crystal structure and biophysical analysis of furfural ...3-pyrridine carboxaldehyde 1,519.52 61.48...

11
Crystal Structure and Biophysical Analysis of Furfural- Detoxifying Aldehyde Reductase from Clostridium beijerinckii Alan F. Scott, a Joel Cresser-Brown, a Thomas L. Williams, a Pierre J. Rizkallah, b Yi Jin, a Louis Y.-P. Luk, a Rudolf K. Allemann a a School of Chemistry, Cardiff University, Cardiff, United Kingdom b Institute of Infection & Immunology, School of Medicine, Cardiff University, Cardiff, United Kingdom ABSTRACT Many aldehydes, such as furfural, are present in high quantities in lignocellulose lysates and are fermentation inhibitors, which makes biofuel produc- tion from this abundant carbon source extremely challenging. Cbei_3974 has re- cently been identified as an aldo-keto reductase responsible for partial furfural resis- tance 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 identi- fied residues involved in substrate binding, and an activity screen revealed the sub- strate tolerance of the enzyme. Hydride transfer, which is partially rate limiting un- der 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 sug- gesting 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 fermen- tation found in lignocellulose lysates. The data contained here will serve as a guide for protein engineers to develop improved enzyme variants that would impart fur- fural resistance to the microorganisms used in biofuel production and thus lead to enhanced biofuel yields from this sustainable resource. KEYWORDS aldehyde reductase, biofuel, detoxification, furfural tolerance, heavy enzyme, dynamic coupling, isotope effect, lignocellulose A n 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 (4–6). 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 en- zymes 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 Citation Scott AF, Cresser-Brown J, Williams TL, Rizkallah PJ, Jin Y, Luk LY-P, Allemann RK. 2019. Crystal structure and biophysical analysis of furfural-detoxifying aldehyde reductase from Clostridium beijerinckii. Appl Environ Microbiol 85:e00978-19. https://doi.org/10.1128/AEM .00978-19. Editor Isaac Cann, University of Illinois at Urbana-Champaign Copyright © 2019 American Society for Microbiology. All Rights Reserved. Address correspondence to Rudolf K. Allemann, [email protected]. Received 30 April 2019 Accepted 7 May 2019 Accepted manuscript posted online 17 May 2019 Published ENZYMOLOGY AND PROTEIN ENGINEERING crossm August 2019 Volume 85 Issue 15 e00978-19 aem.asm.org 1 Applied and Environmental Microbiology 18 July 2019 on May 10, 2021 by guest http://aem.asm.org/ Downloaded from

Upload: others

Post on 15-Dec-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Crystal Structure and Biophysical Analysis of Furfural ...3-Pyrridine carboxaldehyde 1,519.52 61.48 4-Nitrobenzaldehyde 4,390.44 61.61 Isatin ND Methylglyoxal 2,852.59 335.72 Formaldehyde

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

Alan F. Scott,a Joel Cresser-Brown,a Thomas L. Williams,a Pierre J. Rizkallah,b Yi Jin,a Louis Y.-P. Luk,a Rudolf K. Allemanna

aSchool of Chemistry, Cardiff University, Cardiff, United KingdombInstitute of Infection & Immunology, School of Medicine, Cardiff University, Cardiff, United Kingdom

ABSTRACT Many aldehydes, such as furfural, are present in high quantities inlignocellulose lysates and are fermentation inhibitors, which makes biofuel produc-tion from this abundant carbon source extremely challenging. Cbei_3974 has re-cently been identified as an aldo-keto reductase responsible for partial furfural resis-tance in Clostridium beijerinckii. Rational engineering of this enzyme could enhancethe furfural tolerance of this organism, thereby improving biofuel yields. We reportan extensive characterization of Cbei_3974 and a single-crystal X-ray structure ofCbei_3974 in complex with NADPH at a resolution of 1.75 Å. Docking studies identi-fied residues involved in substrate binding, and an activity screen revealed the sub-strate tolerance of the enzyme. Hydride transfer, which is partially rate limiting un-der physiological conditions, occurs from the pro-R hydrogen of NADPH. Enzymeisotope labeling revealed a temperature-independent enzyme isotope effect of unity,indicating that the enzyme does not use dynamic coupling for catalysis and sug-gesting that the active site of the enzyme is optimally configured for catalysis withthe substrate tested.

IMPORTANCE Here we report the crystal structure and biophysical properties of analdehyde reductase that can detoxify furfural, a common inhibitor of biofuel fermen-tation found in lignocellulose lysates. The data contained here will serve as a guidefor protein engineers to develop improved enzyme variants that would impart fur-fural resistance to the microorganisms used in biofuel production and thus lead toenhanced biofuel yields from this sustainable resource.

KEYWORDS aldehyde reductase, biofuel, detoxification, furfural tolerance, heavyenzyme, dynamic coupling, isotope effect, lignocellulose

An ideal source of carbohydrates for biofuel fermentation is lignocellulose, anabundant waste product which is available at low cost and does not affect food

security (1). Fermentable sugars are most commonly released from lignocellulose usingan acid pretreatment (2). One of the major drawbacks of this method is the release ofaldehydes, organic acids, and phenols, which severely inhibit growth and limit the finalyield of biofuel (3). While it has been observed that Clostridium spp. have a highertolerance against the aldehyde inhibitors furfural and hydroxymethylfurfural than otherorganisms, the high levels of inhibitors found in lignocellulose lysates are neverthelesshugely problematic (4–6).

A recent investigation has identified two genes from Clostridium beijerinckii thatencode enzymes that reduce aldehydes to less-toxic alcohols (7, 8). These enzymesshowed activity against furfural, hydroxymethyl furfural, and benzaldehyde, which areall common fermentation inhibitors (7). Furthermore, the genes encoding these en-zymes are upregulated during furfural stress, suggesting the physiological relevance ofthese enzymes to protect C. beijerinckii (8). One of these enzymes, Cbei_3904, belongsto the short-chain dehydrogenase (SDR) family and the other, Cbei_3974, to the

Citation Scott AF, Cresser-Brown J, Williams TL,Rizkallah PJ, Jin Y, Luk LY-P, Allemann RK. 2019.Crystal structure and biophysical analysis offurfural-detoxifying aldehyde reductase fromClostridium beijerinckii. Appl Environ Microbiol85:e00978-19. https://doi.org/10.1128/AEM.00978-19.

Editor Isaac Cann, University of Illinois atUrbana-Champaign

Copyright © 2019 American Society forMicrobiology. All Rights Reserved.

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

Received 30 April 2019Accepted 7 May 2019

Accepted manuscript posted online 17 May2019Published

ENZYMOLOGY AND PROTEIN ENGINEERING

crossm

August 2019 Volume 85 Issue 15 e00978-19 aem.asm.org 1Applied and Environmental Microbiology

18 July 2019

on May 10, 2021 by guest

http://aem.asm

.org/D

ownloaded from

Page 2: Crystal Structure and Biophysical Analysis of Furfural ...3-Pyrridine carboxaldehyde 1,519.52 61.48 4-Nitrobenzaldehyde 4,390.44 61.61 Isatin ND Methylglyoxal 2,852.59 335.72 Formaldehyde

aldo-keto reductase (AKR) family (7). It is highly desirable to engineer greater catalyticefficiency into these enzymes to more rapidly eliminate toxic aldehydes, therebyenhancing resistance to aldehyde inhibition.

A prerequisite for rational engineering of an enzyme is a thorough understanding ofits mechanism. In this study, the furfural-transforming AKR Cbei_3974 was character-ized to provide valuable information for protein engineers. The substrate specificity,steady-state kinetic parameters, and crystal structure of Cbei_3974 have been deter-mined. In addition, the rate-limiting step of the reaction was identified and thecoupling of dynamic motions to the active site explored.

RESULTSSubstrate profile. It has previously been suggested that Cbei_3974 may be useful

to alleviate the toxicity of furfural during the fermentation of acid-treated lignocelluloselysates (7). NADPH-dependent activity toward furfural was previously reported for thisenzyme, but no characterization of the reaction product was shown (7). To confirmthat the enzyme indeed generates the less-toxic alcohol from the aldehyde, thereaction product from an enzyme-substrate-NADPH incubation was analyzed by gaschromatography-mass spectrometry (GC-MS) in parallel with controls containing noNADPH or no enzyme. After 5 h of incubation, a new compound could be detected onthe GC trace (Fig. 1). The retention time and fragmentation pattern were identical tothose of a commercial standard of furfuryl alcohol. This compound was not detected ineither of the controls, showing that its formation was enzyme catalyzed and NADPHdependent.

Cbei_3974 has previously been shown to also exhibit activity with hydroxymethylfurfural, benzaldehyde, and butyraldehyde (7). To more fully explore the substratescope of the enzyme, a selection of aldehydes, ketones, and alcohols was chosen. Theseputative substrates were incubated at 2 mM with the enzyme and an excess of NADPH,

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 withoutenzyme 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 panelD but doped with furfuryl alcohol; (F) standard of furfuryl alcohol; (G) mass spectrum of reaction product from incubation of furfural, NADPH, and enzyme after5 h; (H) mass spectrum of furfuryl alcohol standard. The full-length GC traces can be found in Fig. S1 in the supplemental material.

Scott et al. Applied and Environmental Microbiology

August 2019 Volume 85 Issue 15 e00978-19 aem.asm.org 2

on May 10, 2021 by guest

http://aem.asm

.org/D

ownloaded from

Page 3: Crystal Structure and Biophysical Analysis of Furfural ...3-Pyrridine carboxaldehyde 1,519.52 61.48 4-Nitrobenzaldehyde 4,390.44 61.61 Isatin ND Methylglyoxal 2,852.59 335.72 Formaldehyde

NADH, or NADP�. The change in cofactor concentration was measured continuously byUV spectroscopy to give the reaction rate (Table 1). Surprisingly, L-glyceraldehyde-3-phosphate gave only negligible activity despite the enzyme sharing 57.7% identity withEscherichia 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 theformation of dihydroxyacetone phosphate under gluconeogenic conditions, when TIMis genetically inactivated (9). Purified YghZ was shown to reduce L-glyceraldehyde-3-phosphate to L-glycerol-3-phosphate, which can be converted to dihydroxyacetone byL-glycerol-3-phosphate dehydrogenase, thus complementing TIM deficiency. Ratherunexpectedly, YghZ is stereospecific for the L-enantiomer of the substrate, whereas theTIM substrate is D-glyceraldehyde-3-phosphate. It was therefore proposed that a spon-taneous reaction may interconvert the two enantiomers (9). On this basis, YghZ andenzymes with similar sequences, including Cbei_3974, are annotated in the KEGGdatabase (https://www.kegg.jp/) as L-glyceraldehyde-3-phosphate reductases (10, 11).Our results demonstrate that this annotation is incorrect for Cbei_3974.

Only minimal activity was found for the previously identified substrates, furfural andbutyraldehyde (7), while the enzyme had no measurable activity with benzaldehyde atthe concentration tested (2 mM). In contrast, 4-pyridinecarboxaldehyde, which differsfrom benzaldehyde only by the presence of a nitrogen atom in the aromatic ring, wasthe most kinetically efficient among all of the substrates examined. Turnover was 17times 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 polarmethylglyoxal, a dialdehyde of the same chain length, gave strong activity 6.8 timesfaster than furfural. No activity could be detected with ketones or alcohols at theconcentrations tested. The enzyme was specific for NADPH, with no activity detectablewith 0.4 mM NADH. To determine whether the differences between substrates werecaused by differences in kcat or Km, steady-state kinetics were measured for the 5 fastestsubstrates, not including 4-nitrobenzaldehyde, which was not soluble enough toachieve saturation. All the aldehyde substrates resulted in Michaelis constants in the

TABLE 1 Substrate screen of activity of Cbei_3974 with putative substrates (2 mM) and0.4 mM NADPH

Substrate Sp act (�mol/min/mg)a

Furfural 416.86 � 4.81Benzaldehyde ND4-Pyrridine carboxaldehyde 4,652.70 � 671.053-Pyrridine carboxaldehyde 1,519.52 � 61.484-Nitrobenzaldehyde 4,390.44 � 61.61Isatin NDMethylglyoxal 2,852.59 � 335.72Formaldehyde NDAcetaldehyde NDPropionaldehyde NDButyraldehyde 307.37 � 23.65Valeraldehyde NDFurfural alcohol NDBenzyl alcohol ND4-Pyridine methanol NDMethanol NDEthanol NDIsopropyl alcohol NDButanol ND2,4-Pentanedione ND4-Acetylpyridine ND2-Butanone NDL-Arabinose NDD-Glucose ND�-Lactose NDL-Glyceraldehyde-3-phosphate 269.71 � 34.05aErrors indicate standard deviation from three repeats. ND, no detectable activity.

Structure and Function of a Furfural Reductase Applied and Environmental Microbiology

August 2019 Volume 85 Issue 15 e00978-19 aem.asm.org 3

on May 10, 2021 by guest

http://aem.asm

.org/D

ownloaded from

Page 4: Crystal Structure and Biophysical Analysis of Furfural ...3-Pyrridine carboxaldehyde 1,519.52 61.48 4-Nitrobenzaldehyde 4,390.44 61.61 Isatin ND Methylglyoxal 2,852.59 335.72 Formaldehyde

millimolar range, but NADPH had higher affinity as indicated by a lower Km of 32 �M(Table 2). The best substrate was 4-pyridine carboxaldehyde, with a kcat of 10 s�1 anda Km of 3.87 mM. This was closely followed by methylglyoxal, with a similar kcat of8.52 s�1 but a higher Km of 12.9 mM. Furfural has previously been shown to have anextraordinarily high kcat of 1.4 � 105 s�1 at 40°C (7). In our experiments the kcat wasmeasured at 19°C and was 2.72 s�1. This was lower than expected, even considering thelower temperature, but is more realistic. The Km of 34.9 mM measured here is inagreement with the literature value (7).

Stereochemistry of NADPH transfer. The hydride transfer in aldehyde reductasesoccurs from either the pro-R or pro-S hydrogen on C-4 of the nicotinamide ring ofNADPH. Typically, short-chain reductases (SDRs) transfer the pro-S hydrogen, whilealdo-keto reductases (AKRs) transfer the pro-R hydrogen (12). To determine the ste-reospecificity of Cbei_3974, the enzyme was incubated with NADPH or (4R)-[4-2H]NADPD with an excess of the substrate 4-pyridinecarboxaldehyde. The reactionproducts were analyzed to determine whether the deuterium had been incorporatedinto the alcohol product or remained on the nicotinamide cofactor. The alcoholreaction product was extracted with chloroform and analyzed by GC-MS. The incuba-tion with (4R)-[4-2H]NADPD gave a product which was 1 atomic mass unit larger thanthat from the incubation with NADPH, consistent with the incorporation of deuterium.In a duplicate reaction, the nucleotide product, NADP�, was purified by ion-exchangechromatography, freeze-dried, and dissolved in D2O. Nuclear magnetic resonance(NMR) analysis of NADP� from the reaction compared with commercial NADP� showedidentical spectra, confirming that the deuterium at the pro-R position had beentransferred from (4R)-[4-2H]NADPD (Fig. 2). Cbei_3974 therefore transfers the pro-Rhydrogen from NADPH, in accordance with other members of the AKR superfamily.

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

Heavy-enzyme KIE. The effect of protein dynamics on catalysis was investigated byheavy-enzyme production, where the nonexchangeable carbon and nitrogen atomswere replaced with their heavy counterparts (13C, 15N) to slow protein motions withoutaffecting 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 ratelimiting, steady-state measurements were used to determine any effect that slowerprotein motions in the heavy enzyme may have on the catalytic step.

Heavy enzyme (15N, 13C) was produced in M9 medium with labeled feedstocks andpurified to homogeneity (see Fig. S3 in the supplemental material). The incorporationof 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). Todetermine if the protein was correctly folded, the circular dichroism (CD) spectrum

TABLE 2 Kinetic measurements for Cbei_3974a

Substrate kcat (s�1) Km (mM) kcat/Km (s�1 M�1)

Furfural 2.72 � 0.27 34.9 � 7.6 78Butyraldehyde 5.16 � 1.04 40.79 � 4.5 1273-Pyridine carboxaldehyde 4.97 � 1.38 15.7 � 2.6 317Methylglyoxal 8.52 � 0.26 12.9 � 1.5 6604-Pyridine carboxaldehyde 10.60 � 0.2 3.87 � 0.34 2,739

NADPH with:4-Pyridine carboxaldehyde 10.0 � 1.02 0.032 � 0.02 312,500Furfural 2.59 � 0.145 0.015 � 0.003 172,666

aAll measurements were made at 19°C. Errors show the standard error from fitting the data to the Michaelis-Menten equation in GraphPad Prism.

Scott et al. Applied and Environmental Microbiology

August 2019 Volume 85 Issue 15 e00978-19 aem.asm.org 4

on May 10, 2021 by guest

http://aem.asm

.org/D

ownloaded from

Page 5: Crystal Structure and Biophysical Analysis of Furfural ...3-Pyrridine carboxaldehyde 1,519.52 61.48 4-Nitrobenzaldehyde 4,390.44 61.61 Isatin ND Methylglyoxal 2,852.59 335.72 Formaldehyde

and melting temperature were recorded and compared for both the “heavy” and“light” enzymes (see Fig. S5 in the supplemental material). The two enzymes gaveidentical spectra and had nearly identical melting temperatures of 62.4°C � 0.1 and63.8°C � 0.2, respectively, indicating that isotopic labeling does not significantly alterprotein 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 enzymeKIE of near unity, implying that there were no mass-dependent effects and thatdynamic coupling was minimal (Fig. 3). Although some authors have proposed thatenzymes use “promoting motions” to drive catalysis (14–17), this result is consistentwith a growing body of literature that shows that dynamic effects become signif-icant only outside physiological conditions and only when poorly tolerated substratesthat necessitate rearrangement of the active site are utilized (18–20). The enzyme doesnot, therefore, use dynamic motions as a part of its catalytic mechanism.

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 productobtained from incubations of Cbei_3974 with (4R)-[4-2H]NADPD and 4-pyridine carboxaldehyde (red)against a standard of NADP� (blue).

Structure and Function of a Furfural Reductase Applied and Environmental Microbiology

August 2019 Volume 85 Issue 15 e00978-19 aem.asm.org 5

on May 10, 2021 by guest

http://aem.asm

.org/D

ownloaded from

Page 6: Crystal Structure and Biophysical Analysis of Furfural ...3-Pyrridine carboxaldehyde 1,519.52 61.48 4-Nitrobenzaldehyde 4,390.44 61.61 Isatin ND Methylglyoxal 2,852.59 335.72 Formaldehyde

A recent study on the thermophilic Geobacillus stearothermophilus alcohol dehydro-genase (BsADH) showed that significant heavy-enzyme KIEs manifest only below itsphysiological temperature (40°C) and only with poor substrates (19). The temperaturedependency of heavy-enzyme KIEs has been suggested to be an indicator of whetheran enzyme is optimized for utilization of a particular substrate (19). The temperaturedependence of the KIE for Cbei_3974-catalyzed reduction of 3-pyridine carboxaldehydewas constant over the temperature range from 11 to 44°C (Fig. 4), suggesting that theactive-site architecture of the enzyme is optimized for this substrate.

Single-crystal X-ray structure. The protein was cocrystallized with NADPH andthe structure solved by molecular replacement using PDB entry 5T79, which is thecrystal structure for STM2406, an AKR from Salmonella enterica serovar Typhimurium ofunknown physiological function but with a substrate profile similar to that ofCbei_3974 (21). The two proteins have 60.91% sequence identity and a root meansquare 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 �-helices and �-strands forming an8-stranded TIM barrel with some extra helices (Fig. 5).

This motif is conserved across the AKR superfamily (22). Both this structure andSTM2406 have an unusual N terminus consisting of a long loop and a �-hairpin. MostAKR structures, including the structure of Coptotermes gestroi AKR1 (another AKRknown to reduce furfural), have a shorter N-terminal tail consisting of only the �-hairpinor, in the case of the AKR7 family, have no N-terminal tail (21, 23, 24). The function ofthis extra sequence is unclear. Conversely, the C terminus is truncated and is lacking aloop that is present in many AKRs, leaving the active site exposed to solvent (24). AKRswhich omit this loop have low catalytic efficiency consistent with the measured

FIG 3 Kinetic isotope effects. Gray circles, kcat values for natural-abundance enzyme with NADPH; bluecircles, enzyme KIEs; red circles, substrate KIEs (NADPH). All measurements were made at 19°C. Errors arestandard deviations from three repeats.

FIG 4 Temperature dependency of the enzyme kinetic isotope effect (kcat for light enzyme/kcat for heavyenzyme) with 3-pyridine carboxaldehyde as a substrate. Error bars show the standard deviation fromthree repeats.

Scott et al. Applied and Environmental Microbiology

August 2019 Volume 85 Issue 15 e00978-19 aem.asm.org 6

on May 10, 2021 by guest

http://aem.asm

.org/D

ownloaded from

Page 7: Crystal Structure and Biophysical Analysis of Furfural ...3-Pyrridine carboxaldehyde 1,519.52 61.48 4-Nitrobenzaldehyde 4,390.44 61.61 Isatin ND Methylglyoxal 2,852.59 335.72 Formaldehyde

millimolar Michaelis constants (21, 25, 26). Deletion of the C-terminal loop from humanaldose reductase AKR1B1 (26), Bacillus subtilis YhdN and YvgN, and Pseudomonasaeruginosa PA1127 (21) resulted in a dramatic loss of catalytic efficiency.

NADPH sits in a mostly open cleft with a hydrophobic center and polar residues atthe ends where the adenine base and nicotinamide ring bind. The adenine base is heldin place by hydrogen bonds to Glu-307 and Asn-308. The nucleotide 2=-phosphate thatdistinguishes NADPH from NADH is hydrogen bonded to Gln-304 and Ser-300. Thediphosphate 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 mobileloop that would strap the cofactor in place across the diphosphate bridge (23, 27). Thelack of density in Cbei_3974 indicates that the region is disordered and suggests thatthe loop is not trapping the cofactor.

The canonical mechanism of AKRs involves hydride transfer from NADPH to thecarbonyl acceptor (12). This is followed by protonation from an active-site tyrosine aspart of a proton relay from histidine and bulk water (12). Neighboring aspartate andlysine residues lower the pKa of tyrosine to enable it to function as an acid (12). In theclose homologue STM2406, the catalytic tetrad consists of Tyr-66, Asp61, Lys-97, andHis-138 (21). All these residues are conserved in Cbei_3974 (identical numbering). It wasnot possible to obtain crystal structures of protein-product complexes due to the lowaffinity of the ligands. Therefore, docking was used to predict the possible binding ofsubstrates. 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 basisof proximity to NADPH and the proposed catalytic residues. These poses are illustratedin Fig. 6. Both substrates are orientated toward the pro-R hydrogen of NADPH,consistent with the experimentally determined stereochemistry. The carbonyl oxygenof 4-pyridine carboxaldehyde is within hydrogen bonding distance of the exocyclicamide 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 beimportant for binding in that enzyme (21). Furfural docked into the active site in asimilar location but with a different orientation, possibly due to its smaller size. Theactive site has many polar residues, which may explain why the more hydrophobicaldehydes such as benzaldehyde are less favored. Asn-65 and Tyr-100 contribute tothe polar surface of the active site, and therefore alteration of these residues tomore hydrophobic ones may help improve activity for hydrophobic substrates. InSTM2406, which has a very similar active site, the variant Asn-65-Met (both enzymeshave the same residue numbering) gave a 341% increase in activity toward

FIG 5 Crystal structure of Cbei_3974. (A) Cartoon representation showing the TIM barrel fold. NADPH canbe seen in gold, with the nicotinamide ring in the central cavity. (B) Surface representation showing theexposed nature of the active site. NADPH is shown in gold. (C) Residues involved in binding NADPH. Redsemicircles identify hydrophobic interactions, and residues involved in hydrogen bonds are shown inblue. The green numbers indicate hydrogen bond distance. The figure was prepared with LigPlot� (42).

Structure and Function of a Furfural Reductase Applied and Environmental Microbiology

August 2019 Volume 85 Issue 15 e00978-19 aem.asm.org 7

on May 10, 2021 by guest

http://aem.asm

.org/D

ownloaded from

Page 8: Crystal Structure and Biophysical Analysis of Furfural ...3-Pyrridine carboxaldehyde 1,519.52 61.48 4-Nitrobenzaldehyde 4,390.44 61.61 Isatin ND Methylglyoxal 2,852.59 335.72 Formaldehyde

3-pyridinecarboxaldehyde compared with that of the wild type and a 2-fold decreasein Km (21). Alterations of Tyr-100 to aspartate, leucine, isoleucine, or valine mostlyresulted in insoluble proteins, while Tyr-100-Ala showed decreased activity, but thismay have been due to a loss of steric bulk by replacing a phenyl group with a hydrogenatom (21).

DISCUSSION

Cbei_3974, an enzyme that could putatively address the problem of aldehydetoxicity in biofuel fermentation from lignocellulose (7), has been extensively character-ized and its crystal structure solved. Though annotated as an L-glyceraldehyde-3-phosphate reductase, it shows only low activity toward this substrate. Instead, itpreferentially catalyzes reactions with 4-pyridine carboxaldehyde and methylglyoxal, ofwhich only the latter is likely to be naturally occurring inside the cell. Methylglyoxal isa toxic product formed from dihydroxy-acetone-phosphate by methylglyoxal synthaseto release phosphate (29). Detoxification of methylglyoxal by Clostridium results in theformation 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 regionof 12 mM. The low affinity may be part of a mechanism to conserve NADPH, asdepletion of NADPH can be just as lethal as aldehyde accumulation (31), but this isunlikely given that the 50% lethal dose (LD50) of methylglyoxal is likely to be signifi-cantly less than the Km, rendering the enzyme useless for detoxification. It is alsopossible that the enzyme requires an interaction partner or a posttranslational modi-fication for activity or that the experimental conditions were not optimal. An exami-nation of the C. beijerinckii genome in the KEGG database (10, 11) shows that the

FIG 6 Docking of 4-pyridine carboxaldehyde (A and B) and furfural (C and D) into the active site. Substrates are shown inorange. (A and C) Hydrophobic surface rendering; (B and D) identification of residues involved in binding (gray) andcatalysis (green).

Scott et al. Applied and Environmental Microbiology

August 2019 Volume 85 Issue 15 e00978-19 aem.asm.org 8

on May 10, 2021 by guest

http://aem.asm

.org/D

ownloaded from

Page 9: Crystal Structure and Biophysical Analysis of Furfural ...3-Pyrridine carboxaldehyde 1,519.52 61.48 4-Nitrobenzaldehyde 4,390.44 61.61 Isatin ND Methylglyoxal 2,852.59 335.72 Formaldehyde

cbei_3974 gene is not part of a biosynthetic gene cluster, but it is adjacent to a genefor a putative MerR transcription regulator. These typically respond to environmentalstimuli to upregulate stress response proteins (32). In conclusion, the natural substrateof 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 thepro-R hydrogen of NADPH and is partially rate limiting. Isotopically labeled heavyenzyme (13C, 15N) gave kcat constants identical to those obtained with the natural-abundance enzyme, indicating that the slower protein motions in the heavy enzymedid not impact the catalytic step. This shows that the enzyme does not use “promotingmotions” to drive the chemical transformation.

The enzyme kinetic isotope effect was independent of temperature, suggesting thatthe enzyme’s active site is optimally configured for the use of the tested substrates; thephysiological substrate is therefore likely to be structurally similar (19).

The crystal structure of Cbei_3974 revealed a typical AKR structure based around aTIM barrel fold and is essentially the same as that of STM2406 (21). Docking of4-pyridinecarboxaldehyde and furfural revealed residues that may be involved insubstrate binding. Although no enzyme variants were generated in this study, theseresidues represent targets for future work to generate an improved enzyme for moreefficient detoxification of furfural.

MATERIALS AND METHODSMaterials. 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]glucosewere obtained from Goss Scientific, Cheshire, UK. Furfural was obtained from VWR, methylglyoxal fromApollo Scientific, and 3-pyridine carboxaldehyde from Acros. All other chemicals were obtained fromSigma-Aldrich.

Crystallography. Cbei_3974 was overproduced in BL21(DE3) cells and purified by Ni affinity chro-matography as previously described (7). The protein was dialyzed against 10 mM HEPES-NaOH (pH7.5)–300 mM NaCl and concentrated to 10 mg/ml. Crystallization trials were performed using thesitting-drop vapor diffusion method by mixing 0.5 �l protein stock solution and 0.5 �l of a seed stockwith 0.5 �l reservoir solution. The seed stock was obtained from microcrystals grown in 100 mMmorpholinepropanesulfonic acid (MOPS) (pH 7.3)–13% polyethylene glycol (PEG) 8000 –750 mM NH4Clwith 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 tothe 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) onbeamline I04-1 and integrated with XDS (34) in the xia2 package (35). The data were scaled, reduced, andanalyzed with AIMLESS and TRUNCATE in CCP4i (36). The structure was solved by molecular replacementwith PHASER (37) using coordinates from PDB no. 5T79 as a searching model (21). The structure modelwas adjusted with COOT (38) and refined with REFMAC5 (39). Graphical representations were preparedin 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 inTable 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 wereconverted into PDBQT format using AudoDockTools1.5.6 and docked using AutoDock Vina (28). Graph-ical 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 �g/ml carbenicillin overnight at37°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 toan optical density at 600 nm (OD600) of 1.0 at 37°C and 220 rpm. Cultures were cooled to 12°C, and0.4 mM IPTG (isopropyl-�-D-thiogalactopyranoside) was added to induce gene expression. Cells wereharvested at 4,000 rpm after 16 h of growth at 12°C and 200 rpm. The protein was purified as previouslydescribed (7).

Substrate screen. Cbei_3974 (0.585 �M), 0.4 mM NADPH, NADH, or NADP�, and 2 mM putativesubstrate were mixed in 20 mM potassium phosphate buffer, pH 7.0. The subsequent change in NADPHconcentration was monitored at 340 nm (�340 � 6,220 M�1 cm�1) for 1 min, using a Shimadzu UV-2401PCspectrophotometer in 5-mm quartz cuvettes, to give the reaction rate.

Structure and Function of a Furfural Reductase Applied and Environmental Microbiology

August 2019 Volume 85 Issue 15 e00978-19 aem.asm.org 9

on May 10, 2021 by guest

http://aem.asm

.org/D

ownloaded from

Page 10: Crystal Structure and Biophysical Analysis of Furfural ...3-Pyrridine carboxaldehyde 1,519.52 61.48 4-Nitrobenzaldehyde 4,390.44 61.61 Isatin ND Methylglyoxal 2,852.59 335.72 Formaldehyde

Enzyme kinetics. Kinetic parameters were determined with 117 nM Cbei_3974 with one of the twosubstrates; one was held at a saturating level, while the other was varied across a concentration rangeof 0 to 75 mM (aldehydes) or 0 to 0.2 mM (NADPH). Rates were measured as described above and thedata fitted to the Michaelis-Menten equation using GraphPad Prism version 7.00 for Windows (GraphPadSoftware, 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-Siquadrupole time-of-flight mass spectrometer coupled to a Waters Acquity H-Class ultraperformanceliquid chromatography (UPLC) system. The column was an Acquity UPLC protein BEH C4 (300 Å, 1.7 �mby 2.1 mm by 100 mm) operated in reverse phase and held at 60°C. The gradient employed was 95% Ato 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 version4.1. Deconvolution of protein charged states was obtained using the maximum entropy 1 processingsoftware.

GC-MS analysis of reaction products. Mixtures containing 5 �M Cbei_3974, 4.6 mM NADPH or(4R)-[4-2H]NADPD, and 28 mM aldehyde in a final volume of 40 �l 20 mM potassium phosphate (pH 7.0)were incubated at 40°C for 1 h. Aliquots (4 �l) were removed at time zero and at 5 or 7 h and quenchedwith 1 ml of chloroform, which also served to extract the organic molecules. A 5-�l aliquot of the organiclayer 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°Cfor 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 furfurylalcohol was run on the GC-MS.

1H NMR of nicotinamide cofactors. Cbei_3974 (1 �M), 2 mM (4R)-[4-2H]NADPD, and 25 mM4-pyridine carboxaldehyde were incubated at 37°C for 3 h. NADP� was purified from the reaction on aSAX-10 column using published methodology (33), freeze-dried, and redissolved in D2O. Commercialstandards of NADPH and NADP� were also dissolved in D2O. NMR spectra of standards and reactionproduct were collected on a Bruker 400 instrument.

CD. Circular dichroism (CD) measurements were performed on an Applied Photophysics Chirascanspectrometer using 7 �M protein in 20 mM potassium phosphate buffer (pH 7.0)�20% glycerol. Spectrawere recorded over a temperature range of 5 to 85°C from 200 nm to 400 nm. The melting temperaturewas 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 DataBank (http://www.rcsb.org/pdb/) with accession code 6HG6.

SUPPLEMENTAL MATERIALSupplemental material for this article may be found at https://doi.org/10.1128/AEM

.00978-19.SUPPLEMENTAL FILE 1, PDF file, 0.4 MB.

ACKNOWLEDGMENTSWe 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.

REFERENCES1. Isikgor FH, Becer CR. 2015. Lignocellulosic biomass: a sustainable plat-

form for the production of bio-based chemicals and polymers. PolymChem 6:4497– 4559. https://doi.org/10.1039/C5PY00263J.

2. Kumar P, Barrett DM, Delwiche MJ, Stroeve P. 2009. Methods for pre-treatment of lignocellulosic biomass for efficient hydrolysis and biofuelproduction. Ind Eng Chem Res 48:3713–3729. https://doi.org/10.1021/ie801542g.

3. Liu ZL, Blaschek HP. 2010. Biomass conversion inhibitors and in situdetoxification, 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 agricul-tural residues: impact of degradation products on Clostridium beijer-inckii growth and butanol fermentation. Biotechnol Bioeng 97:1460 –1469. https://doi.org/10.1002/bit.21373.

5. Zhang Y, Han B, Ezeji TC. 2012. Biotransformation of furfural and5-hydroxymethyl furfural (HMF) by Clostridium acetobutylicum ATCC824 during butanol fermentation. Nat Biotechnol 29:345–351. https://doi.org/10.1016/j.nbt.2011.09.001.

6. Ujor V, Agu CV, Gopalan V, Ezeji TC. 2014. Glycerol supplementation of

the growth medium enhances in situ detoxification of furfural by Clos-tridium beijerinckii during butanol fermentation. Appl Microbiol Bio-technol 98:6511– 6521. https://doi.org/10.1007/s00253-014-5802-8.

7. Zhang Y, Ujor V, Wick M, Ezeji TC. 2015. Identification, purification andcharacterization of furfural transforming enzymes from Clostridium bei-jerinckii NCIMB 8052. Anaerobe 33:124 –131. https://doi.org/10.1016/j.anaerobe.2015.03.005.

8. Zhang Y, Ezeji TC. 2013. Transcriptional analysis of Clostridium beijer-inckii NCIMB 8052 to elucidate role of furfural stress during acetonebutanol ethanol fermentation. Biotechnol Biofuels 6:66. https://doi.org/10.1186/1754-6834-6-66.

9. Desai KK, Miller BG. 2008. A metabolic bypass of the triosephosphateisomerase reaction. Biochemistry 47:7983–7985. https://doi.org/10.1021/bi801054v.

10. Kanehisa M, Sato Y, Kawashima M, Furumichi M, Tanabe M. 2016. KEGGas a reference resource for gene and protein annotation. Nucleic AcidsRes 44:D457–D462. https://doi.org/10.1093/nar/gkv1070.

11. Kanehisa M, Furumichi M, Tanabe M, Sato Y, Morishima K. 2017. KEGG:new perspectives on genomes, pathways, diseases and drugs. NucleicAcids Res 45:D353–D361. https://doi.org/10.1093/nar/gkw1092.

Scott et al. Applied and Environmental Microbiology

August 2019 Volume 85 Issue 15 e00978-19 aem.asm.org 10

on May 10, 2021 by guest

http://aem.asm

.org/D

ownloaded from

Page 11: Crystal Structure and Biophysical Analysis of Furfural ...3-Pyrridine carboxaldehyde 1,519.52 61.48 4-Nitrobenzaldehyde 4,390.44 61.61 Isatin ND Methylglyoxal 2,852.59 335.72 Formaldehyde

12. Penning TM. 2015. The aldo-keto reductases (AKRs): overview. Chem BiolInteract 234:236 –246. https://doi.org/10.1016/j.cbi.2014.09.024.

13. Kholodar SA, Ghosh AK, Kohen A. 2017. Measurement of enzyme isotopeeffects. Methods Enzymol 596:43– 83. https://doi.org/10.1016/bs.mie.2017.06.033.

14. Silva RG, Murkin AS, Schramm VL. 2011. Femtosecond dynamicscoupled to chemical barrier crossing in a Born-Oppenheimer enzyme.Proc Natl Acad Sci U S A 108:18661–18665. https://doi.org/10.1073/pnas.1114900108.

15. Kipp DR, Silva RG, Schramm VL. 2011. Mass-dependent bond vibrationaldynamics influence catalysis by HIV-1 protease. J Am Chem Soc 133:19358 –19361. https://doi.org/10.1021/ja209391n.

16. Pudney CR, Guerriero A, Baxter NJ, Johannissen LO, Waltho JP, Hay S,Scrutton NS. 2013. Fast protein motions are coupled to enzymeH-transfer reactions. J Am Chem Soc 135:2512–2517. https://doi.org/10.1021/ja311277k.

17. Antoniou D, Ge X, Schramm VL, Schwartz SD. 2012. Mass modulation ofprotein dynamics associated with barrier crossing in purine nucleosidephosphorylase. J Phys Chem Lett 3:3538 –3544. https://doi.org/10.1021/jz301670s.

18. Luk LYP, Loveridge EJ, Allemann RK. 2014. Different dynamical effects inmesophilic and hyperthermophilic dihydrofolate reductases. J Am ChemSoc 136:6862. https://doi.org/10.1021/ja502673h.

19. Behiry E, Ruiz-Pernia JJ, Luk L, Tunon I, Moliner V, Allemann RK. 2018.Isotope substitution of promiscuous alcohol dehydrogenase revealsorigin of substrate preference in transition state. Angew Chem Int edEngl 57:3128 –3131. https://doi.org/10.1002/anie.201712826.

20. Ruiz-Pernía JJ, Behiry E, Luk LYP, Loveridge EJ, Tuñón I, Moliner V,Allemann RK. 2016. Minimization of dynamic effects in the evolutionof dihydrofolate reductase. Chem Sci 7:3248 –3255. https://doi.org/10.1039/c5sc04209g.

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 of3-hydroxybutanal to 1,3-butanediol: structural and biochemical stud-ies. Appl Environ Microbiol 83:e03172-16. https://doi.org/10.1128/AEM.03172-16.

22. Sanli G, Dudley JI, Blaber M. 2003. Structural biology of the aldo-ketoreductase family of enzymes: catalysis and cofactor binding. CellBiochem Biophys 38:79 –101. https://doi.org/10.1385/CBB:38:1:79.

23. Tramontina R, Franco Cairo JPL, Liberato MV, Mandelli F, Sousa A, SantosS, Rabelo SC, Campos B, Ienczak J, Ruller R, Damásio ARL, Squina FM.2017. The Coptotermes gestroi aldo-keto reductase: a multipurposeenzyme for biorefinery applications. Biotechnol Biofuels 10:4. https://doi.org/10.1186/s13068-016-0688-6.

24. Barski OA, Tipparaju SM, Bhatnagar A. 2008. The aldo-keto reductasesuperfamily and its role in drug metabolism and detoxification. DrugMetab Rev 40:553– 624. https://doi.org/10.1080/03602530802431439.

25. Ehrensberger AH, Wilson DK. 2004. Structural and catalytic diversity inthe two family 11 aldo-keto reductases. J Mol Biol 337:661– 673. https://doi.org/10.1016/j.jmb.2004.01.059.

26. Bohren KM, Grimshaw CE, Gabbay KH. 1992. Catalytic effectiveness ofhuman aldose reductase. Critical role of C-terminal domain. J Biol Chem267:20965–20970.

27. Wilson DK, Bohren KM, Gabbay KH, Quiocho FA. 1992. An unlikely sugarsubstrate site in the 1.65 A structure of the human aldose reductase

holoenzyme implicated in diabetic complications. Science 257:81– 84.https://doi.org/10.1126/science.1621098.

28. Trott O, Olson AJ. 2009. AutoDock Vina: improving the speed andaccuracy of docking with a new scoring function, efficient optimization,and multithreading. J Comput Chem 31:455– 461. https://doi.org/10.1002/jcc.21334.

29. Huang K, Rudolph FB, Bennett GN. 1999. Characterization of methylg-lyoxal synthase from Clostridium acetobutylicum ATCC 824 and its usein the formation of 1,2-propanediol. Appl Environ Microbiol 65:3244 –3247.

30. Liyanage H, Kashket S, Young M, Kashket ER. 2001. Clostridium beijer-inckii and Clostridium difficile detoxify methylglyoxal by a novel mech-anism involving glycerol dehydrogenase. Appl Environ Microbiol 67:2004 –2010. https://doi.org/10.1128/AEM.67.5.2004-2010.2001.

31. Miller EN, Jarboe LR, Yomano LP, York SW, Shanmugam KT, Ingram LO.2009. Silencing of NADPH-dependent oxidoreductase genes (yqhD anddkgA) in furfural-resistant ethanologenic Escherichia coli. Appl EnvironMicrobiol 75:4315– 4323. https://doi.org/10.1128/AEM.00567-09.

32. Brown NL, Stoyanov JV, Kidd SP, Hobman JL. 2003. The MerR family oftranscriptional regulators. FEMS Microbiol Rev 27:145–163. https://doi.org/10.1016/S0168-6445(03)00051-2.

33. Loveridge EJ, Allemann RK. 2011. Effect of pH on hydride transfer byEscherichia coli dihydrofolate reductase. Chembiochem 12:1258 –1262.https://doi.org/10.1002/cbic.201000794.

34. Kabsch W. 2010. XDS. Acta Crystallogr D Biol Crystallogr 66:125–132.https://doi.org/10.1107/S0907444909047337.

35. Winter G. 2010. xia2: an expert system for macromolecular crystallogra-phy data reduction. J Appl Crystallogr 43:186 –190. https://doi.org/10.1107/S0021889809045701.

36. Winn MD, Ballard CC, Cowtan KD, Dodson EJ, Emsley P, Evans PR, KeeganRM, 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 DBiol Crystallogr 67:235–242. https://doi.org/10.1107/S0907444910045749.

37. McCoy AJ, Grosse-Kunstleve RW, Adams PD, Winn MD, Storoni LC, ReadRJ. 2007. Phaser crystallographic software. J Appl Crystallogr 40:658 – 674. https://doi.org/10.1107/S0021889807021206.

38. Emsley P, Cowtan K. 2004. Coot: model-building tools for moleculargraphics. Acta Crystallogr D Biol Crystallogr 60:2126 –2132. https://doi.org/10.1107/S0907444904019158.

39. Murshudov GN, Vagin AA, Dodson EJ. 1997. Refinement of macromolecularstructures by the maximum-likelihood method. Acta Crystallogr D BiolCrystallogr 53:240–255. https://doi.org/10.1107/S0907444996012255.

40. Pettersen EF, Goddard TD, Huang CC, Couch GS, Greenblatt DM, MengEC, Ferrin TE. 2004. UCSF Chimera—a visualization system for explor-atory research and analysis. J Comput Chem 25:1605–1612. https://doi.org/10.1002/jcc.20084.

41. Krieger E, Vriend G. 2014. YASARA View—molecular graphics for all devic-es—from smartphones to workstations. Bioinformatics 30:2981–2982.https://doi.org/10.1093/bioinformatics/btu426.

42. Laskowski RA, Swindells MB. 2011. LigPlot�: multiple ligand-protein inter-action diagrams for drug discovery. J Chem Inf Model 51:2778–2786.https://doi.org/10.1021/ci200227u.

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 substanceand compound databases. Nucleic Acids Res 44:D1202–D1203. https://doi.org/10.1093/nar/gkv951.

Structure and Function of a Furfural Reductase Applied and Environmental Microbiology

August 2019 Volume 85 Issue 15 e00978-19 aem.asm.org 11

on May 10, 2021 by guest

http://aem.asm

.org/D

ownloaded from