In most organisms, IspD and IspF are separate monofunctional enzymes, but in some bacteria, including Desulfovibrio vulgaris, the corresponding genes are fused, resulting in a single IspDF protein . The IspDF protein has two distinct domains, each responsible for one of the two enzymatic activities .
IspD Domain: Catalyzes the reaction of 2-C-methyl-D-erythritol 4-phosphate cytidylyltransferase .
IspF Domain: Catalyzes the reaction of 2-C-methyl-D-erythritol 2,4-cyclodiphosphate synthase . This enzyme catalyzes the conversion of 4-diphosphocytidyl-2-C-methyl-D-erythritol 2-phosphate (CDPME2P) to 2-C-methyl-D-erythritol 2,4-cyclodiphosphate (MEcDP) with the release of cytidine 5'-diphosphate (CMP) .
The Desulfovibrio vulgaris Hildenborough, where IspDF is found, is a model organism for studying the energy metabolism of sulfate-reducing bacteria (SRB) . It is also useful in understanding the economic impacts of SRB, including biocorrosion of metal infrastructure and bioremediation of toxic metal ions .
The MEP pathway involves several enzymatic steps :
The first step is catalyzed by 1-deoxy-D-xylulose-5-phosphate synthase (DXS).
Followed by 1-deoxy-D-xylulose 5-phosphate reductoisomerase (DXR/IspC).
Then, 2-C-methyl-D-erythritol 4-phosphate cytidyl transferase (IspD).
4-(cytidine 5′-diphospho)-2- C-methyl-D-erythritol kinase (IspE).
2- C-methyl-D-erythritol 2,4-cyclodiphosphate synthase (IspF).
2- C-methyl-D-erythritol 2,4-cyclodiphosphate reductase (IspG).
Recombinant IspDF enzyme can be produced in various host organisms, including E. coli, yeast, baculovirus, or mammalian cells . The purity of the recombinant enzyme is typically ≥ 85%, as determined by SDS-PAGE .
The ispG and ispH genes encode for the 2-C-methyl-D-erythritol 2,4-cyclodiphosphate reductase and the 4-hydroxy-3-methylbut-2-enyl diphosphate reductase, respectively . Flavodoxins were identified as compatible redox partners for the two DXP iron-sulfur enzymes, the suitable reductases included ferrodoxin, and flavodoxin NADP+ reductases ("FNRs") . The crucial determinant of heterogeneous function was solubility of an FNR in yeast cytoplasm . In general, if an FNR is soluble then it is likely to be compatible with flavodoxins .
The second enzyme of the MEP pathway, DXR, catalyzes a two-step reaction: the $$Mg^{2+}$$-triggered rearrangement of DXP into a non-isolable aldehyde and its concomitant NADPH-dependent reduction into MEP . DXR is a particularly well-described enzyme of the MEP pathway, with numerous protein crystal structures from several organisms, including important pathogens (e.g., E. coli, M. tuberculosis, Y. pestis) . In 1980, fosmidomycin, a phosphonic acid antibiotic, was identified as an inhibitor of DXR .
KEGG: dvu:DVU1454
STRING: 882.DVU1454
The bifunctional nature of IspD/IspF in D. vulgaris represents an interesting case of protein evolution where two enzymatic functions have been combined into a single polypeptide chain, in contrast to separate IspD and IspF enzymes found in many other bacterial species. This fusion offers several research advantages:
Pathway efficiency: The fusion potentially increases the efficiency of the MEP pathway by channeling intermediates between active sites without releasing them into the cytoplasm
Evolutionary perspective: Provides insights into the evolution of metabolic pathways and enzyme architecture
Drug target potential: Offers a unique target for antimicrobial development, as inhibitors could potentially disrupt two enzymatic steps simultaneously
Structure-function relationships: Allows study of how two catalytic domains maintain their individual functions while being part of the same protein
Researchers investigating this bifunctional enzyme should consider comparing its catalytic efficiency with the individual IspD and IspF enzymes from other organisms to understand potential advantages conferred by this fusion .
For optimal recombinant expression of D. vulgaris IspD/IspF:
Expression System:
Host: E. coli BL21(DE3) or similar expression strains
Vector: pET-based vectors with T7 promoter system
Tags: N-terminal His6-tag or alternative affinity tags (consider TEV protease cleavage site)
Culture Conditions:
Media: LB or 2×YT supplemented with appropriate antibiotics
Temperature: Induction at lower temperatures (16-18°C) overnight after reaching OD600 of 0.6-0.8
IPTG concentration: 0.1-0.5 mM (lower concentrations favor soluble protein)
Supplementation: Consider adding 0.5-1 mM ZnCl2 to the culture medium as the IspF domain requires zinc for activity
Lysis Buffer Components:
25-50 mM Tris-HCl, pH 8.0
100-300 mM NaCl
5-10% glycerol
1 mM DTT or 2-5 mM β-mercaptoethanol
Protease inhibitor cocktail
0.1-0.5 mM ZnCl2
Consider 0.1% Triton X-100 to aid solubilization
Careful optimization of these parameters is essential as the bifunctional nature of the enzyme may present folding challenges .
A multi-step purification strategy is recommended to obtain high-activity recombinant D. vulgaris IspD/IspF:
If His-tagged: Ni-NTA or IMAC chromatography
Buffer: 25 mM Tris-HCl pH 8.0, 300 mM NaCl, 10% glycerol
Imidazole gradient: 20 mM (wash), 50-300 mM (elution)
Include 0.1 mM ZnCl2 in all buffers to maintain IspF domain integrity
TEV protease cleavage (1:50 ratio, overnight at 4°C)
Second IMAC to remove cleaved tag and TEV protease
Q-Sepharose column (anion exchange)
Buffer: 25 mM Tris-HCl pH 8.0, 5% glycerol
NaCl gradient: 0-500 mM
Superdex 200 column
Buffer: 25 mM Tris-HCl pH 8.0, 100 mM NaCl, 5% glycerol, 1 mM DTT, 0.1 mM ZnCl2
Final Storage Conditions:
Concentrate to 1-5 mg/mL
Add glycerol to 25-50% final concentration
Flash-freeze in liquid nitrogen
Store at -80°C
Purity should be assessed by SDS-PAGE (>85% purity is generally acceptable), and activity of both domains should be verified using appropriate enzymatic assays to ensure the bifunctional enzyme has maintained its dual catalytic capabilities .
To separately measure the IspD and IspF activities in the bifunctional enzyme:
IspD Activity Assay:
Reaction mixture:
50 mM Tris-HCl (pH 8.0)
5 mM MgCl2
1-2 mM DTT
100-200 μM CTP
100-200 μM MEP
1-5 μg purified enzyme
Detection methods:
Coupled assay: Monitor release of pyrophosphate using a commercially available pyrophosphate detection kit
HPLC method: Separate and quantify CDP-ME formation using a C18 column with UV detection at 254 nm
Radioactive assay: Use [α-32P]CTP and measure CDP-ME formation by TLC or scintillation counting
IspF Activity Assay:
Reaction mixture:
50 mM MOPS (pH 8.0)
5 mM MgCl2
100 μM CDP-ME2P (substrate)
1 mM phosphatase inhibitor
1-5 μg purified enzyme
Detection methods:
Spectrophotometric method: Monitor CMP release using an auxiliary enzyme system that converts CMP to uridine with subsequent UV detection
Mass spectrometry: Analyze ME-CPP formation directly using an Agilent 6210 mass spectrometer after reaction termination with 10 mM EDTA
HPLC method: Detect ME-CPP formation using appropriate column separation
When analyzing the bifunctional enzyme, it's crucial to run appropriate controls and consider the potential for substrate channeling between domains. For comprehensive characterization, determine kinetic parameters (Km, kcat, kcat/Km) for both activities and compare them to the individual enzymes from other species .
Based on related research with IspD and IspF enzymes from various bacterial species, the expected kinetic parameters for D. vulgaris bifunctional IspD/IspF activities typically fall within these ranges:
| Parameter | Substrate | Expected Range | Notes |
|---|---|---|---|
| Km | MEP | 20-100 μM | May vary with assay conditions |
| Km | CTP | 50-200 μM | ATP can sometimes substitute but with lower efficiency |
| kcat | - | 1-10 s-1 | Temperature dependent (30-37°C) |
| kcat/Km (MEP) | MEP | 104-105 M-1s-1 | Indicator of catalytic efficiency |
| Optimal pH | - | 7.5-8.5 | Often peaks around pH 8.0 |
| Parameter | Substrate | Expected Range | Notes |
|---|---|---|---|
| Km | CDP-ME2P | 10-50 μM | Often lower than IspD substrates |
| kcat | - | 0.1-5 s-1 | Generally slower than IspD activity |
| kcat/Km | CDP-ME2P | 103-104 M-1s-1 | Lower than reported for M. tuberculosis IspF |
| Zinc dependency | - | Required | Activity diminishes significantly with EDTA |
Important considerations:
The bifunctional nature may affect kinetic parameters compared to the individual enzymes
Substrate channeling between domains might result in apparent kinetic parameters that differ from those of separate enzymes
Optimal conditions for simultaneously measuring both activities may require compromise conditions
The kcat/Km value of M. tuberculosis IspF was reported as 5.4×10-4 μM-1min-1, which might serve as a comparison point
When characterizing newly purified recombinant enzyme, establish full kinetic profiles under standardized conditions and compare them with published values for related enzymes .
While the precise crystal structure of D. vulgaris bifunctional IspD/IspF has not been fully determined in the provided search results, comparative structural analysis can be inferred from related enzymes:
Structural Organization:
Individual IspD enzymes typically form dimers with each monomer comprising a single Rossmann fold domain
Individual IspF enzymes form homotrimers with a zinc-binding site at each active site
The bifunctional enzyme likely maintains these structural motifs while connecting them via a linker region
Domain Architecture:
N-terminal region: Contains the IspD domain with the characteristic nucleotide-binding Rossmann fold
Central linker region: Provides flexibility between domains
C-terminal region: Contains the IspF domain that belongs to the IspF family
Key Structural Features:
Active Sites:
IspD active site: Contains conserved residues for MEP and CTP binding
IspF active site: Features a zinc-binding site with conserved histidine and aspartate residues
Quaternary Structure:
The bifunctional enzyme likely forms oligomeric structures to maintain the functional states of both domains
IspF domain may still form trimeric arrangements while being part of the fusion protein
Substrate Channeling:
The fusion nature suggests possible substrate channeling between domains
The linker region length and flexibility would be crucial determinants
Structural Differences from Individual Enzymes:
Potential constraints on domain movement due to the covalent linkage
Possible alterations in oligomerization patterns
Interface regions that may affect substrate access or product release
For definitive structural characterization, X-ray crystallography or cryo-EM studies would be necessary, focusing particularly on the domain interface and potential substrate-channeling pathways .
Zinc plays a critical role in the function of the IspF domain within the bifunctional IspD/IspF enzyme:
Structural Role:
Catalytic Functions:
Substrate Binding:
Zinc coordinates with the phosphate groups of CDP-ME2P
This coordination positions the substrate optimally for catalysis
Activation of Water:
Zinc acts as a Lewis acid, activating a water molecule for nucleophilic attack
This facilitates the cyclization reaction converting CDP-ME2P to ME-CPP
Transition State Stabilization:
The zinc ion helps stabilize negative charges that develop during the reaction
Experimental Evidence:
Treatment with EDTA (a metal chelator) significantly reduces or abolishes IspF activity
Addition of Zn2+ ions restores activity to the metal-depleted enzyme
When expressing and purifying recombinant IspD/IspF, supplementation with ZnCl2 (0.1-1 mM) is recommended in culture media and buffers
Practical Considerations for Researchers:
Include 0.1-0.5 mM ZnCl2 in all purification and storage buffers
Avoid high concentrations of reducing agents that might displace zinc
When measuring IspF activity, ensure zinc is present in assay buffers
Control experiments with EDTA can confirm zinc dependency
The presence of zinc should be confirmed in purified enzyme preparations, potentially using atomic absorption spectroscopy or inductively coupled plasma mass spectrometry (ICP-MS) .
Introducing targeted mutations into the D. vulgaris ispDF gene requires specialized techniques due to the anaerobic nature of this organism and its genetic characteristics. Here are methodological approaches:
1. Markerless Deletion System Approach:
The most effective method utilizes the markerless deletion system developed for D. vulgaris Hildenborough:
Step 1: Generate a Δupp host strain (like JW710) that is resistant to 5-fluorouracil (5-FU)
Step 2: Create a suicide plasmid containing:
Upstream and downstream regions of the target mutation site
The wild-type upp gene under control of a constitutive promoter (e.g., aph(3′)-II)
Spectinomycin resistance marker
Your desired mutation within the ispDF gene
Step 3: Introduce the plasmid via electroporation and select for spectinomycin resistance
Step 4: Verify single-crossover integration via PCR
Step 5: Grow without spectinomycin to allow second recombination
Step 6: Select for 5-FU resistance (loss of upp)
Step 7: Screen colonies for the desired mutation by PCR and sequencing
2. Site-Directed Mutagenesis Protocol:
For structure-function studies, use this approach to create specific amino acid substitutions:
Design primers containing your desired mutation with 15-20 bp flanking sequences
Use a template containing the ispDF gene from D. vulgaris
Perform PCR with high-fidelity polymerase
Treat with DpnI to digest methylated template DNA
Transform into E. coli
Verify mutations by sequencing
Transfer the mutated gene back to D. vulgaris using the markerless system above
Critical Considerations:
D. vulgaris has a restriction-modification system that can decrease transformation efficiency
Consider using a restriction-deficient strain (e.g., JW7035) that has ~100-1000× higher transformation efficiency
The transformation efficiency for D. vulgaris is typically low (2-5 transformants/μg DNA for wild-type)
All manipulations must account for the organism's anaerobic requirements
This approach allows for precise genetic manipulation without antibiotic marker retention, enabling sequential mutations if needed .
Based on structural and functional studies of IspD and IspF enzymes from other organisms, several critical domains and residues in the D. vulgaris bifunctional enzyme merit targeted investigation:
IspD Domain Critical Regions:
CTP-Binding Site:
Conserved glycine-rich motif (likely within first 100 amino acids)
Lysine residues involved in phosphate coordination
Target mutations: Conserved basic residues that interact with CTP phosphates
MEP-Binding Pocket:
Residues interacting with the erythritol moiety
Target mutations: Conserved polar residues in the catalytic pocket
Catalytic Residues:
Conserved aspartate or glutamate residues that may coordinate magnesium
Target mutations: Acidic residues in the active site
IspF Domain Critical Regions:
Zinc-Binding Site:
Conserved histidines and aspartate that coordinate the zinc ion
Based on homology, likely includes residues in the C-terminal portion
Target mutations: His→Ala substitutions to disrupt zinc binding
CDP-ME2P Binding Pocket:
Residues interacting with cytidine moiety
Residues stabilizing the cyclization transition state
Target mutations: Conserved aromatic residues that stack with cytidine
Dimer/Trimer Interface:
Residues involved in oligomerization
Target mutations: Hydrophobic residues at subunit interfaces
Domain Interface Region:
Linker Region:
Flexibility may impact substrate channeling
Target mutations: Proline insertions to reduce flexibility or linker length modifications
Interdomain Contacts:
Potential residues participating in communication between domains
Target mutations: Interface residues to disrupt potential allosteric regulation
Experimental Approach Table:
| Target Region | Suggested Mutations | Expected Effect | Analytical Method |
|---|---|---|---|
| CTP binding | K→A in conserved motifs | Decreased IspD activity | CTP binding assay, IspD activity |
| Zinc binding | H→A in C-terminal domain | Decreased IspF activity | Zinc content analysis, IspF activity |
| Linker region | Deletion or insertion | Altered substrate channeling | Compare kinetics of full reaction vs. individual steps |
| Oligomerization interfaces | Hydrophobic→charged | Altered quaternary structure | Size exclusion chromatography, activity assays |
These structure-function studies would provide valuable insights into how the bifunctional enzyme coordinates its dual catalytic activities and whether substrate channeling occurs between domains .
While specific inhibitors targeting the D. vulgaris bifunctional IspD/IspF have not been extensively documented in the provided search results, we can extrapolate from studies on related IspD and IspF enzymes:
IspD Inhibitors:
Cytidine Nucleotide Analogs:
Mechanism: Competitive inhibition by mimicking CTP substrate
Examples: Cytosine derivatives with modified ribose or phosphate groups
Typical IC50 range: 1-100 μM
MEP-Mimetics:
Mechanism: Competitive inhibition at the MEP-binding site
Examples: Erythritol derivatives with phosphonate groups
Binding features: Interact with conserved residues in the MEP-binding pocket
Fosmidomycin Derivatives:
While fosmidomycin primarily targets DXR (an earlier enzyme in the MEP pathway), some derivatives have shown activity against IspD
Mechanism: Likely mixed inhibition
IspF Inhibitors:
Fragment-Based Hits:
FOL7185: Identified as a fragment hit binding to IspD and IspE
FOL955: Co-crystallized with Burkholderia pseudomallei IspF (PDB: 3QHD)
FOL535: Co-crystallized with BpIspF (PDB: 3K14)
Imidazole Compounds:
Designed based on fragment hit FOL955
Likely mechanism: Interaction with zinc-binding site
Imidazothiazole Compounds:
Designed based on FOL535
Modifications include replacing ethyl ester with various substituents
Some designed to engage both zinc and cytidine binding sites
L-Tryptophan Hydroxamate:
Used as a standard in BpIspF assays
Mechanism: Likely involves zinc coordination
Bifunctional Inhibition Potential:
The bifunctional nature of D. vulgaris IspD/IspF presents unique opportunities for inhibitor design:
Dual-Target Inhibitors:
Compounds designed to simultaneously inhibit both catalytic functions
Potential for increased potency due to proximity of active sites
Substrate-Channeling Disruptors:
Compounds that interfere with the transfer of CDP-ME between domains
May bind at the domain interface
Allosteric Inhibitors:
Target regions that affect communication between domains
Could potentially modulate one activity through binding to the other domain
For developing selective inhibitors against D. vulgaris IspD/IspF, researchers should consider screening the above-mentioned compound classes while focusing on unique structural features of this bifunctional enzyme .
For effective identification of novel inhibitors targeting the D. vulgaris bifunctional IspD/IspF enzyme, a multi-tiered screening approach is recommended:
Primary Screening Assays:
High-Throughput Spectrophotometric Assays:
For IspD activity: Pyrophosphate-coupled assay using commercial pyrophosphate detection systems
For IspF activity: CMP release detection using auxiliary enzymes and spectrophotometric readout
Advantages: Quick, continuous measurement, suitable for 96/384-well formats
Throughput: 10,000-100,000 compounds per day
Fluorescence-Based Assays:
Dansyl-containing fluorescent compounds that interact with the binding sites
FRET-based assays to detect conformational changes upon inhibitor binding
Advantages: Higher sensitivity, fewer false positives from colored compounds
Throughput: Similar to spectrophotometric assays
Secondary Confirmation Assays:
Mass Spectrometry-Based Assays:
Direct detection of reaction products (CDP-ME or ME-CPP)
Lower throughput but higher specificity
Example protocol: Reaction mixtures containing 50 mM MOPS (pH 8.0), 5 mM MgCl2, 100 μM substrate, and test compound, incubated with enzyme, terminated with EDTA, then analyzed by MS
Thermal Shift Assays:
Measure changes in protein thermal stability upon inhibitor binding
Quick verification of direct binding
Can distinguish between inhibitors targeting different domains
Tertiary Mechanistic Characterization:
Enzyme Kinetics:
Determine inhibition mechanisms (competitive, non-competitive, uncompetitive)
Generate Ki values and investigate time-dependent inhibition
Biophysical Methods:
Surface Plasmon Resonance (SPR): Direct measurement of binding kinetics
Isothermal Titration Calorimetry (ITC): Thermodynamic parameters of binding
X-ray Crystallography: Structural confirmation of binding mode
Fragment-Based Approaches:
Given the success with fragment hits like FOL7185, FOL955, and FOL535 for related enzymes:
Fragment Library Screening:
Start with low molecular weight compounds (150-300 Da)
Screen at higher concentrations (100 μM - 1 mM)
Use NMR or X-ray crystallography to confirm binding
Fragment Growing/Linking:
Develop fragments that bind to adjacent sites
Link or merge fragments to create more potent inhibitors
Computational Approaches:
Virtual Screening:
Homology model-based if crystal structure unavailable
Dock compound libraries against both active sites
Structure-Based Design:
Leverage knowledge from related IspD and IspF structures
Focus on unique features of the bifunctional enzyme
For optimal results, combine these approaches in an integrated workflow, starting with high-throughput methods for initial screening, followed by confirmatory assays and mechanistic studies for promising hits .
The bifunctional IspD/IspF enzyme plays a significant role in D. vulgaris pathogenicity through several mechanisms:
Metabolic Contribution to Virulence:
Isoprenoid Biosynthesis:
IspD/IspF is essential for the MEP pathway producing isoprenoid precursors
Isoprenoids are critical for:
Cell membrane integrity and function
Electron transport chain components (menaquinones)
Cell wall biosynthesis
Survival in Host Environments:
Enables adaptation to varying nutrient conditions
Supports growth in anaerobic gut environments
Direct Pathogenic Mechanisms:
Hydrogen Sulfide (H2S) Production:
D. vulgaris produces H2S as a metabolic byproduct
H2S disrupts gut epithelial morphology and function
Higher H2S levels have been observed in D. vulgaris-treated mice compared to controls
Metabolic functions supported by the MEP pathway indirectly contribute to H2S production capacity
Inflammatory Response Induction:
D. vulgaris transplantation in mice causes:
Gut inflammation
Disruption of gut barrier function
Reduced levels of short-chain fatty acids (SCFAs)
Increased expression of inflammatory cytokines (IL-1β, iNOS, TNF-α)
Decreased expression of anti-inflammatory markers (IL-10, arginase 1)
Microbiome Disruption:
D. vulgaris significantly alters gut microbiota composition
Decreases relative abundance of SCFA-producing bacteria
Stimulates growth of Akkermansia muciniphila, possibly via H2S production
These changes may be supported by metabolic adaptability enabled by the MEP pathway
Experimental Evidence:
In mouse models, D. vulgaris exacerbates DSS-induced colitis
It causes persistent epithelial damage and reduced mucus levels
It aggravates DSS-induced damage to gut epithelial barriers by decreasing expression of E-cadherin, Occludin, and ZO-1
These pathogenic effects depend on D. vulgaris metabolic activities, which are supported by essential pathways including the MEP pathway
While the direct contribution of IspD/IspF to these mechanisms has not been specifically isolated in the provided research, its essential role in isoprenoid biosynthesis suggests it is necessary for D. vulgaris survival and pathogenicity. Targeting this enzyme could potentially attenuate D. vulgaris virulence by compromising its metabolic capabilities .
The bifunctional IspD/IspF enzyme offers a valuable target for genetic manipulation to study D. vulgaris in complex microbial communities, providing insights into its ecological roles and interactions:
Genetic Manipulation Strategies:
Conditional Expression Systems:
Replace native ispDF promoter with inducible promoters
Allow controlled expression under specific conditions
Methodology: Use the markerless deletion system with upp counterselection
Reporter Gene Fusions:
Create transcriptional/translational fusions with reporter genes (GFP, luciferase)
Monitor ispDF expression in different environments or community contexts
Methodology: Insert reporter gene downstream of ispDF while maintaining function
Gene Knockdown Approaches:
Create partial loss-of-function variants through targeted mutations
Develop antisense RNA constructs targeting ispDF mRNA
Methodology: Introduce mutations in catalytic residues using site-directed mutagenesis
Applications in Microbial Ecology Research:
Tracking D. vulgaris in Complex Communities:
Experimental design: Create fluorescent reporter strains with ispDF promoter-driven fluorescence
Application: Track spatial distribution and metabolic activity in multi-species biofilms
Analysis: Correlate ispDF expression with community structure using microscopy and flow cytometry
Understanding Metabolic Interactions:
Experimental design: Create conditional ispDF mutants with varying expression levels
Application: Study how D. vulgaris isoprenoid metabolism affects community composition
Analysis: Use 16S rRNA amplicon sequencing to track microbiome shifts when ispDF is modulated
Host-Microbe Interaction Studies:
Experimental design: Create D. vulgaris strains with mutated ispDF affecting catalytic efficiency
Application: Colonize gnotobiotic animals with mutant strains
Analysis: Measure inflammatory markers, gut barrier integrity, and microbiome composition
Competitive Fitness Assessment:
Experimental design: Co-culture wild-type and ispDF-modified strains in different environments
Application: Determine fitness costs of ispDF mutations
Analysis: Use quantitative PCR with strain-specific primers to track population dynamics
Technical Implementation Table:
| Research Goal | Genetic Manipulation Approach | Key Methodological Considerations | Analytical Methods |
|---|---|---|---|
| Community tracking | Fluorescent reporter fusion | Ensure reporter doesn't disrupt function | Fluorescence microscopy, flow cytometry |
| Metabolic impact | Conditional expression | Tight regulation of inducible promoter | Metabolomics, 16S sequencing |
| Virulence studies | Catalytic site mutations | Target residues that reduce but don't eliminate function | Host response measurements, histology |
| Ecological niche | Competitive assays with marked strains | Ensure genetic markers don't affect fitness | qPCR, amplicon sequencing |
Practical Considerations:
Use the JW7035 strain with higher transformation efficiency (100-1000× greater than wild-type)
Account for D. vulgaris' anaerobic requirements in all experimental procedures
Consider potential polar effects of genetic manipulations on downstream genes
Validate all genetic constructs through sequencing and functional assays before community studies
These approaches capitalize on the markerless genetic exchange system developed for D. vulgaris, allowing for sophisticated genetic manipulations without antibiotic marker retention .
The D. vulgaris bifunctional IspD/IspF enzyme presents an interesting case for comparative analysis with homologous enzymes from other bacteria:
Structural Organization Comparison:
| Organism | Enzyme Structure | Molecular Weight | Oligomeric State | Unique Features |
|---|---|---|---|---|
| D. vulgaris | Bifunctional IspD/IspF | 41.7 kDa | Likely complex oligomer | Fusion of two enzymatic functions |
| E. coli | Separate IspD and IspF | IspD: ~25 kDa, IspF: ~17 kDa | IspD: dimer, IspF: trimer | Individual proteins, potential for complex formation |
| M. tuberculosis | Separate IspD and IspF | Similar to E. coli | Similar to E. coli | Lower kcat/Km for IspF (5.4×10-4 μM-1min-1) |
| B. subtilis | Separate IspD and IspF | Similar to E. coli | Similar to E. coli | No detectable protein complex formation among IspD, IspE, and IspF |
Functional Differences:
Catalytic Efficiency:
Separate enzymes may have different kinetic properties compared to the bifunctional enzyme
Substrate channeling in the bifunctional enzyme potentially increases efficiency
M. tuberculosis IspF shows relatively low catalytic efficiency
Protein-Protein Interactions:
Studies with B. subtilis enzymes show no detectable complex formation among IspD, IspE, and IspF
D. vulgaris bifunctional enzyme inherently links IspD and IspF functions
E. coli may form transient complexes of pathway enzymes
Regulatory Mechanisms:
Differential regulation possible between bifunctional and separate enzymes
D. vulgaris likely has coordinated expression and regulation of both functions
Separate enzymes allow independent regulation in other bacteria
Evolutionary Implications:
Gene Fusion Events:
The bifunctional enzyme likely arose through gene fusion events
May represent adaptation to specific ecological niches
Potential for improved metabolic efficiency through proximity of catalytic domains
Taxonomic Distribution:
Bifunctional arrangement appears less common than separate enzymes
May correlate with specific metabolic adaptations in sulfate-reducing bacteria
Sequence Conservation:
The 395-amino acid D. vulgaris IspD/IspF sequence shows recognizable domains corresponding to both IspD and IspF functions, with the C-terminal section belonging to the IspF family. Comparative sequence analysis would likely show:
Higher conservation in catalytic residues
Variable regions in domain interfaces
Unique linker regions in the bifunctional enzyme
The bifunctional nature of D. vulgaris IspD/IspF represents an interesting case of potential evolutionary optimization, possibly providing advantages through substrate channeling or coordinated regulation of consecutive pathway steps .
Analyzing the genomic context of the ispDF gene in D. vulgaris provides valuable insights into its regulation, evolution, and functional relationships within metabolic networks:
Genomic Organization:
Operon Structure:
Determine if ispDF is part of a larger operon containing other MEP pathway genes
Identify co-transcribed genes that may reveal functional associations
Compare with E. coli where separate ispD and ispF genes are often found in different genomic locations
Regulatory Elements:
Analyze upstream regions for promoter elements and transcription factor binding sites
Look for conserved DNA motifs that may be involved in regulation
One study derived conserved DNA motifs from potential promoter regions of putative D. vulgaris regulons, although most were apparently unique compared to E. coli
Comparative Genomic Analysis:
Taxonomic Distribution:
Compare ispDF arrangement across related Desulfovibrio species
Determine if the bifunctional arrangement is conserved within Deltaproteobacteria
Map the evolutionary history of gene fusion events
Synteny Analysis:
Examine conservation of gene order in regions flanking ispDF
Identify genomic rearrangements that may have led to the current organization
Compare with other sulfate-reducing bacteria
Functional Genomic Insights:
Co-Expression Patterns:
Analyze transcriptomic data to identify genes co-regulated with ispDF
Look for condition-specific expression patterns that reveal functional importance
Traditional laboratory studies of sulfate-reducing bacteria have focused on biochemistry, but genomic sequences now allow insights into metabolic and regulatory networks
Metabolic Context:
Map connections between the MEP pathway and other metabolic pathways in D. vulgaris
Identify potential metabolic dependencies or regulatory cross-talk
Understand how isoprenoid biosynthesis integrates with core metabolism
Regulatory Network Analysis:
Transcription Factor Binding:
The GlpR binding site (similar to E. coli) was identified in D. vulgaris
Most regulatory motifs in D. vulgaris appear unique compared to E. coli
Some expected orthologs for regulatory proteins have not been recognized in D. vulgaris
Response to Environmental Conditions:
Analyze expression changes under different growth conditions
Identify factors that upregulate or downregulate ispDF expression
Connect to the organism's ecological niche and pathogenic potential
Practical Applications:
Genetic Tool Development:
Use promoter elements for development of expression systems
Identify suitable regions for genetic manipulation
D. vulgaris has seen significant improvements in genetic manipulation capabilities in recent years
Target Validation:
Assess essentiality based on genomic context and lack of redundant pathways
Evaluate the ispDF gene as a potential antimicrobial target
Understanding the genomic context of ispDF provides crucial insights for both fundamental research and applied studies targeting this enzyme, helping to place it within the broader metabolic and regulatory framework of D. vulgaris .
CRISPR-based technologies offer powerful new approaches for studying ispDF function in D. vulgaris, though they require adaptation for this anaerobic organism:
CRISPR-Cas9 Gene Editing Approaches:
Genome Editing Strategy:
Design guide RNAs (gRNAs) targeting specific regions of ispDF
Deliver CRISPR components via electroporation using optimized protocols for D. vulgaris
Consider using the restriction-deficient strain JW7035 for higher transformation efficiency
Include appropriate homology-directed repair (HDR) templates for precise modifications
Specific Applications:
Domain-specific knockouts: Target either IspD or IspF domain while keeping the other intact
Point mutations: Create catalytic residue mutations to study domain-specific functions
Tagged versions: Insert epitope or fluorescent tags for localization and interaction studies
Promoter modifications: Alter expression levels by modifying regulatory regions
CRISPR Interference (CRISPRi) for Conditional Regulation:
System Design:
Implement dCas9 (catalytically dead Cas9) under control of inducible promoter
Design gRNAs targeting the ispDF promoter or coding region
Optimize for D. vulgaris codon usage and regulatory elements
Experimental Applications:
Tunable repression: Create partial knockdown phenotypes
Temporal control: Study effects of ispDF depletion at different growth phases
Spatial regulation: Control expression in different microenvironments
CRISPR Activation (CRISPRa) for Upregulation:
System Components:
dCas9 fused to transcriptional activators adapted for D. vulgaris
gRNAs targeting ispDF promoter region
Applications:
Overexpression studies: Evaluate effects of increased IspD/IspF levels
Metabolic engineering: Potentially enhance isoprenoid production
Stress response analysis: Test if upregulation confers survival advantages
Base and Prime Editing Applications:
Precise Nucleotide Modifications:
Implement cytosine or adenine base editors for targeted mutations without DSBs
Create specific codon changes to alter enzyme properties
Applications:
Structure-function analysis: Create libraries of variants with specific amino acid changes
Domain interface modifications: Alter interdomain communication
Regulatory element tuning: Modify promoter strength through precise nucleotide changes
Technical Implementation Considerations:
| Approach | Technical Requirements | Key Adaptations for D. vulgaris | Expected Challenges |
|---|---|---|---|
| CRISPR-Cas9 editing | Cas9, gRNA, repair template | Anaerobic conditions for all steps, optimized transformation protocol | Low transformation efficiency, high off-target effects |
| CRISPRi | dCas9, gRNA expression system | D. vulgaris-compatible inducible promoters | Achieving sufficient repression levels |
| CRISPRa | dCas9-activator fusion, gRNAs | Engineering activator domains functional in D. vulgaris | Potential toxicity of overexpression |
| Base editing | Base editor variants, gRNAs | Efficiency testing in anaerobic conditions | Limited editing window may restrict targetable sites |
The patent information (search result ) indicates developing CRISPR systems for modifying bacterial genomic sequences, suggesting these approaches are becoming more feasible for organisms like D. vulgaris. Researchers should optimize protocols specifically for this anaerobic bacterium, potentially using the improved markerless genetic exchange system as a foundation .
Several cutting-edge technologies hold promise for deepening our understanding of the D. vulgaris bifunctional IspD/IspF enzyme:
Advanced Structural Biology Approaches:
Cryo-Electron Microscopy (Cryo-EM):
Application: Determine high-resolution structures of the complete bifunctional enzyme
Advantage: Captures multiple conformational states without crystallization constraints
Key insight potential: Visualize dynamic domain movements during catalysis
Implementation: Single-particle analysis with preferably >300kV microscopes
AlphaFold2 and Machine Learning Structure Prediction:
Application: Generate accurate structural models of the bifunctional enzyme
Advantage: Complements experimental structures, predicts flexible regions
Key insight potential: Model domain interfaces and substrate channeling pathways
Implementation: Use multiple sequence alignments from diverse Desulfovibrio species
Time-Resolved X-ray Crystallography:
Application: Capture intermediates during catalytic cycles
Advantage: Provides dynamic "snapshots" of enzyme function
Key insight potential: Elucidate transition states for inhibitor design
Implementation: X-ray free-electron laser (XFEL) facilities for femtosecond imaging
Advanced Functional and Interaction Analysis:
Single-Molecule Förster Resonance Energy Transfer (smFRET):
Application: Track conformational dynamics between domains
Advantage: Observes heterogeneous behaviors masked in ensemble measurements
Key insight potential: Directly observe substrate channeling between domains
Implementation: Strategic fluorophore placement at domain interfaces
Hydrogen-Deuterium Exchange Mass Spectrometry (HDX-MS):
Application: Map protein dynamics and ligand-induced conformational changes
Advantage: Provides regional dynamics information without structure determination
Key insight potential: Identify allosteric networks between domains
Implementation: Compare exchange patterns in various ligand-bound states
Native Mass Spectrometry:
Application: Determine oligomeric states and complex formation
Advantage: Preserves non-covalent interactions
Key insight potential: Understand quaternary structure arrangements
Implementation: Analyze intact enzyme complexes with substrates/inhibitors
Emerging Chemical Biology Approaches:
Chemoproteomics:
Application: Identify binding sites and off-targets of inhibitors
Advantage: Whole-proteome scope of interactions
Key insight potential: Discover unexpected cross-reactivity of inhibitors
Implementation: Activity-based protein profiling with clickable probes
DNA-Encoded Libraries (DELs):
Application: Screen billions of compounds for binding
Advantage: Massive compound diversity with minimal material
Key insight potential: Discover novel chemical scaffolds targeting specific domains
Implementation: Target enzyme immobilization and selection protocols
Fragment-Based Drug Discovery with NMR:
Application: Identify fragment hits against both domains
Advantage: Maps precise binding locations of weakly binding fragments
Key insight potential: Develop domain-specific inhibitors
Implementation: 19F NMR and protein-observed experiments
Systems Biology and Metabolic Integration: