This article addresses the critical challenge of synthetic intractability in natural product development, a major bottleneck in harnessing their potential for drug discovery.
This article addresses the critical challenge of synthetic intractability in natural product development, a major bottleneck in harnessing their potential for drug discovery. Aimed at researchers, scientists, and drug development professionals, it provides a comprehensive roadmap from foundational concepts to advanced applications. We explore the computational and chemical roots of intractability, detail cutting-edge methodological solutions like Fragment-Based Drug Discovery (FBDD) and CâH activation, and offer troubleshooting strategies for optimization. The scope also covers rigorous validation techniques and comparative analysis of traditional versus modern approaches, synthesizing key insights to guide the development of previously 'undruggable' targets into viable clinical candidates.
Q1: What does it mean for a problem in network biology to be NP-hard? An NP-hard problem is at least as hard as the hardest problems in the NP (Non-deterministic Polynomial time) class [1]. In practical terms for researchers, this means that as your biological network (like a protein-protein interaction or metabolic network) grows in size, the time required to find an exact solution increases exponentially rather than polynomially. For example, finding a minimum set of genes that control a metabolic pathway is often an NP-hard problem.
Q2: I need to find an optimal set of drug targets in a metabolic network. Why does my computation time become unmanageable? You are likely facing an NP-hard problem. The number of possible subsets of genes or proteins to test grows combinatorially with the size of your network. Verifying a given solution might be quick, but exhaustively checking all possible solutions to find the best one is computationally intractable for large networks [1]. This is a classic characteristic of NP-hard problems.
Q3: Are all complex problems in network biology NP-hard? No. Many complex problems are tractable (in class P), meaning they can be solved in polynomial time [1]. However, key problems central to natural product research are NP-hard, such as identifying the optimal scaffold for a synthetic pathway or finding the most influential nodes in a large gene regulatory network. Determining your problem's complexity class is the first troubleshooting step.
Q4: What practical strategies can I use to overcome synthetic intractability in my research? Since exact solutions for NP-hard problems are often impractical, researchers employ strategies like:
Q5: How can I verify that my heuristic solution for a network biology problem is reliable? While you cannot easily check if a heuristic found the best solution, you can validate its biological plausibility and robustness. Techniques include:
Symptoms:
Diagnosis: The analysis task is likely NP-hard. The algorithm is probably attempting to find an exact solution by exploring a solution space that grows exponentially.
Resolution:
Symptoms:
Diagnosis: This is a common indicator of using heuristic solvers on a complex, multi-modal fitness landscape, which is typical for NP-hard problems. Different algorithms may get stuck in different local optima.
Resolution:
Symptoms:
Diagnosis: The algorithm's memory usage is likely growing exponentially or quadratically with network size, which is unsustainable for large networks.
Resolution:
Objective: To find a near-optimal set of critical nodes (e.g., drug targets) in a large-scale biological network using a greedy heuristic.
Workflow:
Methodology:
Objective: To experimentally validate computationally predicted drug targets from an NP-hard optimization in a microbial system.
Workflow:
Methodology:
The following table details key materials and computational tools used in the featured experiments.
| Item Name | Function in Research | Specification / Notes |
|---|---|---|
| Gurobi Optimizer | Solver for mathematical programming problems (MIP, QP). | Used to find exact solutions for small-scale instances of NP-hard problems; provides a benchmark for heuristics. |
| Cytoscape | Open-source platform for network visualization and analysis. | Used to import, visualize, and pre-process biological network data before formal computational analysis. |
| NetworkX Library | Python package for the creation, manipulation, and study of complex networks. | Used to implement custom heuristic algorithms and perform network topology analysis (e.g., centrality, connectivity). |
| Specific Enzyme Inhibitors | Compounds used to experimentally perturb predicted targets in a metabolic network. | Must be highly specific to the predicted enzyme target to minimize off-target effects during validation. |
| LC-MS System | Analytical chemistry technique for identifying and quantifying metabolites. | Used in metabolomic validation protocols to measure the biochemical outcome of target inhibition. |
The table below summarizes key complexity classes and their relevance to computational biology, providing a framework for diagnosing computational challenges [1].
| Complexity Class | Key Characteristic | Example in Network Biology |
|---|---|---|
| P | Easily solvable in polynomial time [1]. | Calculating the shortest path between two nodes in a network. |
| NP | "Yes" answers can be verified in polynomial time, but finding a solution may be hard [1]. | Verifying that a given set of genes is a vertex cover for a protein interaction network. |
| NP-hard | At least as hard as the hardest problems in NP. All NP-hard problems are not in NP [1]. | Finding the smallest possible set of genes that is a vertex cover (Vertex Cover Problem). |
| NP-complete | A problem that is both NP and NP-hard; these are the hardest problems in NP [1]. | The Boolean Satisfiability Problem (SAT), which can model many network logic problems. |
What does 'synthetically intractable' mean in natural product chemistry? A compound is deemed synthetically intractable when its complex molecular architecture presents overwhelming challenges for efficient chemical synthesis in a laboratory. These hurdles can include densely packed functional groups, numerous chiral centers, or unusually reactive and unstable structures that make traditional synthetic routes too long, inefficient, or low-yielding to be practical for development [2] [3].
Why is overcoming synthetic intractability important for drug discovery? Natural products are a historic and enduring source of chemical information for medicine [3]. From 1981 to 2019, 41.9% of all new FDA-approved drugs were derived from natural sources [4]. Overcoming synthetic intractability is crucial because it unlocks access to these biologically validated, complex scaffolds that often possess unique therapeutic activities, such as in oncology, antimicrobials, and antifungals [4].
My natural product target has a very rigid, complex core. What synthetic strategy should I consider? For complex, rigid cores, you should investigate strategies that prioritize the essential functional elements for bioactivity. Biology-Oriented Synthesis (BIOS) is particularly valuable here, as it uses the natural product's core as a "privileged" starting point for designing a more synthetically accessible library of analogues that retain the core's biological relevance [3].
I've identified a promising fragment hit, but it's synthetically challenging to elaborate. What can I do? This is a common challenge. The strategy of Fragment-Based Drug Discovery (FBDD), as pioneered by companies like Astex Pharmaceuticals, directly addresses this. It involves investing in innovative synthetic organic chemistry methodologies specifically designed to elaborate polar, unprotected fragments. Early consideration of synthetic feasibility is as critical as optimizing binding affinity for progressing a fragment hit [2].
What analytical tools can help me characterize complex natural product mixtures without full isolation? Advanced NMR and mass spectrometry platforms are designed for this exact challenge.
Potential Cause: Excessive Structural Complexity and Steric Hindrance The target molecule may possess a high density of functional groups or stereocenters in a confined space, leading to severe steric hindrance that prevents key reactions from proceeding.
Solution Checklist:
Potential Cause: Limited Quantity of Isolated Material or Complex NMR Spectra Traditional 1D NMR may be insufficient for determining the relative configuration of multiple chiral centers in a complex molecule, especially in a mixture.
Solution Checklist:
Potential Cause: Inefficient Prioritization of Novel Chemistry from Complex Extracts Time and resources are wasted on the isolation and structural elucidation of already-known natural products.
Solution Checklist:
Potential Cause: The fragment core lacks straightforward synthetic handles for chemical elaboration, making optimization prohibitively difficult.
Solution Checklist:
This protocol is designed for the untargeted analysis and dereplication of natural products in complex mixtures using 2D NMR.
1. Sample Preparation and Data Acquisition:
2. MADByTE Data Processing Workflow: The following diagram illustrates the core steps of the MADByTE platform for creating and comparing chemical features from 2D-NMR data:
1H-13C connectivity (HSQC) with 1H-1H scalar coupling (TOCSY) to define discrete substructures present in each sample.This protocol uses quantum mechanics to calculate NMR chemical shifts for structural validation, with parameters developed for terpenes.
1. Conformational Search and Selection:
2. NMR Calculation and Scaling:
13C nuclear magnetic shielding constants (Ï) for the selected conformers using the GIAO method at the mPW1PW91/6-31G(d) level of theory, applying Boltzmann statistics at 298 K.a and b are the slope and intercept from the linear regression.The workflow for this computational protocol is summarized below:
| Method | Key Theory | Best For | Key Advantage | Key Disadvantage |
|---|---|---|---|---|
| Parameterized Protocol [6] | GIAO-DFT (mPW1PW91) | Specific classes (e.g., Terpenes) | High accuracy for the class it was designed for, affordable cost. | Requires development of a class-specific scaling factor. |
| SMART Platform [5] | Convolutional Neural Networks | Broad, single compounds | Identifies structurally similar molecules from a large library; does not require a single structure as a starting point. | Limited by the coverage of its reference database. |
| Tool / Resource | Function in Research | Example / Key Feature |
|---|---|---|
| MADByTE Platform [5] | Untargeted analysis of complex NP mixtures via 2D-NMR. | Groups samples by shared NMR spin systems; no proprietary database required. |
| Fragment Libraries [2] | Provides starting points for FBDD against challenging targets. | Follows the "Rule of 3" (MW <300, low lipophilicity); high ligand efficiency. |
| Quantum Chemistry Software [6] | Calculates NMR parameters to validate proposed structures. | Uses GIAO-DFT methods (e.g., in Gaussian 09) with empirical scaling. |
| DrugBank [7] | Database of drug and drug target information. | Provides detailed drug mechanisms, structures, and target data for dereplication. |
| SciFinder [8] | Comprehensive database for chemical literature and substances. | Essential for searching known compounds and reactions to plan synthesis. |
FAQ 1: What makes a protein-protein interface considered 'undruggable'? PPIs have often been classified as 'undruggable' because their interfaces are typically large, flat, and lack deep, well-defined binding pockets, which makes it difficult for small molecules to bind with high affinity [9] [10]. Unlike traditional targets like enzymes, PPI interfaces often do not have endogenous small-molecule ligands to serve as a starting point for drug design [10].
FAQ 2: What are the main strategies for targeting 'undruggable' PPIs? Two primary strategies have emerged. The first is to target allosteric sitesâregions topologically distinct from the PPI interfaceâto modulate the interaction [9]. The second involves using advanced modalities like Targeted Protein Degradation (TPD), which uses small molecules to tag proteins for degradation by the cell's own proteolytic systems, thus overcoming the need to directly inhibit a difficult binding site [11].
FAQ 3: How can computational tools help in overcoming the druggability gap? Computer-Aided Drug Design (CADD) and artificial intelligence (AI) can significantly accelerate the discovery of PPI modulators. AI models like AlphaFold can predict protein structures with high accuracy, aiding in druggability assessments and structure-based drug design [12]. Furthermore, virtual screening and machine learning can help identify potential allosteric sites and optimize lead compounds [12] [13].
FAQ 4: What is synthetic lethality and how does it relate to targeted cancer therapy? Synthetic lethality occurs when the simultaneous disruption of two genes leads to cell death, while disruption of either gene alone does not [14] [15]. This concept is exploited in cancer therapy to selectively target cancer cells with a specific mutation (e.g., in BRCA1/2) by inhibiting its synthetic lethal partner (e.g., PARP), leaving healthy cells relatively unharmed [14] [15].
FAQ 5: Why are PPI stabilizers more challenging to develop than inhibitors? PPI stabilizers, which enhance the interaction between two proteins, present a more complex challenge than inhibitors. Stabilizers often act allosterically, and their binding site may not be readily apparent. They must also be identified under conditions that favor the stabilized complex, which is more difficult to screen for compared to disruptive inhibitors [13].
Problem: The target PPI interface is shallow and lacks obvious pockets for small-molecule binding, leading to low-affinity hits.
Solution & Workflow: Implement an integrated strategy that combines computational pocket prediction with experimental techniques to identify and validate cryptic or transient binding sites.
Recommended Protocol:
The following workflow outlines this integrated approach to tackle flat PPI interfaces:
Problem: Directly targeting a PPI interface has failed; you need to identify alternative, allosteric sites to modulate the interaction.
Solution & Workflow: Systematically discover and characterize allosteric sites that can inhibit or stabilize the PPI upon ligand binding.
Recommended Protocol:
The diagram below illustrates the strategic decision process for PPI modulation:
Problem: You need to identify genes that are synthetically lethal with a specific cancer mutation to discover new, selective therapeutic targets.
Solution & Workflow: Use combinatorial CRISPR-Cas9 screening to systematically knock out gene pairs in a high-throughput manner.
Recommended Protocol:
The table below summarizes a proposed classification system for PPI druggability based on computational assessment with SiteMap, which can help set realistic expectations at the start of a project [10].
| Druggability Class | Dscore Range | Binding Site Characteristics | Example PPI Targets |
|---|---|---|---|
| Very Druggable | > 1.00 | Well-defined, deep pocket; high hydrophobicity [10] | Bcl-2, Bcl-xL [10] |
| Druggable | 0.89 â 1.00 | Significant pocket character; amenable to ligand binding [10] | HDM2, XIAP [10] |
| Moderately Druggable | 0.80 â 0.88 | Shallower, less enclosed pocket [10] | MDMX, VHL [10] |
| Difficult / Poorly Druggable | < 0.80 | Flat, featureless, and hydrophilic interface [10] | IL-2, ZipA [10] |
| Reagent / Technology | Function / Application | Key Utility in Overcoming Intractability |
|---|---|---|
| Combinatorial CRISPR Libraries [15] | High-throughput screening of synthetic lethal gene pairs. | Enables systematic identification of context-specific genetic vulnerabilities in cancer cells. |
| Fragment Libraries [13] | Screening low molecular weight compounds to bind sub-pockets. | Useful for mapping the bindable surface of large, flat PPI interfaces. |
| PROTACs (Proteolysis-Targeting Chimeras) [11] | Bifunctional molecules that recruit a protein to an E3 ubiquitin ligase for degradation. | Modality shifts the goal from inhibition to degradation, targeting previously "undruggable" proteins. |
| AlphaFold & RosettaFold [12] [13] | AI-based protein structure prediction tools. | Provides reliable 3D models for targets with no experimental structure, enabling computational screening and druggability assessment. |
| SiteMap [10] | Computational tool for predicting and scoring binding sites on proteins. | Quantifies druggability (Dscore) to prioritize PPI targets and guides medicinal chemistry efforts. |
| DNA-Encoded Libraries (DELs) [11] | Technology for screening vast numbers of compounds by linking each molecule to a DNA barcode. | Allows ultra-high-throughput screening of chemical space against purified protein targets to find initial hits. |
| Glisoprenin B | Glisoprenin B, CAS:144376-63-6, MF:C45H82O6, MW:719.1 g/mol | Chemical Reagent |
| But-1-en-3-ynyl-benzene | But-1-en-3-ynyl-benzene, CAS:146276-26-8, MF:C10H8, MW:128.17 g/mol | Chemical Reagent |
The majority-leaves minority-hubs (mLmH) topology is a fundamental architectural principle observed in virtually all molecular interaction networks (MINs), irrespective of organism or physiological context [16] [17]. This structure is characterized by an overwhelming majority (~80%) of 'leaf' genes that interact with only 1-3 other genes, and a small minority (~6%) of 'hub' genes that interact with at least 10 or more partners [16]. This case study explores the compelling hypothesis that the mLmH topology is not merely a byproduct of evolution but an adaptive solution to circumvent fundamental computational intractability in biological systems.
The underlying problem, formalized as the Network Evolution Problem (NEP), is computationally equivalent to the well-known (\mathcal{NP})-complete Knapsack Optimization Problem (KOP) [16]. In simple terms, an evolving biological system faces a problem of immense computational complexity when trying to determine the optimal set of genes to conserve, mutate, or delete to maximize beneficial interactions and minimize damaging ones network-wide. The emergence of the mLmH topology provides a sufficient and, assuming (\mathcal{P} \neq \mathcal{NP}), necessary condition for evolving systems to efficiently navigate this intractable optimization landscape [16].
Molecular Interaction Networks (MINs) are graphs where nodes represent biological molecules (e.g., proteins, genes, metabolites), and edges represent direct physical or functional interactions between them [16] [17]. Analyzing their topologyâthe arrangement of nodes and edgesâis a cornerstone of network biology [17].
The mLmH Topology (also historically referred to as 'scale-free-like') describes the specific, non-random connectivity pattern where a few nodes possess a very high number of connections, while the vast majority have very few. It is crucial to note that this study uses the term mLmH to sidestep the controversy surrounding strict power-law distributions in biological networks and focuses on the overarching pattern itself [16].
Computational Intractability refers to computational problems for which no efficient, exact algorithm is known, and the time required to find a solution grows exponentially with the problem size. The NEP is one such problem [16].
This section addresses common computational and conceptual challenges researchers face when studying network topology and intractability.
FAQ 1: Our model of a synthetic network does not converge to an mLmH topology. What could be wrong?
FAQ 2: How can we distinguish an adaptive mLmH topology from a non-adaptive byproduct?
FAQ 3: How do we map a real-world biological dataset onto the NEP framework?
The diagram below illustrates this mapping and scoring logic.
This protocol allows researchers to test the hypothesis that mLmH arises as an adaptation to computational intractability.
Objective: To generate synthetic MINs with mLmH topology using an evolutionary algorithm with an NEP-based fitness function.
Workflow Overview:
Materials & Computational Reagents:
Step-by-Step Procedure:
Objective: To quantify the mLmH topology in a real-world molecular network and compare its degree distribution to the synthetic networks generated in Protocol 4.1.
Step-by-Step Procedure:
The following tables summarize the core quantitative data supporting the mLmH adaptation hypothesis, derived from the analysis of 25 large-scale molecular interaction networks [16].
Table 1: Characteristic Composition of mLmH-Possessing Networks
| Node Category | Degree Range | Average Percentage of Nodes | Proposed Functional Role in NEP |
|---|---|---|---|
| Leaf Genes | 1 - 3 | ~80% | Specialized functions; low-cost optimization units within the knapsack problem. |
| Hub Genes | ⥠10 | ~6% | System integration and stability; critical but costly variables in the optimization. |
| Intermediate | 4 - 9 | ~14% | Transitional or multi-functional roles. |
Table 2: Key Parameters for NEP Evolutionary Simulation
| Parameter | Symbol | Typical Value/Range | Description |
|---|---|---|---|
| Interaction Potency | (\rho) | 1 (default) | The strength/weight of a single interaction in benefit/damage calculations [16]. |
| Damage Threshold | (\tau) | User-defined | The maximum tolerable level of total network damage in the fitness function [16]. |
| Penalty Weight | (\alpha) | User-defined | A multiplier that scales the penalty for exceeding the damage threshold [16]. |
Table 3: Essential Resources for mLmH and NEP Research
| Research Reagent / Resource | Type | Function & Application in Research |
|---|---|---|
| Curated Interaction Databases (e.g., BioGRID, STRING, KEGG, BRENDA) [17] | Data | Source of empirically validated molecular interactions for building and validating network models. |
| Graph Analysis Software (e.g., NetworkX, Cytoscape) | Software | For constructing, visualizing, and computing topological metrics (degree, centrality) on networks. |
| Evolutionary Algorithm Library (e.g., DEAP, custom Python/Scripts) | Software | To implement the NEP fitness function and run the selection-mutation cycles for simulation. |
| Oracle Advice Phenotypic Datasets (e.g., from GEO, knockout phenotype databases) | Data | Provides the experimental basis for assigning promotion/inhibition states to genes in the NEP model. |
| High-Per Computing (HPC) Cluster | Hardware | Facilitates the computationally intensive task of running large-scale evolutionary simulations and solving NEP across many generations. |
| Derrone | Derrone, MF:C20H16O5, MW:336.3 g/mol | Chemical Reagent |
| Triacsin C | Triacsin C, CAS:76896-80-5, MF:C11H17N3O, MW:207.27 g/mol | Chemical Reagent |
Synthetic intractabilityâthe formidable challenge of efficiently constructing complex natural product scaffoldsâoften stymies drug discovery efforts. Fragment-Based Drug Discovery (FBDD) provides a powerful strategy to circumvent this impasse. Instead of attempting to synthesize intricate natural product mimics directly, FBDD begins with small, simple chemical fragments (molecular weight typically â¤300 Da) that bind weakly to a biological target [18] [19]. These fragments serve as efficient starting points that are progressively grown or combined into potent, lead-like compounds [20]. This approach investigates a larger chemical space with fewer compounds and is applicable to challenging biological targets, including those involved in amino acid metabolism and other pathways targeted by natural products [21] [22]. By starting small and building complexity in a structured way, FBDD offers a rational path to overcome the synthetic hurdles inherent in natural product-based drug design.
Successful implementation of FBDD relies on a core set of specialized reagents, libraries, and methodologies. The table below summarizes the key components of the FBDD toolkit.
Table 1: Key Research Reagent Solutions and Methodologies in FBDD
| Item | Function/Description | Key Characteristics |
|---|---|---|
| Fragment Library | A curated collection of small molecules for screening [18] [19] [23]. | Molecular weight â¤300 Da; follows "Rule of Three" (ClogP â¤3, H-bond donors & acceptors â¤3, rotatable bonds â¤3); high solubility [18] [19]. |
| Poised Fragment Library | A specialized fragment library designed for rapid optimization [23]. | Contains points of diversity for derivatization; includes analogue series for early SAR [23]. |
| 19F-NMR Probe | A spectroscopic probe for ligand-observed NMR screening [20] [19]. | Used to detect and quantify weak fragment binding to the target protein. |
| Synpro Orange Dye | A fluorescent dye used in Differential Scanning Fluorimetry (DSF) [18]. | Binds hydrophobic regions of denatured protein; measures protein thermal stability (Tm) shifts. |
| Biosensor Chips | Solid surfaces for immobilizing biological targets in Surface Plasmon Resonance (SPR) [18]. | Enable real-time, label-free measurement of binding kinetics and affinity. |
| Isotopically Labeled Protein (15N, 13C) | Protein sample for protein-observed NMR screening [19]. | Allows monitoring of target protein signals to map fragment binding sites. |
| Hydrocortisone Hemisuccinate | Hydrocortisone Hemisuccinate, CAS:83784-20-7, MF:C25H36O9, MW:480.5 g/mol | Chemical Reagent |
| Arohynapene A | Arohynapene A, CAS:154445-08-6, MF:C18H22O3, MW:286.4 g/mol | Chemical Reagent |
Principle: DSF (or thermal shift assay) detects fragment binding by measuring the increase in the target protein's thermal stability. A fluorescent dye binds to hydrophobic patches exposed upon protein denaturation, and a positive binding event is indicated by an increase in the melting temperature (Tm) [18].
Protocol:
Principle: SPR measures binding interactions in real-time without labels by detecting changes in the refractive index at a sensor surface where the target protein is immobilized [18].
Protocol:
Principle: This critical phase involves using high-resolution structural data, primarily from X-ray crystallography, to guide the chemical elaboration of a weakly binding fragment into a potent lead compound [18] [19].
Protocol:
Problem: A high rate of false positives in fragment screening, particularly with DSF, is a common issue that can derail a project.
Troubleshooting Guide:
Problem: Weak binding affinity is expected at the start of FBDD, but the path to optimization can be unclear.
Troubleshooting Guide:
Problem: Fragment solubility is a common bottleneck, as screening often requires mM concentrations to detect weak binding.
Troubleshooting Guide:
The following diagram illustrates the core iterative cycle of a Fragment-Based Drug Discovery campaign, from initial screening to optimized lead compound.
This diagram conceptualizes how the FBDD approach provides a solution to the problem of synthetic intractability in natural product-inspired drug discovery.
The tables below consolidate critical quantitative parameters and data used to guide and evaluate FBDD campaigns.
Table 2: Key Physicochemical Parameters and Metrics in FBDD
| Parameter | Target Value / Range | Significance |
|---|---|---|
| Fragment Molecular Weight | ⤠300 Da [18] [19] | Ensures low molecular complexity and high ligand efficiency from the start. |
| Fragment Affinity (K_D) | µM to mM range [18] [19] | Weak binding is expected for initial hits and is sufficient to begin optimization. |
| Ligand Efficiency (LE) | > 0.3 kcal/mol per heavy atom [23] | Measures binding efficiency relative to size; a key metric during optimization. |
| Lipophilic Ligand Efficiency (LLE) | Monitored during optimization [23] | Balances potency and lipophilicity; helps avoid overly hydrophobic molecules. |
| Rule of Three | MW < 300, ClogP ⤠3, HBD ⤠3, HBA ⤠3 [18] [19] | A guideline for designing fragment libraries to ensure good solubility and drug-like properties. |
Table 3: Comparison of Primary Fragment Screening Methods
| Method | Throughput | Sample Consumption | Key Information Provided | Primary Limitation |
|---|---|---|---|---|
| Differential Scanning Fluorimetry (DSF) | Medium-High [18] | Low (µM protein) [18] | Thermal Shift (ÎTm) | Susceptible to false positives; no structural info [18]. |
| Surface Plasmon Resonance (SPR) | Medium [18] | Low (for immobilization) [18] | Binding affinity (KD), kinetics (kon, k_off) | Requires immobilization; can be sensitive to bulk effects. |
| Nuclear Magnetic Resonance (NMR) | Low-Medium | High (mg protein) | Binding confirmation, binding site mapping | Low throughput; requires significant protein. |
| X-ray Crystallography | Low (becoming higher) [19] | High (mg protein & crystals) | Atomic-resolution structure of complex | Requires crystallizable protein; lower throughput. |
| Isothermal Titration Calorimetry (ITC) | Low [18] | High (mg protein) [18] | Binding affinity (K_D), stoichiometry (n), enthalpy (ÎH) | Low throughput; high protein consumption [18]. |
The direct functionalization of carbon-hydrogen (CâH) bonds has emerged as a transformative strategy in organic synthesis, particularly for constructing complex natural products. This approach represents a paradigm shift from traditional step-intensive routes toward more economical and atom-efficient disconnections for CâC bond formation [24]. For researchers in natural product development, CâH activation provides powerful tools to overcome synthetic intractability by enabling late-stage functionalization and streamlining access to complex molecular architectures [25]. While traditional synthesis often requires pre-functionalized starting materials and protecting group manipulations, CâH activation allows direct conversion of inert CâH bonds into valuable functionalities, significantly enhancing synthetic efficiency [24]. This technical support document addresses common experimental challenges and provides troubleshooting guidance for implementing CâH activation methodologies in natural product synthesis.
Understanding the precise terminology is crucial for effective communication and experimental design:
The historical classification of CâH activation into distinct mechanistic categories is evolving toward a continuum model based on the degree of charge transfer during the transition state [26]. The classical mechanisms can be understood as special cases within this continuum:
This mechanistic continuum ranges from electrophilic to nucleophilic character, governed by the overall difference in charge transfer during the transition state rather than the formal oxidation state of the metal [26]. The key mechanisms include:
Problem: Reactions show poor conversion despite apparent standard conditions.
Diagnosis and Solutions:
Table 1: Troubleshooting Low Conversion in CâH Activation
| Observation | Potential Cause | Solution Approach | Verification Method |
|---|---|---|---|
| No reaction | Catalyst decomposition | Use fresh catalyst batches; exclude oxygen with rigorous Schlenk techniques | Test catalyst activity with standard reaction |
| Slow initiation | Catalyst pre-activation required | Add initiators (e.g., benzoquinone, Cu salts) or pre-warm catalyst | Monitor reaction start with in situ IR |
| Incomplete conversion | Catalyst poisoning by impurities | Purify substrates (chromatography, recrystallization); use distilled solvents | Analyze substrate purity by NMR/HPLC |
| Variable yields between batches | Moisture sensitivity | Dry glassware, molecular sieves, anhydrous solvents | Karl Fischer titration of solvents |
Experimental Protocol for Oxygen-Sensitive Reactions:
Problem: Lack of desired selectivity in CâH functionalization.
Diagnosis and Solutions:
Table 2: Addressing Selectivity Challenges
| Selectivity Type | Governing Factors | Optimization Strategies | Representative Examples |
|---|---|---|---|
| Regioselectivity | Electronic effects, steric bias, directing groups | Install weakly-coordinating directing groups; leverage inherent substrate bias; adjust steric bulk of ligands | 2-phenylpyridine derivatives [27]; Directed borylation [27] |
| Chemoselectivity | Relative bond strengths, catalyst specificity | Tune catalyst electronics; use redox-active directing groups; employ sequential functionalization | Palladium-catalyzed CâH activation/cyclization cascades [25] |
| Stereoselectivity | Chiral environment, catalyst control | Employ chiral ligands; use chiral carboxylic acids in CMD; design substrates with element of chirality | Asymmetric synthesis of (â)-deoxoapodine [25] |
Protocol for Directing Group Optimization:
Problem: Reactions work on model systems but fail with complex natural product scaffolds.
Solutions:
Q1: How can I distinguish between different CâH activation mechanisms experimentally?
A: Use a combination of techniques:
Q2: What are the most common catalyst decomposition pathways and how can I prevent them?
A: Primary decomposition pathways include:
Q3: My CâH activation works stoichiometrically but not catalytically. What should I investigate?
A: This typically indicates issues with catalyst turnover:
Q4: How can I apply CâH activation to late-stage natural product functionalization without affecting sensitive functionalities?
A: Implementation strategies include:
Table 3: Key Reagents for CâH Activation Methodologies
| Reagent Category | Specific Examples | Function | Application Notes |
|---|---|---|---|
| Catalyst Precursors | Pd(OAc)â, Pd(TFA)â, [RuClâ(p-cymene)]â, [RhCpClâ]â, CpIr(CO)â | Generate active catalytic species | Acetate sources often facilitate CMD; TFA useful for electrophilic pathways |
| Oxidants | AgOAc, AgâCOâ, Cu(OAc)â, PhI(OAc)â, benzoquinone, Oâ (balloon) | Re-oxidize reduced catalyst | Silver salts often best for Pd; Cu systems cheaper; Oâ most atom-economical |
| Directing Groups | Pyridine, pyrazole, amide, carboxylic acid, oxime, N-oxide | Control regioselectivity via coordination | Weaker coordinating groups often provide broader scope |
| Additives | PivOH, AdCOâH, CsOPiv, Cu(OPiv)â, Mg(OTf)â | Accelerate CâH cleavage, enhance selectivity | Carboxylates crucial for CMD; Lewis acids activate electrophiles |
| Solvents | Toluene, DCE, 1,4-dioxane, TFE, DMF, HFIP | Medium for reaction, can influence mechanism | Apolar solvents enhance coordination; fluorinated alcohols facilitate electrophilic pathways |
Based on Tokuyama's synthesis of (â)-deoxoapodine [25]:
Materials: PdIâ (10 mol%), KâPOâ (2.0 equiv.), KNTfâ (1.5 equiv.), norbornene (1.2 equiv.), dry toluene, alkyl iodide substrate
Procedure:
Troubleshooting: If reaction stalls, verify norbornene quality (distill before use) and exclude oxygen. For acid-sensitive substrates, replace KâPOâ with CsOAc.
Protocol for oxidant-free conditions [28]:
Cell Configuration: Undivided cell, graphite anode (6 cm²), Pt cathode, n-BuâNPFâ (0.1 M electrolyte)
Typical Procedure:
Advantages: Eliminates stoichiometric oxidants, mild conditions, tunable by potential
This systematic approach enables efficient development of CâH activation methodologies for complex natural product synthesis. By addressing common experimental challenges through targeted troubleshooting and strategic reagent selection, researchers can effectively implement these transformative methods to overcome synthetic intractability in their synthetic campaigns.
The development of therapeutics from Natural Active Products (NAPs) is often hampered by the challenge of synthetic intractability. Many NAPs possess complex chemical structures that make derivative synthesis for labeled approachesâsuch as attaching biotin or fluorescent tagsâa time-consuming process that risks altering their native biological activity [29] [30]. Label-free target identification methods have emerged as powerful tools to overcome this hurdle. These techniques do not require chemical modification of the small molecule, thereby preserving its natural structure and function, and directly identify protein targets by detecting the biophysical consequences of ligand-binding events [31] [29]. This technical support center details the application, troubleshooting, and protocols for four key label-free methods: DARTS, CETSA, LiP-MS, and SPROX, providing a critical toolkit for advancing NAP drug discovery.
The table below summarizes the core principles, standard sample types, and primary applications of these four key methodologies to help you select the appropriate technique.
| Method | Core Principle | Common Sample Types | Typical Readout | Main Application |
|---|---|---|---|---|
| DARTS [29] [32] | Ligand binding protects the target protein from proteolysis. | Cell lysates, purified proteins [32] | SDS-PAGE/Western Blot, Mass Spectrometry [32] | Initial target validation and identification [32] |
| CETSA [31] [32] | Ligand binding increases the thermal stability of the target protein, raising its melting temperature ((T_m)). | Live cells, cell lysates [32] | Western Blot, Mass Spectrometry (CETSA-MS) [32] | Target engagement in a near-physiological context [32] |
| LiP-MS [31] | Ligand binding alters the protein's susceptibility to proteolysis, changing the peptide digestion profile. | Cell lysates, complex protein mixtures | Mass Spectrometry (Peptide mapping) | Proteome-wide target and binding site identification [31] |
| SPROX [31] [29] | Ligand binding increases the protein's resistance to chemical denaturation and oxidation. | Cell lysates, complex protein mixtures | Mass Spectrometry | Proteome-wide target identification based on thermodynamic stability [31] |
Q: Our CETSA western blot data shows high background and nonspecific protein aggregation. What steps can we take to optimize this?
Q: When should we use live cells versus cell lysates for CETSA?
Q: We are unable to see a clear protective effect in our DARTS experiment. What is the most critical parameter to optimize?
Q: Can DARTS be used for proteome-wide screening?
Q: What is the key difference between LiP-MS and SPROX in what they detect?
Q: Our LiP-MS/SPROX experiment yielded a long list of candidate hits. How can we prioritize targets for validation?
This protocol assesses target engagement in a simplified, cell-free system [32].
Workflow Diagram: CETSA in Cell Lysates
Step-by-Step Guide:
This protocol is based on the principle of ligand-induced protection from proteolysis [32].
Workflow Diagram: DARTS
Step-by-Step Guide:
The table below lists key reagents and their critical functions for successfully implementing these label-free methods.
| Reagent / Material | Function & Importance | Key Considerations |
|---|---|---|
| Cell Lysate | Source of native protein targets for DARTS, LiP-MS, SPROX, and lysate-based CETSA. | Use non-denaturing lysis buffers. Pre-clear by centrifugation. Protein concentration should be consistent across samples. |
| Live Cells | Essential for physiologically relevant CETSA to study target engagement in a cellular context. | Ensure high cell viability. Consider compound solubility and potential cytotoxicity during incubation. |
| Pronase/Thermolysin | Non-specific proteases for DARTS and LiP-MS. | Requires extensive titration. Source and lot-to-lot variability can be high; optimize for each new batch. |
| Chemical Denaturants (e.g., GuHCl) | Used in SPROX to unfold proteins and expose methionine residues to oxidation. | Prepare fresh, high-purity stock solutions. Accurately prepare the concentration gradient. |
| Mass Spectrometer | Core instrument for DARTS-MS, CETSA-MS, LiP-MS, and SPROX for proteome-wide, unbiased target discovery. | Requires expertise in liquid chromatography (LC) and tandem MS (MS/MS) operation and data analysis. |
| High-Quality Antibodies | For specific detection of target proteins in Western blot-based CETSA and DARTS. | Validate specificity and sensitivity for the target protein. Must work well for denatured protein (CETSA). |
| Methionine Oxidation Reagent | Hydrogen peroxide ((H2O2)) is typically used in SPROX to oxidize methionine residues in unfolded regions. | Reaction time and concentration must be carefully controlled to achieve limited oxidation. |
| Strictosidinic Acid | Strictosidinic Acid, CAS:150148-81-5, MF:C26H32N2O9, MW:516.5 g/mol | Chemical Reagent |
| Chrymutasin B | Chrymutasin B|Novel Antitumor Antibiotic|RUO | Chrymutasin B is a novel chartreusin-related antitumor antibiotic for research use only (RUO). Not for human, veterinary, or household use. |
Choosing the right method depends on your experimental goals, resources, and the biological context. The following diagram outlines a decision pathway to guide your selection.
FAQ 1: What are the primary advantages of integrating computational design with high-throughput crystallography in FBDD?
Integrating computational design with high-throughput crystallography creates a powerful synergy in Fragment-Based Drug Discovery (FBDD). Computational methods, including fragment informatics, high-throughput docking, and de novo design, significantly improve the efficiency and success rate of lead discovery and optimization [22]. They can be used independently or in parallel with experimental FBDD. When a high-resolution crystal system is available (typically diffracting to <2.5 Ã ), high-throughput crystallography provides unambiguous confirmation of fragment binding and generates detailed information about the protein-fragment interaction within the 3D protein structure [33]. This structural data directly feeds back into and refines the computational models, creating a virtuous cycle of design and experimental validation.
FAQ 2: Our fragment hits have poor electron density maps, making binding modes hard to interpret. How can we resolve this?
This is a common challenge when detecting weak binders. The PanDDA (Pan Dataset Density Analysis) algorithm is specifically designed to overcome this. It helps amplify the signal of weak fragment binders in electron density maps that would be difficult to interpret using conventional methods [33]. Furthermore, ensuring your experimental setup meets key requirements is crucial: a high-resolution crystal system (diffracting to <2.5 Ã ), robust crystals that can tolerate at least 10-30% DMSO for soaking, and crystal form uniformity, which is essential for PanDDA analysis [33].
FAQ 3: Why is synthetic intractability a major bottleneck in FBDD, and what are the emerging solutions?
Synthetic intractability occurs when promising fragment hits contain complex polar pharmacophores that are difficult to elaborate using traditional synthetic chemistry. A meta-analysis of 131 fragment-to-lead (F2L) examples revealed that in ~80% of cases, growth originates from an aromatic or aliphatic carbon, and more than 50% of the bonds formed are carbon-carbon bonds [34]. This makes robust CâH functionalisation methods that tolerate innate polar functionality critical for progress. Emerging solutions include:
FAQ 4: How do we select the most promising fragments from a screen for further investment?
Fragment assessment should be multi-faceted. Key parameters to consider include [38]:
| Potential Cause | Diagnostic Steps | Recommended Solution |
|---|---|---|
| Inadequate chemical diversity | Analyze the physicochemical properties (MW, logP, rotatable bonds) of your library against the "Rule of Three" [39]. | Curate or acquire a library that emphasizes functional group diversity and is biased toward planar, achiral heterocycles, or consider libraries richer in sp3-centres to target distinct sites [39]. |
| Insufficient fragment solubility | Check fragment solubility in crystal soak buffers (needs to be ~10 mM from ~0.1 M DMSO stocks) [39]. | Pre-filter the library for high solubility. Use cocktail soaking with the number of fragments per cocktail dictated by the required concentration and the DMSO tolerance of the crystals. |
| Non-robust or low-resolution crystals | Determine the reproducibility of crystal growth and the typical resolution limit of diffraction. | Optimize the crystal system. Use robotic plate-based screening or microfluidic platforms for economical sampling. Consider protein engineering, removal of flexible regions, or in-situ proteolysis to improve crystallizability [39]. |
| Potential Cause | Diagnostic Steps | Recommended Solution |
|---|---|---|
| Synthetic intractability of growth vectors | Analyze the fragment's growth vectors and the innate polar functional groups required for binding [34]. | Employ synthetic methods tolerant to polar groups, such as CâH functionalisation or photoredox catalysis, to avoid extensive protecting-group strategies [34] [35]. |
| Poor choice of elaboration strategy | Review the binding mode. Are there two adjacent fragments to link? Is there a larger compound that can be deconstructed into a merged fragment? | Let the structural biology guide the strategy. Use computational tools like FastGrow/SeeSAR for structure-based fragment growing, or generative AI models like FragmentGPT for intelligent linking and merging [38] [37]. |
| Loss of binding affinity upon growing | Determine if the added groups are causing steric clashes or disrupting key interactions. | Use computational docking to screen proposed extensions in silico before synthesis. Prioritize molecules that maintain high Ligand Efficiency (LE) during optimization [22] [39]. |
This protocol outlines the steps for screening a fragment library using high-throughput X-ray crystallography [33].
Principle: Large numbers of protein crystals are grown and individually soaked in fragment solutions. The resulting X-ray diffraction data is collected and analyzed to identify bound fragments and determine their precise binding modes.
Materials:
Method:
This protocol describes a computational method for linking or merging confirmed fragment hits using a generative AI model [37].
Principle: Conditioned on the 3D structures of two fragment-protein complexes, a deep learning model generates chemically valid linkers that connect the fragments or merges their overlapping substructures, simultaneously optimizing for drug-likeness.
Materials:
Method:
Table 1: Analysis of polar functional groups in fragment-protein binding and fragment elaboration based on 131 published F2L case studies (2015-2019) [34]
| Parameter | Statistical Finding | Implication for FBDD |
|---|---|---|
| Most Common Polar Binding Groups | NâH groups (35%); Aromatic nitrogen atoms (23%); Carbonyl oxygen atoms (22%) | Design fragment libraries to maximize the presence of these efficient binding groups. |
| Fragments with Conserved Polar Interactions | 93% of fragments had at least one polar interaction conserved in the lead. | Highlights the importance of identifying and preserving the "minimal pharmacophore" during elaboration. |
| Origin of Growth Vectors | ~80% of growth originated from an aromatic or aliphatic carbon. | Underscores the critical need for synthetic methods that functionalize CâH bonds in the presence of polar groups. |
| Type of Bonds Formed During Elaboration | >50% of bonds formed were carbon-carbon bonds. | Confirms that CâC bond-forming reactions are pivotal for FBDD campaigns. |
Table 2: Essential tools and resources for an integrated computational and crystallographic FBDD workflow
| Tool / Reagent | Function / Description | Example / Vendor |
|---|---|---|
| Rule of 3 Fragment Library | A collection of small molecules (MW <300) designed for high screening hit rates and efficient exploration of chemical space. | Maybridge Ro3 Diversity Fragment Library; Zenobia Therapeutics Libraries [39]. |
| Synchrotron Beamline Access | High-intensity X-ray source for rapid collection of diffraction data from hundreds of fragment-soaked crystals. | Diamond Light Source (UK); Canadian Light Source (Canada) [33]. |
| PanDDA (Software) | Specialized algorithm for analyzing crystallographic data to detect the weak binding signals of low-occupancy fragments. | Pan Dataset Density Analysis [33]. |
| Chemical Space Navigation Software | Software to search ultra-large, synthetically accessible virtual compound catalogs for analogs and expansions of a confirmed fragment hit. | InfiniSee (BioSolveIT); Enamine's REAL Space [38]. |
| Structure-Based Design Software | A visual, interactive dashboard for computational chemists to design and rapidly evaluate fragment-growing ideas in a 3D structure. | SeeSAR with FastGrow (BioSolveIT) [38]. |
| AI-Based Fragmentation Tool | A digital method to segment lead-like molecules into novel, AI-derived fragments for building generative libraries. | DigFrag [36]. |
| Unified Generative Model | An AI model capable of performing fragment growing, linking, and merging in a single framework, optimized for multiple pharmaceutical properties. | FragmentGPT [37]. |
| Pyripyropene B | Pyripyropene B, CAS:151519-44-7, MF:C32H39NO10, MW:597.7 g/mol | Chemical Reagent |
| Icopezil | Icopezil | Icopezil is a selective acetylcholinesterase (AChE) inhibitor for research applications. This product is for Research Use Only (RUO). Not for human or veterinary use. |
Integrated FBDD Workflow
Solving Synthetic Intractability
Q1: How does CâH activation provide a strategic advantage in the total synthesis of complex natural products?
CâH activation offers a significant strategic advantage by improving step- and atom economy, enabling more concise and efficient synthetic routes. Traditional cross-coupling reactions require pre-functionalized starting materials (e.g., organic halides and organometallic reagents), which adds extra steps for installation and purification. CâH activation bypasses this need, allowing direct transformation of inert CâH bonds into desired functional groups. This is particularly powerful in late-stage functionalization, where complex molecular scaffolds can be diversified without de novo synthesis. Within natural product synthesis, this methodology has enabled novel retrosynthetic disconnections and the construction of challenging architectures, such as the polyhydroazocine ring in lundurines and the pentacyclic core of Aspidosperma alkaloids, with remarkable efficiency [40] [25].
Q2: What are the key mechanistic paradigms in transition metal-mediated CâH activation?
The cleavage of CâH bonds by transition metals can occur via several mechanistic pathways. The prevailing mechanism is often determined by the metal, its oxidation state, the ligands, and the substrate.
It is important to note that these mechanisms exist on a reactivity continuum, governed by the degree of electron transfer between the metal and the CâH bond [26] [41].
Q3: Why is sustainability a challenge in CâH activation, and what are the emerging solutions?
While inherently more atom-economical than traditional cross-coupling, many CâH activation methodologies rely on precious metals (e.g., Pd, Rh, Ir), stoichiometric metal oxidants (e.g., Ag(I), Cu(II) salts), and hazardous solvents, which present environmental and economic challenges for large-scale application. Research is actively addressing these limitations by focusing on:
| Potential Cause | Investigation & Diagnostic Steps | Proposed Solution |
|---|---|---|
| Catalyst Deactivation | Check for catalyst precipitation or black Pd(0) formation. Test if reaction yield decreases over time. | Ensure anoxic conditions; use a glovebox or Schlenk line. Add a stoichiometric oxidant (e.g., Cu(OAc)â, AgOAc) to re-oxidize Pd(0) to Pd(II) [41]. |
| Insufficient Oxidant | Monitor reaction by TLC or LC-MS; reaction may stall after initial conversion. | Increase the equivalence of oxidant or use a more potent one (e.g., AgâCOâ acts as both base and oxidant). For aerobic reactions, ensure proper Oâ bubbling [25] [41]. |
| Incorrect Additive | The reaction is highly sensitive to the carboxylate anion. | Screen different carboxylate additives (e.g., pivalate, benzoate, adamantane-1-carboxylate) which act as critical bases in the CMD mechanism [26] [41]. |
| Solvent Incompatibility | The solvent may be coordinating and occupying coordination sites on the metal. | Switch to a non-coordinating solvent (e.g., toluene, DCE) or one that can facilitate proton transfer (e.g., hexafluoroisopropanol - HFIP) [25]. |
| Potential Cause | Investigation & Diagnostic Steps | Proposed Solution |
|---|---|---|
| Weak Coordinating Directing Group | The substrate may not be chelating effectively with the metal. | Strengthen the directing group (e.g., from ketone to picolinamide) or employ a transient directing group strategy. |
| Inherent Substrate Bias | The inherent electronic and steric bias of the substrate overpowers the directing effect. | Use a tailored ligand or template that can override the substrate's inherent selectivity. For example, a template can be designed for meta-selective CâH functionalization [25] [41]. |
| Competing Mechanisms | Electrophilic metalation might occur at the most electron-rich position instead of the directed position. | Tweak the catalyst/oxidant system. For instance, using a cationic Pd(II) catalyst may favor directed CMD over electrophilic pathways [26]. |
| Potential Cause | Investigation & Diagnostic Steps | Proposed Solution |
|---|---|---|
| Oxidative Damage | Look for byproducts from over-oxidation or decomposition of sensitive groups (e.g., aldehydes, free amines). | Lower the reaction temperature. Employ a milder oxidant (e.g., benzoquinone instead of silver salts). Protect sensitive functional groups if necessary. |
| Lewis Acidic Conditions | Many metal triflates formed in situ can be strongly Lewis acidic. | Change the counterion of the catalyst or additive (e.g., from triflate to acetate). Add a mild Lewis base inhibitor. |
| High Reactivity of Product | The initial product may be more reactive than the starting material, leading to double functionalization or decomposition. | Monitor the reaction closely and stop it at partial conversion. Consider using a protecting group on the product's reactive site. |
The following table details key reagents commonly used in Pd-catalyzed CâH activation campaigns.
| Reagent Category | Specific Examples | Function & Rationale |
|---|---|---|
| Palladium Catalysts | Pd(OAc)â, Pd(TFA)â, [Cp*RhClâ]â, [Ru(p-cymene)Clâ]â | The source of the transition metal center that performs the CâH cleavage and subsequent functionalization. Pd(II) is a common pre-catalyst for many transformations [25] [41]. |
| Oxidants | Cu(OAc)â, AgOAc, AgâCOâ, Benzoquinone, Oâ (air) | Re-oxidizes the reduced metal (e.g., Pd(0) to Pd(II)) to turn over the catalytic cycle. Choice impacts efficiency and functional group tolerance [42] [41]. |
| Carboxylate Additives | AgOPiv, NaOAc, CsOPiv | Often acts as a critical base in the Concerted Metalation-Deprotonation (CMD) mechanism. The pivalate (OPiv) anion is particularly effective [26] [41]. |
| Solvents | Toluene, 1,2-Dichloroethane (DCE), Trifluoroethanol (TFE), HFIP | The medium must dissolve reagents and often plays a role in stabilizing transition states. HFIP is especially useful for facilitating proton transfer processes [25]. |
| Directing Groups (DGs) | 8-Aminoquinoline, Picolinamide, Pyrazole, Native Functional Groups (e.g., carboxylic acids) | Coordinates to the metal center, bringing it into proximity to a specific CâH bond, thereby controlling regioselectivity. The trend is toward using simpler or native functional groups as DGs [25] [28]. |
| 2-Ethylhexyl diphenyl phosphate | 2-Ethylhexyl diphenyl phosphate, CAS:1241-94-7, MF:C20H27O4P, MW:362.4 g/mol | Chemical Reagent |
| Isopropamide Iodide | Isopropamide Iodide, CAS:71-81-8, MF:C23H33IN2O, MW:480.4 g/mol | Chemical Reagent |
This protocol is adapted from key literature on the synthesis of Aspidosperma alkaloids and lundurines, which feature pivotal Pd-catalyzed CâH activation/cyclization steps [40] [25].
Objective: To achieve an intramolecular CâH alkenylation for the construction of a fused carbo- or heterocyclic system.
Reaction Mechanism Diagram:
Step-by-Step Procedure:
Reaction Setup: In an argon-filled glovebox or under an inert atmosphere using standard Schlenk techniques, charge a dry reaction vial with:
Reaction Execution: Seal the vial and heat the reaction mixture to 90 °C with vigorous stirring for 12-16 hours. Monitor reaction progress by TLC or LC-MS.
Work-up: After cooling to room temperature, dilute the mixture with ethyl acetate (10 mL) and filter through a pad of Celite to remove metallic precipitates. Wash the filter cake thoroughly with ethyl acetate.
Purification: Concentrate the filtrate under reduced pressure. Purify the crude residue by flash column chromatography on silica gel to obtain the desired cyclized product.
Key Notes:
The process of transforming initial, weakly-binding molecular fragments into potent lead compounds presents significant synthetic challenges. Success hinges on deploying integrated strategies that combine targeted chemical synthesis with rigorous computational and biological evaluation. The table below summarizes the core strategies and their documented applications in modern drug discovery.
Table 1: Integrated Strategies for Fragment Elaboration and Lead Optimization
| Strategy Name | Key Principle | Reported Outcome/Application |
|---|---|---|
| Diversity-Oriented-Target-Focused-Synthesis (DOTFS) [43] | Integrates focused-library design, virtual screening, and robotic synthesis to automate hit-to-lead optimization. | Validation of bromodomain inhibitors with affinity improved by several orders of magnitude [43]. |
| Two-Phase Fragment Elaboration [44] | Initial optimization of fragment hits followed by systematic fragment growth to increase potency and enable structure-based design. | Discovery of two lead series of PRMT5/MTA inhibitors, leading to the clinical candidate MRTX1719 [44]. |
| Modular 3D Elaboration Platform [45] | Uses rigid, sp3-rich bicyclic building blocks to systematically elaborate 2D fragments into lead-like 3D compounds. | Streamlined discovery of a novel, selective 69 nM inhibitor of Janus kinase 3 (JAK3) [45]. |
A systematic approach is critical for diagnosing and resolving experimental failures in the lab. The following workflow provides a general framework for troubleshooting.
Frequently Asked Questions
Q1: I have successfully synthesized a new compound series, but the biological activity is much weaker than expected. What should I investigate first?
Q2: My fragment elaboration relies on a key coupling reaction that consistently gives low yields, stalling my project. How can I proceed?
Q3: The computational model I am using for virtual screening is intractable for my large compound library. What approximation methods are available?
This protocol describes an integrated strategy for automating hit-to-lead optimization via fragment growing.
1. Design Focused Virtual Library: - Input: Start with an "activated fragment" â the substructure known to bind the target. - Reaction Selection: Choose a set of one-step, medicinally relevant chemical transformations from an encoded reaction library. - Virtual Coupling: Combine the activated fragment with a diverse collection of functionalized building blocks using the selected in silico reactions to generate a large virtual compound library.
2. Virtual Screening: - Employ computational methods (e.g., molecular docking, scoring functions) to rank the virtual compounds based on predicted affinity and properties. - Select a top-ranking, structurally diverse subset for synthesis.
3. Robotic Synthesis: - Utilize automated, robotic synthesis platforms to perform the pre-selected one-step reactions and synthesize the target compounds from the virtual library.
4. Automated In Vitro Evaluation: - Use high-throughput automated systems to test the synthesized compounds for biological activity (e.g., binding affinity, inhibition potency) against the target.
5. Data Analysis and Iteration: - Analyze the results to establish structure-activity relationships (SAR). - Use the findings to design the next generation of compounds and repeat the process.
This protocol details the specific case study that led to the discovery of the clinical candidate MRTX1719.
Phase 1: Fragment Hit Optimization - Starting Point: Obtain multiple crystal structures of fragment hits bound to the target (PRMT5/MTA complex). - Initial SAR: Synthesize close analogs of the original fragments to explore immediate chemical space around the hit. The goal is to make small changes to understand which parts of the fragment are critical for binding and to achieve an initial, modest increase in potency. - Synthetic Tractability: Concurrently, assess the ease of synthesis for different fragment cores to ensure a viable path forward for large-scale elaboration.
Phase 2: Systematic Fragment Growth - Structure-Based Design: Using the structural information from X-ray co-crystals, identify specific vectors on the optimized fragment core that can be extended into unexplored regions of the protein's binding pocket. - Growth and Evaluation: Systematically synthesize compounds where the fragment is grown along these vectors. This involves designing and synthesizing a series of compounds that explore different geometries and functional groups. - Lead Series Identification: Evaluate the grown compounds for potency, selectivity, and other drug-like properties. This process led to the identification of two distinct lead series, one of which was successfully advanced to the clinical candidate MRTX1719.
The following table lists key reagents and materials used in the advanced fragment elaboration and optimization strategies described in this guide.
Table 2: Key Reagent Solutions for Fragment Elaboration
| Reagent / Material | Function / Application | Specific Example / Note |
|---|---|---|
| Bifunctional 3D Building Blocks [45] | Provides rigid, 3D scaffolds with synthetic handles for programmable fragment elaboration. | Commercially available from Key Organics. Example: Cyclopropane-based structures with a protected amine and a cyclopropyl MIDA boronate. |
| Cyclopropyl MIDA Boronate [45] | A synthetic handle for Suzuki-Miyaura cross-coupling, enabling the rapid introduction of aromatic systems to the 3D core. | Used to connect the 3D building block to aryl bromides, diversifying the compound structure. |
| Activated Fragments [43] | The core substructure, derived from a fragment hit, that contains a functional group for chemical elaboration. | Serves as the starting point for virtual library generation and automated synthesis in the DOTFS approach. |
| Functionalized Building Blocks [43] | A diverse collection of chemical reagents designed to react with the activated fragment. | Combined with the activated fragment using in silico reactions to create a virtual library for screening. |
| Robotic Synthesis Platform [43] | Automated system for high-throughput, de novo synthesis of designed compound libraries. | Enables the rapid and efficient translation of virtual hits into real compounds for testing. |
| Benztropine | Benztropine, MF:C21H25NO, MW:307.4 g/mol | Chemical Reagent |
| Coumalic acid | Coumalic acid, CAS:500-05-0, MF:C6H4O4, MW:140.09 g/mol | Chemical Reagent |
This technical support center provides targeted troubleshooting guides and FAQs to help researchers overcome common experimental challenges in CâH functionalization, a critical tool for overcoming synthetic intractability in natural product development.
Problem: Reactions with terminal alkenes yield the undesired E-alkene isomer instead of the desired Z-alkene, leading to difficult separations and low yields of the target product.
Root Cause: Under conventional reaction conditions, the mono-sulfonium adduct predominates as the major product, which favors formation of the E-alkene [50].
Solution: Implement a paired electrolysis approach to selectively generate and process 1,2-bis-sulfonium intermediates [50].
Step-by-Step Protocol:
Validation: This method has been successfully scaled to decagram scale using inexpensive electrodes and demonstrates excellent functional group compatibility with various terminal alkenes [50].
Problem: Functionalization occurs at non-specific sites in molecules with multiple similar C-H bonds, resulting in mixture of products that are difficult to separate.
Root Cause: Most organic compounds contain multiple C-H bonds with similar properties, and traditional catalysts lack the precision to distinguish between them [51].
Solution: Utilize dirhodium catalysts that create a flexible, bowl-shaped microenvironment enabling induced fitting and secondary noncovalent interactions [51].
Step-by-Step Protocol:
Key Insight: This approach mimics enzymatic control by leveraging the catalyst's three-dimensional structure to distinguish between similar C-H bonds [51].
Problem: Reactions fail to proceed or proceed too slowly with unactivated alkane substrates due to the chemical inertness of hydrocarbon C-H bonds.
Root Cause: The inherent low polarity of these bonds and their high bond dissociation energies make them difficult to activate [26].
Solution: Understand and manipulate the continuum of C-H activation mechanisms to match the electronic requirements of your specific substrate [26].
Protocol for Mechanism Evaluation:
Advanced Consideration: The traditional segregation of mechanisms (oxidative addition, Ï-bond metathesis, etc.) is being replaced by a continuum model based on charge transfer characteristics [26].
Table 1: Key Optimization Parameters for Z-Selective Alkene Functionalization
| Parameter | Suboptimal Condition | Optimized Condition | Impact on Z-Selectivity |
|---|---|---|---|
| Sulfonium Intermediate | Mono-sulfonium adduct | 1,2-bis-sulfonium intermediate | Major improvement (E to Z preference) |
| Reaction Setup | Conventional conditions | Paired electrolysis in undivided cell | Enables bis-adduct formation |
| Electrode Materials | Specialty electrodes | Graphite/stainless steel | Practical scalability maintained |
| Workup Protocol | Standard chromatography | Simple recrystallization | >99% stereopurity achievable |
| Scale | Milligram scale | Decagram scale | Industrial applicability demonstrated |
Table 2: Comparison of C-H Activation Mechanisms for Saturated Hydrocarbons
| Mechanism Type | Key Characteristics | Typical Metals | Best For | Limitations |
|---|---|---|---|---|
| Oxidative Addition | Metal inserts into C-H bond, cleaving it; oxidizes metal | Late transition metals (Ir, Rh) | Unactivated alkanes | Requires low-valent metal centers |
| Electrophilic Activation | Electrophilic metal attacks hydrocarbon, displacing proton | Pd, Pt, Au, Hg | Aromatic systems | Limited to electron-rich systems |
| Ï-Bond Metathesis | Four-center transition state; bonds break/form simultaneously | Early transition metals, lanthanides | Alkane functionalization | Limited functional group tolerance |
| Concerted Metalation-Deprotonation (CMD) | Metal interacts with C-H bond while base facilitates deprotonation | Pd with carboxylate bases | Directed C-H functionalization | Requires coordinating groups |
Q: What is the fundamental difference between C-H activation and C-H functionalization?
A: In precise terminology, C-H activation refers specifically to a mechanistic step involving direct cleavage of a C-H bond through interaction with a transition metal, resulting in a new carbon-metal bond. C-H functionalization describes the overall process of replacing a C-H bond with another element or functional group, which is typically preceded by a C-H activation event [26].
Q: How can I achieve enantioselective C-H functionalization for chiral natural product synthesis?
A: Enantioselective C-H functionalization requires creating a chiral environment around the metal center. Recent advances using dirhodium complexes demonstrate how flexible, bowl-shaped microenvironments can create enantiodiscrimination through induced fitting and noncovalent interactions with substrates, similar to enzymatic control [51].
Q: Why do my C-H functionalization reactions often give mixtures of products with saturated hydrocarbons?
A: This is a fundamental challenge because most organic compounds contain multiple C-H bonds with similar bond dissociation energies and reactivity. The solution lies in implementing strategies that can distinguish between these similar bonds, such as catalyst-controlled site-selectivity through noncovalent interactions or using directing groups to position the catalyst near specific C-H bonds [26] [51].
Q: What practical methods exist for separating Z/E isomers after alkene functionalization?
A: Traditional separation of Z/E isomers can be challenging. The Z-selective C-H functionalization approach using bis-sulfonium intermediates addresses this directly - the resulting Z-alkenyl thianthrenium salts exhibit high crystallinity, allowing for isolation of nearly stereopure products via simple recrystallization rather than difficult chromatographic separations [50].
Q: How can I improve the atom economy of my C-H functionalization reactions?
A: The potential atom economy of C-H activation/functionalization reactions is often limited by the need for stoichiometric reagents, particularly oxidants. To address this, explore electrochemical approaches (which use electrons as reagents), catalytic methods that regenerate active species, and systems that minimize stoichiometric additives [50] [26].
Table 3: Essential Research Reagent Solutions for C-H Functionalization
| Reagent/Material | Function | Application Examples |
|---|---|---|
| Thianthrene (TT) | Forms key bis-sulfonium intermediates | Z-selective alkene functionalization [50] |
| Dirhodium Catalysts | Bowl-shaped catalysts for site-selectivity | Enantioselective C-H functionalization of arylcyclohexanes [51] |
| Carboxylate Bases | Critical for CMD mechanisms | Concerted metalation-deprotonation reactions [26] |
| Electrochemical Cells | Enables paired electrolysis strategies | Generation of reactive intermediates without chemical oxidants [50] |
| Borylation Reagents | Convert C-H bonds to C-B bonds | Functional group interconversion via versatile boronic esters [27] |
| Butylboronic Acid | Butylboronic Acid, CAS:4426-47-5, MF:C4H11BO2, MW:101.94 g/mol | Chemical Reagent |
These troubleshooting guides and FAQs provide actionable solutions to common challenges in C-H functionalization, enabling more efficient synthesis of complex natural products and pharmaceutical targets.
Within natural product development research, a significant challenge is synthetic intractabilityâthe difficulty in chemically synthesizing or modifying complex natural products for drug development. Fragment-Based Drug Discovery (FBDD) provides a powerful strategy to overcome this by starting with simple, synthetically tractable molecular fragments that mimic key substructures of complex natural products. These fragments, typically weighing <300 Da, bind weakly to therapeutic targets, making the detection of their binding a central technical challenge in biophysical assay development. This technical support center provides targeted guidance to enhance the sensitivity of your biophysical assays, enabling robust detection of these weak interactions and accelerating the progression of novel therapeutics derived from natural product inspiration.
Q1: Our initial Fragment-Based Drug Discovery (FBDD) screen using a thermal shift assay yielded no stabilizing hits. What are the primary factors we should investigate?
Q2: We have identified potential fragment hits using a ligand-observed NMR method, but we are concerned about false positives. What is the best practice for confirmation?
Q3: Our target protein is difficult to express and purify, limiting the amount available for large-scale biophysical screening. Which techniques and strategies should we prioritize?
Q4: What are the key characteristics of a high-quality fragment library for sensitive detection assays?
This protocol outlines a robust, multi-technique cascade to identify and validate fragment hits with high confidence, specifically designed to overcome weak binding challenges [53].
1. Preliminary Screening with Differential Scanning Fluorimetry (DSF)
2. Hit Validation by NMR Spectroscopy
3. Hit Characterization by Isothermal Titration Calorimetry (ITC) and X-ray Crystallography
The following workflow diagram illustrates this three-stage cascade.
Choosing the right assay is critical for detecting weak fragment interactions. The table below compares the key biophysical techniques used in FBDD.
Table 1: Biophysical Assay Comparison for Fragment Screening
| Technique | Typical Throughput | Information Gained | Key Advantage | Key Limitation | Protein Consumption |
|---|---|---|---|---|---|
| Differential Scanning Fluorimetry (DSF) [53] | Medium to High | Binding confirmation (ÎTm) | Low cost, medium-throughput | Indirect measure of binding | Low |
| NMR Spectroscopy [54] | Medium | Binding confirmation, binding site mapping | Can detect very weak binders; provides structural info | Low throughput; may require isotopic labeling | Medium to High |
| Surface Plasmon Resonance (SPR) [55] [56] | Medium to High | Affinity (KD), kinetics (kon, koff) | Label-free, provides kinetic data | Requires immobilization, which can affect activity | Low (after immobilization) |
| Isothermal Titration Calorimetry (ITC) [55] [57] | Low | Affinity (KD), stoichiometry (n), thermodynamics (ÎH, ÎS) | Label-free, provides full thermodynamic profile | Low throughput, high protein consumption | High |
| Microscale Thermophoresis (MST) [55] [57] | Medium | Affinity (KD), binding confirmation | Low sample volume, works in complex solutions | Requires fluorescent labeling or intrinsic protein fluorescence | Low |
Table 2: Key Research Reagent Solutions for Biophysical Assays
| Item | Function in Assay | Key Considerations |
|---|---|---|
| Fragment Library [52] [53] | A collection of 500-1000 low molecular weight compounds for screening. | Must have high aqueous solubility (â¥1 mM) and adhere to the "Rule of Three." Quality control is critical. |
| SYPRO Orange Dye [53] | Fluorescent dye used in DSF that binds hydrophobic patches exposed during protein unfolding. | Concentration must be optimized for each protein target to achieve a strong signal-to-noise ratio. |
| Biacore Sensor Chips [56] | Gold-coated surfaces for immobilizing the target protein in SPR assays. | Choice of chip type (e.g., CM5 for amine coupling) depends on the properties of the target protein. |
| Deuterated Solvents & NMR Tubes [54] | Used for preparing samples for NMR spectroscopy. | High-quality, matched NMR tubes are essential for obtaining consistent and reproducible results. |
| nano-ITC Cells | Sample cells for Isothermal Titration Calorimetry designed to minimize protein consumption. | Requires careful loading to avoid introducing air bubbles, which can disrupt the measurement. |
When experiments fail, a systematic approach to troubleshooting is required. The following decision diagram can guide your investigation for two common scenarios: low hit rates and poor data quality.
Q1: Our natural product-derived macrocycle shows high potency but poor solubility and oral bioavailability. What multidisciplinary strategies can we employ to optimize it?
A1: Optimizing macrocycles requires a combination of computational, medicinal, and synthetic approaches. You can employ several strategies:
Q2: Our project involves a challenging protein-protein interaction (PPI) target. Small molecules have failed, and biologics are not suitable. What approach should we consider?
A2: Macrocycles are an excellent structural class for targeting PPIs due to their ability to pre-organize functional groups over a larger surface area.
Q3: We are struggling to reproduce the yield and purity of a key synthetic step from a published procedure for a complex natural product analog. How can we troubleshoot this?
A3: Reproducibility issues are common in complex synthesis. A systematic troubleshooting protocol is essential.
Q4: How can computational methods be practically integrated into a hit-to-lead optimization campaign to accelerate the project?
A4: Computational chemistry should be integrated at multiple stages to focus experimental efforts.
This guide outlines a step-by-step protocol for diagnosing and addressing a failed chemical reaction, based on standard organic chemistry practice [62].
This workflow integrates multidisciplinary strategies to overcome the common challenges of synthetic intractability and poor drug-likeness in natural product-derived macrocycles [60] [58] [59].
| Technique | Primary Function | Application in Natural Product Development | Example Tool/Platform |
|---|---|---|---|
| Virtual Screening | Rapidly screen billions of compounds in silico to identify novel hits. | Identify synthetic macrocycle starting points or bioisosteres for complex natural product fragments [58]. | NVidia GPU-accelerated platforms [58] |
| Molecular Dynamics (MD) Simulations | Model the physical movements of atoms and molecules over time. | Understand target flexibility and binding site dynamics to inform the design of more selective macrocycles [58]. | State-of-the-art molecular modeling platforms [58] |
| Quantitative Structure-Activity Relationship (QSAR) | Build predictive models that correlate molecular structure to biological activity. | Model non-standard activity data to predict the potency of new macrocyclic analogues before synthesis [58]. | Machine Learning/AI platforms [58] |
| Retrosynthetic Analysis | Deconstruct a target molecule into simpler, available starting materials. | Design feasible synthetic routes for complex natural product scaffolds and their analogs [61]. | Various commercial and open-source software [58] |
This table details essential materials and technologies used in modern, multidisciplinary drug discovery projects focused on complex synthetic targets.
| Item/Technology | Function in Drug Discovery | Application Note |
|---|---|---|
| High-Throughput Experimentation (HTE) | Automated platform for rapidly testing thousands of chemical reaction conditions. | Invaluable for optimizing difficult synthetic steps, such as macrocyclization reactions, by screening catalysts, ligands, and solvents in parallel [63]. |
| Fragment Libraries | Collections of low molecular weight compounds used for screening. | Provides starting points for Fragment-Based Drug Design (FBDD), which is particularly useful for targeting challenging protein-protein interactions with macrocycles [58] [61]. |
| Engineered Enzymes for Biocatalysis | Re-engineered enzymes used as selective and sustainable catalysts. | Enables difficult chiral syntheses and functional group transformations under mild conditions, which is crucial for complex natural product analogs [64] [63]. |
| Flow Chemistry Systems | Continuous flow reactors for performing chemical synthesis. | Improises safety and control for exothermic reactions, allows for use of unstable intermediates, and facilitates reaction scaling [63]. |
| Computer-Aided Drug Design (CADD) Software | Integrated software suites for molecular modeling, docking, and in silico prediction. | The central tool for computational chemists to design new molecules, predict their properties, and prioritize synthetic targets [58] [61]. |
Fragment-Based Drug Discovery (FBDD) is a methodology that begins with identifying very small, low molecular weight compounds (fragments) that bind weakly to target proteins. These fragments typically have molecular weights less than 300 Da and exhibit minimalistic structure while maintaining high ligand efficiency [2]. A significant challenge in FBDD is synthetic tractabilityâmany promising fragment hits cannot be progressed into lead compounds due to difficulties in chemically elaborating them. Astex Pharmaceuticals has pioneered strategies to overcome this fundamental obstacle, creating a platform that successfully transforms weak-binding fragments into clinically viable drugs [2] [65].
The Pyramid platform developed by Astex represents a structured approach to FBDD that systematically addresses synthetic challenges through integrated methodologies [65]. This technical support center distills Astex's practical strategies into actionable troubleshooting guides and protocols for researchers facing similar synthetic challenges in natural product development and drug discovery.
The following table details key reagents and methodologies central to Astex's FBDD platform for addressing synthetic tractability [2]:
Table: Key Research Reagent Solutions for Synthetic Tractability
| Research Reagent/Methodology | Function in Overcoming Synthetic Intractability |
|---|---|
| Polar, Unprotected Fragments | Provides starting points with multiple growth vectors for chemical elaboration while maintaining solubility. |
| High-Throughput X-ray Crystallography | Reveals precise binding modes and optimal growth vectors to guide rational synthetic elaboration. |
| Innovative Synthetic Organic Chemistry | Develops novel routes specifically designed for challenging fragment elaborations. |
| Multidisciplinary Team Integration | Combines structural biology, computational chemistry, and synthetic chemistry expertise for iterative design. |
| "Rule of Three" Compliant Libraries | Ensures fragment quality with molecular weight <300 Da, cLogP â¤3, and H-bond donors/acceptors â¤3. |
Problem Statement: Initial fragment hits bind to the target but lack apparent synthetic handles for elaboration, stalling the hit-to-lead process.
Diagnosis and Solution:
Problem Statement: Desired fragment elaborations are synthetically challenging or require too many steps, making optimization impractical.
Diagnosis and Solution:
Table: Astex's Solutions for Synthetic Intractability
| Challenge | Traditional Approach | Astex's Improved Approach | Key Benefit |
|---|---|---|---|
| Limited Growth Vectors | Focus on affinity alone | Structural guidance of growth vectors | Rational design with clear synthetic pathways |
| Flat, 2D Fragments | Use commercial fragment libraries | DOS-derived 3D fragments [66] | Better coverage of chemical space & vectors |
| Polar Functionality | Protection/deprotection | Methodologies for unprotected fragments [2] | Reduced synthetic steps & improved properties |
Problem Statement: The transition from initial fragment hits to viable lead compounds is slow with high attrition rates.
Diagnosis and Solution:
Purpose: To systematically optimize fragment hits using high-resolution structural information [2].
Workflow:
Key Materials:
Purpose: To create novel, three-dimensional fragment collections with enhanced synthetic tractability and multiple growth vectors [66].
Workflow (Build/Couple/Pair Algorithm):
Key Materials:
Astex's strategy differs fundamentally through its proactive rather than reactive approach to synthetic tractability. Where traditional FBDD often treats synthetic feasibility as a downstream filter, Astex integrates synthetic considerations at every stage: from library design focused on fragments with multiple growth vectors, through structural biology that identifies synthetically accessible elaboration vectors, to dedicated investment in innovative synthetic methodologies specifically for challenging fragment chemotypes [2]. This integrated approach is embodied in their Pyramid platform and has demonstrated success in producing clinical candidates like ribociclib (Kisqali) and erdafitinib (Balversa) [65].
Astex has invested significantly in developing novel synthetic methodologies specifically tailored for polar, unprotected fragments that traditional synthetic approaches struggle with [2]. These include:
Astex's approach demonstrates that synthetic tractability and drug-like properties are synergistic rather than competing goals. Key balancing strategies include:
Strategic partnerships with pharmaceutical companies and academic institutions provide access to complementary expertise, resources, and risk-sharing that are crucial for addressing complex synthetic challenges [2] [67]. Astex's collaborations with companies like MSD, Merck, and AstraZeneca have been integral to their business model, allowing them to:
This technical support center provides troubleshooting guides and FAQs to help researchers overcome common challenges in validating novel drug targets and Mechanisms of Action (MoAs), particularly within the context of natural product development.
Problem: A natural product shows promising bioactivity in an initial screen but fails in subsequent validation stages.
Solution:
Problem: Experimental data supporting a proposed Mechanism of Action is not reproducible across different assay formats or model systems.
Solution:
Problem: Many potential targets are identified, but resources are limited.
Solution: Employ a structured assessment framework like the GOT-IT recommendations. The table below summarizes key quantitative and qualitative factors for prioritization [71]:
| Assessment Area | Key Guiding Questions for Prioritization | Data to Collect |
|---|---|---|
| Target-Disease Link | What genetic, proteomic, or pharmacological evidence links the target to the human disease? [68] | Human genetic association data (e.g., GWAS), differential expression in patient samples, literature evidence. |
| Druggability & Safety | Is the target a member of a protein class with known pharmacology? Are there pre-existing safety concerns? [71] | Structural data for binding pockets, tissue expression distribution, knockout mouse phenotype data. |
| Differentiation Potential | Does modulating this target offer a potential advantage over existing therapies? [71] | In vitro/vivo efficacy data compared to standard of care, biomarker strategy for patient stratification. |
| Assayability | Can robust in vitro and in vivo assays be developed to screen for and characterize compound activity? [68] | Availability of recombinant protein, cell-based reporter assays, and pharmacodynamic biomarkers. |
Insufficient validation of drug targets in the early stages of development has been strongly linked to costly Phase II clinical trial failures. Effective early-stage validation and proof-of-concept studies are critical for reducing this attrition rate [68].
The "fragment" approach is a powerful strategy. Identify the core, biologically active substructure (privileged fragment) of the complex natural product. You can then:
The table below details key reagents and materials essential for experiments in target and MoA validation.
| Research Reagent / Tool | Function in Validation |
|---|---|
| Cellular Thermal Shift Assay (CETSA) | Measures drug-target engagement inside intact cells by detecting ligand-induced thermal stabilization of the target protein [68]. |
| Chemical Probes for Chemical Proteomics | Engineered small molecules used to pull down and identify protein targets from a complex proteome-wide mixture, aiding in deconvoluting targets for natural products [68]. |
| qPCR Assays | Examines the expression profiles of specific genes to provide insights into how drug treatments affect transcriptional pathways [68]. |
| siRNA/shRNA Libraries | Enables genome-wide or pathway-focused gene knockdown to assess the phenotypic consequences of target inhibition and validate its role in a disease process [68]. |
| Xenograft Mouse Models | Provides a manageable in vivo system for validating the therapeutic effect of targeting a specific molecule or pathway in a human tumor context [68]. |
The diagram below outlines a logical, multi-stage workflow for validating a novel target identified from a natural product, incorporating strategies to address synthetic intractability.
This diagram illustrates a generalized signaling pathway that can be perturbed to validate a novel MoA. Researchers can map their specific target and predicted effects onto this framework.
The discovery and development of therapeutics from natural products often face the significant challenge of synthetic intractabilityâthe difficulty in chemically synthesizing or modifying complex natural molecules. This technical support center provides a comparative framework for two primary hit-identification strategies, Fragment-Based Drug Discovery (FBDD) and Traditional High-Throughput Screening (HTS), to help researchers select the optimal path for their specific natural product development projects. This guide offers troubleshooting advice and detailed protocols to navigate the common pitfalls associated with these approaches.
HTS is a well-established paradigm that involves the rapid experimental testing of hundreds of thousands to millions of diverse, drug-like compounds (typically with molecular weights of 400-650 Da) in automated, miniaturized assays to identify initial "hits" [73] [74]. Its primary strength lies in the ability to quickly identify potent chemical matter with a reasonable likelihood of success [75].
FBDD is a complementary approach that involves screening smaller libraries (typically 1,000-3,000 compounds) of low molecular weight fragments (MW <300 Da) [73]. These fragments follow the "Rule of 3" (see Table 1) and are characterized by low complexity and weak binding affinity. The strategy focuses on identifying efficient, initial binding interactions, which are then optimized into lead compounds through structural guidance [75] [74].
The choice between HTS and FBDD is target-dependent and influenced by project goals, available resources, and the specific characteristics of the target itself [73]. The table below summarizes the core differentiating factors.
Table 1: Direct Comparison of HTS and FBDD Key Parameters
| Parameter | High-Throughput Screening (HTS) | Fragment-Based Drug Discovery (FBDD) |
|---|---|---|
| Library Size | Large (100,000 - 1,000,000+ compounds) [73] | Small (1,000 - 20,000 fragments) [75] [73] |
| Compound Properties | Drug-like; MW ~400-650 Da [73] | Fragment-like; MW <300 Da, follows "Rule of 3" [73] |
| Typical Hit Potency | More potent (e.g., µM range) [75] | Weak binders (e.g., 0.1 - 1.0 mM K~i~/K~d~) [75] |
| Primary Readout | Functional activity in biochemical/cellular assays [74] | Direct binding measured by biophysical techniques [75] [73] |
| Structural Information | Not inherent; may be added later | Core component; relies on X-ray crystallography/NMR [75] [73] |
| Typical Hit Rate | ~1% [73] | Higher hit rates, but with lower initial potency [76] |
| Chemical Space Coverage | Sparse sampling of lead/drug-like space [75] | More comprehensive sampling of fragment space [75] |
| Key Advantage | Rapidly identifies potent, cell-active compounds [75] | Systematically probes active site; high-quality starting points [75] [77] |
The fundamental difference in strategy between HTS and FBDD is illustrated in the following workflows.
This protocol is adapted for a 384-well plate format to screen a large compound library against a enzymatic target [74] [78].
Assay Miniaturization and Validation:
Compound Dispensing:
Reagent Addition and Incubation:
Signal Detection and Analysis:
SPR is a gold-standard biophysical method for detecting the direct binding of fragments to a target protein [73] [74].
Sensor Chip Preparation:
Fragment Screening Run:
Data Analysis and Hit Confirmation:
Success in HTS and FBDD campaigns relies on specific reagents and instrumentation. The following table details key solutions for your experiments.
Table 2: Key Research Reagent Solutions for HTS and FBDD
| Reagent/Instrument | Function | Application Context |
|---|---|---|
| 384/1536-well Microplates | Miniaturized assay vessels to enable high-throughput testing [78]. | HTS |
| HTS Compound Library | A curated collection of 100,000s of drug-like small molecules for screening [73]. | HTS |
| Fragment Library | A collection of 1,000-3,000 Rule-of-3 compliant small fragments [73]. | FBDD |
| Surface Plasmon Resonance (SPR) | Label-free technique to detect and quantify real-time binding kinetics of fragments [73] [74]. | FBDD |
| Protein-based NMR | Gold-standard biophysical tool providing atomic-resolution insights into protein-ligand interactions in solution [79]. | FBDD |
| X-ray Crystallography | Determines the 3D atomic structure of a target protein bound to a fragment, guiding optimization [75] [73]. | FBDD |
| Differential Scanning Fluorimetry (DSF) | Measures protein thermal stability shift upon ligand binding; a lower-cost binding assay [73] [74]. | FBDD |
| Automated Liquid Handlers | Robotics for accurate and reproducible dispensing of reagents and compounds in microplates [74]. | HTS & FBDD |
The field is rapidly evolving with the integration of Artificial Intelligence (AI). AI-driven molecular fragmentation techniques are enhancing the representation of compounds as a "chemical language," which can improve the design and optimization of fragments [77] [80] [81]. Furthermore, AI and deep learning are anticipated to accelerate the optimization of fragment hits into leads by simultaneously considering activity, selectivity, and drug-like properties [74].
Q1: Our HTS campaign against a natural product target yielded no viable hits. What went wrong and what should we do next?
Q2: Our confirmed fragment hits have very weak affinity (>>100 µM). How can we realistically develop these into a lead?
Q3: We are a small lab with a novel target. Should we invest in building an HTS infrastructure or focus on FBDD?
Q4: How does the "Rule of 3" for fragments differ from Lipinski's "Rule of 5" for drug-like compounds?
Table 3: Comparison of Rule-of-3 and Rule-of-5 Criteria
| Property | Rule-of-3 (for Fragments) | Rule-of-5 (for Drug-like Compounds) |
|---|---|---|
| Molecular Weight (MW) | ⤠300 Da | < 500 Da |
| cLogP | ⤠3 | ⤠5 |
| Hydrogen Bond Donors (HBD) | ⤠3 | ⤠5 |
| Hydrogen Bond Acceptors (HBA) | ⤠3 | ⤠10 |
| Number of Rings | - | - |
| Rotatable Bonds | ⤠3 | - |
Both HTS and FBDD are powerful, complementary strategies in the modern drug discovery toolkit. For research focused on overcoming the synthetic intractability of natural products, FBDD offers a particularly compelling approach by starting with simple, synthetically accessible fragments and using structural biology to guide their rational optimization into novel lead compounds. By understanding the strengths, requirements, and methodologies of each approach, researchers can strategically deploy them to accelerate the development of new therapeutics.
FAQ 1: What is the core philosophical difference between CâH activation and classical cross-coupling?
The fundamental difference lies in the starting materials and the concept of synthetic pre-functionalization.
FAQ 2: Why has CâH activation seen slower adoption in total synthesis compared to cross-coupling?
Despite its potential, CâH activation is often perceived as less reliable for several reasons [83]:
FAQ 3: In which scenarios does CâH activation provide a clear economic advantage?
CâH activation becomes strategically powerful in these contexts:
FAQ 4: Can CâH activation and cross-coupling be complementary?
Absolutely. They are not mutually exclusive. A robust synthetic plan may use cross-coupling to build a core scaffold reliably in the early stages, and then employ CâH activation for late-stage diversification and introduction of delicate functionalities that would be incompatible with pre-halogenation conditions [83].
Problem: Poor regiocontrol leads to a mixture of mono- and poly-functionalized products.
Solutions:
Problem: Low conversion or catalyst deactivation.
Solutions:
The following tables summarize key economic and practical differences between the two synthetic paradigms.
Table 1: Strategic and Economic Profile Comparison
| Parameter | Classical Cross-Coupling | CâH Activation |
|---|---|---|
| Typical Pre-functionalization | Required (e.g., halide, triflate) | Not required |
| Step Count | Higher (includes halidation/installation) | Lower (more step-economical) [83] |
| Atom Economy | Lower (generates halide waste from installation) | Theoretically higher [26] |
| Regioselectivity Control | High (defined by halide position) | Challenging; requires strategies like DGs [83] |
| Late-Stage Applicability | Can be difficult (sensitive to FG tolerance) | High (powerful for diversification) [84] |
| Typical Catalyst Metals | Pd, Ni, Cu, Fe [82] [85] | Pd, Rh, Ru, Ni, Fe, Co [82] [84] |
Table 2: Common Catalyst and Ligand Systems
| Reaction Type | Typical Catalyst | Common Ligands | Key Function |
|---|---|---|---|
| Suzuki Coupling | Pd(PPhâ)â, Pd(dppf)Clâ | Triarylphosphines (PPhâ), Buchwald-type biarylphosphines | Facilitates oxidative addition & reductive elimination [82] |
| Buchwald-Hartwig Amination | Pdâ(dba)â, Pd(OAc)â | Bulky dialkylbiarylphosphines (e.g., XPhos, SPhos) | Accelerates reductive elimination, stabilizes Pd center [82] |
| Directed CâH Activation | [Cp*RhClâ]â, Pd(OAc)â | Often ligand-free or uses pivalate as an internal base (in CMD mechanism) [27] [26] | -- |
| Undirected CâH Activation | Pd(TFA)â, Fe porphyrins | -- | -- |
This is a workhorse reaction for forming C(sp²)âC(sp²) bonds.
Workflow Diagram: Suzuki-Miyaura Cross-Coupling
Materials & Procedure:
This protocol highlights the use of a directing group (DG) to achieve regiocontrol.
Workflow Diagram: Directed C-H Arylation
Materials & Procedure:
Table 3: Key Reagent Solutions for CâH Activation and Cross-Coupling
| Reagent / Material | Function / Explanation |
|---|---|
| Palladium(II) Acetate (Pd(OAc)â) | A versatile, widely used source of Pd(0) or Pd(II) for both cross-coupling and CâH activation catalysis [82]. |
| Buchwald Ligands (e.g., SPhos, XPhos) | Bulky, electron-rich phosphine ligands that form highly active LPd(0) species, enabling coupling of unactivated aryl chlorides and facilitating challenging reductive eliminations [82]. |
| Tetrakis(triphenylphosphine)palladium(0) (Pd(PPhâ)â) | A common Pd(0) source for Suzuki and Stille cross-coupling reactions [85]. |
| Silver Salts (AgOAc, AgâCOâ) | Commonly used as stoichiometric oxidants in Pd-catalyzed CâH functionalization cycles to re-oxidize Pd(0) back to Pd(II). They can also act as halide scavengers [27]. |
| Cesium Pivalate (CsOPiv) | A common base in the Concerted Metalation-Deprotonation (CMD) mechanism for CâH activation. The pivalate acts as an internal base to accept the proton during CâH cleavage [27] [26]. |
| N-Heterocyclic Carbene (NHC) Precursors | Ligands that are strong Ï-donors, often used to stabilize electron-rich metal centers. They are particularly effective for coupling sterically hindered substrates [82]. |
| Iridium-based Photoredox Catalysts (e.g., [Ir(ppy)â]) | Used in metallaphotoredox catalysis, a modern hybrid approach that merges cross-coupling with photoredox catalysis to activate otherwise inert coupling partners [82]. |
Identifying the protein target of a therapeutic compound is a foundational step in drug discovery. For researchers working with complex Natural Products (NPs), this stage is particularly challenging. Many NPs are synthetically intractable; their intricate chemical structures make them difficult to modify without altering their biological activity. This creates a significant hurdle for traditional label-based methods, which require the covalent attachment of a tag to the small molecule. This technical support article compares the performance of label-free and label-based target identification methods, providing troubleshooting guides and detailed protocols to help you select and optimize the right approach for your natural product research, thereby overcoming the barrier of synthetic intractability.
The following table summarizes the core principles, key advantages, and common challenges associated with major target identification methods.
Table 1: Comparison of Label-Based and Label-Free Target Identification Methods
| Method | Core Principle | Key Advantages | Common Challenges & Limitations |
|---|---|---|---|
| Affinity-Based Pull-Down [29] [86] | A tagged molecule (e.g., biotin) is used to affinity-purify binding partners from a complex mixture. | High specificity; direct isolation of target complexes; well-established protocols. | Requires chemical modification (tag/linker) which can alter bioactivity; time-consuming probe synthesis [29]. |
| Photoaffinity Tagging (PAL) [86] | A photoreactive probe forms a permanent, covalent bond with its target protein upon UV irradiation. | "Captures" transient/weak interactions; reduces false positives from wash steps. | Requires complex probe design & synthesis; potential for non-specific cross-linking [29]. |
| Cellular Thermal Shift Assay (CETSA) [29] [31] [87] | Ligand binding increases protein thermal stability, measured via the aggregation temperature (Tagg). | Works in intact cells (physiological relevance); no need for molecule modification [88]. | Requires specific antibodies or MS readout; may miss proteins with small thermal shifts [88]. |
| Drug Affinity Responsive Target Stability (DARTS) [29] [31] [88] | Ligand binding protects a protein from proteolytic degradation. | Experimentally simple; no specialized equipment; no molecule modification. | Can yield false positives from protease substrate preferences; semi-quantitative [88]. |
| Stability of Proteins from Rates of Oxidation (SPROX) [31] [88] | Ligand binding reduces the rate of chemical denaturation and oxidation of methionine residues. | Can detect interactions not affecting thermal stability. | Limited to proteins containing methionine; typically used with cell lysates [88]. |
This is a foundational label-based protocol for identifying direct binding partners from a cell lysate [86].
Key Research Reagent Solutions:
Step-by-Step Workflow:
This label-free method leverages the increased resistance to proteolysis upon ligand binding [88].
Key Research Reagent Solutions:
Step-by-Step Workflow:
This label-free method detects target engagement in a cellular context by measuring ligand-induced thermal stabilization [87] [88].
Key Research Reagent Solutions:
Step-by-Step Workflow:
The following diagram illustrates the core CETSA workflow.
Answer: For NPs that are difficult to chemically modify, label-free methods are the preferred starting point. Methods like DARTS and CETSA use the native molecule, completely avoiding complex synthetic chemistry and the risk of altering its bioactivity [29] [30]. Use label-based methods (e.g., affinity pull-down) only if you have a clear site on the NP for linker attachment that is known not to affect its activity, or if you need the high specificity of direct isolation for downstream validation.
Answer: Not necessarily. Consider these potential issues:
Answer: Observing many stabilized proteins is common and can indicate both direct binding and downstream effects on protein complexes or pathways. To prioritize candidates:
Answer: High background is a common challenge. Implement these controls and strategies:
The following table lists key reagents and their critical functions for setting up the target identification methods discussed.
Table 2: Key Research Reagent Solutions for Target Identification
| Reagent / Solution | Primary Function | Key Considerations for Natural Products |
|---|---|---|
| Biotin-Avidin/Streptavidin System | High-affinity capture and isolation of target proteins in pull-down assays. | The linker length and attachment chemistry are critical to avoid steric hindrance and loss of NP activity [86]. |
| Photoactivatable Cross-linkers (e.g., Diazirines, Benzophenones) | Forms irreversible covalent bonds between the NP probe and its target upon UV irradiation. | Diazirines are often preferred for their smaller size and better stability. Probe design and synthesis are complex [86]. |
| Broad-Spectrum Proteases (e.g., Pronase, Thermolysin) | Digests unfolded proteins in the DARTS assay. | The type and concentration of protease must be optimized for each system to reveal ligand-induced stabilization [88]. |
| Multiplexed Mass Spectrometry Tags (e.g., TMT, iTRAQ) | Enables simultaneous quantification of protein abundance across multiple samples in TPP/CETSA workflows. | Allows for high-throughput, unbiased target discovery across the entire proteome [31]. |
| Label-Free Detection Systems (e.g., BLI, SPR instruments) | Measures binding kinetics (ka, kd) and affinity (KD) without labels. | Octet (BLI) systems can often handle cruder samples with less purification, speeding up validation [89]. |
This technical support center provides troubleshooting guides and FAQs to help researchers overcome specific experimental challenges in natural product development.
Q1: Our team is debating whether to prioritize speed or quality in our natural product development pipeline. How can we balance both effectively?
The perceived trade-off between speed and quality is a false dichotomy. High-performing teams achieve both by integrating quality into every development stage. Research from the DevOps Research and Assessment team (DORA) confirms that elite teams deploy frequently and maintain high reliability [90].
Q2: Which specific metrics should we track to quantify improvements in our development speed?
To measure delivery speed, focus on metrics that quantify how quickly value is delivered and where bottlenecks exist. The following table summarizes key speed metrics [90]:
| Metric | Description & Calculation | Target |
|---|---|---|
| Deployment Frequency | How often code is released to production or end-users [90]. | Elite teams deploy on-demand or daily [90]. |
| Lead Time for Changes | Duration from code commit to its deployment in production [90]. | Shorter times indicate faster feedback and delivery [90]. |
| Cycle Time | Duration from the start of development work to its deployment [90]. | Shorter cycles indicate greater efficiency and fewer bottlenecks [90]. |
| Throughput | The amount of work completed by a team over a given time period [90]. | Higher throughput of meaningful deliverables indicates productive teams [90]. |
Q3: What are the critical quality metrics for ensuring our lead compounds and processes are reliable and stable?
Quality metrics ensure that rapid delivery does not compromise the reliability of your outputs, which is critical in drug development. Key metrics to monitor include [90]:
| Metric | Description & Calculation | Target |
|---|---|---|
| Change Failure Rate | The percentage of deployments causing production incidents or requiring rollbacks [90]. | A low rate indicates reliable releases and adequate testing [90]. |
| Mean Time to Restore (MTTR) | The average time required to recover from a production failure or incident [90]. | A low MTTR demonstrates effective response and resilience capabilities [90]. |
| Defect Rate | The number of bugs or issues identified post-release [90]. | Low and stable rates indicate good quality control during development [90]. |
| Right First Time (RFT) | The percentage of products or outputs that meet quality standards without requiring rework [90]. | A high RFT rate reflects precision and reliability in your production process [90]. |
Q4: We are experiencing significant bottlenecks in the total synthesis of complex natural products like antibiotics. Are there modern strategies to improve efficiency?
Traditional solution-phase synthesis can be labor-intensive. Solid-phase synthesis is a powerful strategy to circumvent tedious isolation and purification procedures, replacing them with simple filtrations [91].
This approach has enabled the synthesis of over 80 daptomycin analogs for comprehensive structure-activity relationship studies [91].
Q5: Our organization is adopting a more formal "enhanced approach" to drug substance development, as in ICH Q11. What does this mean for our control strategy?
An "enhanced approach" under ICH Q11 uses risk management and extensive scientific knowledge to establish a robust control strategy, moving beyond a traditional fixed-parameter system [92].
Problem: A low percentage of synthesis attempts yield the target compound that passes purity and identity checks on the first attempt, leading to costly rework.
Investigation & Resolution Flowchart:
Problem: Difficulty in identifying and controlling process-related and product-related impurities during the purification of natural products or their synthetic intermediates.
Investigation & Resolution Flowchart:
Essential materials and reagents for modern natural product synthesis and analysis.
| Item | Function & Application |
|---|---|
| Functionalized Resins for Solid-Phase Synthesis | Insoluble polymer supports (e.g., Trityl-chloride resin) for anchoring growing molecules, enabling rapid filtration-based purification after each reaction step [91]. |
| Protected Amino Acid Building Blocks | Non-proteinogenic amino acids (e.g., Fmoc-Kynurenine, 3-methylglutamic acid derivatives) are crucial for synthesizing complex natural product peptides like Daptomycin [91]. |
| Coupling Reagents | Reagents such as HATU, DIC, or HBTU that activate carboxylic acids for amide bond formation, essential for peptide coupling in both solid-phase and solution-phase synthesis [91]. |
| Catalysts for Key Transformations | Specialized catalysts (e.g., Pd(PPhâ)â for deallylation, chiral catalysts for asymmetric synthesis) to enable specific, high-yielding chemical transformations [91]. |
| Deallylation Cocktail | A mixture of Tetrakis(triphenylphosphine)palladium(0) and phenylsilane used for the orthogonal removal of allyl-based protecting groups on solid support [91]. |
| Analytical Standards | Highly purified compounds for use as references in HPLC, LC-MS, and NMR to confirm the identity and purity of synthetic intermediates and final products [92]. |
| Chromatography Media | Media for preparative HPLC and flash column chromatography (e.g., C18 silica, normal-phase silica gel) for the final purification of synthetic natural products [91]. |
Overcoming synthetic intractability is not a singular breakthrough but a strategic integration of foundational understanding, innovative methodologies, rigorous optimization, and robust validation. The synergy between computational biology, advanced synthetic tactics like CâH activation, and powerful screening platforms like FBDD is redefining the possible in natural product-based drug discovery. These approaches collectively provide a roadmap to navigate the complex chemical space of natural products, transforming them from synthetic challenges into tractable leads. The future of this field lies in further refining these toolsâdeveloping more predictive computational models, next-generation CâH activation catalysts with broader applicability, and even more sensitive label-free validation techniques. This progress will undoubtedly unlock new therapeutic avenues, enabling the targeting of intricate biological pathways and the development of first-in-class medicines for diseases that currently lack effective treatments.