You are on page 1of 7

Analytical Biochemistry 544 (2018) 80–86

Contents lists available at ScienceDirect

Analytical Biochemistry
journal homepage: www.elsevier.com/locate/yabio

A continuous assay for L-talarate/galactarate dehydratase using circular T


dichroism
Nicole M. Eastona,1, Sarah A.E. Aboushawareba,1, Stephen L. Bearnea,b,∗
a
Department of Biochemistry & Molecular Biology, Dalhousie University, Halifax, Nova Scotia, B3H 4R2, Canada
b
Department of Chemistry, Dalhousie University, Halifax, Nova Scotia, B3H 4R2, Canada

A R T I C L E I N F O A B S T R A C T

Keywords: L-Talarate/galactarate dehydratase (TGD) is a member of the enolase superfamily of enzymes and catalyzes the
Dehydratase dehydration of either meso-galactarate or L-talarate to form 5-keto-4-deoxy-D-glucarate (5-KDG). To facilitate
Assay study of this enzyme and other galactarate dehydratases, a continuous circular dichroism-based assay has been
Circular dichroism developed. Using recombinant enzyme from Salmonella typhimurium (StTGD), the rates of StTGD-catalyzed
Kinetics
conversion of m-galactarate to 5-KDG were determined by following the change in ellipticity at 323 nm. The
m-Galactarate
apparent molar ellipticity ([θ]323) for the 5-KDG formed was determined to be 202 ± 2 deg cm2 dmol−1, which
5-Keto-4-deoxy-D-glucarate
was used to convert observed rates (Δθ/Δt) into concentration-dependent rates (Δc/Δt). The kinetic parameters
Km, kcat, and kcat/Km were 0.38 ± 0.05 mM, 4.8 ± 0.1 s−1, and 1.3 ( ± 0.2) × 104 M−1s−1, respectively. These
values are in excellent agreement with those published previously [Yew, W.S. et al. (2007) Biochemistry 46,
9564–9577] using a coupled assay system. To demonstrate the utility of the assay, the inhibition constant
(Ki = 10.7 ± 0.4 mM) was determined for the competitive inhibitor tartronate. The continuous CD-based assay
offers a practical and efficient alternative method to the coupled assay that requires access to 5-KDG aldolase,
and to the labor-intensive, fixed-time assays.

Introduction aldonic and aldaric acid substrates. These features of the subgroup, and
indeed the superfamily, have made annotation of function difficult
dehydratase (EC 4.2.1.42, TGD1) is a member
L-Talarate/galactarate [8–10] and numerous detailed studies to understand the structure-
of the mandelate racemase (MR) subgroup of the enolase superfamily. function relationships among the members of the MR subgroup have
Members of the MR subgroup share with other members of the enolase focused on delineating the substrate specificity of the enzymes
superfamily a common bidomain structure (i.e., (β/α)7β-barrel domain [5–7,11–20]. In addition to delineating the biochemical roles of dehy-
and an α+β capping domain) and a common partial reaction (i.e., the dratases in carbohydrate metabolism, these enzymes have received at-
metal-assisted, Brønsted base-catalyzed abstraction of the α-proton tention in recent years because the dehydratase-catalyzed formation of
from a carboxylate substrate to form an enol(ate) intermediate) [1–4]. 2-ketoaldaric acids affords a route to deoxygenating carbohydrates in
TGD catalyzes the dehydration of either meso-galactarate or L-talarate the biomass for use as biofuels, and to producing precursors for the
to form 5-keto-4-deoxy-D-glucarate (5-KDG) as illustrated in Scheme 1 synthesis of value-added chemicals [21,22]. Such studies require con-
[5]. Despite the fact that the MR subgroup is named for its archetype venient assays for the enzymes, preferably ones that are direct and
MR, most enzymes in the subgroup, such as TGD, catalyze a dehydra- continuous. Several assays have been developed for galactarate dehy-
tion reaction of a sugar acid substrate rather than a racemization re- dratases, including a coupled assay that utilizes 5-keto-4-deoxy-D-glu-
action [6,7]. Although the catalytic machinery responsible for over- carate aldolase and L-lactate dehydrogenase as the coupling enzymes
coming the kinetic and thermodynamic barriers accompanying [5], and fixed-time assays that involve detection of the product through
deprotonation of a weak carbon acid substrate and the overall struc- derivatization with semicarbazide [12,16,23] or treatment with peri-
tural fold have been conserved within the MR subgroup, divergent odate and thiobarbituric acid [24]. In addition, the enzyme-catalyzed
evolution has led to a variety of dehydratases that act on specific hydrogen-deuterium exchange at the α-position of the substrate has

Abbreviations: CD, circular dichroism; (His)6, N-terminal hexahistidine tag; 5-KDG, 5-keto-4-deoxy-D-glucarate; MR, mandelate racemase; TGD, L-talarate/galactarate dehydratase;
StTGD, TGD from Salmonella typhimurium; Tris, tris(hydroxymethyl)aminomethane

Corresponding author. Department of Biochemistry & Molecular Biology, Dalhousie University, Halifax, Nova Scotia, B3H 4R2, Canada.
E-mail address: sbearne@dal.ca (S.L. Bearne).
1
Both authors contributed equally to the work.

https://doi.org/10.1016/j.ab.2017.12.015
Received 29 August 2017; Received in revised form 30 November 2017; Accepted 11 December 2017
Available online 14 December 2017
0003-2697/ © 2017 Elsevier Inc. All rights reserved.
N.M. Easton et al. Analytical Biochemistry 544 (2018) 80–86

Scheme 1. StTGD-catalyzed dehydration of m-galactarate and L-talarate to yield 5-KDG.

been used to follow the reaction by 1H NMR spectroscopy [5,12] and respectively. The amplification parameters were adapted from those
polarimetry has been used to follow the change in optical rotation reported by Gerlt and co-workers [5] as follows: initial denaturation at
[5,12,16]. Some of these assay methods have their own particular 98 °C for 30 s, 40 cycles of 94 °C for 60 s for denaturation, thermal
disadvantages. For example, the coupled assay is limited by the fact that gradient of 45–60 °C for 75 s for annealing, 68 °C for 120 s for extension,
the aldolase is not readily available, specific conditions must be met for and 72 °C for 144 s for final extension. The amplification product,
optimal functioning of the coupling enzymes, and the assay may not be purified using a QIAquick Gel Extraction Kit (Qiagen, Toronto, ON),
amenable to inhibition studies because of the potential for structural and the pET-15b vector were then digested separately in a stepwise
similarity of inhibitors to the substrate for the coupling enzyme. Neither manner with NdeI for 2 h followed by BamHI for 2 h at 37 °C, according
the fixed-time assays nor the 1H NMR-based assay permit rapid de- to the manufacturer's instructions. Subsequently, the digestion products
termination of initial rates, and the latter may have a solvent kinetic were purified as mentioned above and ligation with T4 DNA ligase
isotope effect when the reaction is conducted in D2O. However, re- (Invitrogen/Thermo Fisher, Waltham, MA) was conducted overnight at
cognition that m-galactarate has the meso configuration and is optically 16 °C following the manufacturer's instructions.
inactive, while the product 5-KDG is optically active (αD25 = +45°) The pET-15b-StTGD plasmid encodes a fusion protein bearing an N-
[5], led us to develop a circular dichroism (CD)-based assay for TGD terminal His6-tag. Chemically-competent E. coli DH5α cells were
activity. This method is related to the polarimetric assay and could be transformed with the plasmid using heat shock, and glycerol stocks
beneficial if a polarimeter is not available. Using the TGD from Sal- were prepared and stored at −80 °C using standard protocols [25]. The
monella typhimurium (StTGD) and m-galactarate as the substrate, we sequence of the insert was verified by commercial DNA sequencing
show that the CD-based assay offers a quick, practical, and effective (Robarts Research Institute, London, ON) to ensure incorporation of the
method for monitoring TGD activity. insert and the absence of any mutations. Competent E. coli BL21 (DE3)
cells were also transformed with the pET-15b-StTGD plasmid to be used
Materials and methods for protein expression, and glycerol stocks were prepared and stored at
−80 °C.
General
Enzyme purification
Tartronic acid was obtained from Alfa Aesar (Tweksbury, MA).
Mucic acid (m-galactaric acid) and all other chemicals were purchased StTGD was overexpressed and purified using a protocol similar to
from Sigma-Aldrich Canada Ltd. (Oakville, ON). Synthetic deoxy-oli- that described in the Novagen manual [26]. Four starter cultures, each
gonucleotide primers were purchased from Integrated DNA containing sterile Luria-Bertani (LB) medium (5 mL), ampicillin (25 μg/
Technologies (Coralville, IA). Restriction endonucleases were pur- mL), and the glycerol stock (10 μL), were incubated overnight at 37 °C
chased from New England Biolabs (Ipswich, MA). An S1000 Thermal with continuous shaking at 250 rpm. These starter cultures were then
Cycler (Bio-Rad Laboratories, Mississauga, ON) was used for poly- added to sterile LB medium (10 mL per 1 L) containing ampicillin
merase chain reactions (PCR). A Branson Sonifier 250 was used for (25 μg/mL), which was then incubated for 2 h at 37 °C followed by
sonication. His·Bind resin was purchased from Novagen (Madison, WI). 24 h at 27 °C (without induction by isopropyl β-D-1-thiogalactopyrano-
All other chemicals were reagent grade or better. Assays were con- side (IPTG) [5]) with continuous shaking at 250 rpm. The cells were
ducted using a Jasco J-810 spectropolarimeter (Jasco, Inc., Easton, harvested using centrifugation (3000 × g, 10 min, 4 °C). The cell pellet
MD). 1H NMR analyses were conducted using a Bruker AV-500 spec- was re-suspended in 30 mL of ice-cold binding buffer [Tris-HCl buffer
trometer at the Nuclear Magnetic Resonance Research Resource (NMR3, (20 mM, pH 7.9) containing NaCl (0.5 M) and imidazole (5 mM)]. Fol-
Dalhousie University). lowing sonication on ice (5 × 30 s bursts with cooling for 1 min be-
tween bursts), the cell lysate was then clarified by ultracentrifugation
Cloning (110,000 × g, 30 min, 4 °C), and then passed through a Ni2+-charged
His·bind resin-packed column (10-mL columns packed to 2.5 mL) at
The open reading frame encoding StTGD (GI: 16766982) was cloned 4 °C. The column was subsequently washed with binding buffer
from genomic DNA of Salmonella typhimurium LT2 using Phusion High- (25 mL), wash buffer [15 mL, Tris-HCl (20 mM, pH 7.9) containing NaCl
Fidelity DNA Polymerase (New England Biolabs, Ipswich, MA). The (0.5 M) and imidazole (60 mM)], and strip buffer [7 mL, Tris-HCl
PCR-based amplification reaction was conducted as reported by Gerlt (20 mM, pH 7.9) containing NaCl (0.5 M) and EDTA (100 mM)]. The
and co-workers [5] using the forward (5′-GTGATTATCAGGAGAAAAC majority of the protein eluted in the strip buffer and the purity was
ATATGGCTTTAAGCGCGAATTCCG-3′) and reverse (5′-GATTCCCGCCA assessed using 10% acrylamide SDS-PAGE followed by staining with
GGATCCTTAAGGGCGTTTGCCAAATTCAC-3′) primers indicated, where Coomassie blue R-250. StTGD was then dialyzed against assay buffer
the underlined bases correspond to NdeI and BamHI recognition sites, [Tris-HCl (50 mM, pH 8.0) containing MgCl2 (10 mM)] for 24 h at 4 °C

81
N.M. Easton et al. Analytical Biochemistry 544 (2018) 80–86

[5]. The enzyme preparation was subsequently diluted 2-fold by addi- tartronic acid (0.0, 8.0, 16.0, and 24.0 mM; pH adjusted to 8.0) in Tris-
tion of assay buffer and stored at −20 °C. Enzyme concentrations were HCl buffer (50 mM, pH 8.0) containing MgCl2 (10 mM). The reaction
determined using either the Bio-Rad protein assay (Bio-Rad Labora- was initiated by addition of StTGD (final concentration of 25 ng/μL).
tories, Mississauga, ON, Canada) with BSA standards, or by measuring Nonlinear regression analysis was used to fit equation (2) to initial rate
the absorbance at 280 nm using the extinction coefficient for the re- data to obtain the kinetic parameters followed by re-plotting of the
combinant protein of 59,930 M−1cm−1 calculated using the ExPASY apparent Km/Vmax values against the inhibitor concentration to esti-
ProtParam tool [27]. The N-terminal His6-tag was not removed. mate the value of Ki in accord with equation (2). Inhibition assays were
conducted in triplicate and the reported Ki value is the average and the
Determination of assay wavelength error is the standard deviation.
Vmax [S]
To determine the appropriate wavelength for following the con- vi =
version of m-galactarate to 5-KDG, CD spectra (wavelength range: (
Km 1 +
[I]
Ki ) + [S] (2)
250–400 nm) were obtained for the assay buffer, m-galactarate (6.0 mM
in assay buffer), StTGD (25 ng/μL in assay buffer), and then for a
mixture of TGD with m-galactarate for 5 min-intervals up to 60 min Results and discussion
while incubating the reaction at 25 °C. A quartz cuvette with a 0.5-cm
light-path was used and the final volume of the reaction was 1.0 mL. Direct continuous enzyme assays offer numerous advantages be-
cause they permit rapid determination of initial rates and do not limit
Molar ellipticity assay conditions to those that might be dictated by the other enzymes
employed in a coupled assay system. Consequently, we sought to de-
To convert the observed changes in ellipticity to changes in con- velop a direct continuous assay for TGD to facilitate studies on this
centration, the apparent molar ellipticity of 5-KDG was determined. m- member of the enolase superfamily, as well as on other galactarate
Galactarate (2.0–6.0 mM) was incubated at 25 °C with StTGD (50 ng/ dehydratases. The open reading frame encoding the enzyme from
μL) in Tris-HCl buffer (50 mM, pH 8.0) containing MgCl2 (10 mM) and Salmonella typhimurium was cloned into a pET-15b vector as reported
the reaction was followed at 323 nm in quartz cuvette with a 0.5-cm previously by Gerlt and co-workers [5]. The enzyme, bearing an N-
light-path until the reaction was complete. Product analysis was con- terminal His6-tag, was purified using metal ion affinity chromato-
ducted by incubating a reaction mixture (25-mL total volume) that graphy. We noted that the His6-tagged enzyme was quite prone to
contained m-galactarate (initial concentration of 6.0 mM), StTGD precipitation on the His·Bind resin, necessitating occasional removal of
(50 μg/mL), and MgCl2 (10.0 mM) in Tris-HCl buffer (50 mM, pH 8.0) at the precipitate during the addition of cell lysate to the column. Al-
37 °C until the reaction was complete as judged by the change in el- though a substantial amount of the desired protein was lost in the
lipticity at 323 nm. The protein was then removed by centrifugal fil- precipitate (Fig. 1, lane 2), the yield (∼7 mg/L culture) of pure protein
tration using an Amicon Ultra-15 filtration device with a 10-kDa MWCO (Fig. 1, lane 8) was satisfactory for subsequent enzymatic studies. The
that had been pre-washed with H2O and NaOH (0.1 M). The pH was precipitation problem was not ameliorated by the inclusion of MgCl2
then adjusted to ∼5 using HCl and the sample was applied to a column (5 mM) in the buffer [5], so magnesium was not included in the buffers
(1.4 × 15 cm) containing Dowex-50-X8 (Na+-form) to remove the Tris used during the purification, except for the final dialysis buffer. Finally,
buffer. Fractions containing the product were pooled and the solvent we note that the enzyme was only stable for about 2–3 weeks at −20 °C.
removed by lyophilization. The resulting powder was dissolved in H2O For a CD-based assay to be feasible, the substrate and product
containing 10% D2O and analyzed using 1H and 13C NMR spectroscopy should exhibit substantially different ellipticities at the assay wave-
and mass spectrometry. length so that the change in ellipticities can be observed during the
course of the enzyme-catalyzed reaction. Since m-galactarate is opti-
TGD assay cally inactive, we surmised that the optically active 5-KDG should af-
ford a CD signal that could be used to follow the conversion of m-ga-
Stock solutions of the substrate were prepared by first dissolving lactarate to 5-KDG. Although StTGD catalyzes a competing
both m-galactarate and Tris base in water, followed by dropwise addi- epimerization reaction between the two substrates m-galactarate and L-
tion of a solution of MgCl2, and then adjusting the pH to 8.0 using HCl. talarate along with the dehydration reaction, m-galactarate was chosen
The conversion of m-galactarate to 5-KDG was followed by measuring
the change in ellipticity at 323 nm for 3–5 min at 25 °C. To determine
the concentrations of StTGD that gave linear time courses, the assay was
conducted using m-galactarate concentrations of 0.1 mM and 6.0 mM
with varying enzyme concentrations (2.0–30.0 ng/μL). For routine as-
says, reaction mixtures contained StTGD (25 ng/μL), m-galactarate
(0.1–6.0 mM), Tris-HCl buffer (50 mM, pH 8.0), and MgCl2 (10 mM) in
a total volume of 1.0 mL. Nonlinear regression analysis was used to fit
equation (1) to initial rate data to obtain the kinetic parameters. Assays
were conducted in triplicate and the reported values are the averages
and the error is the standard deviation.
Vmax [S]
vi =
Km + [S] (1)

Fig. 1. SDS-PAGE gel (10%) showing the purification of StTGD. Lane 1, molecular weight
Inhibition studies marker; lane 2, precipitate; lane 3, crude cell lysate; lane 4, flow-through; lane 5, binding
buffer eluant; lane 6, wash buffer eluant; lane 7; strip buffer eluant; and lane 8, purified
Initial velocities were measured as described above. Reactions StTGD post-dialysis. The calculated MW of StTGD based on the amino acid sequence is
46,564 Da.
contained m-galactarate (0.25, 0.50, 1.0, 2.5, and 5.0 mM) and

82
N.M. Easton et al. Analytical Biochemistry 544 (2018) 80–86

Fig. 2. CD spectral changes during the StTGD-catalyzed conversion of m-galactarate to 5-


KDG at 25 °C. CD spectra were recorded between 250 and 400 nm for the assay reaction
mixture containing m-galactarate (6.0 mM) and StTGD (25 ng/μL) in Tris-HCl buffer
(50 mM, pH 8.0) containing MgCl2 (10 mM) at various time periods up to 60 min. The
final volume of the reaction was 1.0 mL and the light path was 0.5 cm. The spectra of the
buffer, m-galactarate (6.0 mM in buffer), and StTGD (25 ng/μL in buffer) did not differ
from the zero time spectrum shown in the figure.

as the substrate because Gerlt and co-workers had previously reported


that 30% of L-talarate undergoes epimerization to form m-galactarate,
whereas epimerization of m-galactarate to L-talarate is negligible [5]. In
addition, m-galactarate (mucic acid) is commercially available.

Molar ellipticity determination

To determine the appropriate wavelength for following the reaction,


CD spectra (wavelength range: 250–400 nm) for the assay buffer, m-
galactarate, StTGD, and the reaction mixture of TGD with m-galactarate
at various times were obtained (Fig. 2). As m-galactarate was con-
sumed, the growth of a signal in the CD spectrum with a maximum
ellipticity at 323 nm was observed. Since m-galactarate exhibited neg-
ligible ellipticity at this wavelength, 323 nm was chosen as the optimal
wavelength at which to follow the reaction. Analysis of the StTGD- Fig. 3. Determination of the apparent molar ellipticity for 5-KDG. Progress curves (A) are
catalyzed reaction product using 1H and 13C NMR spectroscopy de- shown for complete StTGD-catalyzed (50 ng/μL) conversion of various initial con-
monstrated complete conversion to 5-KDG with the resulting 1H and centrations of m-galactarate (2.0 mM, ◯; 3.0 mM, △; 4.0 mM, ▽; 5.0 mM, □; and
13
C NMR spectra matching that of the published spectrum of 5-KDG 6.0 mM, ◇) to 5-KDG. The final changes in ellipticity were plotted as a function of the
initial concentration of m-galactarate (B) and the molar ellipiticity ([θ]) determined from
(see Supporting Information Fig. S1) [28]. In addition, low- and high-
the slope of the line. The experiment was conducted in triplicate and a representative trial
resolution mass spectrometry also confirmed dehydration of the sub-
is shown. The average slope was 1.01 ± 0.01 mdeg/mM (l = 0.5 cm), which gave
strate (see Supporting Information Figs. S2 and S3). By following the [θ]323 = 202 ± 2 deg cm2 dmol−1.
complete conversion of various initial concentrations of m-galactarate
to 5-KDG (Fig. 3A), the apparent molar ellipticity of 5-KDG at 323 nm
Δc (Δθ/Δt )
([θ]323) was determined to be 202 ( ± 2) deg cm2 dmol−1, using vi {in mM/s} = = = 0.99(Δθ/Δt ){in mdeg/s}
Δt [θ] l (4)
equation (3), where [θ] is the molar ellipticity (deg cm2 dmol−1), θ is
the observed ellipticity (deg), c is the concentration (M), and l is the
light-path length (cm).
Assay development
θ= [θ] lc (3)
Linear progress curves were obtained over the first 5 min and the
The molar ellipticity could then be used to convert observed initial values of Δθ/Δt were used to calculate the initial velocities as outlined
rates (i.e., Δθ/Δt in units of mdeg/s) into concentration-dependent above (data not shown). The validity of the continuous assay was es-
terms using equation (4), where l = 0.5 cm. tablished by demonstrating that the initial velocities for conversion of

83
N.M. Easton et al. Analytical Biochemistry 544 (2018) 80–86

Table 1
Kinetic parameters for the StTGD-catalyzed conversion of m-galactarate to 5-KDG.

Source Km (mM) kcat (s−1) kcat/Km (M−1s−1)b

This studya 0.38 ± 0.05 4.8 ± 0.1 1.3 ( ± 0.2) × 104


Ref. [5] 0.33 ± 0.07 3.6 ± 0.2 1.1 × 104

a
Assays were conducted in triplicate with the concentration of StTGD equal to 25 ng/μL.
Errors are standard deviations.
b
The kcat/Km value and associated error was calculated from the values of kcat and Km
determined individually by fitting of equation (1) to the initial velocity data.

compromised by a requirement for substrate concentrations in the high


mM range due to their typically high Km values or by high concentra-
tions of inhibitors that exhibit weak binding affinities [29,30]. The
present assay offers a particular advantage in this regard since the
optimal assay wavelength is 323 nm, well above the absorbance of most
buffer components and small molecule ligands.

Kinetic parameters

Kinetic studies were conducted using the His6-tagged StTGD (Fig. 4)


and the kinetic parameters were determined by fitting equation (1) to
the initial velocity data (Table 1). The values of the kinetic parameters
are in excellent agreement with those obtained by Gerlt and co-workers
[5]. These authors used StTGD from which the N-terminal His6-tag had
been cleaved, while the present study was conducted with His6-tagged
enzyme, demonstrating that the tag has a negligible effect on the ob-
served kinetic parameters of StTGD.

Inhibition studies

To demonstrate the utility of this assay, we conducted an inhibition


study with tartronate. Previously, tartronate had been shown to be a
good competitive inhibitor of MR, binding with an inhibition constant
of 1.8 mM [31]. Interestingly, the X-ray crystal structure of the MR-
tartronate complex revealed that the glycolate moiety of tartronate
chelated the Mg2+ and that the distal carboxylate bridged the side
chains of the two active site Brønsted acid-base catalysts Lys 166 and
His 297. Since the active site architecture of MR is similar to that of
StTGD, which also contains Lys and His residues with similar orienta-
tions [5], we anticipated that tartronate should also inhibit StTGD.
Fig. 4. Effect of enzyme concentration on the rate of the StTGD-catalyzed conversion of Indeed, as the representative Lineweaver-Burk plot (Fig. 5) demon-
m-galactarate to 5-KDG (A) and Michaelis-Menten plot (B). Panel A: The concentrations of
strates, tartronate is a competitive inhibitor of StTGD with a Ki value of
m-galactarate were 0.1 mM (▾) and 6.0 mM (▴). Other conditions are as described in the
10.7 ± 0.4 mM. Interestingly, it is a much weaker inhibitor of StTGD
Materials and methods. The initial velocities show a linear dependence on enzyme con-
centration between 2.0 and 30 ng/μL. Therefore, the concentration of StTGD chosen (Ki/Km ≈ 28) than MR (Ki/Km ≈ 2) [31]. This difference may arise
within this range for standard kinetic analyses was 25 ng/μL. Panel B: Representative from subtle differences in either the precise location or the prontona-
Michaelis-Menten plot for the StTGD-catalyzed conversion of m-galactarate to 5-KDG. The tion states of the active site Brønsted acid-base catalysts that facilitate a
values of Km and Vmax are 0.38 ± 0.05 mM and 2.56 ± 0.05 μM/s, respectively, and the dehydration reaction in the case of StTGD versus a racemization reac-
corresponding kinetic constants are given in Table 1. The concentration of StTGD was tion in the case of MR.
25 ng/μL and the experimental conditions were as described in the Materials and
In summary, we have developed a convenient, direct continuous
methods.
CD-based assay for following the TGD-catalyzed conversion of m-ga-
lactarate to 5-KDG. The advantages that this assay method offers over
m-galactarate to 5-KDG varied linearly with StTGD concentrations the existing assays are that initial rate measurements may be obtained
(2.0–30.0 ng/μL) at both low (0.1 mM) and high (6.0 mM) substrate much more rapidly than is possible using fixed-time assays (e.g.,
concentrations (Fig. 4A). Consequently, 25 ng/μL was chosen as the semicarbazide derivatization or NMR methods) and it is much less
standard concentration of StTGD to be employed in the assay. complex than a coupled assay, which is often limited by the availability
One disadvantage of CD-based assays is that they are limited by the and specific properties of the coupling enzyme(s). Furthermore, the
total absorbance of the sample, especially when the assay wavelength utility of the assay has been demonstrated through determination of the
is < 250 nm since the increase in the total absorbance of the reaction inhibition constant for the competitive inhibitor tartronate. In addition
mixture arising from elevated concentrations of buffer components and to being extremely useful for studying the enzymology of StTGD, this
ligands can substantially decrease the signal-to-noise ratio. For ex- assay should also be generally applicable to the assay of other ga-
ample, CD-based assays of amino acid racemases are often lactarate dehydratases.

84
N.M. Easton et al. Analytical Biochemistry 544 (2018) 80–86

References

[1] P.C. Babbitt, M.S. Hasson, J.E. Wedekind, D.R. Palmer, W.C. Barrett, G.H. Reed,
I. Rayment, D. Ringe, G.L. Kenyon, J.A. Gerlt, The enolase superfamily: a general
strategy for enzyme-catalyzed abstraction of the α-protons of carboxylic acids,
Biochemistry 35 (1996) 16489–16501.
[2] P.C. Babbitt, J.A. Gerlt, Understanding enzyme superfamilies. Chemistry as the
fundamental determinant in the evolution of new catalytic activities, J. Biol. Chem.
272 (1997) 30591–30594.
[3] J.A. Gerlt, P.C. Babbitt, Divergent evolution of enzymatic function: mechanistically
diverse superfamilies and functionally distinct suprafamilies, Annu. Rev. Biochem.
70 (2001) 209–246.
[4] J.A. Gerlt, P.C. Babbitt, I. Rayment, Divergent evolution in the enolase superfamily:
the interplay of mechanism and specificity, Arch. Biochem. Biophys. 433 (2005)
59–70.
[5] W.S. Yew, A.A. Fedorov, E.V. Fedorov, S.C. Almo, J.A. Gerlt, Evolution of enzymatic
activities in the enolase superfamily: L-talarate/galactarate dehydratase from
Salmonella typhimurium LT2, Biochemistry 46 (2007) 9564–9577.
[6] W.S. Yew, A.A. Fedorov, E.V. Fedorov, J.F. Rakus, R.W. Pierce, S.C. Almo,
J.A. Gerlt, Evolution of enzymatic activities in the enolase superfamily: L-fuconate
dehydratase from Xanthomonas campestris, Biochemistry 45 (2006) 14582–14597.
[7] J.F. Rakus, A.A. Fedorov, E.V. Fedorov, M.E. Glasner, J.E. Vick, P.C. Babbitt,
S.C. Almo, J.A. Gerlt, Evolution of enzymatic activities in the enolase superfamily:
D-Mannonate dehydratase from Novosphingobium aromaticivorans, Biochemistry 46
(2007) 12896–12908.
[8] J.A. Gerlt, K.N. Allen, S.C. Almo, R.N. Armstrong, P.C. Babbitt, J.E. Cronan,
D. Dunaway-Mariano, H.J. Imker, M.P. Jacobson, W. Minor, C.D. Poulter,
F.M. Raushel, A. Sali, B.K. Shoichet, J.V. Sweedler, The enzyme function initiative,
Biochemistry 50 (2011) 9950–9962.
[9] J.A. Gerlt, P.C. Babbitt, M.P. Jacobson, S.C. Almo, Divergent evolution in the en-
olase superfamily: strategies for assigning functions, J. Biol. Chem. 287 (2012)
29–34.
[10] A.M. Schnoes, S.D. Brown, I. Dodevski, P.C. Babbitt, Bias in the experimental an-
notations of protein function and their effect on our understanding of protein
function space, PLoS Comput. Biol. 5 (2009) e1000605.
[11] S. Ghasempur, S. Eswaramoorthy, B.S. Hillerich, R.D. Seidel, S. Swaminathan,
S.C. Almo, J.A. Gerlt, Discovery of a novel L-lyxonate degradation pathway in
Pseudomonas aeruginosa PAO1, Biochemistry 53 (2014) 3357–3366.
[12] F.P. Groninger-Poe, J.T. Bouvier, M.W. Vetting, C. Kalyanaraman, R. Kumar,
S.C. Almo, M.P. Jacobson, J.A. Gerlt, Evolution of enzymatic activities in the en-
olase superfamily: galactarate dehydratase III from Agrobacterium tumefaciens C58,
Biochemistry 53 (2014) 4192–4203.
[13] A.M. Gulick, B.K. Hubbard, J.A. Gerlt, I. Rayment, Evolution of enzymatic activities
in the enolase superfamily: crystallographic and mutagenesis studies of the reaction
catalyzed by D-glucarate dehydratase from Escherichia coli, Biochemistry 39 (2000)
4590–4602.
[14] A.M. Gulick, B.K. Hubbard, J.A. Gerlt, I. Rayment, Evolution of enzymatic activities
in the enolase superfamily: identification of the general acid catalyst in the active
site of D-glucarate dehydratase from Escherichia coli, Biochemistry 40 (2001)
10054–10062.
[15] J.F. Rakus, A.A. Fedorov, E.V. Fedorov, M.E. Glasner, B.K. Hubbard, J.D. Delli,
P.C. Babbitt, S.C. Almo, J.A. Gerlt, Evolution of enzymatic activities in the enolase
superfamily: L-rhamnonate dehydratase, Biochemistry 47 (2008) 9944–9954.
[16] J.F. Rakus, C. Kalyanaraman, A.A. Fedorov, E.V. Fedorov, F.P. Mills-Groninger,
R. Toro, J. Bonanno, K. Bain, J.M. Sauder, S.K. Burley, S.C. Almo, M.P. Jacobson,
Fig. 5. Inhibition of StTGD by tartronate. Panel A: Representative Lineweaver-Burk plot J.A. Gerlt, Computation-facilitated assignment of the function in the enolase su-
demonstrating competitive inhibition of StTGD activity by tartronate. The concentrations perfamily: a regiochemically distinct galactarate dehydratase from Oceanobacillus
of m-galactarate ranged between 0.25 and 5.00 mM and the concentrations of tartronate iheyensis, Biochemistry 48 (2009) 11546–11558.
[17] D.J. Wichelecki, B.M. Balthazor, A.C. Chau, M.W. Vetting, A.A. Fedorov,
were 0.0 mM (●), 8.0 mM (▴), 16.0 mM (▾), and 24.0 mM (▪). Other conditions are as
E.V. Fedorov, T. Lukk, Y.V. Patskovsky, M.B. Stead, B.S. Hillerich, R.D. Seidel,
described in the Materials and methods. Panel B: Replot of the apparent Km/Vmax values S.C. Almo, J.A. Gerlt, Discovery of function in the enolase superfamily: D-manno-
obtained using nonlinear regression analysis as a function of tartronate concentration. nate and D-gluconate dehydratases in the D-mannonate dehydratase subgroup,
The Ki value is 10.7 ± 0.4 mM. Biochemistry 53 (2014) 2722–2731.
[18] D.J. Wichelecki, D.C. Graff, N. Al-Obaidi, S.C. Almo, J.A. Gerlt, Identification of the
in vivo function of the high-efficiency D-mannonate dehydratase in Caulobacter
Acknowledgment crescentus NA1000 from the enolase superfamily, Biochemistry 53 (2014)
4087–4089.
[19] D.J. Wichelecki, J.A. Vendiola, A.M. Jones, N. Al-Obaidi, S.C. Almo, J.A. Gerlt,
We thank the Natural Sciences and Engineering Research Council Investigating the physiological roles of low-efficiency D-mannonate and D-gluconate
(NSERC) of Canada for a Discovery Grant (S.L.B.; RGPIN-2016-05083) dehydratases in the enolase superfamily: pathways for the catabolism of L-gulonate
and an Undergraduate Summer Research Award (N.M.E.). N.M.E. and and L-idonate, Biochemistry 53 (2014) 5692–5699.
[20] W.S. Yew, A.A. Fedorov, E.V. Fedorov, B.M. Wood, S.C. Almo, J.A. Gerlt, Evolution
S.A.E.A. were recipients of a Faye Sobey Undergraduate Research
of enzymatic activities in the enolase superfamily: D-tartrate dehydratase from
Award and a Nova Scotia Graduate Scholarship, respectively. We also Bradyrhizobium japonicum, Biochemistry 45 (2006) 14598–14608.
thank Dr. John Rohde (Dalhousie University) for kindly providing [21] K. Matsubara, R. Köhling, B. Schönenberger, T. Kouril, D. Esser, C. Bräsen,
genomic DNA from Salmonella typhimurium, and Dr. Mike Lumsden B. Siebers, R. Wohlgemuth, One-step synthesis of 2-keto-3-deoxy-D-gluconate by
biocatalytic dehydration of D-gluconate, J. Biotechnol. 191 (2014) 69–77.
(NMR3, Dalhousie University) for his assistance with NMR experiments. [22] J.M. Carsten, A. Schmidt, V. Sieber, Characterization of recombinantly expressed
dihydroxy-acid dehydratase from Sulfobus solfataricus-A key enzyme for the con-
version of carbohydrates into chemicals, J. Biotechnol. 211 (2015) 31–41.
[23] B.K. Hubbard, M. Koch, D.R. Palmer, P.C. Babbitt, J.A. Gerlt, Evolution of enzy-
Appendix A. Supplementary data matic activities in the enolase superfamily: characterization of the (D)-glucarate/
galactarate catabolic pathway in Escherichia coli, Biochemistry 37 (1998)
Supplementary data related to this article can be found at http://dx. 14369–14375.
[24] H.J. Blumenthal, T. Jepson, Asymmetric dehydration of galactarate by bacterial
doi.org/10.1016/j.ab.2017.12.015.

85
N.M. Easton et al. Analytical Biochemistry 544 (2018) 80–86

galactarate dehydratase, Biochem. Biophys. Res. Commun. 17 (1964) 282–287. threo-hexarate, J. Biol. Chem. 287 (2012) 17662–17671.
[25] J. Sambrook, E.F. Fritsch, T. Maniatis, Molecular Cloning, Cold Spring Harbor [29] M. Harty, M. Nagar, L. Atkinson, C.M. LeGay, D.J. Derksen, S.L. Bearne, Inhibition
Laboratory Press, Plainview, New York, 1989, pp. 1.21–21.52. of serine and proline racemases by substrate-product analogues, Bioorg. Med.
[26] Novagen, pET System Manual, seventh ed., (1997), pp. 18–64. TB055. Chem. Lett 23 (2013) 390–393.
[27] E. Gasteiger, A. Gattiker, C. Hoogland, I. Ivanyi, R.D. Appel, A. Bairoch, ExPASy: the [30] M. Pal, S.L. Bearne, Inhibition of glutamate racemase by substrate-product analo-
proteomics server for in-depth protein knowledge and analysis, Nucleic Acids Res. gues, Bioorg. Med. Chem. Lett 24 (2014) 1432–1436.
31 (2003) 3784–3788. [31] M. Nagar, A.D. Lietzan, M. St. Maurice, S.L. Bearne, Potent inhibition of mandelate
[28] M. Andberg, H. Maaheimo, H. Boer, M. Penttilä, A. Koivula, P. Richard, racemase by a fluorinated substrate-product analogue with a novel binding mode,
Characterization of a novel Agrobacterium tumefaciens galactarolactone cycloi- Biochemistry 53 (2014) 1169–1178.
somerase enzyme for direct conversion of D-galactarolactone to 3-deoxy-2-keto-L-

86

You might also like