Note: While we prioritize shipping the format currently in stock, please specify your format preference during order placement for customized preparation.
Note: All proteins are shipped with standard blue ice packs unless dry ice shipping is specifically requested. Please contact us in advance for dry ice shipping arrangements; additional fees will apply.
The specific tag type is determined during production. If you require a particular tag, please inform us; we will prioritize its incorporation.
Catalyzes a trans-dehydration reaction via an enolate intermediate.
KEGG: pmm:PMM0387
STRING: 59919.PMM0387
3-Dehydroquinate dehydratase (DHQD, aroQ) catalyzes the third step in the shikimate pathway, converting 3-dehydroquinic acid (DHQ) to 3-dehydroshikimic acid (DHS). This enzymatic reaction is crucial for the biosynthesis of aromatic amino acids and folates, which are essential metabolites for bacterial survival. In P. marinus, this pathway provides precursors for the synthesis of chorismate, which subsequently leads to the biosynthesis of phenylalanine, tyrosine, and tryptophan .
The reaction catalyzed is specifically a dehydration, eliminating water from DHQ to form DHS, which serves as a critical intermediate in the multi-step shikimate pathway. The enzyme belongs to type II DHQDs, which catalyze anti-dehydration through an enolate intermediate, forming a Schiff base with a conserved lysine residue in the active site .
Prochlorococcus marinus subsp. pastoris, also known as strain CCMP1986 or MED4, has one of the smallest genomes of any photosynthetic organism, consisting of a single circular chromosome of only 1,657,990 bp containing 1,796 predicted protein-coding genes . Within this compact genome, the aroQ gene exists as part of the cyanobacterium's essential metabolic machinery.
Unlike many other bacterial species where aroQ frequently occurs as a fusion with other aromatic pathway genes (such as pheA or tyrA), in P. marinus it appears to exist as a standalone gene. This differs from organisms like Erwinia herbicola where aroQ has been found to have a cleavable signal peptide and is located in the periplasmic compartment .
The aroQ gene in P. marinus is functionally connected to other aromatic pathway genes in its metabolic network, as demonstrated by protein-protein interaction data:
| Protein | Function | Interaction Score with aroK |
|---|---|---|
| aroB | 3-dehydroquinate synthetase | 0.997 |
| aroA | 5-enolpyruvylshikimate-3-phosphate synthase | 0.982 |
| aroE | Shikimate 5-dehydrogenase | 0.980 |
| aroQ | 3-dehydroquinate dehydratase II | 0.834 |
| aroC | Chorismate synthase | 0.815 |
This network of interactions highlights aroQ's integration into the larger shikimate pathway machinery .
The aroQ gene in Prochlorococcus marinus encodes a Type II dehydroquinate dehydratase (DHQD). DHQDs are classified into two distinct types based on their structure, oligomeric state, and catalytic mechanism:
Type I DHQD (aroD): Catalyzes syn-dehydration through a covalent imine intermediate. These enzymes have an (α/β)8 fold and exist as homodimers .
Type II DHQD (aroQ): Catalyzes anti-dehydration by forming a Schiff base with a conserved lysine residue through an enolate intermediate. These enzymes exist as homododecamers containing a flavodoxin fold .
The Type II DHQDs, including the one from P. marinus, form complex quaternary structures consisting of four trimers with three interfacial active sites . This structural arrangement is important for the enzyme's function and stability.
AroQ proteins (Type II DHQDs) belong to a structurally distinct class that differs significantly from Type I DHQDs. Based on structural studies of similar enzymes, AroQ proteins typically:
Form homododecameric structures consisting of four trimers
Contain a flavodoxin fold in each monomer
Have three interfacial active sites per trimer
Possess a distinctive lid loop that can adopt open and closed conformations
A periplasmic subclass of AroQ has been identified in some bacteria, such as Salmonella typhimurium and Pseudomonas aeruginosa, which are approximately twice the size of cytoplasmic AroQ proteins due to a carboxy-terminal extension of unknown function .
While the specific structure of P. marinus AroQ has not been directly reported in the search results, structural comparison with homologs from other bacteria can provide insight into its likely structural features. The quaternary structure plays a crucial role in the enzyme's stability and function, with interactions between subunits contributing to the formation of catalytically active interfaces.
Type II DHQDs, including the AroQ from P. marinus, catalyze the conversion of DHQ to DHS through an anti-elimination mechanism. The catalytic process involves:
Binding of the substrate (DHQ) in the active site
Formation of a Schiff base between the substrate and a conserved lysine residue
Generation of an enolate intermediate
Anti-elimination of water to form DHS
Studies on similar Type II DHQDs have identified key residues involved in catalysis. For example, in Corynebacterium glutamicum DHQD (CgDHQD), residues such as R19, S103, and P105 play important roles in substrate binding and catalysis . Specifically:
R19 likely interacts with the carboxyl group of DHQ
S103 is positioned near the 5-hydroxyl group of DHQ
P105 is a distinctive residue in some Corynebacterium species that affects enzyme activity
Replacement of P105 with isoleucine or valine (conserved in other DHQDs) caused approximately 70% decrease in activity, while replacement of S103 with threonine increased activity by 10% . These structure-function relationships provide valuable insights into the catalytic mechanism of Type II DHQDs.
While specific inhibitors of P. marinus aroQ have not been directly reported in the search results, studies on similar Type II DHQDs provide valuable insights into potential inhibition mechanisms:
Citrate inhibition: In CgDHQD, citrate has been identified as an inhibitor that binds to the active site with a half-opened lid loop . The crystal structure of CgDHQD with bound citrate was determined at a resolution of 1.80 Å, revealing details of this inhibitor-enzyme interaction.
Competitive inhibitors: Compounds structurally similar to the substrate (DHQ) or product (DHS) may act as competitive inhibitors by occupying the active site.
Transition state analogs: Compounds that mimic the transition state of the reaction are potential potent inhibitors of the enzyme.
Understanding inhibition mechanisms is important for:
Developing tools to study enzyme function
Potential antimicrobial drug development, as the shikimate pathway is absent in mammals
Elucidating the structural basis of substrate specificity
For researchers investigating P. marinus aroQ, testing known inhibitors of related DHQDs would be a valuable approach to characterizing its inhibition profile.
Based on available information for similar recombinant proteins from P. marinus and other organisms, several expression systems have been successfully employed:
Escherichia coli-based expression: The most commonly used system for recombinant protein production. For similar enzymes like CgDHQD, expression in E. coli BL21(DE3) T1R strain using a pET30a vector with a C-terminal 6x-His tag has proven successful . Protein expression can be induced with IPTG (typically 0.5 mM) at reduced temperatures (e.g., 291 K for 20 hours) to enhance protein solubility.
Yeast expression systems: According to the search results, some commercial recombinant proteins from P. marinus are produced in yeast expression systems . Yeast can provide advantages for proteins requiring eukaryotic post-translational modifications, although this is generally less critical for bacterial proteins.
For optimal expression of recombinant P. marinus aroQ, researchers should consider:
Codon optimization for the expression host
Selection of appropriate promoters (e.g., T7 for E. coli)
Inclusion of purification tags (His-tag, GST-tag, etc.)
Optimization of induction conditions (temperature, inducer concentration, duration)
Use of specialty strains designed to express proteins that may be toxic or form inclusion bodies
A multi-step purification strategy is recommended to obtain high-purity recombinant P. marinus aroQ, based on successful approaches used for similar enzymes:
Initial capture: Affinity chromatography using Ni-NTA agarose for His-tagged proteins. After cell lysis by ultrasonication and clarification by centrifugation, the soluble fraction is applied to the Ni-NTA column. Washing with buffer containing low imidazole concentrations (e.g., 20 mM) removes weakly bound proteins, followed by elution with higher imidazole concentrations (e.g., 300 mM) .
Intermediate purification: Ion-exchange chromatography (e.g., HiTrap Q FF) can separate proteins based on charge differences, removing contaminants with different isoelectric points .
Polishing step: Size exclusion chromatography (e.g., HiPrep 26/60 Sephacryl S-300 HR column) for final purification, which separates proteins based on molecular size and shape .
Quality control: The purified protein should be assessed by:
SDS-PAGE to verify purity and molecular weight
Western blotting if specific antibodies are available
Activity assays to confirm enzymatic function
Mass spectrometry to verify protein identity
Using this approach, researchers have achieved high-purity preparations of similar enzymes. For example, CgDHQD was purified to a state where a single band was visible by SDS-PAGE after Coomassie blue staining .
Based on information from commercial recombinant proteins and similar enzymes, the following recommendations can be made for maximizing the stability and shelf life of recombinant P. marinus aroQ:
Buffer: Typically Tris-based (40-50 mM, pH 7.5-8.0)
Cryoprotectant: 20-50% glycerol for frozen storage
Optional additives: Small amounts of reducing agents (e.g., DTT, β-mercaptoethanol) may help maintain activity
For lyophilized protein: Reconstitute in deionized sterile water to a concentration of 0.1-1.0 mg/mL
Addition of glycerol to a final concentration of 5-50% is recommended for long-term storage
Centrifugation of the vial before opening helps bring contents to the bottom
Avoid repeated freeze-thaw cycles
Aliquot the protein solution into single-use volumes before freezing
Store working stocks at 4°C only for short periods (≤1 week)
Avoid exposure to extreme pH, high temperatures, or strong denaturants
By following these guidelines, researchers can maximize the stability and activity of recombinant P. marinus aroQ during storage and experimental use.
Prochlorococcus marinus has unique adaptations compared to typical cyanobacteria, and these differences extend to its metabolic pathways. Regarding the aroQ gene:
Genomic context: In P. marinus, aroQ exists within a highly compact genome, reflecting the organism's evolutionary trajectory toward genome minimization . This contrasts with many other cyanobacteria that have larger genomes.
Light adaptation influence: P. marinus strains are adapted to either high-light or low-light conditions, which affects their genomic content. The high-light-adapted strains like P. marinus subsp. pastoris (MED4) have undergone more extensive genome reduction compared to low-light-adapted strains .
Protein interaction network: In P. marinus, aroQ interactions with other shikimate pathway enzymes (aroB, aroA, aroE, aroC) as well as with aromatic amino acid biosynthesis enzymes (tyrA, pheA) have been identified through computational approaches . This interaction network highlights the integration of aroQ in the metabolic framework specific to this organism.
Evolutionary conservation: While the shikimate pathway is generally conserved across bacteria, specific adaptations in enzyme sequence, regulation, and cellular localization can differ. In some bacteria, periplasmic forms of aroQ have been identified, whereas in others, the enzyme is cytoplasmic or exists as a fusion protein with other enzymatic domains .
The evolution of aroQ in P. marinus likely reflects the environmental pressures faced by this organism in nutrient-limited oceanic environments, where efficient metabolism with minimal genetic content provides a selective advantage.
The evolution of two distinct types of dehydroquinate dehydratases (Type I/aroD and Type II/aroQ) across bacterial species reflects diverse selective pressures and evolutionary processes:
Structural and mechanistic divergence: Type I and Type II DHQDs catalyze the same reaction but through different mechanisms:
Domain fusions and protein evolution: AroQ has been an extremely popular partner for fusion with other aromatic pathway genes, highlighting evolutionary forces driving the co-localization of functionally related enzymes . Examples include:
aroQ- pheA fusion in gamma and beta proteobacteria
aroQ- tyrA fusion in enteric bacteria
aroQ- aroD fusion in Clostridium acetobutylicum
aroQ- aroA fusion in Bacillus subtilis, Staphylococcus, and Deinococcus
Functional advantages: In B. subtilis, the AroQ domain functions not only as a catalytic domain for chorismate mutase but also as the allosteric domain for the fused AroA domain, providing a physical basis for sequential feedback inhibition .
Cellular localization adaptations: The emergence of periplasmic AroQ proteins with signal peptides and C-terminal extensions represents another evolutionary trajectory, potentially related to metabolic compartmentalization .
These evolutionary patterns suggest that DHQDs have undergone significant diversification in response to metabolic requirements, regulatory pressures, and cellular organization across different bacterial lineages.
While direct evidence for horizontal gene transfer (HGT) of aroQ in Prochlorococcus marinus is not explicitly stated in the search results, several factors suggest this possibility:
Distribution of periplasmic aroQ: The erratic phylogenetic distribution of periplasmic AroQ (designated as *AroQ) may be explained by various mechanisms, including "massive gene loss, horizontal transfer or independent evolution of a signal peptide and carboxy-terminal extension" . This suggests HGT as one potential mechanism for the spread of this gene variant across unrelated bacteria.
Genomic diversity in Prochlorococcus: Despite having small individual genomes, Prochlorococcus strains collectively show remarkable genetic diversity. The pangenome of Prochlorococcus contains more than 80,000 genes , indicating substantial genetic exchange within this genus.
Ecosystem context: Prochlorococcus exists in microbial communities with diverse bacteria and phages, providing opportunities for genetic exchange through various HGT mechanisms.
Gene content variation: Differences in gene content between high-light and low-light adapted Prochlorococcus strains could potentially be influenced by HGT events in addition to gene loss processes.
To definitively determine if aroQ in P. marinus resulted from HGT would require comprehensive phylogenetic analyses comparing:
Gene phylogeny versus species phylogeny
GC content and codon usage patterns compared to the genomic average
Presence of nearby mobile genetic elements or HGT signatures
Distribution patterns across related and unrelated bacterial species
Such analyses could reveal whether aroQ in P. marinus represents an ancestral gene or one acquired through horizontal transfer.
Based on methods used for similar dehydroquinate dehydratases, the following protocol can be optimized for measuring the enzymatic activity of recombinant P. marinus aroQ:
Substrate: 3-dehydroquinic acid potassium salt (CAS No. 494211-79-9), with concentrations ranging from 50 to 2,000 μM for kinetic studies
Enzyme concentration: Approximately 20 nM (may require optimization)
Detection method: Monitor the increased absorbance of 3-dehydroshikimic acid (DHS) at 234 nm (ε = 1.2 × 10^4 M^-1cm^-1) using an ultraviolet spectrophotometer
Temperature: Typically room temperature (25°C) or physiologically relevant temperature
Reaction initiation: Pre-incubate buffer and substrate for one minute, then add enzyme to initiate the reaction
Enzyme stability: Ensure the enzyme remains stable throughout the measurement period
Linear range: Determine the linear range of the assay with respect to time and enzyme concentration
Controls: Include appropriate controls:
No-enzyme control to account for non-enzymatic conversion
Heat-inactivated enzyme control
Known inhibitor control (e.g., citrate) to confirm specific activity
Data collection: Collect multiple time points to calculate initial velocity accurately
HPLC: For more sensitive detection or confirmation of product identity
Coupled enzyme assays: If direct spectrophotometric measurement is challenging
Mass spectrometry: For detailed product analysis or reaction mechanism studies
By optimizing these conditions, researchers can obtain reliable measurements of P. marinus aroQ enzymatic activity for kinetic characterization and inhibitor studies.
Site-directed mutagenesis is a powerful approach for investigating enzyme catalytic mechanisms by systematically altering key residues. For P. marinus aroQ, the following methodology can be implemented:
Identification of target residues: Based on sequence alignments with well-characterized DHQDs, homology modeling, and/or crystal structures, identify conserved residues likely involved in:
Substrate binding
Catalysis
Maintenance of active site architecture
Conformational changes during catalysis
Mutagenesis method: Use a QuikChange site-directed mutagenesis kit or similar approach to introduce specific mutations . This typically involves:
Designing complementary primers containing the desired mutation
PCR amplification of the entire plasmid
DpnI digestion to remove template DNA
Transformation into competent E. coli cells
Mutation strategies:
Conservative substitutions: Replace residues with chemically similar amino acids (e.g., Lys→Arg, Asp→Glu)
Non-conservative substitutions: Replace with functionally distinct amino acids
Alanine scanning: Systematically replace residues with alanine to assess their contribution
Validation: Confirm mutations by DNA sequencing before proceeding to protein expression and purification
Kinetic characterization: Determine Km, kcat, and kcat/Km values for each mutant compared to wild-type
Substrate specificity: Test alternative substrates to assess changes in specificity
pH-activity profiles: Determine if mutations alter the pH optimum or dependence
Inhibitor sensitivity: Test sensitivity to known inhibitors
Structural studies: When possible, determine crystal structures of key mutants
Example mutations based on similar DHQDs:
Findings from CgDHQD suggest the following mutations might be informative:
Conserved arginine residues involved in substrate binding
Serine residues near the 5-hydroxyl group of DHQ
Unique proline residues that may affect activity
Lysine residues involved in Schiff base formation
By systematically analyzing the effects of these mutations on enzyme activity and structure, researchers can elucidate the catalytic mechanism of P. marinus aroQ and identify key residues essential for function.
Understanding the quaternary structure of P. marinus aroQ is crucial for elucidating its function, as Type II DHQDs typically form complex oligomeric assemblies. Several complementary techniques can be employed:
X-ray crystallography: The gold standard for high-resolution structural determination. Crystal structures of similar DHQDs have been determined at resolutions of 1.8-2.0 Å, revealing detailed quaternary arrangements . Key steps include:
Crystallization screening using sparse-matrix approaches
Optimization of crystallization conditions
Data collection at synchrotron sources
Phase determination and structure refinement
Cryo-electron microscopy (cryo-EM): Particularly useful for large protein complexes, allowing visualization of quaternary structure without crystallization.
Chemical crosslinking combined with mass spectrometry: Identifies residues in close proximity, providing information about subunit interfaces.
Differential scanning calorimetry (DSC): Measures thermal stability, which can differ between monomeric and oligomeric forms.
Circular dichroism (CD): Assesses secondary structure content and thermal stability.
Native PAGE: Provides information about the native oligomeric state and homogeneity.
For comprehensive characterization of P. marinus aroQ quaternary structure, a combination of these techniques would provide complementary information about its oligomeric state, assembly mechanism, and structural stability under various conditions.
Proper analysis of enzyme kinetic data for P. marinus aroQ requires rigorous methodological approaches to obtain accurate Michaelis-Menten parameters:
Substrate range: Use 8-12 substrate concentrations ranging from 0.2×Km to 5×Km (typically 50-2,000 μM for DHQDs)
Enzyme concentration: Use the minimum concentration that gives reliable rate measurements (approximately 20 nM for similar DHQDs)
Initial velocity conditions: Ensure measurements are taken during the linear phase of the reaction (<10% substrate conversion)
Replication: Perform at least triplicate measurements at each substrate concentration
Initial velocity calculation: Determine the rate (v) at each substrate concentration [S] from the linear portion of progress curves
Primary data visualization: Plot reaction velocity (v) versus substrate concentration [S] to visualize the hyperbolic relationship
Non-linear regression analysis: Fit data directly to the Michaelis-Menten equation:
v = Vmax × [S] / (Km + [S])
using non-linear regression software (e.g., GraphPad Prism, Origin, R)
Linear transformations for verification and visualization:
Lineweaver-Burk plot (1/v vs. 1/[S])
Eadie-Hofstee plot (v vs. v/[S])
Hanes-Woolf plot ([S]/v vs. [S])
Parameter extraction: Determine:
Km (Michaelis constant): Substrate concentration at half-maximal velocity
Vmax (maximum velocity): Asymptotic maximum rate at saturating substrate
kcat (turnover number): Calculate as Vmax/[E]total
kcat/Km (catalytic efficiency): Often the most relevant parameter for comparing enzyme variants
Statistical analysis: Report 95% confidence intervals for all parameters and evaluate goodness of fit (R² values)
Substrate inhibition: If velocity decreases at high [S], fit to the modified equation:
v = Vmax × [S] / (Km + [S] + [S]²/Ki)
Sigmoidal kinetics: If the enzyme shows cooperativity, use the Hill equation:
v = Vmax × [S]^h / (K'+ [S]^h)
Non-ideal behavior: Address potential issues like substrate depletion, product inhibition, or enzyme instability by appropriate controls and corrections
By following these methodological approaches, researchers can obtain reliable kinetic parameters for P. marinus aroQ that facilitate comparison with other DHQDs and assessment of mutagenesis effects.
Multiple computational methods can be integrated to predict substrate specificity of P. marinus aroQ from its amino acid sequence:
Multiple sequence alignment (MSA): Align the P. marinus aroQ sequence with well-characterized DHQDs to identify conserved residues in the active site. Tools like Clustal Omega, MUSCLE, or T-Coffee can be used.
Motif identification: Search for conserved motifs associated with substrate binding using tools like MEME, PROSITE, or InterProScan.
Phylogenetic analysis: Construct phylogenetic trees to identify the closest characterized homologs, which may share similar specificity profiles. This helps place P. marinus aroQ within the context of known DHQD subfamilies.
Conservation analysis: Use tools like ConSurf to map conservation scores onto protein structures, highlighting functionally important residues.
Homology modeling: Generate a 3D structural model of P. marinus aroQ using templates from closely related DHQDs with known structures. Programs like SWISS-MODEL, Phyre2, or I-TASSER are suitable for this purpose.
Active site analysis: Identify the binding pocket and catalytic residues using CASTp, SiteMap, or similar tools. Compare with known DHQD structures to identify similarities and differences that might affect specificity.
Molecular docking: Dock potential substrates into the predicted active site using software like AutoDock, GOLD, or Glide to assess binding poses and energetics.
Molecular dynamics simulations: Perform MD simulations of the enzyme-substrate complex to assess stability and dynamics of binding interactions.
Quantitative structure-activity relationship (QSAR): Develop models relating enzyme sequence features to substrate specificity patterns.
Machine learning methods: Use supervised learning approaches trained on known DHQD data to predict specificity of novel enzymes.
Network analysis: Examine the genomic context and metabolic network to predict likely substrates based on pathway relationships.
Case example from research:
Structural comparison of CgDHQD with a homolog from Streptomyces coelicolor revealed differences in the terminal regions, lid loop, and active site that affect substrate binding . Particularly, CgDHQD possesses a distinctive proline residue (P105) not conserved in other DHQDs. Replacement of this residue significantly decreased enzyme activity, highlighting how computational identification of unique residues can guide experimental work .
By integrating these computational approaches, researchers can generate testable hypotheses about P. marinus aroQ substrate specificity and guide experimental design for biochemical characterization.
Integrating crystallographic structural data with biochemical findings provides comprehensive insights into P. marinus aroQ function through a multidisciplinary approach:
Structure determination and analysis:
Solve the crystal structure of P. marinus aroQ at high resolution
Identify key structural features: active site architecture, quaternary structure, mobile elements
When multiple structures are available (e.g., with/without ligands), analyze conformational changes
Structure-guided biochemical investigation:
Design site-directed mutagenesis experiments targeting residues identified in the structure
Perform enzyme kinetics on wild-type and mutant proteins
Investigate substrate specificity and inhibitor binding
Analyze the effects of pH, temperature, and buffer conditions on structure stability
Iterative refinement:
Use biochemical data to inform additional structural studies
Crystallize enzyme-substrate or enzyme-inhibitor complexes based on biochemical findings
Target mobile regions or alternative conformations identified through biochemistry
Example integration from similar DHQDs:
In studies of CgDHQD, researchers determined two crystal structures :
Wild-type CgDHQD with citrate (inhibitor) at 1.80 Å resolution
CgDHQD R19A mutant with DHQ (substrate) complexed at 2.00 Å resolution
These structures revealed:
The DHQ-binding site and interaction patterns
An unusual binding mode of citrate inhibitor with a half-opened lid loop
Structural variations compared to homologs from other species
Based on these structural insights, biochemical experiments were designed:
Mutations of specific residues (P105, S103) identified from structural comparisons
Activity assays showing ~70% decrease in activity when P105 was replaced with isoleucine or valine
10% increase in activity when S103 was replaced with threonine
The crystallography data provided the following parameters, which were essential for interpreting biochemical results:
| Parameter | Wild-type with Citrate | R19A Mutant with DHQ |
|---|---|---|
| Resolution (Å) | 1.80 | 2.00 |
| R-work / R-free | 15.8 / 20.3 | 14.2 / 16.9 |
| Number of protein atoms | 4372 | 6389 |
| B-factors (protein) | 25.1 | 15.5 |
| B-factors (ligand) | 32.5 (citrate) | 15.6 (DHQ) |
This integration of structural and biochemical data provided a comprehensive understanding of how specific residues contribute to enzyme function, explaining variation in reaction efficiency due to structural differences .