Indeed, tight feedback regulation of metabolic flux is particularly important for the shikimate pathway, because it metabolizes a significant portion of organic carbon for the biosynthesis of energetically costly aromatic compounds in bacteria, archaea, fungi, algae, plants, and protozoan parasites (
3Shikimic Acid: Metabolism and Metabolites.
,
4- Herrmann K.M.
- Weaver L.M.
The shikimate pathway.
,
5- Roberts F.
- Roberts C.W.
- Johnson J.J.
- Kyle D.E.
- Krell T.
- Coggins J.R.
- Coombs G.H.
- Milhous W.K.
- Tzipori S.
- Ferguson D.J.
- Chakrabarti D.
- McLeod R.
Evidence for the shikimate pathway in apicomplexan parasites.
). The pathway starts with the condensation of phosphoenolpyruvate and erythrose-4-phosphate to form 3-deoxy-
d-
arabino-heptulosonate 7-phosphate (DAHP) catalyzed by DAHP synthase (DS). Six subsequent enzymatic steps lead to the biosynthesis of the branch-point metabolite chorismate, the last common precursor of the aromatic amino acids and other essential aromatic compounds. The first committed step toward
l-phenylalanine and
l-tyrosine is the conversion of chorismate to prephenate (
Fig. 1A), catalyzed by chorismate mutase (CM). Because DS and CM are the central nodes of the shikimate pathway, organisms have developed a variety of strategies both at the genetic (
4- Herrmann K.M.
- Weaver L.M.
The shikimate pathway.
) and protein (
6The diversity of allosteric controls at the gateway to aromatic amino acid biosynthesis.
,
7- Fan Y.
- Cross P.J.
- Jameson G.B.
- Parker E.J.
Exploring modular allostery via interchangeable regulatory domains.
,
8- Sterritt O.W.
- Kessans S.A.
- Jameson G.B.
- Parker E.J.
A pseudoisostructural type II DAH7PS enzyme from Pseudomonas aeruginosa: alternative evolutionary strategies to control shikimate pathway flux.
,
9- Nazmi A.R.
- Lang E.J.M.
- Bai Y.
- Allison T.M.
- Othman M.H.
- Panjikar S.
- Arcus V.L.
- Parker E.J.
Interdomain conformational changes provide allosteric regulation en route to chorismate.
,
10- Bai Y.
- Lang E.J.M.
- Nazmi A.R.
- Parker E.J.
Domain cross-talk within a bifunctional enzyme provides catalytic and allosteric functionality in the biosynthesis of aromatic amino acids.
) level to regulate these enzymes. In
M. tuberculosis, CM (MtCM, a dimer encoded by
Rv0948c,
aroQδ) has evolved to transiently interact with DS (MtDS, a tetramer encoded by
Rv2178c,
aroG) to form a heterooctameric complex (
Fig. 1B). Only the complexed, but not the free dimeric MtCM, is responsive to feedback regulation by Phe and Tyr (
1- Munack S.
- Roderer K.
- Ökvist M.
- Kamarauskaite˙ J.
- Sasso S.
- van Eerde A.
- Kast P.
- Krengel U.
Remote control by inter-enzyme allostery: a novel paradigm for regulation of the shikimate pathway.
,
11- Sasso S.
- Ökvist M.
- Roderer K.
- Gamper M.
- Codoni G.
- Krengel U.
- Kast P.
Structure and function of a complex between chorismate mutase and DAHP synthase: efficiency boost for the junior partner.
).
MtCM represents the structurally simple AroQ
δ subclass of CMs composed of two intertwined three-helix subunits (
Fig. 1C) (
1- Munack S.
- Roderer K.
- Ökvist M.
- Kamarauskaite˙ J.
- Sasso S.
- van Eerde A.
- Kast P.
- Krengel U.
Remote control by inter-enzyme allostery: a novel paradigm for regulation of the shikimate pathway.
,
11- Sasso S.
- Ökvist M.
- Roderer K.
- Gamper M.
- Codoni G.
- Krengel U.
- Kast P.
Structure and function of a complex between chorismate mutase and DAHP synthase: efficiency boost for the junior partner.
). Other prototypical fold variants within the structurally and evolutionarily related α-helical AroQ class of CMs (
12Evolutionary cycles for pericyclic reactions - Or why we keep mutating mutases.
) comprise the CM domain of the bifunctional CM-prephenate dehydratase from
Escherichia coli (EcCM, subclass AroQ
α (
13- Lee A.Y.
- Karplus P.A.
- Ganem B.
- Clardy J.
Atomic structure of the buried catalytic pocket of Escherichia coli chorismate mutase.
)), the elaborate eukaryotic 12-helical CM from
Saccharomyces cerevisiae (ScCM, AroQ
β (
14- Sträter N.
- Schnappauf G.
- Braus G.
- Lipscomb W.N.
Mechanisms of catalysis and allosteric regulation of yeast chorismate mutase from crystal structures.
,
15- Xue Y.
- Lipscomb W.N.
- Graf R.
- Schnappauf G.
- Braus G.
The crystal structure of allosteric chorismate mutase at 2.2 Å resolution.
)), and the secreted enzyme from
M. tuberculosis (*MtCM, AroQ
γ (
16- Sasso S.
- Ramakrishnan C.
- Gamper M.
- Hilvert D.
- Kast P.
Characterization of the secreted chorismate mutase from the pathogen Mycobacterium tuberculosis.
)). In stark contrast to the α, β, and γ CM subclasses, MtCM utilizes an arginine residue (Arg
46) instead of an otherwise absolutely conserved lysine to promote the electrostatic catalysis (
17- Warshel A.
- Sharma P.K.
- Kato M.
- Xiang Y.
- Liu H.
- Olsson M.H.M.
Electrostatic basis for enzyme catalysis.
) of the Claisen rearrangement of chorismate (
Fig. 1,
D and
E) (
11- Sasso S.
- Ökvist M.
- Roderer K.
- Gamper M.
- Codoni G.
- Krengel U.
- Kast P.
Structure and function of a complex between chorismate mutase and DAHP synthase: efficiency boost for the junior partner.
). Furthermore, MtCM is shorter at the C terminus and lacks the active site residue homologous to Gln
88 in EcCM (
18- Liu D.R.
- Cload S.T.
- Pastor R.M.
- Schultz P.G.
Analysis of active site residues in Escherichia coli chorismate mutase by site-directed mutagenesis.
,
19- Zhang S.
- Kongsaeree P.
- Clardy J.
- Wilson D.B.
- Ganem B.
Site-directed mutagenesis of monofunctional chorismate mutase engineered from the E. coli P-protein.
), Glu
109 in *MtCM (
20- Ökvist M.
- Dey R.
- Sasso S.
- Grahn E.
- Kast P.
- Krengel U.
1.6 Å crystal structure of the secreted chorismate mutase from Mycobacterium tuberculosis: novel fold topology revealed.
), and Glu
246 in ScCM (
21- Schnappauf G.
- Sträter N.
- Lipscomb W.N.
- Braus G.H.
A glutamate residue in the catalytic center of the yeast chorismate mutase restricts enzyme activity to acidic conditions.
). Despite these dramatic deviations from the consensus active site, MtCM is capable of catalyzing the conversion of chorismate to prephenate with a high catalytic efficiency (
kcat/
Km = 2.4 × 10
5m−1 s
−1) (
11- Sasso S.
- Ökvist M.
- Roderer K.
- Gamper M.
- Codoni G.
- Krengel U.
- Kast P.
Structure and function of a complex between chorismate mutase and DAHP synthase: efficiency boost for the junior partner.
). However, this can only be achieved in the presence of MtDS. On its own, MtCM is a mediocre enzyme catalyzing the reaction by two orders of magnitude less efficiently (
kcat/
Km of 1.8 × 10
3m−1 s
−1) than the prototypical CMs of the other subclasses (
16- Sasso S.
- Ramakrishnan C.
- Gamper M.
- Hilvert D.
- Kast P.
Characterization of the secreted chorismate mutase from the pathogen Mycobacterium tuberculosis.
,
18- Liu D.R.
- Cload S.T.
- Pastor R.M.
- Schultz P.G.
Analysis of active site residues in Escherichia coli chorismate mutase by site-directed mutagenesis.
,
21- Schnappauf G.
- Sträter N.
- Lipscomb W.N.
- Braus G.H.
A glutamate residue in the catalytic center of the yeast chorismate mutase restricts enzyme activity to acidic conditions.
). In fact, the poor activity of the MtCM dimer is essential for effective shikimate pathway regulation, exerted through inter-enzyme allostery in
M. tuberculosis. We (
1- Munack S.
- Roderer K.
- Ökvist M.
- Kamarauskaite˙ J.
- Sasso S.
- van Eerde A.
- Kast P.
- Krengel U.
Remote control by inter-enzyme allostery: a novel paradigm for regulation of the shikimate pathway.
,
11- Sasso S.
- Ökvist M.
- Roderer K.
- Gamper M.
- Codoni G.
- Krengel U.
- Kast P.
Structure and function of a complex between chorismate mutase and DAHP synthase: efficiency boost for the junior partner.
) and others (
22- Blackmore N.J.
- Nazmi A.R.
- Hutton R.D.
- Webby M.N.
- Baker E.N.
- Jameson G.B.
- Parker E.J.
Complex formation between two biosynthetic enzymes modifies the allosteric regulatory properties of both: an example of molecular symbiosis.
,
23- Jiao W.
- Blackmore N.J.
- Nazmi A.R.
- Parker E.J.
Quaternary structure is an essential component that contributes to the sophisticated allosteric regulation mechanism in a key enzyme from Mycobacterium tuberculosis.
) have shown that binding of the allosteric feedback inhibitors Phe and Tyr to MtDS induces MtCM-MtDS complex dissociation and thereby a shift from high to low intracellular CM activity, providing tight control over cytoplasmic aromatic amino acid concentrations.
Discussion
In this work, we subjected the mediocre MtCM to four cycles of directed evolution to explore the catalytic potential of this natural enzyme (
Fig. 3). We have applied a powerful CM selection system that couples the enzyme's activity to bacterial viability under stringent selective conditions to efficiently identify mutations that improve the catalytic prowess of MtCM. However, the limitations of the original selection approach were already reached after cycle II, when the target enzyme had gained enough proficiency to fully complement the metabolic defect of the selection strain. To increase the selection stringency for the following rounds, we successfully implemented structural perturbation-compensation strategies. These consisted of temporarily crippling the catalytic activity of the evolving MtCM variants followed by augmenting it again through the introduction of new and different mutations. Removal of the deliberately installed lesions in the further-evolved (destabilized) variants resulted in most cases in better catalysts than the starting points.
In the course of the directed evolution experiments we identified MtCM variant s4.15, which is, with a kcat/Km of 4.7 × 105m−1 s−1, twice as efficient as the MtDS-activated WT enzyme. The crystal structure of N-s4.15 implies that its improved activity is due to a combination of pre-positioning active site residues for efficient substrate and transition state binding, tighter packing of the active site, and an overall stabilization of the fold, as reflected by the increase in the enzyme's melting temperature.
Of all the beneficial substitutions, the introduction of Pro
52 and Asp
55 had the biggest effect by reshaping the catalytically important H1-H2 loop. The induced register shift of residues 54 and 55 places the backbone of residue 55 in an optimal position to interact with the ligand's hydroxyl group, whereas the Asp
55 side chain may pre-align Arg
18 of the other protomer for better substrate binding. Interestingly, the change from
52TRLV
55 to
52PRLD
55 resembles the corresponding sequences in the AroQ
α subclass EcCM (
45PVRD
48), the AroQ
γ subclass *MtCM (
66PIED
69), and the AroQ
β subclass ScCM, which has Pro
174 and Asn
194 at the homologous positions (
Fig. 1E). The latter three CMs achieve high catalytic efficiencies without DS interactions, suggesting that having Pro
52 and Asp
55 in the H1-H2 loop is an important feature of an autonomously proficient catalytic machinery. In contrast, the fact that native MtCM and many other δ-subclass CMs (
11- Sasso S.
- Ökvist M.
- Roderer K.
- Gamper M.
- Codoni G.
- Krengel U.
- Kast P.
Structure and function of a complex between chorismate mutase and DAHP synthase: efficiency boost for the junior partner.
,
31- Roderer K.
- Neuenschwander M.
- Codoni G.
- Sasso S.
- Gamper M.
- Kast P.
Functional mapping of protein-protein interactions in an enzyme complex by directed evolution.
) use Thr
52 can be rationalized by structural arguments. Threonine (like the frequently selected serine;
Fig. 4A) can adopt a similar conformation as proline by H-bonding to the main chain nitrogen. In contrast to proline, this conformation is, however, not permanently fixed for threonine (or serine), allowing for greater conformational sampling. Such a temporary “kink-potential” is probably crucial for DS-dependent activity switching of AroQ
δ-subclass CMs (
1- Munack S.
- Roderer K.
- Ökvist M.
- Kamarauskaite˙ J.
- Sasso S.
- van Eerde A.
- Kast P.
- Krengel U.
Remote control by inter-enzyme allostery: a novel paradigm for regulation of the shikimate pathway.
,
2- Burschowsky D.
- Thorbjørnsrud H.V.
- Heim J.B.
- Fahrig-Kamarauskaitė J.
- Würth-Roderer K.
- Kast P.
- Krengel U.
Inter-enzyme allosteric regulation of chorismate mutase in Corynebacterium glutamicum: structural basis of feedback activation by Trp.
,
11- Sasso S.
- Ökvist M.
- Roderer K.
- Gamper M.
- Codoni G.
- Krengel U.
- Kast P.
Structure and function of a complex between chorismate mutase and DAHP synthase: efficiency boost for the junior partner.
).
In N-s4.15, the conformation of the H1-H2 loop enforced by Pro
52 and Asp
55 might, upon substrate binding, favor appropriate positioning of Arg
46 for electrostatic catalysis (
17- Warshel A.
- Sharma P.K.
- Kato M.
- Xiang Y.
- Liu H.
- Olsson M.H.M.
Electrostatic basis for enzyme catalysis.
) like in MtDS-activated MtCM (
11- Sasso S.
- Ökvist M.
- Roderer K.
- Gamper M.
- Codoni G.
- Krengel U.
- Kast P.
Structure and function of a complex between chorismate mutase and DAHP synthase: efficiency boost for the junior partner.
). However, complex formation with MtDS provides this active conformation by a different mechanism than in autonomous N-s4.15. Rather than taking advantage of an ideally prearranged H1-H2 loop, native MtCM involves its C terminus. By hooking onto MtDS, the C terminus exposes Leu
88, which in turn binds to Leu
54, inducing a register shift that extends to residue 55, and concomitantly repositions Arg
46 for catalysis (
11- Sasso S.
- Ökvist M.
- Roderer K.
- Gamper M.
- Codoni G.
- Krengel U.
- Kast P.
Structure and function of a complex between chorismate mutase and DAHP synthase: efficiency boost for the junior partner.
).
The diffusion barrier is generally regarded as the only obvious limitation for the
kcat/
Km of “perfect” enzyme catalysis (
42Enzyme kinetics and molecular evolution.
). However, whereas some enzyme-catalyzed reactions (
43- Blacklow S.C.
- Raines R.T.
- Lim W.A.
- Zamore P.D.
- Knowles J.R.
Triosephosphate isomerase catalysis is diffusion controlled.
,
) reach the diffusion-limited apparent second-order rate constant (
kcat/
Km) of 10
8–10
9m−1 s
−1, a global analysis (
45- Bar-Even A.
- Noor E.
- Savir Y.
- Liebermeister W.
- Davidi D.
- Tawfik D.S.
- Milo R.
The moderately efficient enzyme: evolutionary and physicochemical trends shaping enzyme parameters.
) of catalytic parameters revealed that the “average enzyme” only has a
kcat/
Km of ∼10
5m−1 s
−1. It is of significant interest to elucidate the factors behind nature's reasons for making seemingly less-than-perfect catalysts: is it the difficulty of the chemistry, or the intrinsically limited potential of proteins to evolve for a given catalytic task, or the lack of sufficient selection pressure?
Even though the activity of our top-evolved MtCM variant N-s4.15 is still 2–3 orders of magnitude below the diffusion limit, it has reached essentially the same activity level as the best natural CMs characterized so far (
Fig. 7). A similar upper value for
kcat/
Km was recently confirmed for a set of previously unknown CMs that were sampled from 1130 natural AroQ sequences of phylogenetically widely diverse organisms (
46- Russ W.P.
- Figliuzzi M.
- Stocker C.
- Barrat-Charlaix P.
- Socolich M.
- Kast P.
- Hilvert D.
- Monasson R.
- Cocco S.
- Weigt M.
- Ranganathan R.
An evolution-based model for designing chorismate mutase enzymes.
). The systematic decrease of improvements per evolutionary round and the inability to go beyond ∼10
6m−1 s
−1 despite applying sophisticated evolutionary strategies might indicate that we have approached an intrinsic threshold for the evolution of the enzymatic Claisen rearrangement of chorismate.
This interpretation is supported by a recent site-directed mutagenesis study of MtDS, the partner enzyme of MtCM. Guided by statistical coupling analysis, Parker and co-workers (
47- Jiao W.
- Fan Y.
- Blackmore N.J.
- Parker E.J.
A single amino acid substitution uncouples catalysis and allostery in an essential biosynthetic enzyme in Mycobacterium tuberculosis.
) generated an MtDS variant (Y131A) impaired in allosteric regulation by the pathway's end products Phe, Tyr, and Trp. Moreover, this MtDS variant lent an unexpected activity boost to WT MtCM, with a reported catalytic efficiency
kcat/
Km of 7 × 10
5m−1 s
−1 for MtCM when in the large heterooctameric complex, matching the level for the evolved stand-alone MtCM variant N-s4.15 (
47- Jiao W.
- Fan Y.
- Blackmore N.J.
- Parker E.J.
A single amino acid substitution uncouples catalysis and allostery in an essential biosynthetic enzyme in Mycobacterium tuberculosis.
). The origins of this activity enhancement still need to be elucidated, as MtDS residue 131 is positioned ∼30 Å from the MtCM-MtDS interface. It was speculated that tweaked subunit positioning and conformational changes at the interface and/or an altered dynamic equilibrium might have affected the catalytic parameters (
47- Jiao W.
- Fan Y.
- Blackmore N.J.
- Parker E.J.
A single amino acid substitution uncouples catalysis and allostery in an essential biosynthetic enzyme in Mycobacterium tuberculosis.
).
That
kcat/
Km of typical natural CMs hardly reaches 10
6m−1 s
−1 prompted several investigations into the nature of the rate-limiting steps. Kinetic isotope effects determined for an AroQ
α-subclass CM and the structurally unrelated AroH-class BsCM (
Fig. 7) showed that the chemistry of the [3,3]-sigmatropic rearrangement of chorismate is largely rate-determining for BsCM (
48- Gustin D.J.
- Mattei P.
- Kast P.
- Wiest O.
- Lee L.
- Cleland W.W.
- Hilvert D.
Heavy atom isotope effects reveal a highly polarized transition state for chorismate mutase.
) but not for the AroQ
α enzyme (
49- Addadi L.
- Jaffe E.K.
- Knowles J.R.
Secondary tritium isotope effects as probes of the enzymic and nonenzymic conversion of chorismate to prephenate.
). The AroQ
α result implies a kinetically significant transition state prior to the chemical step, possibly involving ligand complexation or protein conformational changes. Viscosity-variation experiments established that diffusive processes also partially limit the reaction rate of BsCM (
50- Mattei P.
- Kast P.
- Hilvert D.
Bacillus subtilis chorismate mutase is partially diffusion-controlled.
). The fact that its
kcat/
Km is still far below the diffusion limit suggests that a rare conformation of the flexible chorismate and/or the enzyme is required for a catalytically productive binding event (
50- Mattei P.
- Kast P.
- Hilvert D.
Bacillus subtilis chorismate mutase is partially diffusion-controlled.
,
51- Guilford W.J.
- Copley S.D.
- Knowles J.R.
On the mechanism of the chorismate mutase reaction.
).
In addition to the intrinsic physicochemical limitations (
45- Bar-Even A.
- Noor E.
- Savir Y.
- Liebermeister W.
- Davidi D.
- Tawfik D.S.
- Milo R.
The moderately efficient enzyme: evolutionary and physicochemical trends shaping enzyme parameters.
) for the Claisen rearrangement of chorismate, evolutionary scenarios must be considered, too. Global data analysis suggests that evolutionary pressure causes natural enzymes to become more proficient until a balance is reached between an organism's metabolic needs and the cost for producing and improving a catalyst (
45- Bar-Even A.
- Noor E.
- Savir Y.
- Liebermeister W.
- Davidi D.
- Tawfik D.S.
- Milo R.
The moderately efficient enzyme: evolutionary and physicochemical trends shaping enzyme parameters.
). As a result, most enzymes found in nature are mediocre rather than functioning at the maximum catalytic efficiency (
52- Newton M.S.
- Arcus V.L.
- Gerth M.L.
- Patrick W.M.
Enzyme evolution: innovation is easy, optimization is complicated.
). Moreover, natural CMs have evolved to work optimally under physiological conditions including, for instance, the presence of matching upstream and downstream enzymatic processes or the actual substrate and product concentrations within a cell. Because our selection system depends on prevailing intracellular constraints, we speculate that the top-evolved variants have reached a level of catalytic proficiency that would be difficult to exceed by further
in vivo evolution.
In fact, the MtCM-derived top-evolved N-s4.15 rivals the highest-ever reported CM efficiencies even with its very simple AroQ
δ fold, built from the shortest primary sequences known for CMs. At the same time, the evolution experiment demonstrated that the compromised natural enzyme MtCM already possesses all functional groups required for efficient catalysis despite lacking otherwise absolutely conserved active site residues. The relatively facile evolutionary trajectory to boost activity by 2–3 orders of magnitude implies that AroQ
δ subclass CMs on their own evolved to be intentionally poor natural catalysts for enabling inter-enzyme allosteric regulation by switching between mediocre and highly active states (
1- Munack S.
- Roderer K.
- Ökvist M.
- Kamarauskaite˙ J.
- Sasso S.
- van Eerde A.
- Kast P.
- Krengel U.
Remote control by inter-enzyme allostery: a novel paradigm for regulation of the shikimate pathway.
,
2- Burschowsky D.
- Thorbjørnsrud H.V.
- Heim J.B.
- Fahrig-Kamarauskaitė J.
- Würth-Roderer K.
- Kast P.
- Krengel U.
Inter-enzyme allosteric regulation of chorismate mutase in Corynebacterium glutamicum: structural basis of feedback activation by Trp.
,
11- Sasso S.
- Ökvist M.
- Roderer K.
- Gamper M.
- Codoni G.
- Krengel U.
- Kast P.
Structure and function of a complex between chorismate mutase and DAHP synthase: efficiency boost for the junior partner.
,
22- Blackmore N.J.
- Nazmi A.R.
- Hutton R.D.
- Webby M.N.
- Baker E.N.
- Jameson G.B.
- Parker E.J.
Complex formation between two biosynthetic enzymes modifies the allosteric regulatory properties of both: an example of molecular symbiosis.
,
23- Jiao W.
- Blackmore N.J.
- Nazmi A.R.
- Parker E.J.
Quaternary structure is an essential component that contributes to the sophisticated allosteric regulation mechanism in a key enzyme from Mycobacterium tuberculosis.
). We expect that analogous rigorous evolutionary studies of natural enzymes will elucidate capabilities and limitations of numerous biocatalysts and shed light on yet undiscovered allosteric control mechanisms. Furthermore, the ease of finding beneficial mutations gives fundamental insights into the development of drug resistance and the evolution of enzyme function.
Experimental procedures
Materials and general procedures
Plasmid DNA purification from
E. coli cultures was performed using the Genomed Jetquick spin columns (Brunschwig AG, Switzerland), NucleoSpin cups (Macherey-Nagel, Germany), or ZR-Miniprep Classic kit (Zymo Research). DNA from PCR, restriction digestions, and ligations was purified either directly from the reactions (using the DNA Clean and Concentrator kit-5 from Zymo Research, Jetquick Spin columns from Genomed, or NucleoSpin cups from Macherey-Nagel) or after agarose gel electrophoresis in TAE (40 m
m Tris base, 20 m
m acetic acid, and 1 m
m EDTA, pH 8.25) buffer (using the Zymoclean
TM Gel DNA Recovery kit from Zymo Research or Jetquick Spin columns from Genomed). DNA concentration was determined spectrophotometrically using NanoDrop (Thermo Fisher Scientific). DNA manipulations were performed using standard procedures (
53Molecular Cloning: A Laboratory Manual.
) or according to manufacturer recommendations. All cloned PCR–amplified fragments were checked for undesired mutations by sequence analysis. Sanger DNA sequencing and oligonucleotide synthesis were generally performed by Microsynth AG (Switzerland). Oligonucleotides were purified by desalting for routine primers and by HPLC for degenerate primers by Microsynth AG. Restriction endonucleases, Phusion DNA-polymerase, and T4 DNA ligase (for 16-h library ligations at 16 °C) were purchased from New England Biolabs. pFPhe was from Bachem Holding AG (Switzerland). Chorismate for enzymatic assays was produced following a published protocol (
54- Grisostomi C.
- Kast P.
- Pulido R.
- Huynh J.
- Hilvert D.
Efficient in vivo synthesis and rapid purification of chorismic acid using an engineered Escherichia coli strain.
). Other chemicals were purchased from Sigma-Aldrich/Fluka.
Bacterial strains and plasmids
E. coli strain KA12 (F
−, λ
−, Δ
(srlR-recA)306::Tn
10 (Tet
R), Δ(
pheA-tyrA-aroF),
thi-1,
endA1,
hsdR17 (r
K-, m
K+), Δ(
argF-lac)205(U169),
supE44) (
30- Kast P.
- Asif-Ullah M.
- Jiang N.
- Hilvert D.
Exploring the active site of chorismate mutase by combinatorial mutagenesis and selection: the importance of electrostatic catalysis.
) was used for cloning, protein production, and
in vivo assays.
E. coli strain KA13 (genotype as KA12, additionally carrying λ (DE3) [UV5 P
lac-expressed T7 RNA polymerase gene,
imm21, Δ
nin5,
Sam7 (
int−)]) (
55UGA read-through artifacts - When popular gene expression systems need a pATCH.
,
56- MacBeath G.
- Kast P.
- Hilvert D.
A small, thermostable, and monofunctional chorismate mutase from the archaeon Methanococcus jannaschii.
) was used for overproduction of MtDS using plasmid pKTDS-HN (
11- Sasso S.
- Ökvist M.
- Roderer K.
- Gamper M.
- Codoni G.
- Krengel U.
- Kast P.
Structure and function of a complex between chorismate mutase and DAHP synthase: efficiency boost for the junior partner.
). The plasmid pKIMP-UAUC was used in selection experiments (
30- Kast P.
- Asif-Ullah M.
- Jiang N.
- Hilvert D.
Exploring the active site of chorismate mutase by combinatorial mutagenesis and selection: the importance of electrostatic catalysis.
). It carries the chloramphenicol resistance gene (
cat) and provides genes
tyrA* and
pheC encoding monofunctional forms of prephenate dehydrogenase (PDH) from
Erwinia herbicola and prephenate dehydratase (PDT) from
Pseudomonas aeruginosa, respectively. Plasmid pT7POLTS (
31- Roderer K.
- Neuenschwander M.
- Codoni G.
- Sasso S.
- Gamper M.
- Kast P.
Functional mapping of protein-protein interactions in an enzyme complex by directed evolution.
) was used for protein overproduction in strain KA12. It carries
cat, the p15A origin of replication, and the genes for both the tetracycline repressor (
tetR) and the P
tet-controlled T7 RNA polymerase fused to a C-terminal SsrA degradation tag. The latter allows for reduction of the uninduced intracellular T7 RNA polymerase concentration by targeting it to the ClpXP protease system, thus alleviating potential toxic effects of genes due to leaky expression during the growth phase (
31- Roderer K.
- Neuenschwander M.
- Codoni G.
- Sasso S.
- Gamper M.
- Kast P.
Functional mapping of protein-protein interactions in an enzyme complex by directed evolution.
).
Plasmids pKECMB-W (
57- MacBeath G.
- Kast P.
- Hilvert D.
Probing enzyme quaternary structure by combinatorial mutagenesis and selection.
), pKSS (
58pKSS - A second-generation general purpose cloning vector for efficient positive selection of recombinant clones.
), pKTCTET-0 (
31- Roderer K.
- Neuenschwander M.
- Codoni G.
- Sasso S.
- Gamper M.
- Kast P.
Functional mapping of protein-protein interactions in an enzyme complex by directed evolution.
), pKSS-TM4 (
31- Roderer K.
- Neuenschwander M.
- Codoni G.
- Sasso S.
- Gamper M.
- Kast P.
Functional mapping of protein-protein interactions in an enzyme complex by directed evolution.
), pKTNTET (
31- Roderer K.
- Neuenschwander M.
- Codoni G.
- Sasso S.
- Gamper M.
- Kast P.
Functional mapping of protein-protein interactions in an enzyme complex by directed evolution.
), pKTCMM-H (
11- Sasso S.
- Ökvist M.
- Roderer K.
- Gamper M.
- Codoni G.
- Krengel U.
- Kast P.
Structure and function of a complex between chorismate mutase and DAHP synthase: efficiency boost for the junior partner.
), and pHS10-3p5 (
38Structure, function, and directed evolution of shikimate pathway enzymes from Mycobacterium tuberculosis.
) were described previously. Plasmid pMG248 (3681 bp) contains a 5′-truncated nonfunctional
aroQδ gene fragment. It was assembled by ligating the 3462-bp BglII-BlpI fragment from pMG210 (
16- Sasso S.
- Ramakrishnan C.
- Gamper M.
- Hilvert D.
- Kast P.
Characterization of the secreted chorismate mutase from the pathogen Mycobacterium tuberculosis.
) with a 219-bp BglII/BlpI-digested PCR product generated with primers 300-TEMP (5′-TAAGAT
GCTCAGCGAGATCGACCGGCTAGA, restriction sites underlined) and 301-TEMP (5′-TAA
AGATCTGTGACCGAGGCGGCCACGGCCCAAT) using plasmid pMG242 (
31- Roderer K.
- Neuenschwander M.
- Codoni G.
- Sasso S.
- Gamper M.
- Kast P.
Functional mapping of protein-protein interactions in an enzyme complex by directed evolution.
) as template.
3M. Gamper, unpublished results.
Acceptor plasmids for aroQδ gene libraries
The initial acceptor vector pKTNTET-0 for the first
aroQδ gene libraries codes for an N-terminal Met-hexahistidine tag linked by Ser-Ser-Gly to the Met
5-Ser
39 fragment of MtCM. pKTNTET-0 was constructed from the 2703-bp and 1052-bp fragments obtained from a quadruple restriction endonuclease digestion of pKTCTET-0 (
31- Roderer K.
- Neuenschwander M.
- Codoni G.
- Sasso S.
- Gamper M.
- Kast P.
Functional mapping of protein-protein interactions in an enzyme complex by directed evolution.
) with enzymes NdeI, BlpI, NheI, and AscI and two correspondingly digested PCR products generated with oligonucleotides 383-dNhe-S (5′-GCTAA
TCTAGAAGCACGCCATAGTGACTG) and 384-Hind-N (5′-CCTAA
GCTCAGCATAAGCTTCCGCAGCCACTAGTCATTATTAGTGGTGGTGGT) on pKTCTET-0 (PCR product cut with XbaI and BlpI, 209 bp) and oligonucleotides 352-CTLIB-S (5′-CCTGTT
CATATGCACCATCATCATCACCACTCTT) and 385-NheAsc-N (5′-GGTAA
GGCGCGCCCGCTAGCCATCCGGGCCTTGCCGAT) on pKSS-TM4 (PCR product digested with NdeI and AscI, 170 bp). A four-fragment ligation yielded the desired 4134 bp pKTNTET-0, thereby removing the unwanted NheI restriction site in the stuffer fragment.
Acceptor vectors for L1.13, L4.4, and L4.7 libraries were constructed by ligating an NheI/HindIII-digested 1261-bp stuffer fragment from pKTNTET-0 with the correspondingly digested 2873-bp fragment of plasmids pKTNTET-L1.13, pKTNTET-L4.4, and pKTNTET-L4.7, containing the genes of the selected variants L1-13, L4-4, and L4-7 (
Table S2), respectively. This yielded the 4134-bp acceptor vectors pKTNTET-0-L1.13, pKTNTET-0-L4.4, and pKTNTET-0-L4.7.
To avoid a second AscI site in the acceptor vectors for the L3.6, L3.7, and L3.8 libraries, the AscI/HindIII-digested 1253-bp stuffer fragment from pKTNTET-0 was ligated with the correspondingly digested 2833-bp fragment from pKTNTET-L3.6, pKTNTET-L3.7, and pKTNTET-L3.8, containing the genes of the selected variants L3.6, L3.7, and L3.8 (
Table S2), respectively, yielding the 4086-bp acceptor vectors pKTNTET-0-L3.6, pKTNTET-0-L3.7, and pKTNTET-0-L3.8, respectively.
The acceptor vector pKTNTET-dHis-0 for the untagged version of the aroQδ genes was constructed by cloning the PCR fragment generated with oligonucleotides 131-TERM (5′-CCCTCAAGACCCGTTTAGA) and 533-pKTNTET-dHis (5′-AGATATACATATGCTCGAGTCCCAACCTGTCC), using pKTNTET-0 as the template. The PCR product was restriction-digested with NdeI and SpeI, and the resulting 1378-bp fragment was ligated with the 2726-bp fragment of the accordingly digested pKTNTET-0. The resulting 4104-bp pKTNTET-dHis-0 vector was sequenced using oligonucleotide 60-T7Pro (5′-TAATACGACTCACTATAGGG).
The acceptor vector pKTNTET-RepA for the LS4RepA and LS4RepA2 libraries, which encoded the RepA degradation tag MNQSFISDILYADIES (
59- Butz M.
- Neuenschwander M.
- Kast P.
- Hilvert D.
An N-terminal protein degradation tag enables robust selection of highly active enzymes.
,
60- Díaz-Lopéz T.
- Lages-Gonzalo M.
- Serrano-Lopéz A.
- Alfonso C.
- Rivas G.
- Díaz-Orejas R.
- Giraldo R.
Structural changes in RepA, a plasmid replication initiator, upon binding to origin DNA.
,
61Two peptide sequences can function cooperatively to facilitate binding and unfolding by ClpA and degradation by ClpAP.
), was based on pKTNTET-0 as backbone and PCR template. A first PCR product for overlap extension was generated using primers 131-TERM and 538-pKTNTET-Fw-RepA (5′-TTAGCGATATTCTGTATGCGGATATTGAATCC
CTCGAGTCCCAACCTGT), which contained the XhoI site but excluded the Met codon following the RepA sequence. A second PCR product was obtained using primers 372-TetPro (5′-AGCTCTAATGCGCTGTTAATCACT) and 539-pKTNTet-Rev-RepA (5′-ATACAGAATATCGCTAATAAAGCTCTGGTT
CATATGTATATCTCCTTC) providing an NdeI site. The resulting 257-bp and 1479-bp PCR products that had a 17-bp overlapping sequence were assembled using the external 372-TetPro and 131-TERM primers to give a 1719-bp fragment. After digestion with SpeI and NdeI, the resulting 1423-bp fragment was ligated with the correspondingly digested 2726-bp fragment from pKTNTET-0 to obtain vector pKTNTET-RepA (4149 bp).
Cassette mutagenesis libraries and selection
Cassette mutagenesis for the evolutionary cycle I was performed with oligonucleotides 368-Lp-TRLVfw (5′-ATCGGCAAGGCCCGGATGGCTAGCGGTGGCNNKNNKNNKNNKCATAGTCGGGAGATGAAGGTCATCGAAC) and 386-LpLib-N2 (5′-GGTTAAAGCTTCCGCAGCCACTAGTTATTAGTGACCGAGGCGGCCACGGCCCAAT) on template pMG248 (carrying a 5′-truncated nonfunctional aroQδ gene fragment). The NheI/HindIII-digested 148-bp library fragment was ligated to the accordingly cut 2873-bp pKTNTET-0 acceptor fragment.
For cycle II mutagenesis, PCR fragments were generated using 390-CT-LGHrv-2 (5′-GGTTAAAGCTTCCGCAGCCACTAGTTATTANNNNNNNNNTCGACCACGACCAAGACGCAAAAGCAGGATGGCCAGAT), 424-CT-RLGHrv (5′-GGTTAAAGCTTCCGCAGCCACTAGTTATTANNNNNNNNNNNNACCACGACCAAGACGCAAAAGCAGGATGGCCAGAT), or 426-CT7rv (5′-GGTTAAAGCTTCCGCAGCCACTAGTTATTANNNNNNNNNNNNNNNNNNNNNAAGACGCAAAAGCAGGATGGCCAGAT) together with 379-LpconS (5′-GTTCGCTAGCGGAGGTCCACGTCTTGATCATAGTCGGGAGATGAAGGTCATCGAAC) on template pKSS-TM4 (carrying a 3′-truncated aroQδ gene). The crude 163-bp PCR products were digested with NheI and HindIII, and the resulting 148-bp fragments were ligated with the accordingly cut 2873-bp pKTNTET-0 acceptor fragment.
The ligation products were transformed into electrocompetent KA12/pKIMP-UAUC cells (
30- Kast P.
- Asif-Ullah M.
- Jiang N.
- Hilvert D.
Exploring the active site of chorismate mutase by combinatorial mutagenesis and selection: the importance of electrostatic catalysis.
). The suspension of transformed cells was washed three times with 1× M9 salts (6 mg/ml Na
2HPO
4, 3 mg/ml KH
2PO
4, 1 mg/ml NH
4Cl, and 0.5 mg/ml NaCl) (
53Molecular Cloning: A Laboratory Manual.
). For cycle I, the cells were spread on M9c minimal medium plates (1.5% agar), which are based on 1× M9 salts and also contain 0.2% (w/v)
d-(+)-glucose, 1 m
m MgSO
4, 0.1 m
m CaCl
2, 5 μg/ml thiamine-HCl, 5 μg/ml 4-hydroxybenzoic acid, 5 μg/ml 4-aminobenzoic acid, 1.6 μg/ml 2,3-dihydroxybenzoic acid, 20 μg/ml Trp, 100 μg/ml sodium ampicillin, and 20 μg/ml chloramphenicol, but lack an inducer for gene expression. Library sizes were determined from M9c+Phe+Tyr plates (minimal M9c medium additionally containing 20 μg/ml Tyr and 20 μg/ml Phe). For selection in cycle II, the minimal M9c plates additionally contained 100 μ
m pFPhe.
Construction of destabilized Leu24 and Leu31 MtCM variants
Libraries L1, L2, L3, and L4 were constructed by overlap-extension PCR, using plasmid pHS10-3p5 (which encodes the His-tagged 3p5 variant) as the template (
38Structure, function, and directed evolution of shikimate pathway enzymes from Mycobacterium tuberculosis.
). The 143-bp 5′ part of the gene was constructed using the forward primer 352-CTLIB-S (5′-CCTTGTT
CATATGCACCATCATCATCACCACTCTT) and an individual reverse primer for each library that introduced an AscI restriction site, removed the BsaHI site, and randomized specific codons. The reverse primers 505-L24-N (5′-AAACCTC
GGCGCGCCGCTTGACTAACGCAGGATTTCAGCATCMNNACGGTCGATCTCTTCGCGCA, randomizing Leu
24 via the NNK codon), 506-L31-N (5′-AAACCTC
GGCGCGCCGCTTAACMNNAGCGAGGATTTCAGCATCTAGCCGGTCGATCTCTT, randomizing Leu
31 via NNK), 507-L24-L31-N (5′-AAACCTC
GGCGCGCCGCTTAACMNNAGCGAGGATTTCAGCATCMNNACGGTCGATCTCTTCGCGCA, randomizing both Leu
24 and Leu
31 via NNK), and 508-L24-L31-RNK-N (5′-AAACCTC
GGCGCGCCGCTTAACMNYAGCGAGGATTTCAGCATCMNYACGGTCGATCTCTTCGCGCA, randomizing Leu
24 and Leu
31 via RNK, which excludes stop codons) were used for the gene libraries L1, L2, L3, and L4 respectively. For all four libraries, the 202-bp 3′ part of the gene was generated using forward primer 509-AscI-S (5′-AAGC
GGCGCGCCGAGGTTTCCAAGGCCATCGG, introducing an AscI site) and reverse primer 510-noBsaHI-N (5′-TCACAGCTTCCGCAGCC
ACTAGTTATTACATAGCATCCGGACCACGACCAAGAC). The 143-bp and 202-bp PCR fragments were assembled using external primers 352-CTLIB-S and 510-noBsaHI-N, and the resulting 326-bp fragment was restriction-digested with XhoI (coincidentally present in the
aroQδ gene) and SpeI. The obtained 260-bp fragments from L1, L2, L3, and L4 libraries were ligated with the 2761-bp XhoI-SpeI fragments of pKTNTET-0 yielding the 3021-bp library plasmids.
After transformation of electrocompetent KA12/pKIMP-UAUC and washing twice in 1× M9 salts, the transformants were plated onto relaxed-stringency M9c+Phe+Tet500ng/ml minimal agar (containing M9c medium as described above, with 500 ng/ml tetracycline to induce aroQδ gene expression and 20 μg/ml Phe). A small fraction of washed cells were also plated onto nonselective M9c+Phe+Tyr plates for library size estimation. Typically, the aroQδ gene of five clones from these plates was sequenced to determine the quality and mutation rates of the library. Clones from M9c+Phe+Tet500ng/ml plates were picked and streaked out onto higher-stringency plates, such as M9c+Tet80ng/ml, M9c without additives, and M9c+pFPhe40μm (containing 40 μm toxic pFPhe). Clones not growing after 72 h at 30 °C on these stringent plates were sequenced using 131-TERM.
Construction of epL1.13, epL4.4, epL4.7, epL3.6, epL3.7, epL3.8, shL1.13, shL4.4, shL4.7, shL3.6, shL3.7, and shL3.8 libraries, and in vivo selection
To obtain the epPCR libraries (epL), two rounds of epPCR were performed on the relatively short
aroQδ template using the Mutazyme II kit from Stratagene (Agilent Technologies). First, the appropriate DNA of
aroQδ variants from plasmids pKTNTET-L1.13, pKTNTET-L4.4, pKTNTET-L4.7, pKTNTET-L3.6, pKTNTET-L3.7, and pKTNTET-L3.8 (which encode variants L1.13, L4.4, L4.7, L3.6, L3.7, and L3.6, respectively;
Table S2) was amplified using primers 372-TetPro and 131-TERM and 0.5 ng of the template plasmid to yield 591-bp products. The second round of epPCR for the epL3.6, epL3.7, and epL3.8 libraries was done by using 0.5 ng of the 591-bp epPCR product and the 509-AscI-S/131-TERM primer pair to obtain 253-bp products, which were used for cloning after appropriate restriction digestion (see below). For libraries epL1.13, epL4.4, and epL4.7, 0.5 ng of the appropriate 591-bp epPCR products and the 372-TetPro/131-TERM primer pair were used for the second round of epPCR to generate again 591-bp products. In this case, an additional nonmutagenic PCR was performed using primers 509-AscI-S and 131-TERM to obtain the 253-bp products desired for cloning of the epL1.13, epL4.4, and epL4.7 libraries.
DNA shuffling libraries were constructed by shuffling
aroQδ genes carrying beneficial mutations including the genes from the previously constructed plasmids pHS10-3p5 (
aroQδ gene encoding variant PHS10-3p5, mutations T52P, V55D, R87P, L88D, G89A, and H90M), pHS08-3p20 (PHS08-3p20, mutation V62I), pHS08-5p12 (PHS08-5p12, mutation G43V), pHS10-ANp10 (PHS10-ANp10, mutation R82Q), and pHS10-2p14 (PHS10-2p14, mutation D75Y) (
38Structure, function, and directed evolution of shikimate pathway enzymes from Mycobacterium tuberculosis.
). These genes were PCR-amplified using primers 131-TERM and 509-AscI-S to yield 591-bp products. 600 ng of each PCR product were mixed and treated with 1 μg of DNase I until the preferred fragments of ∼50 bp were obtained. These DNA fragments were reassembled by PCR using primers 131-TERM and 509-AscI-S to yield a 253-bp product. If the yield of the reassembly PCR was low, an additional PCR amplification with the same primers was performed.
The 253-bp products of the epPCRs and the reassembly PCR of DNA shuffling for the L1.13, L4.4, and L4.7 libraries were restriction-digested with AcsI and HindIII, and the resulting 188-bp fragments were respectively ligated to the correspondingly digested 2833-bp fragment of acceptor vectors pKTNTET-0-L1.13, pKTNTET-0-L4.4, and pKTNTET-0-L4.7, yielding 3021-bp library plasmids. The 253-bp products of the epPCRs and reassembly PCRs for the L3.6, L3.7, and L3.8 libraries were restriction-digested with AcsI and SpeI, and the resulting 174-bp fragments were ligated to the correspondingly digested 2847-bp fragment of acceptor vectors pKTNTET-0-L3.6, pKTNTET-0-L3.7, and pKTNTET-0-L3.8, respectively, yielding the 3021-bp library plasmids.
The ligations were transformed into electrocompetent KA12/pKIMP-UAUC cells. The cells were washed twice with 1× M9 salts and spread onto the M9c plates (listed by increasing stringency) M9c+Phe+Tet80ng/ml, M9c+Phe+Tet40ng/ml, M9c+Tet80ng/ml, M9c+Tet40ng/ml, M9c, M9c+pFPhe40μm, and M9c+pFPhe100μm (containing 100 μm pFPhe). Library size and quality were determined from platings on nonselective M9c+Phe+Tyr agar. Colonies still growing at the highest selective conditions were picked and sequenced using oligonucleotide 131-TERM.
Reversion of destabilizing mutations from evolved inter-subunit destabilized variants
Overlap-extension PCR was used to remove the L24 and L31 mutations from the further-evolved aroQδ variants. Fragment PCR I was obtained with oligonucleotides 372-TetPro and 512-x31L-Fw (5′-GACGCCGAAATCCTCGCGTTAGTCAAGCGACGCGCTGAGG) on the appropriate pKTNTET-based template. Fragment PCR II was generated with primers 513-x24L-Rev (5′-CGCGAGGATTTCGGCGTCTAGCCGGTCGATCTCTTCGCGC) and 131-TERM, on the same template. PCR I and PCR II fragments were combined using oligonucleotides 372-TetPro and 131-TERM. The resulting PCR products were digested with XhoI and HindIII to obtain the required 349-bp insert, which was ligated with the 2747-bp fragments from the appropriately digested pKTNTET-0.
Design and construction of the truncation libraries
The truncation libraries LdNdC and LdNR85X were based on variant re4.7s11 (s11). Each library offered incremental two-residue deletions between the His-tag (including the adjacent Ser-Ser-Gly-Met-Leu-Glu sequence to retain the linker and XhoI site for in-frame cloning) and the active site residue Arg
18 (
Fig. 3). For library LdNdC, the C terminus was also varied by allowing truncations of one residue at a time until Arg
85 (
Fig. 3). This residue, together with Gly
84, had been shown in previous directed evolution experiments to be conserved and important for catalysis (
31- Roderer K.
- Neuenschwander M.
- Codoni G.
- Sasso S.
- Gamper M.
- Kast P.
Functional mapping of protein-protein interactions in an enzyme complex by directed evolution.
). In addition to the random two-residue deletions in the N-terminal region, library LdNR85X members also lacked the last five residues at the C terminus and had codon 85 randomized (
Fig. 3).
The LdNdC library of truncated s11 genes was constructed from pKTNTET-re4.7s11 (which encodes the His-tagged s11 variant) as a template and a mixture of several forward and reverse primers. The forward primers 521-MtCMi-N-S8 (5′-GTATG
CTCGAGTCCCAACCTGTCCCCGAGATCGACACGC), 522-MtCMi-N-P10 (5′-GTATG
CTCGAGCCTGTCCCCGAGATCGACACGCTGC), 523-MtCMi-N-P12 (5′-GTATG
CTCGAGCCCGAGATCGACACGCTGCGCGAAG), 524-MtCMi-N-I14 (5′-GTATG
CTCGAGATCGACACGCTGCGCGAAGAGATC), 525-MtCMi-N-T16 (5′-GTATG
CTCGAGACGCTGCGCGAAGAGATCGAC), and 526-MtCMi-N-R18 (5′-GTATG
CTCGAGCGCGAAGAGATCGACCGGCTAG) introduced an XhoI restriction site via the Leu-Glu–encoding sequence directly following the MtCM-native codon for Met
5 and the respective N-terminal deletions up to residue Ser
8 (no truncation), Pro
10, Pro
12, Ile
14, Thr
16, and Arg
18. The reverse primers 527-MtCMi-C-M90 (5′-CAGCC
ACTAGTTATTACATAGCATCCGGACCACGACCA), 528-MtCMi-C-A89 (5′-CAGCC
ACTAGTTATTAAGCATCCGGACCACGACCAAGC), 529-MtCMi-C-D88 (5′-CAGCC
ACTAGTTATTAATCCGGACCACGACCAAGAC), 530-MtCMi-C-P87 (5′-CAGCC
ACTAGTTATTACGGACCACGACCAAGACGCAA), 531-MtCMi-C-G86 (5′-CAGCC
ACTAGTTATTAACCACGACCAAGACGCAAAAGC), and 532-MtCMi-C-R85 (5′-CAGCC
ACTAGTTATTAACGACCAAGACGCAAAAGC) introduced a SpeI restriction site, two stop codons, and C-terminal truncations back to Met
90 (no truncation), Ala
89, Asp
88, Pro
87, Gly
86, and Arg
85, respectively (
Fig. 3).
The PCR product for library LdNdCR85X was constructed using the same template and mixtures of forward primers as described for library LdNdC. The reverse primer was 534-MtCMi-C-R85X (5′-CAGCCACTAGTTATTAMNNACCAAGACGCAAAAGCAG), which introduced a SpeI restriction site and two stop codons after Arg85, and randomized Arg85 via the NNK codon. The PCR products for libraries LdNdC and LdNdCR85X, which had different sizes depending on the extent of truncation, were digested with XhoI and SpeI and ligated with the 2761-bp fragment of XhoI/SpeI-digested pKTNTET-0.
Selection of truncated re4.7s11 (s11) variants
The electrocompetent KA12/pKIMP-UAUC cells transformed with the ligation mixtures were plated onto relaxed-stringency M9c+Phe+Tet500ng/ml agar, as described above. Library sizes were calculated from colony growth on nonselective M9c+Phe+Tyr plates, and five clones were sequenced to determine library quality and mutation rate. To confirm a reduced complementation ability of truncated variants, clones from M9c+Phe+Tet80ng/ml plates were picked, inoculated in 100 μl of 1× M9 salts, and drop-spotted onto higher-stringency plates, such as M9c+Phe+Tet40ng/ml, M9c+Tet80ng/ml, M9c+Tet40ng/ml, M9c, and M9c+pFPhe40μm and M9c+pFPhe100μm containing 40 and 100 μm pFPhe, respectively. The clones growing slowest under stringent conditions were picked and sequenced using 131-TERM.
DNA sequencing revealed only one weakly complementing clone (dNdCs1) from library LdNdC that did not grow at higher stringencies (M9c and M9c+pFPhe plates). In contrast, substitutions of Arg85 in library LdNdCR85X were severe enough to drastically reduce in vivo complementation ability even in the absence of N-terminal lesions. Thus, randomizing Arg85 in a variant of s11 lacking the five C-terminal residues yielded variants with reduced activity in vivo more frequently than the progressive deletions at the N- and C-terminal regions for library LdNdC members.
Randomization of truncation libraries based on dNdCs1 and dNdCR85Xs10 and selection experiments
Truncated variants were chosen as templates for the fourth cycle of directed evolution. The genes of variants dNdCs1 (NΔ10, CΔ5;
Table S5) and dNdCR85Xs10 (NΔ0, CΔ5, R85M) were mutagenized via two rounds of epPCR using the Mutazyme II kit. The first round involved primers 372-TetPro and 131-TERM and 0.5 ng of plasmids pKTNTET-dNdCs1 or pKTNTET-dNR85Xs10 encoding the His-tagged version of the target variants. 0.5 ng of the respective 546-bp and 576-bp epPCR products were used for a second epPCR round using the forward primer 542-Fw-His (5′-ACTCTTCTGGTATG
CTCGAG) and the reverse primers 547-Rv-dNdCs1 (5′-GCAGCC
ACTAGTTATTAACG for dNdCs1) and 548-Rv-dNR85s10 (5′-GCAGCC
ACTAGTTATTACAT for dNdCR85Xs10). The XhoI/SpeI-digested 215-bp and 245-bp products were subcloned into the correspondingly digested pKTNTET-0 acceptor vector. The resulting library plasmids carrying the mutagenized dNdCs1 and dNR85s10 genes, respectively, were electroporated into KA12/pKIMP-UAUC. Library size, quality, and mutation rates were determined after plating a fraction of the cells onto nonselective M9c+Phe+Tyr agar plates.
In vivo selection was performed on plates with M9c+Phe+Tet
80ng/ml, M9c+Phe+Tet
40ng/ml, M9c+Tet
80ng/ml, M9c+Tet
40ng/ml, M9c+Tet
10ng/ml, M9c+Tet
5ng/ml, M9c+Tet
2.5ng/ml, M9c, and M9c+pFPhe
40μm and M9c+pFPhe
100μm. Clones that grew fastest under stringent conditions were picked and sequenced using oligonucleotide 131-TERM.
Construction and selection of the RepA-tagged library
To use variant dNdCs4 (NΔ4, CΔ5) as a template for the evolutionary cycle IV, a stepwise evolution protocol was implemented, because this gene already weakly complemented the growth defect of KA12/pKIMP-UAUC. First, an epPCR gene library was generated without the N-terminal His-tag. This was accomplished by mutagenizing the His-tagged dNdCs4 gene (
Table S5) via two rounds of epPCR. First, the gene was amplified with primers 372-TetPro and 131-TERM from plasmid pKTNTET-dNdCs4. For the first epPCR, 0.5 ng of this amplified gene served as a template for primers 372-TetPro and 131-TERM using the Mutazyme II kit. Of the resulting 564-bp mutated PCR products, 0.5 ng was used for a second epPCR round using primers 542-Fw-His and 547-Rv-dNdCs1 to give 259-bp fragments. The XhoI/SpeI-digested 233-bp products (without the sequence encoding for His-tag) were subcloned into the correspondingly digested pKTNTET-dHis-0 vector to generate the 2993-bp library plasmid pool, which was electroporated into KA12/pKIMP-UAUC cells.
Library size, quality, and mutation rates were determined from transformants plated onto M9+Phe+Tyr agar and from five sequenced clones. Selection
in vivo was accomplished by plating onto M9c+Phe+Tet
80ng/ml, M9c+Phe+Tet
40ng/ml, M9c+Tet
80ng/ml, M9c+Tet
40ng/ml, M9c+Tet
10ng/ml, M9c+Tet
5ng/ml, M9c, M9c+pFPhe
40μm, and M9c+pFPhe
100μm. After 3 days at 30 °C, well-growing colonies were picked from the most stringent pates and streaked out onto M9c agar. From these streaks, 192 well-growing clones were inoculated into 100 μl of liquid M9c+Phe+Tyr medium in two 96-well plates and incubated overnight at 37 °C. Eight pools of 24 individual clones grown up in 96-well plates were collected and their plasmid DNA was isolated. The target genes were PCR-amplified from each of the eight pools using oligonucleotides 543-Fw-dHis (5′-GAGATATACATATGCTCGAG) and 547-Rv-dNdCs1. The PCR products were gel-purified, pooled to have 800 ng of each fragment, and subjected to DNA shuffling as described above. The DNase I digested products were assembled by PCR with oligonucleotides 543-Fw-dHis and 547-Rv-dNdCs1 and digested with XhoI and SpeI, and the 233-bp fragments were subcloned into the accordingly digested pKTNTET-RepA vector to generate 3009-bp library plasmids. Thereby, the MtCM variants were fused with the N-terminal RepA protein degradation tag (
59- Butz M.
- Neuenschwander M.
- Kast P.
- Hilvert D.
An N-terminal protein degradation tag enables robust selection of highly active enzymes.
) with the goal of lowering the intracellular enzyme concentration. After electroporation into KA12/pKIMP-UAUC, the resulting Ls4RepA library transformants were plated on M9c+Phe+Tet
80ng/ml, M9c+Phe+Tet
40ng/ml, M9c+Tet
80ng/ml, M9c+Tet
40ng/ml, and M9c plates for selection and on M9c+Phe+Tyr plates for determination of library size, quality, and mutation rates.
The best-complementing genes were then again shuffled to create library LS4RepA2, still carrying the RepA tag. For this, two 96-well plates were inoculated as described above with the clones growing on M9c+Tet80ng/ml and M9c+Tet40ng/ml plates. Again, eight pools of individual clones were prepared, plasmid DNA was extracted, and the genes were PCR-amplified using oligonucleotides 544-Fw-RepA (5′-GCGGATATTGAATCCCTCGAG) and 547-Rv-dNdCs1. The PCR products were shuffled, assembled using primers 544-Fw-RepA and 547-Rv-dNdCs1, and cloned into the pKTNTET-RepA vector as described above for the Ls4RepA library. The resulting Ls4RepA2 library was transformed into KA12/pKIMP-UAUC cells and plated onto M9c, M9c+pFPhe40μm, and M9c+pFPhe100μm agar plates for selection and on M9c+Phe+Tyr plates for determination of library size, quality, and mutation rates. 150 clones that complemented on high-stringency plates were purified by streaking them out on M9c+pFPhe100μm agar plates. Twenty-three of the best-growing clones were retransformed into KA12/pKIMP-UAUC, tested on M9c+pFPhe100μm agar plates, and sequenced using oligonucleotide 131-TERM.
Re-elongation of evolved truncated variants
Individual genes for evolved truncated and/or RepA-tagged variants in pKTNTET-based plasmids from libraries LdCdNs1, LdNdCR85Xs10, and Ls4RepA2 were amplified using an appropriate primer pair (summarized in
Table S11 and
Table S12). The PCR fragments were digested with SpeI/XhoI and the 260-bp inserts ligated with the 2761-bp fragment of the correspondingly digested pKTNTET-0 vector.
Providing the evolved variants with the WT N terminus
To remove the His-tag and instead provide the evolved variants with the native N terminus, the WT MtCM gene on plasmid pKTCMM-H (
11- Sasso S.
- Ökvist M.
- Roderer K.
- Gamper M.
- Codoni G.
- Krengel U.
- Kast P.
Structure and function of a complex between chorismate mutase and DAHP synthase: efficiency boost for the junior partner.
) was replaced by the corresponding fragment of the evolved variant. Restriction digestion with XhoI and SpeI of the pKTNTET-based plasmids that carry the genes of the evolved variants s11 and s4.15 yielded the required 260-bp fragments for ligating with the 4547-bp XhoI-SpeI fragment of pKTCMM-H (
11- Sasso S.
- Ökvist M.
- Roderer K.
- Gamper M.
- Codoni G.
- Krengel U.
- Kast P.
Structure and function of a complex between chorismate mutase and DAHP synthase: efficiency boost for the junior partner.
).
Production and purification of His-tagged proteins
N-terminally His-tagged (pKTNTET-encoded) H
6-MtCM variants, wherein the four native N-terminal MNLE residues of MtCM were replaced by the sequence MHHHHHHSSG, were produced in KA12/pT7POLTS cells (
31- Roderer K.
- Neuenschwander M.
- Codoni G.
- Sasso S.
- Gamper M.
- Kast P.
Functional mapping of protein-protein interactions in an enzyme complex by directed evolution.
). The transformants were grown in 500 ml of LB medium containing 150 μg/ml Na-ampicillin and 30 μg/ml chloramphenicol at 30 °C. Gene expression was induced with 2 μg/ml tetracycline when an
A600 of 0.3–0.5 was reached. The crude lysate was obtained as described before (
62- Gamper M.
- Hilvert D.
- Kast P.
Probing the role of the C-terminus of Bacillus subtilis chorismate mutase by a novel random protein-termination strategy.
) (but without the RNase A and DNase I treatment) and provided with 10 m
m imidazole. The sample was loaded onto an equilibrated nickel-nitrilotriacetic acid column containing His-Select Nickel Affinity Gel (Merck, Germany). The MtCM variant was eluted with 250 m
m imidazole in 50 m
m sodium phosphate buffer, pH 8, containing 0.3
m NaCl and dialyzed against 20 m
m potassium phosphate, pH 7.5. Proteins were assessed by SDS-PAGE using the PhastSystem (20% homogeneous gels, GE Healthcare), and their molecular mass (
Table S10) was determined by electrospray ionization MS by the Mass Spectrometry Service at the Laboratory of Organic Chemistry, ETH Zurich.
Production and purification of untagged MtCM variants
The untagged versions of s11 (N-s11; predicted pI = 6.11) and s4.15 (N-s4.15; predicted pI = 6.09), encoded on pKTCMM-H derived plasmids, were produced in E. coli strain KA13. The transformants were grown at 37 °C in LB medium containing 150 μg/ml Na-ampicillin. At an A600 of 0.3–0.6, gene expression was induced by 0.5 mm isopropyl 1-thio-β-d-galactopyranoside and growth continued overnight. After harvesting by centrifugation (5000 rpm for 10 min at 4 °C), the cells were resuspended in 20 mm 1,3-bis[tris(hydroxymethyl)methylamino]propane (BTP) buffer, pH 7.5, and incubated for 1 h on ice with 1 mg/ml lysozyme prior to disruption by sonication. The crude cell lysate was adjusted to 65–70% (w/v) ammonium sulfate and stirred for 1.5–2 h at 4 °C. The precipitate was collected by centrifugation (5880 rpm for 30 min at 4 °C), dissolved in buffer A (20 mm piperazine, pH 9.0), and dialyzed against the same buffer overnight at 4 °C. Dialyzed samples were loaded onto a pre-equilibrated MonoQ column in buffer A. A gradient between 0–30% of buffer B (20 mm piperazine, pH 9, and 1 m NaCl) in buffer A was applied over 50–100 ml at a flow rate of 2 ml/min. The fractions from the MonoQ column were collected, concentrated, and dialyzed against 20 mm MES, pH 5.5 (buffer C). After dialysis, the sample was applied onto a pre-equilibrated MonoS column in buffer C, and the column was washed with 95% buffer C/5% buffer D (20 mm MES, pH 5.5, and 1 m NaCl). The proteins were eluted with a gradient between 5–40% of buffer D in buffer C over 50–100 ml at a flow rate of 2 ml/min. The MonoS column fractions containing the MtCM variants were pooled, concentrated, and loaded onto a Superdex 75 column. Proteins were eluted in 20 mm BTP, pH 7.5, containing 150 mm NaCl. Protein integrity was analyzed by SDS-PAGE and electrospray ionization MS at the Mass Spectrometry Service of the Laboratory of Organic Chemistry, ETH Zurich.
Production and purification of MtDS
The His-tagged MtDS was produced following the previously established protocol (
38Structure, function, and directed evolution of shikimate pathway enzymes from Mycobacterium tuberculosis.
) with minor modifications. A single colony of KA13/pKTDS-HN (
38Structure, function, and directed evolution of shikimate pathway enzymes from Mycobacterium tuberculosis.
) was inoculated into 5 ml of LB containing 150 μg/ml ampicillin (sodium salt) and grown overnight at 30 °C. The resulting pre-culture was used to inoculate M9c minimal medium as described above but containing 150 μg/ml sodium ampicillin and 10 μg/ml each of Trp, Tyr, and Phe. The culture was grown at 30 °C to an
A600 of 0.3–0.5, and protein production was induced with 0.1 m
m salicylate. After incubation at 30 °C for 16 h, the cells were harvested and resuspended in BTP++ buffer consisting of 20 m
m BTP, pH 7.5, 1 m
m tris(2-carboxyethyl)phosphine hydrochloride, 0.2 m
m phosphoenolpyruvate, 0.1 m
m MnCl
2, and 150 m
m NaCl (
11- Sasso S.
- Ökvist M.
- Roderer K.
- Gamper M.
- Codoni G.
- Krengel U.
- Kast P.
Structure and function of a complex between chorismate mutase and DAHP synthase: efficiency boost for the junior partner.
,
63- Webby C.J.
- Baker H.M.
- Lott J.S.
- Baker E.N.
- Parker E.J.
The structure of 3-deoxy-d-arabino-heptulosonate 7-phosphate synthase from Mycobacterium tuberculosis reveals a common catalytic scaffold and ancestry for Type I and Type II enzymes.
).
The cell pellet was treated with 1 mg/ml lysozyme for 30 min on ice and ruptured by sonication. The insoluble cell debris was removed by centrifugation for 20 min at 13,000 rpm (Sorvall rotor SA600) at 4 °C, and if there was a significant amount of cell pellet left, the sonication was repeated. The crude soluble cell extract was subjected to nickel-nitrilotriacetic acid affinity chromatography (Qiagen, Germany). The bound protein was first washed with BTP++ containing 40 mm imidazole and then with BTP++ containing 40 mm imidazole and 2 m NaCl. MtDS was eluted in BTP++ containing 250 mm imidazole and dialyzed overnight against buffer A (20 mm BTP, pH 7.5, containing 0.1 mm MnCl2, 0.2 mm phosphoenolpyruvate, and 1 mm tris(2-carboxyethyl)phosphine hydrochloride)). The sample was further purified by FPLC on a MonoQ HR 10/10 column with buffer A as the running buffer and eluting with a linear gradient of buffer B (buffer A, containing 500 mm NaCl). For storage, a protease inhibitor mixture without EDTA (catalog no. P-8849; Sigma-Aldrich) was added. On average, protein yield was around 1–5 mg/liter of cell culture. The electrophoretic homogeneity of the protein preparations was assessed by SDS–PAGE using the PhastSystem (GE Healthcare).
Enzyme assays
Steady-state kinetics for AroQ
δ variants (with or without the His-tag) were performed at 30 °C in 50 m
m potassium phosphate, pH 7.5, by varying chorismate concentrations between 20 and ∼2000 μ
m. Initial rates were acquired from the initial slopes of absorption decrease curves obtained by measuring chorismate disappearance at 310 nm (ε
310 = 370
m−1 cm
−1). The protein concentration was determined by a calibrated Bradford assay using BSA as a standard (
11- Sasso S.
- Ökvist M.
- Roderer K.
- Gamper M.
- Codoni G.
- Krengel U.
- Kast P.
Structure and function of a complex between chorismate mutase and DAHP synthase: efficiency boost for the junior partner.
). The data were fitted to the Michaelis-Menten equation using KaleidaGraph (Synergy Software, Reading, PA, USA) to derive
kcat and
Km.
kcat is calculated per active site.
Fig. S6 shows the kinetic data and fitted Michaelis-Menten plots for four representative MtCM variants. Where error ranges are given, mean and standard deviations (σ
n−1) were calculated from data from at least two independently produced and isolated enzyme preparations. The kinetic characterization of the AroQ
δ variants in the presence of MtDS was performed at 274 nm as described before (
11- Sasso S.
- Ökvist M.
- Roderer K.
- Gamper M.
- Codoni G.
- Krengel U.
- Kast P.
Structure and function of a complex between chorismate mutase and DAHP synthase: efficiency boost for the junior partner.
) and is detailed in the legend to
Table S8.
CD spectroscopy
Thermal denaturation curves were determined using CD measurements at 222 nm (bandwidth 1 nm) using an Aviv 202 CD spectrometer (Aviv Biomedical, Lakewood, NJ, USA). Approximately 4 μ
m protein in 20 m
m potassium phosphate, pH 7.5, was used in quartz cuvettes with a 0.2-cm path length. After initial equilibration at 10 or 20 °C for 5 min, the sample was heated in 0.2 °C steps with 0.4-min equilibration time and 3-s signal averaging time per data point. The melting was monitored between 10 or 20 °C and 95 °C. The melting temperature (
Tm) was calculated from fitting the mean residue ellipticity (Θ) to a sigmoidal curve between 60 and 95 °C, as exemplified in
Fig. S7 with representative MtCM variants, using the equation shown in the figure's legend.
Crystallization, data processing, and structure refinement
Crystallization of N-s4.15 was performed using hanging drop setups at 20 °C and a protein concentration ranging from 5–11 mg/ml (in 20 m
m BTP, pH 7.5, and 150 m
m NaCl). N-s4.15 crystals grew in conditions containing 0.1
m 2-amino-2-(hydroxymethyl)propane-1,3-diol (Tris-HCl), pH 8.5, 0.2
m trimethylamine N-oxide, and 25% w/v PEG 2000 MME (optimization of JCSG+ crystallization screen, Molecular Dimensions). Diffraction quality crystals were obtained with microseeding. The protein solution was pre-incubated with TSA (from a stock previously synthesized by Dr. Rosalino Pulido as described (
64An improved synthesis of the transition-state analog inhibitor of chorismate mutase.
)) by adding a few flakes of the compound 30 min before crystallization setup. Prior to data collection, the crystal was soaked <5 min in mother liquor additionally containing 20% ethylene glycol and flash-cooled in a nitrogen cryo-stream (100 K).
Diffraction data were collected at beamline BM14 at the European Synchrotron Radiation Facility (ESRF, Grenoble, France). Diffraction and refinement data are summarized in
Table 2. The data were processed and scaled with
XDS (
) and merged using the program
AIMLESS (
66- Evans P.R.
- Murshudov G.N.
How good are my data and what is the resolution?.
) from the
CCP4 program suite (
67- Winn M.D.
- Ballard C.C.
- Cowtan K.D.
- Dodson E.J.
- Emsley P.
- Evans P.R.
- Keegan R.M.
- Krissinel E.B.
- Leslie A.G.W.
- McCoy A.
- McNicholas S.J.
- Murshudov G.N.
- Pannu N.S.
- Potterton E.A.
- Powell H.R.
- et al.
Overview of the CCP4 suite and current developments.
). The structure was solved by molecular replacement with the program
Phaser (
68- McCoy A.J.
- Grosse-Kunstleve R.W.
- Adams P.D.
- Winn M.D.
- Storoni L.C.
- Read R.J.
Phaser crystallographic software.
), using as a search model the WT MtCM structure (PDB ID:
2W1A, (
11- Sasso S.
- Ökvist M.
- Roderer K.
- Gamper M.
- Codoni G.
- Krengel U.
- Kast P.
Structure and function of a complex between chorismate mutase and DAHP synthase: efficiency boost for the junior partner.
)), with TSA ligand, N terminus, C terminus, and H1-H2 loop (Ala
45-Ser
57) removed. The structures were refined by alternating rounds of rigid body refinement with
REFMAC5 (
69- Murshudov G.N.
- Skubák P.
- Lebedev A.A.
- Pannu N.S.
- Steiner R.A.
- Nicholls R.A.
- Winn M.D.
- Long F.
- Vagin A.A.
REFMAC5 for the refinement of macromolecular crystal structures.
,
70- Vagin A.A.
- Steiner R.A.
- Lebedev A.A.
- Potterton L.
- McNicholas S.
- Long F.
- Murshudov G.N.
REFMAC5 dictionary: organization of prior chemical knowledge and guidelines for its use.
) and model building using
Coot (
71- Emsley P.
- Lohkamp B.
- Scott W.G.
- Cowtan K.
Features and development of Coot.
).
PHENIX (
72- Adams P.D.
- Afonine P.V.
- Bunkóczi G.
- Chen V.B.
- Davis I.W.
- Echols N.
- Headd J.J.
- Hung L.W.
- Kapral G.J.
- Grosse-Kunstleve R.W.
- McCoy A.J.
- Moriarty N.W.
- Oeffner R.
- Read R.J.
- Richardson D.C.
- et al.
PHENIX: a comprehensive Python-based system for macromolecular structure solution.
) was used to generate composite OMIT-maps to validate ambiguous regions for avoiding model bias during model building. Despite the presence of TSA in the crystallization condition, the data contained no electron density that could be attributed to this ligand, and it was therefore not modeled. N-terminal residues prior to Leu
11 and C-terminal residues Ala
89 and Met
90 also lack electron density and were thus not included in the structural model deposited in the Protein Data Bank (
73- Berman H.M.
- Westbrook J.
- Feng Z.
- Gilliland G.
- Bhat T.N.
- Weissig H.
- Shindyalov I.N.
- Bourne P.E.
The Protein Data Bank.
) with PDB ID:
5MPV.
Preparation of figures
Structure figures were created using PyMOL (version 2.3.1; Schrödinger LLC) with standard alignment parameters. We adhered to the following color scheme:
green, MtCM in the MtCM-MtDS complex with TSA (PDB ID:
2W1A (
11- Sasso S.
- Ökvist M.
- Roderer K.
- Gamper M.
- Codoni G.
- Krengel U.
- Kast P.
Structure and function of a complex between chorismate mutase and DAHP synthase: efficiency boost for the junior partner.
));
cyan, MtCM in the MtCM-MtDS complex without TSA (PDB ID:
2W19 (
11- Sasso S.
- Ökvist M.
- Roderer K.
- Gamper M.
- Codoni G.
- Krengel U.
- Kast P.
Structure and function of a complex between chorismate mutase and DAHP synthase: efficiency boost for the junior partner.
));
orange, MtCM with malate (PDB ID:
2VKL (
11- Sasso S.
- Ökvist M.
- Roderer K.
- Gamper M.
- Codoni G.
- Krengel U.
- Kast P.
Structure and function of a complex between chorismate mutase and DAHP synthase: efficiency boost for the junior partner.
));
yellow, apo MtCM (PDB ID:
2QBV (
41- Kim S.K.
- Reddy S.K.
- Nelson B.C.
- Robinson H.
- Reddy P.T.
- Ladner J.E.
A comparative biochemical and structural analysis of the intracellular chorismate mutase (Rv0948c) from Mycobacterium tuberculosis H37Rv and the secreted chorismate mutase (y2828) from Yersinia pestis.
)); and
purple/pink, autonomous MtCM variant N-s4.15 (PDB ID:
5MPV, this work).
Article info
Publication history
Published online: December 18, 2020
Received in revised form:
October 8,
2020
Received:
June 22,
2020
Edited by Joseph M. Jez
Footnotes
This article contains supporting information.
Author contributions—J. F.-K. and K. W.-R. data curation; J. F.-K., K. W.-R., H. V. T., and U. K. validation; J. F.-K., K. W.-R., H. V. T., and S. M. investigation; J. F.-K., K. W.-R., and H. V. T. visualization; J. F.-K., K. W.-R., and S. M. methodology; J. F.-K. writing-original draft; J. F.-K., K. W.-R., H. V. T., S. M., U. K., and P. K. writing-review and editing; U. K. and P. K. supervision; U. K. and P. K. funding acquisition; P. K. conceptualization.
Funding and additional information—This study was financed by funds from the University of Oslo (position of H. V. T.), the ETH Zurich (to P. K.), the Norwegian Research Council Grants 216625 and 247730 (to U. K.), and the Swiss National Science Foundation Grants 31003A-116475, 31003A-135651, 31003A-156453, and 310030M_182648 (to P. K.).
Conflict of interest—The authors declare that they have no conflicts of interest with the contents of this article.
Abbreviations—The abbreviations used are: BTP
1,3-bis[tris(hydroxymethyl)methylamino]propane
CMchorismate mutase
DAHP3-deoxy-d-arabino-heptulosonate 7-phosphate
DSDAHP synthase
EcCMCM domain of the bifunctional CM-prephenate dehydratase from Escherichia coli
epPCRerror-prone PCR
MtCMintracellular CM from M. tuberculosis encoded by Rv0948c, aroQδ
*MtCMsecreted CM from M. tuberculosis
MtDSDS from M. tuberculosis
pFPhedl-para-fluoro-phenylalanine
ScCMCM from Saccharomyces cerevisiae
Tettetracycline
Tmmelting temperature
TSAendo-oxa-bicyclic transition state analog of the CM reaction
s11re4.7s11
s4.15s4.15s10es4.15
epLepPCR libraries
ESRFEuropean Synchrotron Radiation Facility
PDHprephenate dehydrogenase
PDTprephenate dehydratase.
Copyright
© 2020 Fahrig-Kamarauskait? et al.