ml. Crystallization trials were performed using the sitting-drop vapor diffusion method by mixing 0.5 μl protein stock solution and 0.5 μl of a seed stock with 0.5 μl reservoir solution. The seed stock was obtained from microcrystals grown in 100 mM morpholinepropanesulfonic acid (MOPS) (pH 7.3)–13% polyethylene glycol (PEG) 8000–750 mM NH4Cl with 2 mM NADPH and 2 mM 4-pyridine methanol. Diffracting crystals were obtained from 90 mM MOPS (pH 7.6)–271 mM NH4Cl–2.7% PEG 8000, with 10 mM NADPH and 2 mM 4-pyridine methanol added to the protein prior to crystallization. The crystals were transferred to cryoprotectant (90 mM MOPS [pH 7.6], 271 mM NH4Cl, 2.7% PEG 8000, 25% ethylene glycol) and flash-frozen with liquid nitrogen.
The X-ray diffraction data were collected at 100 K at Diamond Light Source (Oxfordshire, UK) on beamline I04-1 and integrated with XDS (34) in the xia2 package (35). The data were scaled, reduced, and analyzed with AIMLESS and TRUNCATE in CCP4i (36). The structure was solved by molecular replacement with PHASER (37) using coordinates from PDB no. 5T79 as a searching model (21). The structure model was adjusted with COOT (38) and refined with REFMAC5 (39). Graphical representations were prepared in Chimera (40), PyMOL (The PyMOL Molecular Graphics System, version 1.8.X [Schrödinger, LLC]), YASARA View (41) and LigPlot+ (42). X-ray crystallography data and refinement statistics are given in Table S1 in the supplemental material.

Molecular docking.

Ligand structures were downloaded from PubChem (https://pubchem.ncbi.nlm.nih.gov/) (43) as SDF files and converted into mol2 format using Chimera (40). Ligand and protein were converted into PDBQT format using AudoDockTools1.5.6 and docked using AutoDock Vina (28). Graphical representations were prepared in PyMOL (The PyMOL Molecular Graphics System, Version 1.8.X [Schrödinger, LLC]), Chimera (40), and LigPlot+ (42).

Production of heavy and natural-abundance enzymes.

Arctic Express DE3 cells harboring pET-14b_Cbei3974 were grown in 20 ml M9 minimal medium containing 100 μg/ml carbenicillin overnight at 37°C with shaking. This was diluted 1:50 into 0.5 liters of M9 medium containing either natural-abundance isotopes or [13C]glucose and 15NH4Cl2 for heavy-enzyme production. Cultures were grown to an optical density at 600 nm (OD600) of 1.0 at 37°C and 220 rpm. Cultures were cooled to 12°C, and 0.4 mM IPTG (isopropyl-β-d-thiogalactopyranoside) was added to induce gene expression. Cells were harvested at 4,000 rpm after 16 h of growth at 12°C and 200 rpm. The protein was purified as previously described (7).

Substrate screen.

Cbei_3974 (0.585 μM), 0.4 mM NADPH, NADH, or NADP+, and 2 mM putative substrate were mixed in 20 mM potassium phosphate buffer, pH 7.0. The subsequent change in NADPH concentration was monitored at 340 nm (ε340 = 6,220 M−1 cm−1) for 1 min, using a Shimadzu UV-2401PC spectrophotometer in 5-mm quartz cuvettes, to give the reaction rate.

Enzyme kinetics.

Kinetic parameters were determined with 117 nM Cbei_3974 with one of the two substrates; one was held at a saturating level, while the other was varied across a concentration range of 0 to 75 mM (aldehydes) or 0 to 0.2 mM (NADPH). Rates were measured as described above and the data fitted to the Michaelis-Menten equation using GraphPad Prism version 7.00 for Windows (GraphPad Software, La Jolla, CA, USA). Each data point is the average from 3 repeats. For heavy-enzyme kinetics, a minimum of two data sets of triplicates were collected.

MS.

Liquid chromatography-mass spectrometry (LC-MS) was performed on a Waters Synapt G2-Si quadrupole time-of-flight mass spectrometer coupled to a Waters Acquity H-Class ultraperformance liquid chromatography (UPLC) system. The column was an Acquity UPLC protein BEH C4 (300 Å, 1.7 μm by 2.1 mm by 100 mm) operated in reverse phase and held at 60°C. The gradient employed was 95% A to 35% A over 50 min, where A is H2O with 0.1% HCO2H and B is acetonitrile (ACN) with 0.1% HCO2H. Data were collected in positive ionization mode and analyzed using Waters MassLynx software version 4.1. Deconvolution of protein charged states was obtained using the maximum entropy 1 processing software.

GC-MS analysis of reaction products.

Mixtures containing 5 μM Cbei_3974, 4.6 mM NADPH or (4R)-[4-2H]NADPD, and 28 mM aldehyde in a final volume of 40 μl 20 mM potassium phosphate (pH 7.0) were incubated at 40°C for 1 h. Aliquots (4 μl) were removed at time zero and at 5 or 7 h and quenched with 1 ml of chloroform, which also served to extract the organic molecules. A 5-μl aliquot of the organic layer was injected onto a PerkinElmer Clarus 680 gas chromatograph. The initial temperature was 40°C, which was held for 1 min, and elution was with a gradient rising to 150°C at 15°C/min, holding at 150°C for 1 min. After a 3-min solvent delay, mass spectra were collected over the range of 45 to 200 E+. Control reactions with either enzyme or NADPH omitted were also performed, and a standard of furfuryl alcohol was run on the GC-MS.

1H NMR of nicotinamide cofactors.

Cbei_3974 (1 μM), 2 mM (4R)-[4-2H]NADPD, and 25 mM 4-pyridine carboxaldehyde were incubated at 37°C for 3 h. NADP+ was purified from the reaction on a SAX-10 column using published methodology (33), freeze-dried, and redissolved in D2O. Commercial standards of NADPH and NADP+ were also dissolved in D2O. NMR spectra of standards and reaction product were collected on a Bruker 400 instrument.

CD.

Circular dichroism (CD) measurements were performed on an Applied Photophysics Chirascan spectrometer using 7 μM protein in 20 mM potassium phosphate buffer (pH 7.0)−20% glycerol. Spectra were recorded over a temperature range of 5 to 85°C from 200 nm to 400 nm. The melting temperature was calculated by fitting the data in SigmaPlot (Systat Software, San Jose, CA).

Accession number(s).

The X-ray structure solved in this study was deposited into the Protein Data Bank (http://www.rcsb.org/pdb/) with accession code 6HG6.

ACKNOWLEDGMENTS

We thank Diamond Light Source for access to beamline I04-1 (beamtime code mx1484-11). We gratefully acknowledge helpful discussions with Enas Behiry.
This work was supported by the United Kingdom’s Biotechnology and Biological Sciences Research Council through grants BB/J005266/1 and BB/L020394/1.

Supplemental Material

File (aem.00978-19-s0001.pdf)
ASM does not own the copyrights to Supplemental Material that may be linked to, or accessed through, an article. The authors have granted ASM a non-exclusive, world-wide license to publish the Supplemental Material files. Please contact the corresponding author directly for reuse.

REFERENCES

1.
Isikgor FH, Becer CR. 2015. Lignocellulosic biomass: a sustainable platform for the production of bio-based chemicals and polymers. Polym Chem 6:4497–4559.
2.
Kumar P, Barrett DM, Delwiche MJ, Stroeve P. 2009. Methods for pretreatment of lignocellulosic biomass for efficient hydrolysis and biofuel production. Ind Eng Chem Res 48:3713–3729.
3.
Liu ZL, Blaschek HP. 2010. Biomass conversion inhibitors and in situ detoxification, p 233–259. In Vertes AA, Qureshi N, Blaschek HP, Yukawa H (ed), Biomass to biofuels. Blackwell Publishing Ltd., Oxford, UK.
4.
Ezeji T, Qureshi N, Blaschek HP. 2007. Butanol production from agricultural residues: impact of degradation products on Clostridium beijerinckii growth and butanol fermentation. Biotechnol Bioeng 97:1460–1469.
5.
Zhang Y, Han B, Ezeji TC. 2012. Biotransformation of furfural and 5-hydroxymethyl furfural (HMF) by Clostridium acetobutylicum ATCC 824 during butanol fermentation. Nat Biotechnol 29:345–351.
6.
Ujor V, Agu CV, Gopalan V, Ezeji TC. 2014. Glycerol supplementation of the growth medium enhances in situ detoxification of furfural by Clostridium beijerinckii during butanol fermentation. Appl Microbiol Biotechnol 98:6511–6521.
7.
Zhang Y, Ujor V, Wick M, Ezeji TC. 2015. Identification, purification and characterization of furfural transforming enzymes from Clostridium beijerinckii NCIMB 8052. Anaerobe 33:124–131.
8.
Zhang Y, Ezeji TC. 2013. Transcriptional analysis of Clostridium beijerinckii NCIMB 8052 to elucidate role of furfural stress during acetone butanol ethanol fermentation. Biotechnol Biofuels 6:66.
9.
Desai KK, Miller BG. 2008. A metabolic bypass of the triosephosphate isomerase reaction. Biochemistry 47:7983–7985.
10.
Kanehisa M, Sato Y, Kawashima M, Furumichi M, Tanabe M. 2016. KEGG as a reference resource for gene and protein annotation. Nucleic Acids Res 44:D457–D462.
11.
Kanehisa M, Furumichi M, Tanabe M, Sato Y, Morishima K. 2017. KEGG: new perspectives on genomes, pathways, diseases and drugs. Nucleic Acids Res 45:D353–D361.
12.
Penning TM. 2015. The aldo-keto reductases (AKRs): overview. Chem Biol Interact 234:236–246.
13.
Kholodar SA, Ghosh AK, Kohen A. 2017. Measurement of enzyme isotope effects. Methods Enzymol 596:43–83.
14.
Silva RG, Murkin AS, Schramm VL. 2011. Femtosecond dynamics coupled to chemical barrier crossing in a Born-Oppenheimer enzyme. Proc Natl Acad Sci U S A 108:18661–18665.
15.
Kipp DR, Silva RG, Schramm VL. 2011. Mass-dependent bond vibrational dynamics influence catalysis by HIV-1 protease. J Am Chem Soc 133:19358–19361.
16.
Pudney CR, Guerriero A, Baxter NJ, Johannissen LO, Waltho JP, Hay S, Scrutton NS. 2013. Fast protein motions are coupled to enzyme H-transfer reactions. J Am Chem Soc 135:2512–2517.
17.
Antoniou D, Ge X, Schramm VL, Schwartz SD. 2012. Mass modulation of protein dynamics associated with barrier crossing in purine nucleoside phosphorylase. J Phys Chem Lett 3:3538–3544.
18.
Luk LYP, Loveridge EJ, Allemann RK. 2014. Different dynamical effects in mesophilic and hyperthermophilic dihydrofolate reductases. J Am Chem Soc 136:6862.
19.
Behiry E, Ruiz-Pernia JJ, Luk L, Tunon I, Moliner V, Allemann RK. 2018. Isotope substitution of promiscuous alcohol dehydrogenase reveals origin of substrate preference in transition state. Angew Chem Int ed Engl 57:3128–3131.
20.
Ruiz-Pernía JJ, Behiry E, Luk LYP, Loveridge EJ, Tuñón I, Moliner V, Allemann RK. 2016. Minimization of dynamic effects in the evolution of dihydrofolate reductase. Chem Sci 7:3248–3255.
21.
Kim T, Flick R, Brunzelle J, Singer A, Evdokimova E, Brown G, Joo JC, Minasov GA, Anderson WF, Mahadevan R, Savchenko A, Yakunin AF. 2017. Novel aldo-keto reductases for the biocatalytic conversion of 3-hydroxybutanal to 1,3-butanediol: structural and biochemical studies. Appl Environ Microbiol 83:e03172-16.
22.
Sanli G, Dudley JI, Blaber M. 2003. Structural biology of the aldo-keto reductase family of enzymes: catalysis and cofactor binding. Cell Biochem Biophys 38:79–101.
23.
Tramontina R, Franco Cairo JPL, Liberato MV, Mandelli F, Sousa A, Santos S, Rabelo SC, Campos B, Ienczak J, Ruller R, Damásio ARL, Squina FM. 2017. The Coptotermes gestroi aldo-keto reductase: a multipurpose enzyme for biorefinery applications. Biotechnol Biofuels 10:4.
24.
Barski OA, Tipparaju SM, Bhatnagar A. 2008. The aldo-keto reductase superfamily and its role in drug metabolism and detoxification. Drug Metab Rev 40:553–624.
25.
Ehrensberger AH, Wilson DK. 2004. Structural and catalytic diversity in the two family 11 aldo-keto reductases. J Mol Biol 337:661–673.
26.
Bohren KM, Grimshaw CE, Gabbay KH. 1992. Catalytic effectiveness of human aldose reductase. Critical role of C-terminal domain. J Biol Chem 267:20965–20970.
27.
Wilson DK, Bohren KM, Gabbay KH, Quiocho FA. 1992. An unlikely sugar substrate site in the 1.65 A structure of the human aldose reductase holoenzyme implicated in diabetic complications. Science 257:81–84.
28.
Trott O, Olson AJ. 2009. AutoDock Vina: improving the speed and accuracy of docking with a new scoring function, efficient optimization, and multithreading. J Comput Chem 31:455–461.
29.
Huang K, Rudolph FB, Bennett GN. 1999. Characterization of methylglyoxal synthase from Clostridium acetobutylicum ATCC 824 and its use in the formation of 1,2-propanediol. Appl Environ Microbiol 65:3244–3247.
30.
Liyanage H, Kashket S, Young M, Kashket ER. 2001. Clostridium beijerinckii and Clostridium difficile detoxify methylglyoxal by a novel mechanism involving glycerol dehydrogenase. Appl Environ Microbiol 67:2004–2010.
31.
Miller EN, Jarboe LR, Yomano LP, York SW, Shanmugam KT, Ingram LO. 2009. Silencing of NADPH-dependent oxidoreductase genes (yqhD and dkgA) in furfural-resistant ethanologenic Escherichia coli. Appl Environ Microbiol 75:4315–4323.
32.
Brown NL, Stoyanov JV, Kidd SP, Hobman JL. 2003. The MerR family of transcriptional regulators. FEMS Microbiol Rev 27:145–163.
33.
Loveridge EJ, Allemann RK. 2011. Effect of pH on hydride transfer by Escherichia coli dihydrofolate reductase. Chembiochem 12:1258–1262.
34.
Kabsch W. 2010. XDS. Acta Crystallogr D Biol Crystallogr 66:125–132.
35.
Winter G. 2010. xia2: an expert system for macromolecular crystallography data reduction. J Appl Crystallogr 43:186–190.
36.
Winn MD, Ballard CC, Cowtan KD, Dodson EJ, Emsley P, Evans PR, Keegan RM, Krissinel EB, Leslie AGW, McCoy A, McNicholas SJ, Murshudov GN, Pannu NS, Potterton EA, Powell HR, Read RJ, Vagin A, Wilson KS. 2011. Overview of the CCP 4 suite and current developments. Acta Crystallogr D Biol Crystallogr 67:235–242.
37.
McCoy AJ, Grosse-Kunstleve RW, Adams PD, Winn MD, Storoni LC, Read RJ. 2007. Phaser crystallographic software. J Appl Crystallogr 40:658–674.
38.
Emsley P, Cowtan K. 2004. Coot: model-building tools for molecular graphics. Acta Crystallogr D Biol Crystallogr 60:2126–2132.
39.
Murshudov GN, Vagin AA, Dodson EJ. 1997. Refinement of macromolecular structures by the maximum-likelihood method. Acta Crystallogr D Biol Crystallogr 53:240–255.
40.
Pettersen EF, Goddard TD, Huang CC, Couch GS, Greenblatt DM, Meng EC, Ferrin TE. 2004. UCSF Chimera—a visualization system for exploratory research and analysis. J Comput Chem 25:1605–1612.
41.
Krieger E, Vriend G. 2014. YASARA View—molecular graphics for all devices—from smartphones to workstations. Bioinformatics 30:2981–2982.
42.
Laskowski RA, Swindells MB. 2011. LigPlot+: multiple ligand-protein interaction diagrams for drug discovery. J Chem Inf Model 51:2778–2786.
43.
Kim S, Thiessen PA, Bolton EE, Chen J, Fu G, Gindulyte A, Han L, He J, He S, Shoemaker BA, Wang J, Yu B, Zhang J, Bryant SH. 2016. PubChem substance and compound databases. Nucleic Acids Res 44:D1202–D1203.

Information & Contributors

Information

Published In

cover image Applied and Environmental Microbiology
Applied and Environmental Microbiology
Volume 85Number 151 August 2019
eLocator: e00978-19
Editor: Isaac Cann, University of Illinois at Urbana-Champaign
PubMed: 31101612

History

Received: 30 April 2019
Accepted: 7 May 2019
Published online: 18 July 2019

Permissions

Request permissions for this article.

Keywords

  1. aldehyde reductase
  2. biofuel
  3. detoxification
  4. furfural tolerance
  5. heavy enzyme
  6. dynamic coupling
  7. isotope effect
  8. lignocellulose

Contributors

Authors

School of Chemistry, Cardiff University, Cardiff, United Kingdom
Joel Cresser-Brown
School of Chemistry, Cardiff University, Cardiff, United Kingdom
Thomas L. Williams
School of Chemistry, Cardiff University, Cardiff, United Kingdom
Pierre J. Rizkallah
Institute of Infection & Immunology, School of Medicine, Cardiff University, Cardiff, United Kingdom
Yi Jin
School of Chemistry, Cardiff University, Cardiff, United Kingdom
Louis Y.-P. Luk
School of Chemistry, Cardiff University, Cardiff, United Kingdom
Rudolf K. Allemann
School of Chemistry, Cardiff University, Cardiff, United Kingdom

Editor

Isaac Cann
Editor
University of Illinois at Urbana-Champaign

Notes

Address correspondence to Rudolf K. Allemann, [email protected].

Metrics & Citations

Metrics

Note:

  • For recently published articles, the TOTAL download count will appear as zero until a new month starts.
  • There is a 3- to 4-day delay in article usage, so article usage will not appear immediately after publication.
  • Citation counts come from the Crossref Cited by service.

Citations

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. For an editable text file, please select Medlars format which will download as a .txt file. Simply select your manager software from the list below and click Download.

View Options

Figures

Tables

Media

Share

Share

Share the article link

Share with email

Email a colleague

Share on social media

American Society for Microbiology ("ASM") is committed to maintaining your confidence and trust with respect to the information we collect from you on websites owned and operated by ASM ("ASM Web Sites") and other sources. This Privacy Policy sets forth the information we collect about you, how we use this information and the choices you have about how we use such information.
FIND OUT MORE about the privacy policy