Next Article in Journal
Assessment of Platinum Catalyst in Rice Husk Combustion: A Comparative Life Cycle Analysis with Conventional Methods
Previous Article in Journal
Recent Progress of 2D Pt-Group Metallic Electrocatalysts for Energy-Conversion Applications
Previous Article in Special Issue
A Recurring Misconception Regarding the Fitting and Plotting of Enzyme Kinetics Data Leads to the Loss of Significant Reaction Parameters and Rate Constants
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Multi-Enzyme Synergy and Allosteric Regulation in the Shikimate Pathway: Biocatalytic Platforms for Industrial Applications

1
Department of Chemistry, Pennsylvania State University, University Park, PA 16802, USA
2
Department of Chemistry, COMSATS University Islamabad, Abbottabad Campus, Abbottabad, KPK 22010, Pakistan
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(8), 718; https://doi.org/10.3390/catal15080718
Submission received: 6 July 2025 / Revised: 24 July 2025 / Accepted: 26 July 2025 / Published: 28 July 2025

Abstract

The shikimate pathway is the fundamental metabolic route for aromatic amino acid biosynthesis in bacteria, plants, and fungi, but is absent in mammals. This review explores how multi-enzyme synergy and allosteric regulation coordinate metabolic flux through this pathway by focusing on three key enzymes: 3-deoxy-d-arabino-heptulosonate-7-phosphate synthase, chorismate mutase, and tryptophan synthase. We examine the structural diversity and distribution of these enzymes across evolutionary domains, highlighting conserved catalytic mechanisms alongside species-specific regulatory adaptations. The review covers directed evolution strategies that have transformed naturally regulated enzymes into standalone biocatalysts with enhanced activity and expanded substrate scope, enabling synthesis of non-canonical amino acids and complex organic molecules. Industrial applications demonstrate the pathway’s potential for sustainable production of pharmaceuticals, polymer precursors, and specialty chemicals through engineered microbial platforms. Additionally, we discuss the therapeutic potential of inhibitors targeting pathogenic organisms, particularly their mechanisms of action and antimicrobial efficacy. This comprehensive review establishes the shikimate pathway as a paradigmatic system where understanding allosteric networks enables the rational design of biocatalytic platforms, providing blueprints for biotechnological innovation and demonstrating how evolutionary constraints can be overcome through protein engineering to create superior industrial biocatalysts.

1. Introduction

The shikimate pathway, a critical metabolic route for bacterial, plant, and fungal survival but absent in humans, has become a central subject for both basic biochemical research and industrial advancement [1]. This pathway links primary metabolism to the production of aromatic amino acids and numerous secondary metabolites and enables biological synthesis of pharmaceuticals and sustainable materials. Its operation relies on sophisticated coordination: enzymes function as integrated teams to direct substrate flow efficiently, while allosteric regulation dynamically modulates activity based on cellular demands. These characteristics establish the pathway as a valuable resource for understanding biological regulation as well as for the development of therapeutics, engineered biocatalysts, and green chemistry.
The shikimate pathway’s industrial promise lies in its versatility. Its enzymes have been repurposed for cost effective production of high-value chemicals, such as shikimic acid (a Tamiflu precursor) and muconic acid (a building block for nylon) [2]. Engineered microbes show increasing promise for industrial biotechnology, with recent advances enabling improved production capabilities, though most targets have not yet reached commercial viability [3,4]. Meanwhile, computational tools map allosteric hotspots and predict resistance mutations, guiding the design of smarter drugs and robust enzymes [5]. Together, these advances position the pathway as a testbed for the converging disciplines of systems biology, synthetic biology, and metabolic engineering, to help address global challenges in healthcare, agriculture, and sustainability.
This review explores how multi-enzyme synergy and allostery regulates the pathway, from entry-point control at 3-deoxy-d-arabino-heptulosonate-7-phosphate synthase (DAHPS), through critical branch-point regulation at chorismate mutase (CM), to the sophisticated substrate channeling mechanisms in the tryptophan synthase (TrpS) complex (Figure 1). This analysis uniquely bridges fundamental enzyme structure–allosteric function relationships with cutting-edge engineering strategies, demonstrating how mechanistic insights directly translate into biotechnological innovations and therapeutic interventions. We highlight insights gathered from structural biology, enzyme engineering, and industrial microbiology that can be leveraged towards improving enzymes in this pathway towards bio-based chemical production, as well as how these enzymes might be targeted in the development of new antimicrobials.

The Shikimate Pathway: A Central Hub for Aromatic Metabolism

The shikimate pathway is a foundational metabolic route conserved in plants, bacteria, fungi, and algae, serving as the primary source of aromatic amino acids (phenylalanine, tyrosine, and tryptophan) (Figure 2) and serves as a gateway to diverse secondary metabolites [6]. By channeling carbon from central metabolism via phosphoenolpyruvate (PEP) and erythrose-4-phosphate (E4P) into chorismate, this pathway bridges primary metabolism with the biosynthesis of compounds critical for growth, defense, and communication [1].
Among the core enzymes of the shikimate pathway, DAHPS, CM, and TrpS emerge as focal points for drug discovery and industrial biotechnology due to their regulatory complexity and metabolic versatility. DAHPS catalyzes the committed step by condensing PEP and E4P into DAHP, with tight feedback inhibition from phenylalanine and tyrosine [1]. CM functions as a pivotal branch-point enzyme, isomerizing chorismate to prephenate for phenylalanine and tyrosine biosynthesis, with NMR studies revealing how tyrosine “freezes” the enzyme while tryptophan “unlocks” catalysis, ensuring balanced amino acid synthesis [7,8]. The TrpS bienzyme complex catalyzes the final steps for tryptophan biosynthesis, where the α-subunit cleaves indole-3-glycerol phosphate (IGP) and the β-subunit condenses the released indole with serine via pyridoxal 5′-phosphate (PLP)-dependent mechanisms. A key control mechanism is the shuttling of indole through a hydrophobic channel, which requires allosteric communication to synchronize the α and β catalytic cycles [9].
These enzymes have garnered disproportionate research attention because CM’s physical association with DAHPS in pathogens like Mycobacterium tuberculosis (Mtb) creates allosterically regulated complexes controlling biosynthesis of aromatic amino acids [10], while TrpS’s capacity to generate noncanonical amino acids (ncAAs) through engineering has propelled pharmaceutical and agrochemical applications [11]. DAHPS has been engineered for microbial shikimate overproduction [5], and other pathway enzymes like 5-enolpyruvylshikimate-3-phosphate synthase and shikimate kinase remain essential for flux, but their therapeutic and industrial exploration lags behind CM and TrpS. Understanding the biochemical and biophysical mechanisms of enzymatic function and allosteric regulation helps to set the stage for subsequent discussions on pathway engineering and metabolic bottlenecks.

2. Complex Regulation of DAHPS, Which Catalyzes the First Committed Step in Aromatic Amino Acid Biosynthesis

2.1. Evolutionary Distribution, Structural Classification, and Functional Significance of DAHPS

DAHPS (EC 2.5.1.54) represents the first and rate-limiting enzyme of the shikimate pathway [12] with its catalytic activity directly influencing the production of aromatic amino acids and downstream aromatic metabolites including phenolic compounds, alkaloids, and lignin precursors [13].
DAHPS is structurally diverse across different organisms, reflecting evolutionary adaptations to specific regulatory requirements and cellular environments. The enzyme is broadly classified into several structural types based on domain organization and regulatory mechanisms (Table 1). DAHPS enzymes are categorized into two types: Type I (<40 kDa) and Type II (>50 kDa), with Type I further divided into Iα/Iβ subfamilies (Figure 3). Despite less than 10% sequence similarity, both types share conserved structural features including the TIM (triose phosphate isomerase) barrel fold, active site architecture, and metal ion coordination, indicating common ancestry. Type II enzymes are exemplified by the Mycobacterium tuberculosis (Mtb) DAHPS [14,15]. The quaternary structure of DAHPS consists of two tightly bound dimers, implying that DAHPS is a tetramer although significant variations exist in the specific dimer–dimer interfaces and subunit orientations between different organisms. The tetrameric arrangement is critical for allosteric regulation, as allosteric sites are near subunit interfaces. The X-ray crystal structures of different classes of DAHPS have revealed that conformational changes initiated at allosteric binding sites propagate through the quaternary structure to modulate active site geometry and catalytic efficiency [16].
Bacterial systems typically express multiple DAHPS isoforms, each subject to feedback inhibition by different aromatic amino acids [16]. In Escherichia coli, three distinct isoforms exist: AroF (tyrosine-sensitive), AroG (phenylalanine-sensitive), and AroH (tryptophan-sensitive), each exhibiting unique structural features and regulatory properties [17,18,19]. In Bacillus subtilis, DAHPS exists as part of a bifunctional enzyme containing both DAHPS and CM activities within a single polypeptide chain [20]. The N-terminal domain belongs to the AroQ class of CM and exhibits conformational plasticity that allows accommodation of various substrate analogs and inhibitors.
Unlike bacterial enzymes, plant DAHPSs are subject to redox regulation through reversible disulfide bond formation, allowing coordination with photosynthetic electron transport chains. Plant enzymes typically exist as homodimers or tetramers with molecular weights ranging from 80–160 kDa, significantly larger than their bacterial counterparts. The increased size reflects additional regulatory domains and structural elements required for integration with plant-specific metabolic networks. The Arabidopsis thaliana enzyme contains an N-terminal signal sequence for plastid import and undergoes post-translational processing to generate the mature, catalytically active form [21,22,23].
Archaeal DAHPS represents an evolutionary intermediate between bacterial and eukaryotic forms, exhibiting structural features characteristic of both domains. Thermophilic bacteria such as Thermotoga maritima possess thermally stable DAHPS with enhanced thermal stability [24,25]. Fungal DAHPSs, exemplified by Neurospora crassa, and certain bacterial species like Streptomyces rimosus, belong to distinct structural classes characterized by larger subunit molecular weights and unique regulatory properties, often exhibiting tryptophan sensitivity and possessing structural features distinct from typical bacterial and plant counterparts [26].

2.2. General Catalytic Mechanism and Active Site Architecture of DAHPS

The catalytic mechanism of DAHPS involves a complex series of chemical transformations requiring precise coordination of both substrates within the active site [27]. The active site architecture is highly conserved across different organisms, featuring essential catalytic residues and metal binding sites that facilitate substrate positioning and chemical transformation [14]. However, subtle variations in active site geometry and accessibility contribute to the observed differences in substrate affinity and catalytic efficiency between different structural classes.

2.3. Regulatory Mechanisms and Allosteric Control in DAHPS Enzymes

As a critical regulatory node controlling carbon flux into aromatic amino acid biosynthesis, DAHPS has evolved diverse regulatory mechanisms that reflect the physiological complexity of different organisms (Table 1). DAHPS exhibits at least four allosteric regulatory mechanisms. In bacterial systems, DAHPS regulation primarily relies on allosteric feedback inhibition by aromatic amino acid end products, with the allosteric pattern of control being strongly conserved among member species within a given genus [12]. The second most intricate bacterial regulatory systems are exemplified by Mtb DAHPS, where the enzyme demonstrates sophisticated inter-enzyme allostery through direct interaction with CM. In this system, DAHPS is inhibited by phenylalanine and tyrosine, while tryptophan indirectly activates CM through inter-enzyme allostery, effectively redirecting metabolic flux toward tryptophan biosynthesis. Structural studies on the Mtb DAHPS have revealed the molecular basis of this intricate allostery, with d-amino acid studies identifying specific residues critical for signal transduction [28]. Complex formation between CM and DAHPS represents an example of molecular symbiosis where inter-enzyme allostery creates a novel regulatory paradigm that modifies the allosteric properties of both enzymes and provides sophisticated control over shikimate pathway flux [29]. DAHPS, in Corynebacterium glutamicum, activates CM (CgCM) through inter-enzyme allostery, redirecting flux toward TrpS biosynthesis [30].
Another remarkable example of bacterial DAHPS regulation is found in Geobacillus sp., where evolution has produced a fused DAHPS-CM enzyme that utilizes prephenate, the product of CM, as an intramolecular allosteric inhibitor of DAHPS. [31]. Additional bacterial diversity is seen in systems like Pseudomonas chlororaphis, where PhzC functions as an essential DAHPS variant, and T. maritima, where structural studies have refined understanding of catalytic mechanisms and revealed novel allosteric regulatory features [25,32]. The evolutionary progression of these regulatory systems is evident in the late emergence of DAHPS-Phe, a regulatory isozyme subject to allosteric control by l-phenylalanine, which represents the addition to the isozyme family and is found only within the enteric bacterial lineage. This evolutionary pattern suggests that increasingly sophisticated regulatory mechanisms developed as metabolic networks became more complex [16,33].
Plant DAHPS regulation exhibits significantly greater complexity, incorporating additional regulatory layers that reflect the more sophisticated physiological demands of multicellular photosynthetic organisms. The most distinctive feature of plant DAHPS regulation is redox control through cysteine residues, particularly evident in Arabidopsis, where the enzyme appears to be regulated by the ferredoxin/thioredoxin (TRX) redox control system of the chloroplast [23]. This redox regulation allows plants to coordinate aromatic amino acid biosynthesis with photosynthetic activity and cellular redox status, providing a direct link between energy metabolism and secondary metabolite production. The integration of these regulatory mechanisms positions DAHPS as the gatekeeper of plant aromatic natural product biosynthesis with the regulatory systems that integrate metabolic flux with environmental conditions [34].

2.4. Enzyme Engineering of DAHPS

DAHPS has been extensively engineered to overcome feedback inhibition to increase metabolic flux. Removal of this metabolic ‘brake’ helps to redirect cellular resources toward maximum product formation rather than balanced homeostasis, enabling the higher titers required for industrial biotechnology applications [35]. Engineered DAHPS variants with reduced feedback inhibition have been successfully implemented in modular metabolic engineering strategies for enhanced aromatic amino acid production. For example, mutation in yeast DAHPSD154N works distally from inhibitor binding sites by reducing binding affinity for Phe, thereby improving the metabolic yield of various aromatic compound (tyrosol and salidroside) [36].
Recent structural advances have revealed the molecular basis of allosteric inhibition in DAHPS. X-Ray diffraction studies demonstrate that feedback inhibition involves conformational changes in key residues (Pro148, Gln152, Ile213 in E. coli) altering substrate coordination [19]. Plant DAHPS engineering has revealed unique regulatory mechanisms distinct from bacterial systems. Arabidopsis DAHPS isozymes demonstrate cytosolic retention mediated by 14-3-3 protein interactions under elevated tyrosine conditions [22]. This discovery has opened new avenues for engineering subcellular localization to enhance pathway flux in plant systems.

2.5. Shikimate Pathway Engineering Involving DAHPS

DAHPS pathway engineering has achieved remarkable production titers through systematic metabolic rewiring (Table 1). E. coli systems optimized for precursor balancing (E4P/PEP) achieved 101 g/L shikimate, the highest reported titer via strategic gene deletions (shiA, tyrR) and enhanced precursor availability [3]. The success stems from coordinated overexpression of feedback-resistant aroGD146N variants in B. subtilis combined with systematic removal of competing pathways, paving the way towards the production of menaquinone-7 [37]. Corynebacterium glutamicum has emerged as a superior platform for DAHPS-based production, achieving 37 g/L shikimate through knockout of bypass genes (qsuB, pyk1) and strategic overexpression of rate-limiting enzymes [38]. The organism’s natural tolerance to aromatic compounds and robust central metabolism make it ideal for high-titer production. Advanced bioprocess optimization, including growth-arrested cell reactions, has pushed titers to 141 g/L [39]. Cyanobacterial platforms represent the frontier of sustainable production, with Synechocystis sp. PCC 6803 engineered to redirect more than 30% of fixed CO2 toward the shikimate pathway. The introduction of feedback-resistant DAHPS variants combined with pathway optimization enables carbon-neutral biosynthesis, producing 138 mg/L trans-cinnamic acid under photoautotrophic conditions [40]. Antibiotic production has benefited significantly from shikimate pathway engineering. Amycolatopsis balhimycina engineered with overexpressed DAHPS increased balhimycin titers >4-fold, demonstrating the pathway’s potential in glycopeptide antibiotic production [41].

2.6. Targeting DAHPS in Antimicrobial Drug Discovery

DAHPS has been exploited as an antimicrobial target due to its essential role as the gateway enzyme of the shikimate pathway, which is absent in mammals but critical for bacterial survival and virulence. Since DAHPS is required to produce essential aromatic amino acids, its inhibition prevents sufficient protein synthesis resources, effectively controlling or slowing bacterial growth. It is of note that the DAHPS substrate binding site was previously characterized as “undruggable” due to its high content of charged residues and extremely high overall polarity [42]. However, recent computational screening of phosphoenolpyruvate derivatives as dual inhibitors of DAHPS and 5-enolpyruvylshikimate-3-phosphate (EPSP) synthase have successfully identified nanomolar affinity candidates, with molecular dynamics (MD) simulations and binding free energy calculations revealing consistent binding modes and pharmacophoric properties suitable for multi-target inhibition. This suggests that the perceived undruggability of DAHPS may be overcome through innovative computational approaches [5]. The cyanobacterial antimetabolite 7-deoxy-sedoheptulose has demonstrated the therapeutic potential of targeting DAHPS, showing antimicrobial and herbicidal activity in the low micromolar range by inhibiting 3-dehydroquinate synthase and causing accumulation of toxic pathway intermediates [43]. Structure-based drug design approaches have been facilitated by detailed molecular modeling studies of DAHPS from Pseudomonas fragi, revealing conserved sequence motifs and providing molecular insights beneficial for rational inhibitor design against broad-spectrum bacterial targets [44].
Table 1. Structure, allosteric regulation, and engineering of DAHPS.
Table 1. Structure, allosteric regulation, and engineering of DAHPS.
OrganismStructural ClassificationAllosteric RegulationEngineering EffortsApplicationsRef.
E. coliType I;
Isoforms:
AroF
AroG
AroH
AroF: Tyr (−)
AroG: Phe (−)
AroH: Trp (−)
shikimate production (yield 101 g/L)
Precursor balancing (E4P/PEP)
[3,17,18,19]
B. subtilisType II + CM fusion (AroH/AroQ)CM substrate/product (−)aroGD146N: Metabolic EngineeringMenaquinone-7 production (yield 281.4 mg/L)[20,37]
M. tuberculosisType II + CM
Large subunit, Tetrameric
Inter-enzyme Allostery
Phe/Tyr (−)
Trp (+)
d-amino acid resistance studies
Allosteric hotspot mapping
antimicrobial target[14,15,28]
C. glutamicumType II variantDAHPS activates CM through inter-enzyme allosteryBypass gene knockouts (qsuB, pyk1)
Coordinated pathway engineering
Shikimate production (yield 37 g/L) with bioprocess optimization (yield 141 g/L)[35,38,39]
T. maritimaType Iβ thermostableN-terminal ferredoxin-like (FL) regulatory domain.Thermostability engineering; enhanced thermal adaptationProposed thermophilic bioprocessing[24,25]
Geobacillus sp.Type Iβ + CM fusionIntramolecular control
DAHPS Phe (−)
Coordinated with CM activity (+)
Domain interface engineering
Sterically controlled regulation
[31]
A. thaliana Ferredoxin/thioredoxin-mediated control
Thioredoxin reduction (+)
DAHPSD154N Reduced inhibitor binding affinityPlant metabolic engineering; enhanced secondary metabolite production[21,22,23]
Synechocystis sp. PCC 6803Cyanobacterial Type IEngineered light-dependent control>30% CO2 redirection to shikimate pathwayTrans-cinnamic acid production (yield 138 mg/L)
Carbon-neutral biosynthesis
[40]

3. Structural Diversity of the Branch-Point Enzyme Chorismate Mutase Provides Insights into Allosteric Regulation and Enzyme Engineering

CM (EC 5.4.99.5) catalyzes the intramolecular Claisen rearrangement of chorismate to prephenate in the phenylalanine/tyrosine branch of the shikimate pathway. The enzyme’s evolutionary significance is underscored by its presence across all three domains of life where aromatic amino acid biosynthesis occurs: bacteria, archaea, and eukaryotes (plants and fungi) [6,45]. What makes CM particularly intriguing from a structural biology perspective is the remarkable diversity of protein folds that have evolved to catalyze the same chemical transformation [46]. This structural plasticity, coupled with varying regulatory mechanisms across species, provides an excellent case study for understanding enzyme evolution, structure–function relationships, and the molecular basis of metabolic regulation [47,48].

3.1. Structural Classification and Phylogenetic Distribution of CM Enzymes

CMs can be broadly classified into three major structural families based on their protein fold architectures [45] (Figure 4, Table 2). The AroH Family (α/β-barrel) represents a structurally distinct class of CM featuring trimeric quaternary structure with a pseudo-α/β-barrel fold wherein β-sheets form the core and α-helices are positioned on the exterior of the trimer. The active sites are located between helices. This family is best exemplified by the B. subtilis CM structure [49]. The AroQ family represents the most prevalent structural class of CMs, characterized by an all-α-helical architecture. This family is further partitioned into distinct subclasses [46,49]. The AroQα subclass is comprised of ~110 amino acid structures that function as homodimers, featuring a characteristic three helix bundle per monomer. Proteins from this class form an intertwined homodimer with the active site located at the dimer interface, and this is predominantly found in bacteria and some archaea. The second subclass is AroQβ, which contains larger ~255 amino acid structures functioning as dimers. Protein from this subclass maintain the all-α-helical bundle architecture but with additional structural elements [50,51]. This structural fold exhibits improved thermostability, which may be a reason that it is represented primarily in thermophilic organisms [52].

3.2. Conserved Catalytic Mechanisms of CM Enzymes Across Structural Families

Despite their structural diversity, all CMs catalyze the same pericyclic reaction through similar mechanistic principles. Some common catalytic features include formation of a hydrophobic active site that excludes water, with precise positioning of substrate through specific protein–substrate interactions, followed by the stabilization of the transition state through electrostatic interactions. A conservation of key catalytic residues, despite sequence divergence, is also observed. The reaction in CM proceeds through an oxabicyclic transition state. Active site architecture has been proposed to help pre-organize the substrate to achieve a “chair-like” transition state (Figure 5). Arginine residues provide crucial electrostatic stabilization whereas hydrophobic residues create the necessary environment for the pericyclic reaction [46].

3.3. Allosteric Regulation in CM as a Metabolic Control Mechanism

CMs demonstrate significant diversity in allosteric regulatory mechanisms [48,53] (Table 2). Bacterial CMs can exhibit minimal regulation, specifically, the Mtb CM forms a regulatory complex with DAHPS, where prephenate acts as a dissociable allosteric inhibitor to check flux when aromatic amino acids are abundant [10,54]. Conversely, fungal CMs, like that in Saccharomyces cerevisiae (ScCM), exhibit complex regulation with multiple allosteric effectors. ScCM adopts a proposed two-state equilibrium between more active R and less active T conformations with tryptophan and tyrosine binding stabilizing the R and T conformations, respectively [7,8]. Plant CMs further demonstrate tissue specific and pathogen responsive allosteric regulation [47]. Arabidopsis AtCM1 employs bidirectional allostery wherein tryptophan activates the enzyme, and phenylalanine and tyrosine inhibit it, ensuring proportional allocation of chorismite [55]. This bidirectional regulation ensures balanced phenylalanine/tyrosine and tryptophan synthesis under varying cellular demands.

3.4. Directed Evolution of CM Enzymes

The most extensively studied bacterial system is Mtb CM. Würth-Roderer and colleagues achieved a remarkable enhancement in catalytic efficiency through iterative directed evolution of Mtb CM. Building upon these foundational studies, Thorbjørnsrud and colleagues focused on understanding the structural mechanisms underlying this enhancement, specifically examining the effects of CMT52P/V55D mutations that stabilize the catalytic loops [54,56]. The resulting improvements not only showcase the potential of directed evolution but also provide valuable insights into the structural determinants of catalytic enhancement in CM.
Russ and colleagues leveraged comprehensive evolutionary sequence data to design de novo CM enzymes that exhibit natural-like catalytic properties, demonstrating that evolutionary constraints can be effectively translated into design templates. Their approach represents a paradigm shift from purely random mutagenesis toward evolution informed design, where natural sequence variation guides exploration of the functional sequence space. The success of these methods suggests that laboratory evolution can access beneficial mutations that may be difficult to achieve through purely rational design approaches, particularly when combined with computational methods that can predict and prioritize mutations based on evolutionary likelihood [46].
The scope of CM engineering extends beyond simple catalytic enhancement to well-crafted allosteric control mechanisms. An extensive study demonstrated innovative approaches to enzyme regulation by introducing ncAAs into CM, effectively reengineering protein interfaces to achieve controlled dimerization and catalytic activity. This approach leverages genetic code expansion techniques and strategic ncAAs placement at protein interfaces, where dimerization-dependent enzyme activation can be precisely controlled [57].
While much of the directed evolution work on CM has focused on bacterial systems, recent advances in understanding the structural dynamics and catalytic mechanisms of CM provide a foundation for extending directed evolution approaches to plant and fungal systems, where the enzyme plays equally critical roles in aromatic amino acid biosynthesis [46,47].

3.5. Engineering CM: From Protein Design to Industrial Applications

Enzyme engineering efforts have targeted CM as a critical bottleneck in aromatic amino acid biosynthesis, spanning molecular protein design to complete pathway optimization (Table 2). Multi-isoform regulation systems in plants provide sophisticated templates for industrial applications. AtCM1 demonstrates complex allosteric behavior, being activated by tryptophan while inhibited by phenylalanine and tyrosine, while AtCM3 shows broader effector specificity including cysteine and histidine activation. The study authors also identified that AtCM2 is a non-allosteric form, and through X-ray crystal structure analysis and site-directed mutagenesis, they revealed key residues responsible for differential effector regulation with Gly213 in AtCM1 controlling the presence or absence of allosteric regulation, and Gly149 in AtCM1 versus Asp132 in AtCM3 controlling amino acid effector specificity [55].
Industrial applications have achieved significant production milestones through strategic pathway rewiring. Engineered Corynebacterium glutamicum strains expressing feedback-resistant CS and chorismate lyase (ubiC) produced 36.6 g/L 4-hydroxybenzoic acid with 47.8% C-mol yield, a polymer precursor with substantial commercial value [58]. The deregulated DAHPS (aroGfbr) combined with chorismate lyase enables efficient flux redirection toward industrially relevant compounds. S. cerevisiae engineering with modular overexpression of DAHPS and CM achieved 26.5 g/L salidroside and 9.9 g/L tyrosol, high-value nutraceuticals with applications in cosmetics and functional foods [59]. The multi-modular approach demonstrates the synergy between DAHPS and CM engineering for maximizing pathway flux toward valuable end products. Production of l-tyrosine represents a mature industrial application, with E. coli strains expressing feedback inhibition-insensitive TyrC and CM domain (PheACM) achieving 3 g/L l-tyrosine with 66 mg/g glucose yield [60]. These values represent 46–48% improvement over single-enzyme approaches, highlighting the importance of multi-enzyme optimization. Similarly, Streptomyces venezuelae with overexpressed aroB and aroK elevated chorismate flux, increasing chloramphenicol yields 4.3-fold [61].

3.6. Allostery in Antimicrobial Therapeutics

The metabolic importance of CM also makes it a prime candidate for targeted inhibition in antimicrobial drug development. Recent drug discovery campaigns have yielded diverse chemotypes designed to exploit these structural features and disrupt CM activity in bacterial pathogens. Reddy et al. designed pyrazolo/benzo-triazinones that inhibit CM with IC50 values of 0.4–0.9 μM, leveraging hydrophobic interactions with conserved residues [62,63]. Additionally, the natural product, abscisic acid, repurposed as a fungal CM inhibitor, blocks phenylpyruvate synthesis, suppressing growth [64]. Particularly promising developments have emerged in anti-tubercular drug discovery. Shukla et al. designed pyrazolo pyrimidinones (e.g., cyclopentyl-substituted I) as Mtb CM inhibitors via Wang resin catalyzed sonochemical synthesis. These compounds reduced Mtb viability by >20% at 30 μM in Alamar Blue assays while sparing mammalian cells, highlighting their therapeutic potential [65]. In another study, Kumar et al. synthesized Mtb CM-targeting quinazolin-4 (1H)-ones via eco-friendly Wang-OSO3H catalysis. Top derivatives (e.g., 2-(3,4,5-trimethoxyphenyl)-quinazolin-4 (3H)-one; IC50 = 11.43 μM) interacted with catalytic residues, highlighting aryl substituents as being critical for potency [66]. These collective findings underscore the significant therapeutic potential of CM-targeted antimicrobial agents and provide valuable structural insights for continued drug development efforts against antibiotic resistant pathogens.
Table 2. Structure, allosteric regulation, and engineering of CM.
Table 2. Structure, allosteric regulation, and engineering of CM.
OrganismStructural Classification Allosteric RegulationEngineering EffortsApplications Ref.
E. coli AroQα Multi-enzyme optimization-TyrC + CM domain (PheACM)l-Tyrosine production (3 g/L, 66 mg/g glucose, 46–48% improvement)[60]
M. tuberculosisAroQδ
homodimer with DAHPS complex
Phe (−)Directed evolution-T52P/V55D loop stabilization, 12-fold catalytic enhancementHyperactive variants; validated as a drug target [54,56,65]
S. cerevisiaeAroQ
dimer
Trp (+), Tyr (−)Allosteric reengineering-Interface modificationsControllable dimerization, programmable regulation[7,8,46,53]
S. cerevisiae
(pathway engineered)
Multi-modular systems-ARO7 overexpression + pathway integrationProduction of salidroside (yield 26.5 g/L), tyrosol (yield 9.9 g/L) nutraceuticals[59]
Arabidopsis thalianaAtCM1-3 Complex multi-domain architectureTrp (+), Phe/Tyr (−) (AtCM1)Multi-isoform analysis-AtCM1, AtCM2, AtCM3 characterizationControllable biocatalytic systems[55]
C. glutamicumAroQα
Homodimer
Trp bound DAHPS (+)feedback-resistant CS + chorismate lyase (ubiC)Amino acid production;
4-Hydroxybenzoic acid production (36.6 g/L, 47.8% C-mol yield)
[30,58]

4. Engineering of the Channeling Enzyme Tryptophan Synthase Provides Catalytic Utility

TrpS represents one of the most extensively studied enzyme complexes in biochemistry, serving as a paradigm for understanding allosteric regulation, substrate channeling, and coordinated catalysis in multi-enzyme complexes [67]. TrpS exists as an α2β2 tetrameric complex in most organisms It is composed of two α-subunits (Trpα) and two β-subunits (Trpβ) arranged in a linear α-β-β-α configuration [68]. The α-subunit adopts a TIM barrel fold characteristic of many glycolytic enzymes. The β-subunit, adopts a canonical TIM barrel fold architecture but exhibits significant structural elaboration through supplementary domains, most notably a characteristic insertion domain that functions as a determinant of substrate recognition specificity and serves as a conduit for long-range allosteric signal transduction within the enzyme complex (Figure 6) [69].

4.1. Catalytic Mechanism and Indole Channeling in Tryptophan Synthase

The TrpS enzyme catalyzes the final steps of tryptophan biosynthesis through two coordinated reactions [9]. The α-subunit catalyzes the aldol cleavage IGP to indole and G3P [70]. The β-subunit employs PLP as a cofactor to catalyze the condensation of indole with l-serine to form l-tryptophan [71,72]. This reaction involves multiple intermediates including aldimine, quinonoid, and carbanionic species [73]. The coordination between the α and β catalytic cycles is essential for efficient tryptophan synthesis, as the volatile indole intermediate must be rapidly transferred from the α-subunit to the β-subunit without being lost into the cellular environment.
Indole channeling in TrpS involves several key components including the hydrophobic tunnel, together with the conformational gating and well-coordinated active site dynamics [74,75]. Recent structural studies have demonstrated that the tunnel’s architecture changes dynamically during the catalytic cycle, opening and closing in response to substrate binding and product formation [9,72].

4.2. Evolutionary Distribution and Regulation of TrpS Enzymes

There is remarkable conservation exhibited by TrpS across diverse organisms, reflecting its fundamental and historical importance in amino acid biosynthesis. Phylogenetic analysis reveals that the enzyme complex arose early in the evolution, with separate α- and β-subunits that subsequently underwent coordinated function through gene duplication and specialization events. [76].
TrpS activity is subject to multiple levels of regulation that integrate tryptophan biosynthesis with broader cellular metabolism. Transcriptional regulation of the pathway genes responds to tryptophan availability via an attenuation mechanism and repressor proteins that sense cellular amino acid levels. The trp operon in bacteria represents a classic example of coordinated gene regulation, where tryptophan availability controls expression of all enzymes in biosynthetic pathway [70].
Post-translational regulation involves feedback inhibition by the tryptophan and its metabolites, preventing overproduction when the amino acid levels are sufficient. The enzyme also participates in the metabolic channeling networks that coordinate tryptophan synthesis with other biosynthetic pathways, including nucleotide synthesis, vitamin production, and stress response pathways ensuring that tryptophan production matches cellular demands while avoiding wasteful overproduction [77].

4.3. Allosteric Regulation and Communication Networks in TrpS

Allosteric communication between Trpα and Trpβ operates through multiple pathways including direct protein–protein contacts at the α–β interface, hydrogen bonding networks, and dynamic coupling through the shared tunnel structure. Binding of substrates to either active site induces conformational changes that are propagated to the partner subunit, modulating its catalytic activity and its own substrate affinity. These allosteric effects are reciprocal, with changes in the α-subunit affecting β-subunit activity and vice versa. The communication network involves specific residues that undergo coordinated movements, including loop regions that open and close active sites, and side chains that reorient to optimize catalytic geometry [78,79].

4.4. Directed Evolution of TrpS: From Allosteric Regulation to Industrial Biocatalysis

The directed evolution of TrpS represents an example of how enzymatic constraints imposed by natural allosteric regulation can be systematically overcome through protein engineering (Table 3). While indole channeling is critical to biological function, the intricate regulatory apparatus underlying this mechanism presented significant limitations for biotechnological applications requiring standalone enzyme function and expanded substrate scope.

4.5. Engineering the α-Subunit: Decoupling Allosteric Networks

The α-subunit of TrpS exists in a ligand-mediated equilibrium between open (inactive) and closed (active) conformations, with catalytic efficiency being dramatically enhanced through allosteric activation by the β-subunit [80]. The challenge in α-subunit engineering lies in stabilizing the catalytically competent closed conformation without β-subunit interaction. Advances in understanding long-range allosteric networks within the α-subunit have revealed that intrinsic amino acid networks can be modulated independently of β-subunit binding [81]. Rational engineering approaches targeting key residues involved in allosteric communication networks could potentially decouple alpha-subunit activity from its beta-subunit. Surface exposed residues involved in allosteric networks have shown particular promise with the substitutions enhancing TrpS function in cellular environments via stabilizing active conformations [82].

4.6. Recapitulating Allosteric Activation in Stand-Alone β-Subunit

The β-subunit of TrpS exhibits reduced catalytic activity due to the absence of allosteric activation normally provided by the α-subunit. The groundbreaking work by Buller et al. [83] demonstrated that directed evolution could recapitulate this allosteric activation through mutations that stabilize the catalytically active conformation independent of α-subunit interaction (Table 3). Building upon this foundation, Herger engineered PfTrpβ to accept non-natural substrates such as l-threonine, achieving over 1000-fold enhancement in activity for β-methyltryptophan synthesis. This study showed that the substrate range can be broadened using engineered biocatalysts capable of synthesizing complex molecules from basic precursors, providing a straightforward and scalable approach for producing β-methyl ncAA analogues [84].
Romney et al. further extended this approach by applying directed evolution to Trpβ from P. furiosus and T. maritima to generate a suite of catalysts for the synthesis of previously intractable tryptophan analogues. For the most challenging substrates, such as nitroindoles, the key to improving activity lay in the mutation of a universally conserved and mechanistically important residue, E104, unlocking reactivity toward bulky or electron-deficient indoles that were poorly accepted by wild-type enzymes [85]. The landmark achievement was the engineering of P. furiosus - PfTrpSβ into a standalone catalyst, PfTrpSβ7E6, which represents the first biocatalyst to synthesize bulky β-branched tryptophan analogues in a single step. This was achieved through introduction of active-site and remote mutations that increase the abundance and persistence of a key reactive intermediate, enabling stereoselective synthesis and accessing 27 ncAAs with industrial relevance [86]. Furthermore, directed evolution has enabled Trpβ to act as a latent tyrosine synthase, demonstrating the enzyme’s potential for synthesizing diverse aromatic amino acids beyond its native tryptophan products [87].
The specificity determinants of Trpβ have been systematically explored through structure-function studies. Francis et al. utilized a Salmonella-derived TrpβL166V variant to achieve efficient catalysis with threonine substrates, producing β-methyltryptophans with high efficiency [88]. The universality of activating mutations across Trpβ homologs was demonstrated by Murciano-Calles et al., who showed that mutations conferring enhanced activity could be successfully transferred between Trpβ enzymes sharing ≥57% sequence identity, enabling synthesis of 5-substituted tryptophans across multiple enzyme scaffolds, proving evolutionary insights transcend species barriers [89]. Recent studies highlight Trpβ versatility ncAAs synthesis, with engineered variants capable of incorporating halogenated and alkylated indole analogs [90].
Dick et al. were able to engineer PfTrpβquat variants capable of forming all carbon quaternary centers, a reaction type previously exclusive to synthetic organic chemistry. This breakthrough expanded the synthetic utility of engineered Trpβ beyond natural amino acid analogs to the structurally complex products with pharmaceutical relevance [91], with recent optimization efforts being more focused on improving yields for specific products. Xu et al. developed E. coli TrpS mutants that increased l-5-hydroxytryptophan yields to 86.7% under optimized reaction conditions, demonstrating the potential for high-efficiency biocatalytic processes [92].
The advent of high throughput screening technologies has dramatically accelerated Trpβ evolution. Scheele et al. demonstrated the power of droplet microfluidics combined with aptamer-based sensors to screen over 100,000 Trpβ mutants daily, identifying variants with 100-fold enhanced activity including UV-activatable variants with unique regulatory properties. This ultrahigh throughput approach enables exploration of sequence space that would be prohibitive using conventional screening methods [93]. Continuous evolution platforms have further expanded the toolkit for Trpβ engineering, with OrthoRep-based approaches being used to rapidly evolve Trpβ variants with altered substrate promiscuity, generating sequence-diverse enzyme libraries encompassing novel catalytic activities [94].

4.7. Engineering of the TrpS Complex

While standalone subunit engineering has achieved remarkable success, efforts to engineer the intact α2β2 complex have revealed additional opportunities for optimization (Table 3). Allosteric communication between subunits can be fine-tuned to enhance overall catalytic efficiency while maintaining the advantages of substrate channeling.
Dimethylallyl Trps (DMATS) represent a distinct class of tryptophan-modifying enzymes that have been extensively engineered for expanded prenylation chemistry. These enzymes share a common “αββα” fold but exhibit diverse active site architectures that direct regiospecific prenylation patterns. Eaton et al. characterized DMATrpS1 from Fusarium fujikuroi, revealing remarkable substrate promiscuity that enables both reverse and forward prenylation reactions through structural adaptability mechanisms [95].
The engineering of DMATrpS enzymes has focused on altering regiospecificity and substrate scope. Fan et al. successfully modified FgaPT2K174F specificity through site-directed mutagenesis, shifting preference from tryptophan to tyrosine substrates and achieving the first enzymatic synthesis of C-prenylated l-tyrosine. This demonstrated that relatively modest changes in active site architecture could dramatically alter substrate specificity while maintaining catalytic efficiency [96].
Success in engineering TrpS subunits for standalone function provides a blueprint for liberating other allosterically regulated enzymes from their regulatory constraints via the machine learning approach [97]. Advances in understanding allosteric networks enable more predictive enzyme design, exemplified by engineered TrpS enzymes whose enhanced properties, achieved through directed evolution, are expanding industrial biocatalysis for sustainable chemistry, allowing systematic exploration of natural diversity for improved biocatalytic potential.
Computational studies have revealed the molecular basis of allosteric activation, showing that beneficial mutations restore conformational ensembles resembling the native αββα complex, significantly increasing the population of catalytically active states [98]. Engineered variants now catalyze nitroalkane alkylation, yielding nitroalkyl-tryptophans with >95% enantiomeric excess [99]. Similarly, light-responsive variants demonstrate sophisticated control mechanisms, with azobenzene photochromic switches enabling control over TrpS allostery [100]. This represents a new frontier in controllable biocatalysis.

4.8. Pathway Engineering for Biosynthesis of Tryptophan and Derivatives

Corynebacterium glutamicum has emerged as a premium platform for tryptophan production, achieving 50.5 g/L through systems metabolic engineering approaches [101]. The organism’s superior aromatic compound tolerance and the established industrial fermentation processes position it as the preferred host for large scale amino acid production. E. coli expressing Sphingobacterium soilsilvae Trpβ achieved food grade tryptophan production [102]. TmTrpβ achieved tryptophan analogue production, especially 4-cyano tryptophan with 78% yield [103], while specialized applications include psilocybin production from 4-hydroxyindole, reducing the substrate costs [104]. The PfTrpSβ engineered variants demonstrate enhanced activity with 5-substituted indoles, enabling efficient biocatalytic synthesis of β-branched amino acids and 5-substituted tryptophans with improved yields [105].
Biocatalytic cascade development has expanded synthetic utility beyond single-enzyme applications. Integration with evolved Methylophaga flavin monooxygenase (MaFMOD197E) enables 1288 mg/L indigo production, offering a sustainable alternative to petroleum-based dyes [106]. One-pot d-tryptophan synthesis (>99% ee) through Trpβ integration with aminotransferases demonstrates the enzyme’s compatibility with multi-enzyme systems [107]. Ultrasound-assisted bioconversion represents process innovation, doubling l-cysteine production to 91 g/L by improving substrate permeability [108]. Patent innovations include thermostable Trpβ variants for continuous flow reactors and low-temperature synthesis applications, highlighting the enzyme’s industrial maturation.

4.9. Targeting TrpS in Pathogenic Bacteria with Novel Antimicrobials

Within the shikimate pathway, TrpS has emerged as a particularly attractive target due to its essential role in tryptophan biosynthesis and its sophisticated allosteric regulatory mechanisms that can be exploited for drug design. The allosteric communication between α- and β-subunits is essential for enzymatic efficiency, as disruption of this interface significantly affects catalytic function. Inhibitors belonging to the sulfonamide and sulfolane class were identified targeting the allosteric site in Mtb TrpS [109].
Advanced computational approaches coupled with in vitro studies have also introduced promising α site inhibitors. Through virtual screening methodologies, researchers identified a potent benzamide inhibitor specifically targeting the Mtb Trp α-subunit. This compound exhibited remarkable antimicrobial activity, achieving complete growth inhibition at concentrations of 25 μg/mL, with MD simulations revealing an impressive binding affinity characterized by a ΔGbind of −48.2 kcal/mol [110]. Additionally, ligand-based pharmacophore screening of compounds identified ZINC09150898 as a potent α-subunit inhibitor with a binding score of −32.07 kcal/mol, demonstrating 100% growth inhibition of Mtb-(H37Rv) at nearly 50 μg/mL, with MD simulations confirming stable binding through van der Waals interactions. These computational methods accelerated the identification of lead compounds targeting the α-subunit, later confirmed through wet lab experiments [111].
The most sophisticated and promising approach to TrpS inhibition involves targeting the complex interface between both subunits, disrupting the essential allosteric communication network that coordinates the two catalytic reactions. This strategy has yielded several classes of highly effective antimicrobials that exploit the fundamental requirement for subunit cooperation in tryptophan synthesis. Both sulfolane derivatives and indoline-5-sulfonamides represent successful classes of dual target inhibitors that exploit subunit interface disruption. These compounds bind at critical interaction sites between the α and β-subunits, effectively decoupling the coordinated catalytic mechanism [112]. Additionally, azetidine derivatives such as BRD4592 provide another approach to allosteric inhibition through interface targeting [113].
The development of phenyldiazenyl-propanamides represents a significant advancement in broad-spectrum antimicrobial development. Originally identified through screening against Salmonella TrpS, these compounds target the critical subunit interface with remarkable specificity [114]. Comprehensive MIC assays and high-resolution crystallographic studies have confirmed their mechanism of action, showing effective disruption of tryptophan biosynthesis and validating their potential as scaffolds for the next generation of broad-spectrum antibiotics.
The emergence of inhibitors regulated by light represents a paradigm shift in the antimicrobial drug design, offering unprecedented control over therapeutic activity. Azobenzene based compounds targeting TrpS exhibit switchable activity, with E-configuration isomers showing moderately strong inhibition while Z-configuration isomers display five times weaker inhibitory activity [115]. These photochromic inhibitors achieve their selectivity by binding to cryptic allosteric cavities and inducing noncompetitive inhibition through the disruption of normal allosteric signaling networks. This light-dependent nature of their activity provides a novel mechanism for controlling the antimicrobial action with extraordinary precision, potentially allowing for targeted therapy that minimizes the effects on beneficial microorganisms.
Table 3. Summary of TrpS engineering efforts.
Table 3. Summary of TrpS engineering efforts.
Source Organism(s)VariantApplicationRef.
E. coliTrpSV231A/K382Gl-5-hydroxytryptophan synthesis [92]
E. coli and
M. aminisulfidivorans
TrpS and MaFMOD197EIndigo production[106]
P. furiosusTrpSβmutantsncAAs production;
β-methyltryptophan synthesis
[84]
T. maritimaTmTrpSSynthesis of blue, fluorescent amino acid 4-cyanotryptophan (78% yield) [103]
P. furiosusPfTrpSβ7E6Synthesis of β-branched amino acids; 5-substituted tryptophan analogues[86]
P. furiosusPfTrpSβquatSynthesis of alkylation of 3-substituted oxindoles, ncAAs construction with quaternary stereocenters[91]

5. Conclusions and Future Directions

The shikimate pathway exemplifies how evolution has crafted sophisticated regulatory networks that coordinate metabolic flux through allosteric communication between functionally distinct enzymes. Protein engineering has demonstrated how natural allosteric constraints, while evolutionarily optimized for cellular regulation, can be strategically modified to create superior biocatalysts. The remarkable success in engineering standalone enzyme variants that bypass native regulatory dependencies, while maintaining or enhancing catalytic efficiency, represents a fundamental advance in our ability to harness biological systems for industrial applications.
The integration of structural biology insights with directed evolution methodologies has proven particularly powerful, enabling the rational targeting of allosteric networks to achieve desired functional outcomes. The development of high-throughput screening platforms, including droplet microfluidics and continuous evolution systems, has accelerated the exploration of sequence space beyond what conventional approaches could achieve. These technological advances, combined with computational tools for predicting allosteric hotspots and designing stabilizing mutations, are establishing new paradigms for enzyme engineering that extend beyond the shikimate pathway.
Looking toward future developments, several promising research directions emerge. Machine learning approaches that can predict allosteric network behaviors from sequence information alone hold potential for dramatically accelerating enzyme design cycles. The integration of synthetic biology tools with engineered shikimate pathway enzymes may enable the construction of entirely artificial metabolic networks with programmable regulatory properties. Advanced bioprocess engineering, including the development of continuous flow systems and multi-phase reaction environments, could further enhance the industrial viability of these biocatalytic platforms.
The shikimate pathway also serves another role as a potential target for novel antimicrobials. Here, the development of photo switchable inhibitors and other controllable therapeutic modalities could provide unprecedented precision in targeting pathogenic organisms while preserving beneficial microbiota. Additionally, the pathway’s central role in plant secondary metabolism suggests opportunities for agricultural applications, where engineered variants could enhance crop nutritional content or stress resistance.
The successful engineering of multi-enzyme systems from the shikimate pathway provides a blueprint for addressing similar challenges in other metabolic networks. As our understanding of allosteric communication mechanisms deepens, the principles established through shikimate pathway engineering will likely find broad application across diverse biochemical systems, ultimately contributing to the development of sustainable biotechnological solutions for chemical production, therapeutic development, and environmental remediation.

Author Contributions

Conceptualization, S.K. and D.D.B.; resources, D.D.B.; writing—original draft preparation, S.K.; writing—review and editing, S.K. and D.D.B.; visualization, S.K.; funding acquisition, D.D.B. All authors have read and agreed to the published version of the manuscript.

Funding

Our work on tryptophan synthase and chorismate mutase has been supported by NSF grant MCB 1615032.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding authors.

Acknowledgments

We would like to acknowledge all members of the Boehr lab who have contributed to our efforts on understanding enzymes within the shikimate pathway, including those on indole-3-glycerol phosphate synthase, chorismate mutase and tryptophan synthase.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Herrmann, K.M.; Weaver, L.M. The Shikimate Pathway. Annu. Rev. Plant Biol. 1999, 50, 473–503. [Google Scholar] [CrossRef]
  2. Wu, S.; Chen, W.; Lu, S.; Zhang, H.; Yin, L. Metabolic Engineering of Shikimic Acid Biosynthesis Pathway for the Production of Shikimic Acid and Its Branched Products in Microorganisms: Advances and Prospects. Molecules 2022, 27, 4779. [Google Scholar] [CrossRef] [PubMed]
  3. Lee, H.-N.; Seo, S.-Y.; Kim, H.-J.; Park, J.-H.; Park, E.; Choi, S.-S.; Lee, S.J.; Kim, E.-S. Artificial cell factory design for shikimate production in Escherichia coli. J. Ind. Microbiol. Biotechnol. 2021, 48, kuab043. [Google Scholar] [CrossRef]
  4. Mao, J.; Zhang, H.; Chen, Y.; Wei, L.; Liu, J.; Nielsen, J.; Chen, Y.; Xu, N. Relieving metabolic burden to improve robustness and bioproduction by industrial microorganisms. Biotechnol. Adv. 2024, 74, 108401. [Google Scholar] [CrossRef]
  5. de Oliveira, M.D.; Araujo, J.d.O.; Galúcio, J.M.; Santana, K.; Lima, A.H. Targeting shikimate pathway: In silico analysis of phosphoenolpyruvate derivatives as inhibitors of EPSP synthase and DAHP synthase. J. Mol. Graph. Model. 2020, 101, 107735. [Google Scholar] [CrossRef]
  6. Shende, V.V.; Bauman, K.D.; Moore, B.S. The shikimate pathway: Gateway to metabolic diversity. Nat. Prod. Rep. 2024, 41, 604–648. [Google Scholar] [CrossRef]
  7. Winston, D.S.; Gorman, S.D.; Boehr, D.D. Conformational transitions in yeast chorismate mutase important for allosteric regulation as identified by nuclear magnetic resonance spectroscopy. J. Mol. Biol. 2022, 434, 167531. [Google Scholar] [CrossRef]
  8. Gorman, S.D.; Boehr, D.D. Energy and enzyme activity landscapes of yeast chorismate mutase at cellular concentrations of allosteric effectors. Biochemistry 2019, 58, 4058–4069. [Google Scholar] [CrossRef] [PubMed]
  9. D’Amico, R.N.; Boehr, D.D. Allostery, engineering and inhibition of tryptophan synthase. Curr. Opin. Struct. Biol. 2023, 82, 102657. [Google Scholar] [CrossRef] [PubMed]
  10. 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. J. Biol. Chem. 2015, 290, 18187–18198. [Google Scholar] [CrossRef]
  11. Watkins-Dulaney, E.; Straathof, S.; Arnold, F. Tryptophan synthase: Biocatalyst extraordinaire. ChemBioChem 2021, 22, 5–16. [Google Scholar] [CrossRef] [PubMed]
  12. Jensen, R.A. Tyrosine and Phenylalanine Biosynthesis: Relationship between Alternative Pathways, Regulation and Subcellular Location. In The Shikimic Acid Pathway; Conn, E.E., Ed.; Springer: Boston, MA, USA, 1986; pp. 57–81. [Google Scholar]
  13. Bentley, R. The shikimate pathway--a metabolic tree with many branches. Crit. Rev. Biochem. Mol. Biol. 1990, 25, 307–384. [Google Scholar] [CrossRef]
  14. 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. J. Mol. Biol. 2005, 354, 927–939. [Google Scholar] [CrossRef] [PubMed]
  15. Faponle, A.S.; Fagbohunka, B.S.; Gauld, J.W. Influence of Cysteine 440 on the Active Site Properties of 3-Deoxy-d-Arabino-Heptulosonate 7-Phosphate Synthase in Mycobacterium tuberculosis (MtDAHPS). ACS Omega 2023, 8, 14401–14409. [Google Scholar] [CrossRef] [PubMed]
  16. Light, S.H.; Anderson, W.F. The diversity of allosteric controls at the gateway to aromatic amino acid biosynthesis. Protein Sci. 2013, 22, 395–404. [Google Scholar] [CrossRef]
  17. Ray, J.M.; Yanofsky, C.; Bauerle, R. Mutational analysis of the catalytic and feedback sites of the tryptophan-sensitive 3-deoxy-D-arabino-heptulosonate-7-phosphate synthase of Escherichia coli. J. Bacteriol. 1988, 170, 5500–5506. [Google Scholar] [CrossRef]
  18. McCandliss, R.J.; Poling, M.D.; Herrmann, K.M. 3-Deoxy-D-arabino-heptulosonate 7-phosphate synthase. Purification and molecular characterization of the phenylalanine-sensitive isoenzyme from Escherichia coli. J. Biol. Chem. 1978, 253, 4259–4265. [Google Scholar] [CrossRef]
  19. Cui, D.; Deng, A.; Bai, H.; Yang, Z.; Liang, Y.; Liu, Z.; Qiu, Q.; Wang, L.; Liu, S.; Zhang, Y.; et al. Molecular basis for feedback inhibition of tyrosine-regulated 3-deoxy-d-arabino-heptulosonate-7-phosphate synthase from Escherichia coli. J. Struct. Biol. 2019, 206, 322–334. [Google Scholar] [CrossRef]
  20. Pratap, S.; Dev, A.; Kumar, V.; Yadav, R.; Narwal, M.; Tomar, S.; Kumar, P. Structure of Chorismate Mutase-like Domain of DAHPS from Bacillus subtilis Complexed with Novel Inhibitor Reveals Conformational Plasticity of Active Site. Sci. Rep. 2017, 7, 6364. [Google Scholar] [CrossRef]
  21. Keith, B.; Dong, X.N.; Ausubel, F.M.; Fink, G.R. Differential induction of 3-deoxy-D-arabino-heptulosonate 7-phosphate synthase genes in Arabidopsis thaliana by wounding and pathogenic attack. Proc. Natl. Acad. Sci. USA 1991, 88, 8821–8825. [Google Scholar] [CrossRef]
  22. Kanaris, M.; Poulin, J.; Shahinas, D.; Johnson, D.; Crowley, V.M.; Fucile, G.; Provart, N.; Christendat, D. Elevated tyrosine results in the cytosolic retention of 3-deoxy-d-arabino-heptulosonate 7-phosphate synthase in Arabidopsis thaliana. Plant J. Cell Mol. Biol. 2022, 109, 789–803. [Google Scholar] [CrossRef]
  23. Entus, R.; Poling, M.; Herrmann, K.M. Redox regulation of Arabidopsis 3-deoxy-D-arabino-heptulosonate 7-phosphate synthase. Plant Physiol. 2002, 129, 1866–1871. [Google Scholar] [CrossRef] [PubMed]
  24. Zhou, M. Evolution of DAHP Synthase: From Archaea to Eubacteria; University of Michigan: Ann Arbor, MI, USA, 2007. [Google Scholar]
  25. Shumilin, I.A.; Bauerle, R.; Wu, J.; Woodard, R.W.; Kretsinger, R.H. Crystal structure of the reaction complex of 3-deoxy-D-arabino-heptulosonate-7-phosphate synthase from Thermotoga maritima refines the catalytic mechanism and indicates a new mechanism of allosteric regulation. J. Mol. Biol. 2004, 341, 455–466. [Google Scholar] [CrossRef] [PubMed]
  26. Walker, G.E.; Dunbar, B.; Hunter, I.S.; Nimmo, H.G.; Coggins, J.R. Evidence for a novel class of microbial 3-deoxy-D-arabino-heptulosonate-7-phosphate synthase in Streptomyces coelicolor A3(2), Streptomyces rimosus and Neurospora crassa. Microbiology 1996, 142 Pt 8, 1973–1982. [Google Scholar] [CrossRef] [PubMed]
  27. Heyes, L.C.; Reichau, S.; Cross, P.J.; Jameson, G.B.; Parker, E.J. Structural analysis of substrate-mimicking inhibitors in complex with Neisseria meningitidis 3-deoxy-d-arabino-heptulosonate 7-phosphate synthase—The importance of accommodating the active site water. Bioorganic Chem. 2014, 57, 242–250. [Google Scholar] [CrossRef]
  28. Reichau, S.; Blackmore, N.J.; Jiao, W.; Parker, E.J. Probing the Sophisticated Synergistic Allosteric Regulation of Aromatic Amino Acid Biosynthesis in Mycobacterium tuberculosis Using ᴅ-Amino Acids. PLoS ONE 2016, 11, e0152723. [Google Scholar] [CrossRef]
  29. 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. J. Mol. Biol. 2016, 428, 1237–1255. [Google Scholar] [CrossRef]
  30. Burschowsky, D.; Thorbjørnsrud, H.V.; Heim, J.B.; Fahrig-Kamarauskaite, 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. Biochemistry 2018, 57, 557–573. [Google Scholar] [CrossRef]
  31. Nazmi, A.R.; Lang, E.J.; 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. J. Biol. Chem. 2016, 291, 21836–21847. [Google Scholar] [CrossRef]
  32. Wang, S.; Liu, D.; Bilal, M.; Wang, W.; Zhang, X. Uncovering the Role of PhzC as DAHP Synthase in Shikimate Pathway of Pseudomonas chlororaphis HT66. Biology 2022, 11, 86. [Google Scholar] [CrossRef] [PubMed]
  33. Ahmad, S.; Johnson, J.L.; Jensen, R.A. The recent evolutionary origin of the phenylalanine-sensitive isozyme of 3-deoxy-D-arabino-heptulosonate 7-phosphate synthase in the enteric lineage of bacteria. J. Mol. Evol. 1987, 25, 159–167. [Google Scholar] [CrossRef]
  34. Yokoyama, R.; Kleven, B.; Gupta, A.; Wang, Y.; Maeda, H.A. 3-Deoxy-D-arabino-heptulosonate 7-phosphate synthase as the gatekeeper of plant aromatic natural product biosynthesis. Curr. Opin. Plant Biol. 2022, 67, 102219. [Google Scholar] [CrossRef]
  35. Jayaraman, K.; Trachtmann, N.; Sprenger, G.A.; Gohlke, H. Protein engineering for feedback resistance in 3-deoxy-D-arabino-heptulosonate 7-phosphate synthase. Appl. Microbiol. Biotechnol. 2022, 106, 6505–6517. [Google Scholar] [CrossRef]
  36. Liu, H.; Xiao, Q.; Wu, X.; Ma, H.; Li, J.; Guo, X.; Liu, Z.; Zhang, Y.; Luo, Y. Mechanistic investigation of a D to N mutation in DAHP synthase that dictates carbon flux into the shikimate pathway in yeast. Commun. Chem. 2023, 6, 152. [Google Scholar] [CrossRef]
  37. Yang, S.; Wang, Y.; Cai, Z.; Zhang, G.; Song, H. Metabolic engineering of Bacillus subtilis for high-titer production of menaquinone-7. AIChE J. 2020, 66, e16754. [Google Scholar] [CrossRef]
  38. Park, E.; Kim, H.-J.; Seo, S.-Y.; Lee, H.-N.; Choi, S.-S.; Lee, S.J.; Kim, E.-S. Shikimate metabolic pathway engineering in Corynebacterium glutamicum. J. Microbiol. Biotechnol. 2021, 31, 1305. [Google Scholar] [CrossRef]
  39. Kogure, T.; Kubota, T.; Suda, M.; Hiraga, K.; Inui, M. Metabolic engineering of Corynebacterium glutamicum for shikimate overproduction by growth-arrested cell reaction. Metab. Eng. 2016, 38, 204–216. [Google Scholar] [CrossRef]
  40. Kukil, K.; Lindberg, P. Metabolic engineering of Synechocystis sp. PCC 6803 for the improved production of phenylpropanoids. Microb. Cell Factories 2024, 23, 57. [Google Scholar] [CrossRef] [PubMed]
  41. Goldfinger, V.; Spohn, M.; Rodler, J.-P.; Sigle, M.; Kulik, A.; Cryle, M.J.; Rapp, J.; Link, H.; Wohlleben, W.; Stegmann, E. Metabolic engineering of the shikimate pathway in Amycolatopsis strains for optimized glycopeptide antibiotic production. Metab. Eng. 2023, 78, 84–92. [Google Scholar] [CrossRef] [PubMed]
  42. Frlan, R. An Evolutionary Conservation and Druggability Analysis of Enzymes Belonging to the Bacterial Shikimate Pathway. Antibiotics 2022, 11, 675. [Google Scholar] [CrossRef]
  43. Brilisauer, K.; Rapp, J.; Rath, P.; Schöllhorn, A.; Bleul, L.; Weiß, E.; Stahl, M.; Grond, S.; Forchhammer, K. Cyanobacterial antimetabolite 7-deoxy-sedoheptulose blocks the shikimate pathway to inhibit the growth of prototrophic organisms. Nat. Commun. 2019, 10, 545. [Google Scholar] [CrossRef]
  44. Tapas, S.; Kumar Patel, G.; Dhindwal, S.; Tomar, S. In Silico sequence analysis and molecular modeling of the three-dimensional structure of DAHP synthase from Pseudomonas fragi. J. Mol. Model. 2011, 17, 621–631. [Google Scholar] [CrossRef]
  45. Romero, R.; Roberts, M.; Phillipson, J. Chorismate mutase in microorganisms and plants. Phytochemistry 1995, 40, 1015–1025. [Google Scholar] [CrossRef]
  46. Russ, W.P.; Figliuzzi, M.; Stocker, C.; Barrat-Charlaix, P.; Socolich, M.; Kast, P.; Hilvert, D.; Monasson, R.; Cocco, S.; Weigt, M. An evolution-based model for designing chorismate mutase enzymes. Science 2020, 369, 440–445. [Google Scholar] [CrossRef]
  47. Kroll, K.; Holland, C.K.; Starks, C.M.; Jez, J.M. Evolution of allosteric regulation in chorismate mutases from early plants. Biochem. J. 2017, 474, 3705–3717. [Google Scholar] [CrossRef]
  48. Helmstaedt, K.; Krappmann, S.; Braus, G.H. Allosteric regulation of catalytic activity: Escherichia coli aspartate transcarbamoylase versus yeast chorismate mutase. Microbiol. Mol. Biol. Rev. 2001, 65, 404–421. [Google Scholar] [CrossRef]
  49. Chook, Y.M.; Ke, H.; Lipscomb, W.N. Crystal structures of the monofunctional chorismate mutase from Bacillus subtilis and its complex with a transition state analog. Proc. Natl. Acad. Sci. USA 1993, 90, 8600–8603. [Google Scholar] [CrossRef] [PubMed]
  50. Calhoun, D.H.; Bonner, C.A.; Gu, W.; Xie, G.; Jensen, R.A. The emerging periplasm-localized subclass of AroQ chorismate mutases, exemplified by those from Salmonella typhimurium and Pseudomonas aeruginosa. Genome Biol. 2001, 2, research0030.1. [Google Scholar] [CrossRef] [PubMed]
  51. Xue, Y.; Lipscomb, W.N.; Graf, R.; Schnappauf, G.; Braus, G. The crystal structure of allosteric chorismate mutase at 2.2-A resolution. Proc. Natl. Acad. Sci. USA 1994, 91, 10814–10818. [Google Scholar] [CrossRef] [PubMed]
  52. MacBeath, G.; Kast, P.; Hilvert, D. A Small, Thermostable, and Monofunctional Chorismate Mutase from the Archeon Methanococcus jannaschii. Biochemistry 1998, 37, 10062–10073. [Google Scholar] [CrossRef]
  53. Sträter, N.; Schnappauf, G.; Braus, G.; Lipscomb, W.N. Mechanisms of catalysis and allosteric regulation of yeast chorismate mutase from crystal structures. Structure 1997, 5, 1437–1452. [Google Scholar] [CrossRef]
  54. Fahrig-Kamarauskait, J.; Würth-Roderer, K.; Thorbjørnsrud, H.V.; Mailand, S.; Krengel, U.; Kast, P. Evolving the naturally compromised chorismate mutase from Mycobacterium tuberculosis to top performance. J. Biol. Chem. 2020, 295, 17514–17534. [Google Scholar] [CrossRef] [PubMed]
  55. Westfall, C.S.; Xu, A.; Jez, J.M. Structural evolution of differential amino acid effector regulation in plant chorismate mutases. J. Biol. Chem. 2014, 289, 28619–28628. [Google Scholar] [CrossRef]
  56. Thorbjørnsrud, H.V.; Bressan, L.; Khatanbaatar, T.; Carrer, M.; Würth-Roderer, K.; Cordara, G.; Kast, P.; Cascella, M.; Krengel, U. What Drives Chorismate Mutase to Top Performance? Insights from a Combined In Silico and In Vitro Study. Biochemistry 2023, 62, 782–796. [Google Scholar] [CrossRef] [PubMed]
  57. Koh, M.; Nasertorabi, F.; Han, G.W.; Stevens, R.C.; Schultz, P.G. Generation of an orthogonal protein–protein interface with a noncanonical amino acid. J. Am. Chem. Soc. 2017, 139, 5728–5731. [Google Scholar] [CrossRef]
  58. Kitade, Y.; Hashimoto, R.; Suda, M.; Hiraga, K.; Inui, M. Production of 4-hydroxybenzoic acid by an aerobic growth-arrested bioprocess using metabolically engineered Corynebacterium glutamicum. Appl. Environ. Microbiol. 2018, 84, e02587-17. [Google Scholar] [CrossRef] [PubMed]
  59. Liu, H.; Tian, Y.; Zhou, Y.; Kan, Y.; Wu, T.; Xiao, W.; Luo, Y. Multi-modular engineering of Saccharomyces cerevisiae for high-titre production of tyrosol and salidroside. Microb. Biotechnol. 2021, 14, 2605–2616. [Google Scholar] [CrossRef]
  60. Chávez-Béjar María, I.; Lara Alvaro, R.; López, H.; Hernández-Chávez, G.; Martinez, A.; Ramírez Octavio, T.; Bolívar, F.; Gosset, G. Metabolic Engineering of Escherichia coli for l-Tyrosine Production by Expression of Genes Coding for the Chorismate Mutase Domain of the Native Chorismate Mutase-Prephenate Dehydratase and a Cyclohexadienyl Dehydrogenase from Zymomonas mobilis. Appl. Environ. Microbiol. 2008, 74, 3284–3290. [Google Scholar] [CrossRef]
  61. Vitayakritsirikul, V.; Jaemsaeng, R.; Lohmaneeratana, K.; Thanapipatsiri, A.; Daduang, R.; Chuawong, P.; Thamchaipenet, A. Improvement of chloramphenicol production in Streptomyces venezuelae ATCC 10712 by overexpression of the aroB and aroK genes catalysing steps in the shikimate pathway. Antonie van Leeuwenhoek 2016, 109, 379–388. [Google Scholar] [CrossRef]
  62. Reddy, G.S.; Snehalatha, A.V.; Edwin, R.K.; Hossain, K.A.; Giliyaru, V.B.; Hariharapura, R.C.; Shenoy, G.G.; Misra, P.; Pal, M. Synthesis of 3-indolylmethyl substituted (pyrazolo/benzo) triazinone derivatives under Pd/Cu-catalysis: Identification of potent inhibitors of chorismate mutase (CM). Bioorganic Chem. 2019, 91, 103155. [Google Scholar] [CrossRef]
  63. Reddy, G.S.; Shukla, S.; Bhuktar, H.; Hossain, K.A.; Edwin, R.K.; Giliyaru, V.B.; Misra, P.; Pal, M. Pd/Cu-catalyzed access to novel 3-(benzofuran-2-ylmethyl) substituted (pyrazolo/benzo) triazinone derivatives: Their in silico/in vitro evaluation as inhibitors of chorismate mutase (CM). RSC Adv. 2022, 12, 26686–26695. [Google Scholar] [CrossRef]
  64. Khedr, M.A.; Massarotti, A.; Mohamed, M.E. Rational Discovery of (+) (S) Abscisic Acid as a Potential Antifungal Agent: A Repurposing Approach. Sci. Rep. 2018, 8, 8565. [Google Scholar] [CrossRef]
  65. Shukla, S.; Nishanth Rao, R.; Bhuktar, H.; Edwin, R.K.; Jamma, T.; Medishetti, R.; Banerjee, S.; Giliyaru, V.B.; Shenoy, G.G.; Oruganti, S.; et al. Wang resin catalysed sonochemical synthesis of pyrazolo[4,3-d]pyrimidinones and 2,3-dihydroquinazolin-4(1H)-ones: Identification of chorismate mutase inhibitors having effects on Mycobacterium tuberculosis cell viability. Bioorganic Chem. 2023, 134, 106452. [Google Scholar] [CrossRef]
  66. Kumar, B.S.; Rao, L.V.; Dhananjaya, G.; Kapavarapu, R.; Pal, M. Wang resin mediated unexpected greener access to 2-substituted quinazoline-4(3H)-ones and their evaluation against chorismate mutase. J. Mol. Struct. 2025, 1321, 139931. [Google Scholar] [CrossRef]
  67. Ghosh, R.; Hilario, E.; Chang, C.-E.; Mueller, L.; Dunn, M. Allosteric regulation of substrate channeling: Salmonella typhimurium tryptophan synthase. Front. Mol. Biosci. 2022, 9, 923042. [Google Scholar] [CrossRef] [PubMed]
  68. Hyde, C.C.; Ahmed, S.A.; Padlan, E.A.; Miles, E.W.; Davies, D.R. Three-dimensional structure of the tryptophan synthase alpha 2 beta 2 multienzyme complex from Salmonella typhimurium. J. Biol. Chem. 1988, 263, 17857–17871. [Google Scholar] [CrossRef]
  69. Bahar, I.; Jernigan, R.L. Cooperative fluctuations and subunit communication in tryptophan synthase. Biochemistry 1999, 38, 3478–3490. [Google Scholar] [CrossRef] [PubMed]
  70. Yanofsky, C.; Steven, S.L.L.; Horn, V.; Rowe, J. Structure and Properties of a Hybrid Tryptophan Synthetase; Chain Produced by Genetic Exchange between Escherichia coli and Salmonella typhimurium. Proc. Natl. Acad. Sci. USA 1977, 74, 286–290. [Google Scholar] [CrossRef]
  71. Ngo, H.; Kimmich, N.; Harris, R.; Niks, D.; Blumenstein, L.; Kulik, V.; Barends, T.R.; Schlichting, I.; Dunn, M.F. Allosteric regulation of substrate channeling in tryptophan synthase: Modulation of the L-serine reaction in stage I of the β-reaction by α-site ligands. Biochemistry 2007, 46, 7740–7753. [Google Scholar] [CrossRef] [PubMed]
  72. Boehr, D.D.; D’Amico, R.; O’Rourke, K.; Bosken, Y.; Sahu, D.; Chang, C.-e. Allosteric networks regulate function in multi-enzyme complexes. FASEB J. 2022, 36. [Google Scholar] [CrossRef]
  73. Lane, A.N.; Kirschner, K. The Catalytic Mechanism of Tryptophan Synthase from Escherichia coli: Kinetics of the Reaction of Indole with the Enzyme -l-Serine Complexes. Eur. J. Biochem. 1983, 129, 571–582. [Google Scholar] [CrossRef]
  74. Ngo, H.; Harris, R.; Kimmich, N.; Casino, P.; Niks, D.; Blumenstein, L.; Barends, T.R.; Kulik, V.; Weyand, M.; Schlichting, I.; et al. Synthesis and characterization of allosteric probes of substrate channeling in the tryptophan synthase bienzyme complex. Biochemistry 2007, 46, 7713–7727. [Google Scholar] [CrossRef]
  75. Anderson, K.S.; Miles, E.; Johnson, K.A. Serine modulates substrate channeling in tryptophan synthase. A novel intersubunit triggering mechanism. J. Biol. Chem. 1991, 266, 8020–8033. [Google Scholar] [CrossRef]
  76. Crawford, I.P. Evolution of a biosynthetic pathway: The tryptophan paradigm. Annu. Rev. Microbiol. 1989, 43, 567–600. [Google Scholar] [CrossRef] [PubMed]
  77. Raboni, S.; Bettati, S.; Mozzarelli, A. Tryptophan synthase: A mine for enzymologists. Cell. Mol. Life Sci. CMLS 2009, 66, 2391–2403. [Google Scholar] [CrossRef] [PubMed]
  78. Fan, Y.-X.; McPhie, P.; Miles, E.W. Regulation of tryptophan synthase by temperature, monovalent cations, and an allosteric ligand. Evidence from Arrhenius plots, absorption spectra, and primary kinetic isotope effects. Biochemistry 2000, 39, 4692–4703. [Google Scholar] [CrossRef] [PubMed]
  79. Weber-Ban, E.; Hur, O.; Bagwell, C.; Banik, U.; Yang, L.-H.; Miles, E.W.; Dunn, M.F. Investigation of allosteric linkages in the regulation of tryptophan synthase: The roles of salt bridges and monovalent cations probed by site-directed mutation, optical spectroscopy, and kinetics. Biochemistry 2001, 40, 3497–3511. [Google Scholar] [CrossRef]
  80. Dunn, M.F. Allosteric regulation of substrate channeling and catalysis in the tryptophan synthase bienzyme complex. Arch. Biochem. Biophys. 2012, 519, 154–166. [Google Scholar] [CrossRef]
  81. Axe, J.M.; Boehr, D.D. Long-range interactions in the alpha subunit of tryptophan synthase help to coordinate ligand binding, catalysis, and substrate channeling. J. Mol. Biol. 2013, 425, 1527–1545. [Google Scholar] [CrossRef]
  82. D’Amico, R.N.; Bosken, Y.K.; O’Rourke, K.F.; Murray, A.M.; Admasu, W.; Chang, C.-e.A.; Boehr, D.D. Substitution of a surface-exposed residue involved in an allosteric network enhances tryptophan synthase function in cells. Front. Mol. Biosci. 2021, 8, 679915. [Google Scholar] [CrossRef]
  83. Buller, A.R.; Brinkmann-Chen, S.; Romney, D.K.; Herger, M.; Murciano-Calles, J.; Arnold, F.H. Directed evolution of the tryptophan synthase β-subunit for stand-alone function recapitulates allosteric activation. Proc. Natl. Acad. Sci. USA 2015, 112, 14599–14604. [Google Scholar] [CrossRef] [PubMed]
  84. Herger, M.; van Roye, P.; Romney, D.K.; Brinkmann-Chen, S.; Buller, A.R.; Arnold, F.H. Synthesis of β-branched tryptophan analogues using an engineered subunit of tryptophan synthase. J. Am. Chem. Soc. 2016, 138, 8388–8391. [Google Scholar] [CrossRef]
  85. Romney, D.K.; Murciano-Calles, J.; Wehrmüller, J.E.; Arnold, F.H. Unlocking reactivity of TrpB: A general biocatalytic platform for synthesis of tryptophan analogues. J. Am. Chem. Soc. 2017, 139, 10769–10776. [Google Scholar] [CrossRef] [PubMed]
  86. Boville, C.E.; Scheele, R.A.; Koch, P.; Brinkmann-Chen, S.; Buller, A.R.; Arnold, F.H. Engineered Biosynthesis of β-Alkyl Tryptophan Analogues. Angew. Chem. Int. Ed. Engl. 2018, 57, 14764–14768. [Google Scholar] [CrossRef] [PubMed]
  87. Almhjell, P.J.; Johnston, K.E.; Porter, N.J.; Kennemur, J.L.; Bhethanabotla, V.C.; Ducharme, J.; Arnold, F.H. The β-subunit of tryptophan synthase is a latent tyrosine synthase. Nat. Chem. Biol. 2024, 20, 1086–1093. [Google Scholar] [CrossRef]
  88. Francis, D.; Winn, M.; Latham, J.; Greaney, M.F.; Micklefield, J. An engineered tryptophan synthase opens new enzymatic pathways to β-methyltryptophan and derivatives. ChemBioChem 2017, 18, 382–386. [Google Scholar] [CrossRef] [PubMed]
  89. Murciano-Calles, J.; Romney, D.K.; Brinkmann-Chen, S.; Buller, A.R.; Arnold, F.H. A panel of TrpB biocatalysts derived from tryptophan synthase through the transfer of mutations that mimic allosteric activation. Angew. Chem. Int. Ed. 2016, 55, 11577–11581. [Google Scholar] [CrossRef]
  90. Watkins-Dulaney, E.J.; Dunham, N.P.; Straathof, S.; Turi, S.; Arnold, F.H.; Buller, A.R. Asymmetric alkylation of ketones catalyzed by engineered TrpB. Angew. Chem. 2021, 133, 21582–21587. [Google Scholar] [CrossRef]
  91. Dick, M.; Sarai, N.S.; Martynowycz, M.W.; Gonen, T.; Arnold, F.H. Tailoring tryptophan synthase TrpB for selective quaternary carbon bond formation. J. Am. Chem. Soc. 2019, 141, 19817–19822. [Google Scholar] [CrossRef]
  92. Xu, L.; Li, T.; Huo, Z.; Chen, Q.; Xia, Q.; Jiang, B. Directed evolution improves the enzymatic synthesis of L-5-hydroxytryptophan by an engineered tryptophan synthase. Appl. Biochem. Biotechnol. 2021, 193, 3407–3417. [Google Scholar] [CrossRef]
  93. Scheele, R.A.; Weber, Y.; Nintzel, F.E.; Herger, M.; Kaminski, T.S.; Hollfelder, F. Ultrahigh throughput evolution of tryptophan synthase in droplets via an aptamer sensor. ACS Catal. 2024, 14, 6259–6271. [Google Scholar] [CrossRef] [PubMed]
  94. Rix, G.; Watkins-Dulaney, E.J.; Almhjell, P.J.; Boville, C.E.; Arnold, F.H.; Liu, C.C. Scalable continuous evolution for the generation of diverse enzyme variants encompassing promiscuous activities. Nat. Commun. 2020, 11, 5644. [Google Scholar] [CrossRef] [PubMed]
  95. Eaton, S.A.; Ronnebaum, T.A.; Roose, B.W.; Christianson, D.W. Structural basis of substrate promiscuity and catalysis by the reverse prenyltransferase N-dimethylallyl-l-tryptophan synthase from Fusarium fujikuroi. Biochemistry 2022, 61, 2025–2035. [Google Scholar] [CrossRef]
  96. Fan, A.; Zocher, G.; Stec, E.; Stehle, T.; Li, S.-M. Site-directed mutagenesis switching a dimethylallyl tryptophan synthase to a specific tyrosine C3-prenylating enzyme. J. Biol. Chem. 2015, 290, 1364–1373. [Google Scholar] [CrossRef]
  97. Yang, K.K.; Wu, Z.; Arnold, F.H. Machine-learning-guided directed evolution for protein engineering. Nat. Methods 2019, 16, 687–694. [Google Scholar] [CrossRef]
  98. Maria-Solano, M.A.; Iglesias-Fernández, J.; Osuna, S. Deciphering the Allosterically Driven Conformational Ensemble in Tryptophan Synthase Evolution. J. Am. Chem. Soc. 2019, 141, 13049–13056. [Google Scholar] [CrossRef]
  99. Romney, D.K.; Sarai, N.S.; Arnold, F.H. Nitroalkanes as versatile nucleophiles for enzymatic synthesis of noncanonical amino acids. ACS Catal. 2019, 9, 8726–8730. [Google Scholar] [CrossRef]
  100. Kneuttinger, A.C.; Zwisele, S.; Straub, K.; Bruckmann, A.; Busch, F.; Kinateder, T.; Gaim, B.; Wysocki, V.H.; Merkl, R.; Sterner, R. Light-regulation of tryptophan synthase by combining protein design and enzymology. Int. J. Mol. Sci. 2019, 20, 5106. [Google Scholar] [CrossRef]
  101. Dong, Y.; Chen, Z. Systems metabolic engineering of Corynebacterium glutamicum for efficient l-tryptophan production. Synth. Syst. Biotechnol. 2025, 10, 511–522. [Google Scholar] [CrossRef] [PubMed]
  102. Cao, L.; Zhang, J.; Chen, J.; Li, M.; Chen, H.; Wang, C.; Gong, C. Tryptophan production by catalysis of a putative tryptophan synthase protein. Arch. Microbiol. 2024, 206, 390. [Google Scholar] [CrossRef]
  103. Boville, C.E.; Romney, D.K.; Almhjell, P.J.; Sieben, M.; Arnold, F.H. Improved synthesis of 4-cyanotryptophan and other tryptophan analogues in aqueous solvent using variants of TrpB from Thermotoga maritima. J. Org. Chem. 2018, 83, 7447–7452. [Google Scholar] [CrossRef]
  104. Seibold, P.S.; Dörner, S.; Fricke, J.; Schäfer, T.; Beemelmanns, C.; Hoffmeister, D. Genetic regulation of l-tryptophan metabolism in Psilocybe mexicana supports psilocybin biosynthesis. Fungal Biol. Biotechnol. 2024, 11, 4. [Google Scholar] [CrossRef]
  105. Murciano-Calles, J.; Buller, A.R.; Arnold, F.H. Directed evolution of an allosteric tryptophan synthase to create a platform for synthesis of noncanonical amino acids. In Directed Enzyme Evolution: Advances and Applications; Springer: Berlin/Heidelberg, Germany, 2017; pp. 1–16. [Google Scholar]
  106. Luo, S.; Wei, W.; Wu, J.; Song, W.; Hu, G.; Liu, L. A dual-enzyme cascade for production of indigo from L-tryptophan. Sheng Wu Gong Cheng Xue Bao Chin. J. Biotechnol. 2024, 40, 2444–2456. [Google Scholar]
  107. Parmeggiani, F.; Rué Casamajo, A.; Walton, C.J.; Galman, J.L.; Turner, N.J.; Chica, R.A. One-pot biocatalytic synthesis of substituted d-tryptophans from indoles enabled by an engineered aminotransferase. ACS Catal. 2019, 9, 3482–3486. [Google Scholar] [CrossRef]
  108. Xu, L.; Wu, F.; Li, T.; Zhang, X.; Chen, Q.; Jiang, B.; Xia, Q. Ultrasound-assisted l-cysteine whole-cell bioconversion by recombinant Escherichia coli with tryptophan synthase. Green Process. Synth. 2021, 10, 842–850. [Google Scholar] [CrossRef]
  109. Michalska, K.; Chang, C.; Maltseva, N.I.; Jedrzejczak, R.; Robertson, G.T.; Gusovsky, F.; McCarren, P.; Schreiber, S.L.; Nag, P.P.; Joachimiak, A. Allosteric inhibitors of Mycobacterium tuberculosis tryptophan synthase. Protein Sci. 2020, 29, 779–788. [Google Scholar] [CrossRef]
  110. Naz, S.; Farooq, U.; Ali, S.; Sarwar, R.; Khan, S.; Abagyan, R. Identification of new benzamide inhibitor against α-subunit of tryptophan synthase from Mycobacterium tuberculosis through structure-based virtual screening, anti-tuberculosis activity and molecular dynamics simulations. J. Biomol. Struct. Dyn. 2019, 37, 1043–1053. [Google Scholar] [CrossRef] [PubMed]
  111. Naz, S.; Farooq, U.; Khan, S.; Sarwar, R.; Mabkhot, Y.N.; Saeed, M.; Alsayari, A.; Muhsinah, A.B.; Ul-Haq, Z. Pharmacophore model-based virtual screening, docking, biological evaluation and molecular dynamics simulations for inhibitors discovery against α-tryptophan synthase from Mycobacterium tuberculosis. J. Biomol. Struct. Dyn. 2021, 39, 610–620. [Google Scholar] [CrossRef] [PubMed]
  112. Abrahams, K.A.; Cox, J.A.G.; Fütterer, K.; Rullas, J.; Ortega-Muro, F.; Loman, N.J.; Moynihan, P.J.; Pérez-Herrán, E.; Jiménez, E.; Esquivias, J.; et al. Inhibiting mycobacterial tryptophan synthase by targeting the inter-subunit interface. Sci. Rep. 2017, 7, 9430. [Google Scholar] [CrossRef]
  113. Wellington, S.; Nag, P.P.; Michalska, K.; Johnston, S.E.; Jedrzejczak, R.P.; Kaushik, V.K.; Clatworthy, A.E.; Siddiqi, N.; McCarren, P.; Bajrami, B. A small-molecule allosteric inhibitor of Mycobacterium tuberculosis tryptophan synthase. Nat. Chem. Biol. 2017, 13, 943–950. [Google Scholar] [CrossRef] [PubMed]
  114. Bosken, Y.K.; Ai, R.; Hilario, E.; Ghosh, R.K.; Dunn, M.F.; Kan, S.H.; Niks, D.; Zhou, H.; Ma, W.; Mueller, L.J. Discovery of antimicrobial agent targeting tryptophan synthase. Protein Sci. 2022, 31, 432–442. [Google Scholar] [CrossRef] [PubMed]
  115. Simeth, N.A.; Kinateder, T.; Rajendran, C.; Nazet, J.; Merkl, R.; Sterner, R.; König, B.; Kneuttinger, A.C. Towards photochromic azobenzene-based inhibitors for tryptophan synthase. Chem. A Eur. J. 2021, 27, 2439–2451. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Biological regulation of the shikimate pathway. Systems-level regulatory architecture showing the sophisticated allosteric communication networks governing shikimate pathway flux. DAHPS functions as the pathway entry point with feedback inhibition by aromatic amino acids (Phe, Tyr, Trp). CM serves as the critical branch point control mechanism, balancing Phe/Tyr synthesis with allosteric regulation. TrpS operates as an α2β2 complex with substrate channeling between the α and β active sites. Key regulatory mechanisms include feedback inhibition (orange circles), allosteric activation (green circles), substrate channeling (red arrows), and allosteric regulation networks (dashed lines) that ensure metabolic homeostasis and efficient aromatic amino acid production.
Figure 1. Biological regulation of the shikimate pathway. Systems-level regulatory architecture showing the sophisticated allosteric communication networks governing shikimate pathway flux. DAHPS functions as the pathway entry point with feedback inhibition by aromatic amino acids (Phe, Tyr, Trp). CM serves as the critical branch point control mechanism, balancing Phe/Tyr synthesis with allosteric regulation. TrpS operates as an α2β2 complex with substrate channeling between the α and β active sites. Key regulatory mechanisms include feedback inhibition (orange circles), allosteric activation (green circles), substrate channeling (red arrows), and allosteric regulation networks (dashed lines) that ensure metabolic homeostasis and efficient aromatic amino acid production.
Catalysts 15 00718 g001
Figure 2. Shikimate pathway architecture and metabolic branching networks. Metabolic flow diagram depicting the complete shikimate pathway from central metabolite precursors (PEP and E4P) through DAHPS-catalyzed condensation to form DAHP, followed by five enzymatic steps to chorismate. The pathway branches at chorismate towards the phenylalanine/tyrosine and tryptophan enzyme cascades. CM directs flux toward phenylalanine and tyrosine biosynthesis through prephenate. Among the tryptophan biosynthesis branch, tryptophan synthase (TrpS), which catalyzes the final two steps has emerged as an important chemical biology platform. Enzyme structures shown represent the three-dimensional architectures of the key regulatory enzymes, with molecular structures illustrating substrate and product chemical transformations at each branching point.
Figure 2. Shikimate pathway architecture and metabolic branching networks. Metabolic flow diagram depicting the complete shikimate pathway from central metabolite precursors (PEP and E4P) through DAHPS-catalyzed condensation to form DAHP, followed by five enzymatic steps to chorismate. The pathway branches at chorismate towards the phenylalanine/tyrosine and tryptophan enzyme cascades. CM directs flux toward phenylalanine and tyrosine biosynthesis through prephenate. Among the tryptophan biosynthesis branch, tryptophan synthase (TrpS), which catalyzes the final two steps has emerged as an important chemical biology platform. Enzyme structures shown represent the three-dimensional architectures of the key regulatory enzymes, with molecular structures illustrating substrate and product chemical transformations at each branching point.
Catalysts 15 00718 g002
Figure 3. Evolutionary classification and catalytic mechanism of DAHPS across biological domains. Upper panel shows the three major structural types of DAHPS: Type Iα; Type Iβ-; Type II. Lower panel illustrates the catalytic mechanism showing the coordination of PEP and E4P substrates with the central M2+ cofactor, with key catalytic residues (Cys, His, Lys positioned for aldol-like condensation to form DAHP.
Figure 3. Evolutionary classification and catalytic mechanism of DAHPS across biological domains. Upper panel shows the three major structural types of DAHPS: Type Iα; Type Iβ-; Type II. Lower panel illustrates the catalytic mechanism showing the coordination of PEP and E4P substrates with the central M2+ cofactor, with key catalytic residues (Cys, His, Lys positioned for aldol-like condensation to form DAHP.
Catalysts 15 00718 g003
Figure 4. Structural diversity and evolutionary relationships of CM enzyme families. Representative three-dimensional structures and schematic domain organizations of the five major CM structural classes. AroH exhibits an all-α-helical fold with 4-helix bundles; AroQα features an α/β fold with P domain (homodimer); AroQβ displays a compact α/β fold with TIM barrel-like architecture (homodimer); AroQγ adopts an 8-strand β-barrel fold (monomer); and AroQδ represents a multi-domain architecture with two TIM domains (multidomain complex).
Figure 4. Structural diversity and evolutionary relationships of CM enzyme families. Representative three-dimensional structures and schematic domain organizations of the five major CM structural classes. AroH exhibits an all-α-helical fold with 4-helix bundles; AroQα features an α/β fold with P domain (homodimer); AroQβ displays a compact α/β fold with TIM barrel-like architecture (homodimer); AroQγ adopts an 8-strand β-barrel fold (monomer); and AroQδ represents a multi-domain architecture with two TIM domains (multidomain complex).
Catalysts 15 00718 g004
Figure 5. CM catalytic mechanism and transition state stabilization architecture. Mechanistic representation of the [3,3]-sigmatropic rearrangement converting chorismate to prephenate via a chair-like transition state.
Figure 5. CM catalytic mechanism and transition state stabilization architecture. Mechanistic representation of the [3,3]-sigmatropic rearrangement converting chorismate to prephenate via a chair-like transition state.
Catalysts 15 00718 g005
Figure 6. The reactions and substrate channeling of TrpS. The native α2β2 complex shows the bidirectional reaction mechanism, Key catalytic residues (Glu49, Asp60 in Trpα; Lys87, Asp305, Ser377 in Trpβ) are highlighted. The 25 Å hydrophobic tunnel (purple) provides 100-fold rate enhancement by protecting indole from side reactions and maintaining local concentrations.
Figure 6. The reactions and substrate channeling of TrpS. The native α2β2 complex shows the bidirectional reaction mechanism, Key catalytic residues (Glu49, Asp60 in Trpα; Lys87, Asp305, Ser377 in Trpβ) are highlighted. The 25 Å hydrophobic tunnel (purple) provides 100-fold rate enhancement by protecting indole from side reactions and maintaining local concentrations.
Catalysts 15 00718 g006
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Khan, S.; Boehr, D.D. Multi-Enzyme Synergy and Allosteric Regulation in the Shikimate Pathway: Biocatalytic Platforms for Industrial Applications. Catalysts 2025, 15, 718. https://doi.org/10.3390/catal15080718

AMA Style

Khan S, Boehr DD. Multi-Enzyme Synergy and Allosteric Regulation in the Shikimate Pathway: Biocatalytic Platforms for Industrial Applications. Catalysts. 2025; 15(8):718. https://doi.org/10.3390/catal15080718

Chicago/Turabian Style

Khan, Sara, and David D. Boehr. 2025. "Multi-Enzyme Synergy and Allosteric Regulation in the Shikimate Pathway: Biocatalytic Platforms for Industrial Applications" Catalysts 15, no. 8: 718. https://doi.org/10.3390/catal15080718

APA Style

Khan, S., & Boehr, D. D. (2025). Multi-Enzyme Synergy and Allosteric Regulation in the Shikimate Pathway: Biocatalytic Platforms for Industrial Applications. Catalysts, 15(8), 718. https://doi.org/10.3390/catal15080718

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop