This article provides a comprehensive guide for researchers and drug development professionals on overcoming the primary challenge in natural product-based drug discovery: the poor chemical accessibility of complex natural leads.
This article provides a comprehensive guide for researchers and drug development professionals on overcoming the primary challenge in natural product-based drug discovery: the poor chemical accessibility of complex natural leads. It explores the foundational reasons why these molecules are often difficult to synthesize, details modern computational and experimental methodologiesâincluding fragment-based design, SCAR by Space, and in silico tools like WHALESâto simplify structures while preserving bioactivity. The content further addresses common troubleshooting scenarios for ADMET optimization and provides frameworks for validating synthetic feasibility and comparing lead candidates. By integrating these strategies, scientists can more effectively translate promising natural product hits into viable, synthetically accessible drug candidates.
FAQ 1: What is meant by "chemical accessibility" in natural product drug discovery? Chemical accessibility refers to the ability to obtain a natural product compound in sufficient quantity and purity for comprehensive biological testing and subsequent development. This encompasses the entire process from sourcing the raw biological material, isolating the pure compound from complex mixtures, to having enough material for hit confirmation, lead optimization, and pre-clinical studies. Challenges in any of these steps can halt an otherwise promising drug discovery program [1].
FAQ 2: Why is sourcing natural products a major challenge? Sourcing presents multiple hurdles. The collection of plant or marine organisms can lead to overharvesting and biodiversity loss, raising significant ecological and sustainability concerns. Furthermore, many source organisms, particularly microorganisms from extreme environments, are uncultivable under standard laboratory conditions, making their metabolic products inaccessible. Legal complexities, such as those governed by the Nagoya Protocol, also regulate international access to genetic resources and the fair sharing of benefits, which can complicate collaborations and sourcing from biodiversity-rich regions [2] [3].
FAQ 3: What are the specific technical barriers in the isolation and purification of natural products? The path from a crude extract to a pure, characterized compound is fraught with difficulties. Crude biological extracts are inherently complex mixtures of many compounds, making the separation of individual pure substances a laborious, multi-step process. The quantity of the target compound isolated from the natural source is often minute (milligrams or less), which is insufficient for full biological profiling and development. Additionally, the process of dereplicationâthe early identification of known compounds to avoid re-isolationâis crucial for efficiency but remains a significant technical bottleneck [1] [4].
FAQ 4: How does chemical accessibility impact the progression of a natural product lead? A lack of chemical accessibility directly translates to a high attrition rate in natural product-based drug discovery. Many biologically active extracts identified in initial screenings never progress to a identified lead compound because the active constituent cannot be isolated in usable quantities. Even when a potent lead is identified, insufficient material can prevent the thorough evaluation of its mechanism of action, toxicity, and pharmacokinetic properties, and can stall programs aimed at synthesizing simpler or more potent analogues [1] [3].
FAQ 5: What modern strategies are being used to overcome supply bottlenecks? The field is adopting several innovative strategies to address supply issues:
Issue: An extract shows promising activity in a primary bioassay, but the activity is lost, diminishes, or becomes inconsistent as you fractionate and purify the sample.
Possible Causes & Solutions:
| Cause | Diagnostic Experiments | Solution |
|---|---|---|
| Synergistic Effects: The bioactivity is the result of multiple compounds working together, which are separated during purification. | Recombine purified fractions in different combinations and re-test for activity restoration. | Consider developing a standardized extract instead of pursuing a single compound. Alternatively, focus on a defined mixture of fractions [3]. |
| Compound Instability: The active compound is degrading under the isolation conditions (e.g., pH, light, temperature). | Re-analyze active fractions by LC-MS immediately after purification and again after 24-48 hours to look for decomposition products. | Optimize isolation protocols to use protective conditions (e.g., under nitrogen, in amber glass, at lower temperatures). Add stabilizers if compatible with the assay. |
| Non-Specific Binding: The active compound is binding to labware (e.g., plastic tubes, filtration membranes) or stationary phases during chromatography. | Use different types of labware (e.g., glass, low-binding plastics). Analyze the flow-through and washings from solid-phase extraction for activity. | Use silanized glassware or low-binding plastics. Change chromatography media (e.g., switch from C18 to polymer-based). |
Issue: After significant effort in isolation, you find that your pure compound is already known from published literature, leading to wasted resources.
Possible Causes & Solutions:
| Cause | Diagnostic Experiments | Solution |
|---|---|---|
| Insufficient Pre-screening: Relying solely on a single database or analytical technique (e.g., LC-UV) for dereplication. | Perform high-resolution mass spectrometry (HR-MS) to determine molecular formula and search against specialized NP databases. Use MS/MS molecular networking. | Implement a multi-technique dereplication workflow early in the process. Combine HR-MS, MS/MS fragmentation, and NMR profiling (even on partially purified samples) [4]. |
| Inefficient Use of Databases: Not querying comprehensive or specialized natural product databases. | Search the compound's molecular formula or predicted structure in several databases (e.g., GNPS, NPASS, PubChem) [4]. | Integrate in-silico tools and databases into the discovery pipeline. Use tools like the Global Natural Products Social Molecular Networking (GNPS) platform for comparative analysis of MS/MS data [4]. |
Issue: After successfully cloning a biosynthetic gene cluster (BGC) into a heterologous host, the production titer of the target natural product is negligible or very low.
Possible Causes & Solutions:
| Cause | Diagnostic Experiments | Solution |
|---|---|---|
| Inadequate Gene Expression: The native promoters of the BGC are not recognized efficiently by the heterologous host's transcriptional machinery. | Use RT-PCR to check the transcription levels of key biosynthetic genes. Compare them to levels in the native producer if possible [5]. | Refactor the gene cluster: Replace native promoters with strong, constitutive promoters (e.g., ErmE*) that are well-known in the host system. |
| Absence of Pathway-Specific Regulators: The positive regulatory gene(s) that activate the BGC in the native host may not be present or functional. | Check the BGC sequence for putative regulatory genes. Overexpress them in the heterologous host and monitor production. | Co-express positive regulators: Clone and express the pathway-specific regulatory gene(s) alongside the BGC in the heterologous host [5]. |
| Bottleneck in Biosynthesis: A single enzyme in the pathway may be poorly expressed or inefficient, causing a metabolic bottleneck. | Use RT-PCR and/or proteomics to identify genes/proteins with very low expression levels compared to the rest of the pathway. | Identify and overcome the bottleneck: Co-express the rate-limiting gene(s). For example, co-overexpression of fdmR1 (regulator) and fdmC (ketoreductase) was crucial for improving fredericamycin A production in S. lividans [5]. |
Objective: To rapidly identify known compounds in a biologically active crude extract before committing to large-scale isolation.
Materials:
Methodology:
Objective: To produce a target natural product by expressing its BGC in a genetically tractable heterologous host.
Materials:
Methodology:
The following diagram illustrates the multi-stage process of natural product drug discovery, highlighting the critical points where chemical accessibility can become a bottleneck.
Diagram: NP Drug Discovery Path and Bottlenecks. This flowchart outlines the key stages of natural product-based drug discovery and pinpoints where major chemical accessibility bottlenecks occur, from sourcing to scalable production.
The following table details essential reagents, tools, and technologies used to navigate the challenges of chemical accessibility in natural product research.
| Tool/Reagent | Function & Application in NP Research |
|---|---|
| High-Resolution Mass Spectrometry (HR-MS) | Determines the exact mass of a compound, allowing for the calculation of its molecular formula. Critical for the first step in dereplication and structure elucidation [4]. |
| Global Natural Products Social Molecular Networking (GNPS) | An online platform that allows for the creation of molecular networks from MS/MS data. It enables the rapid comparison of your compounds against a vast library of known spectra, drastically improving dereplication efficiency [4]. |
| Heterologous Host Strains (e.g., S. albus J1074) | Genetically tractable microbial chassis used to express biosynthetic gene clusters from uncultivable or slow-growing source organisms. This is a key strategy for solving sustainable supply issues [5]. |
| Computer-Assisted Structure Elucidation (CASE) | Software that uses NMR and other spectroscopic data to propose chemical structures. It helps to accelerate the challenging process of determining the structure of novel compounds, especially those with complex stereochemistry [4]. |
| antiSMASH | A bioinformatics tool for the genome-wide identification, annotation, and analysis of biosynthetic gene clusters. It is the starting point for most modern genome-mining campaigns [2]. |
| Synthetic Biology Vectors (BACs, Cosmids) | Large-capacity cloning vectors capable of holding the entire DNA sequence of a biosynthetic gene cluster (often 50-150 kb) for transfer into a heterologous host [5]. |
| Constitutive Promoters (e.g., ErmE*) | Strong, always-on promoters used in synthetic biology to "refactor" biosynthetic gene clusters, ensuring high expression of pathway genes in heterologous hosts where native regulators may not function [5]. |
| 3,5-Dihydroxybenzoic Acid | 3,5-Dihydroxybenzoic Acid, CAS:99-10-5, MF:C7H6O4, MW:154.12 g/mol |
| 3-O-Methyltolcapone | 3-O-Methyltolcapone, CAS:134612-80-9, MF:C15H13NO5, MW:287.27 g/mol |
Problem: Difficulty in determining the complete molecular structure of a newly isolated natural product, especially when dealing with large, complex ring systems or flexible chains.
Solution: Employ advanced structural elucidation techniques that can handle complexity and require minimal material.
Problem: Ambiguous or incorrect assignment of stereocenters in a natural product, leading to failed biological activity replication.
Solution: Combine computational predictions with experimental validation.
Problem: The natural source produces the target compound in extremely low yields, insufficient for drug development or comprehensive bioactivity testing.
Solution: Bypass the native producer using synthetic biology and heterologous expression.
FAQ 1: Why is structural elucidation still a major bottleneck in natural product discovery? Structural elucidation remains challenging due to the intrinsic complexity of natural products. They often contain multiple chiral centers, large, fused ring systems, and flexible chains that make determining relative stereochemistry, especially between distal parts of the molecule, difficult with NMR alone. Furthermore, traditional X-ray crystallography requires large, well-formed crystals that are often impossible to grow with the limited quantities of material typically isolated [7].
FAQ 2: Our lead natural product has promising activity but poor solubility and metabolic stability. What are our options? This is a common challenge. The primary strategy is lead optimization through medicinal chemistry [9]. This involves:
FAQ 3: We've identified a promising biosynthetic gene cluster, but it's silent in the lab. How can we activate it? Two primary strategies exist:
FAQ 4: How do natural products and synthetic compounds compare in terms of chemical space and drug discovery potential? Chemoinformatic analyses show that natural products (NPs) occupy a distinct and more diverse region of chemical space compared to synthetic compounds (SCs). NPs are generally larger, more complex, have more chiral centers and oxygen atoms, and contain more non-aromatic rings. SCs, while more numerous, often have higher aromatic ring content and nitrogen/sulfur atoms. Critically, NPs have higher "biological relevance" due to their evolution to interact with biological macromolecules, which is why over 60% of pharmaceuticals are NP-derived or inspired [11] [12].
Table 1: Contribution of Natural Products to Approved Drugs (1981-2010) [9]
| Category | Definition | All Small-Molecule Drugs (%) | Anticancer Drugs (%) |
|---|---|---|---|
| Natural Product (N) | Unmodified natural product | 5.5% | 11.1% |
| Natural Product Derived (ND) | Semi-synthetic derivative | 27.9% | 32.3% |
| Synthetic, NP Pharmacophore (S*) | Synthetic, with NP-inspired active moiety | 5.1% | 11.1% |
| Totally Synthetic (S) | No NP inspiration | 36.0% | 20.2% |
| Total NP-Inspired | Sum of N, ND, S* | ~38.5% | ~54.5% |
Table 2: Comparison of Key Properties: Natural Products vs. Synthetic Compounds [12]
| Property | Natural Products (NPs) | Synthetic Compounds (SCs) |
|---|---|---|
| Molecular Size | Larger and increasing over time (MW, volume, etc.) | Smaller, constrained by drug-like rules |
| Rings | More rings, predominantly non-aromatic | Fewer rings, high proportion of aromatic rings |
| Structural Diversity | Higher scaffold diversity and complexity | Broader synthetic diversity but less unique |
| Biological Relevance | Higher, evolved to interact with biomolecules | Lower, despite larger chemical libraries |
| Chemical Space | More diverse and expanding | More concentrated and constrained |
Table 3: Essential Reagents and Materials for Overcoming NP Research Barriers
| Item | Function/Application | Example Use Case |
|---|---|---|
| Heterologous Host Strains | Genetically tractable chassis for expressing foreign BGCs. | Aspergillus nidulans A1145 ÎEMÎST for fungal clusters; Streptomyces albus for actinobacterial clusters [7] [5]. |
| Pathway-Specific Regulatory Genes | Positive regulators that activate transcription of silent BGCs. | Overexpression of SARP family regulators (e.g., fdmR1) to boost titers of target compounds like Fredericamycin A [5]. |
| Constitutive Promoters | Strong, always-on promoters to drive high-level gene expression. | ErmE* promoter for constitutive expression of biosynthetic or regulatory genes in heterologous hosts [5]. |
| MicroED Platform | Cryo-EM method for determining structures from nano-crystals. | Ab initio structural elucidation of new natural products like Py-469, solving stereochemistry where NMR fails [7]. |
| Machine Learning Models (e.g., NPstereo) | In-silico prediction of stereochemical configuration. | Assigning or correcting the stereochemistry of newly discovered NPs from their planar structure [8]. |
| Specialized Compound Databases | Curated collections of NP structures for mining and prediction. | COCONUT database for training ML models; Dictionary of Natural Products (DNP) for chemoinformatic analysis [12] [8]. |
| Montelukast-d6 | Montelukast-d6, MF:C35H36ClNO3S, MW:592.2 g/mol | Chemical Reagent |
| Pradimicin Q | Pradimicin Q, CAS:141869-53-6, MF:C24H16O10, MW:464.4 g/mol | Chemical Reagent |
Answer: This is a common challenge. The biological relevance of the natural product (NP) scaffold often justifies the optimization effort. Several strategies can be employed:
Answer: Poor PK is a frequent hurdle that can often be overcome through rational structural modification.
Answer: A significant number of NP databases exist, but their accessibility and focus vary. The table below summarizes key open-access resources [15].
Table 1: Selected Open-Access Natural Products Databases
| Database Name | Type / Focus | Approximate Number of Compounds | Key Features |
|---|---|---|---|
| COCONUT | Generalistic Collection | > 400,000 | The largest open collection of non-redundant NPs; available as a downloadable dataset [15]. |
| Various Resources | Thematic (e.g., Traditional Medicine, Geographic) | Varies | Many thematic databases focus on specific geographic regions, taxonomic groups, or traditional medicine applications [15]. |
| ZINC | Commercial Compounds | Includes NPs | Contains collections of commercially available NPs for virtual screening [15]. |
This protocol outlines the design, synthesis, and biological evaluation of a pseudo-natural product (pseudo-NP) library to discover new bioactive chemotypes [14] [6].
1. Design and In Silico Planning
2. Library Synthesis
3. Biological Evaluation
4. Hit Validation & Target Identification
The following workflow diagram illustrates the pseudo-NP discovery process:
This protocol is used to improve the potency and drug-like properties of an initial NP-derived hit [13].
1. Analogue Design
2. Library Synthesis and Profiling
3. Data Analysis and Lead Selection
The following flowchart visualizes the SAR optimization cycle:
Table 2: Key Research Reagent Solutions for NP-Based Drug Discovery
| Reagent / Resource | Function / Application | Examples / Notes |
|---|---|---|
| NP Fragment Libraries | Building blocks for designing pseudo-NP scaffolds or for BIOS. | Curated sets of NP-derived fragments that comply with the "rule of three" for fragments, ensuring favorable properties for library synthesis [14]. |
| Commercial NP Databases | Source of structures and metadata for dereplication and inspiration. | Dictionary of Natural Products (DNP), MarinLit. These are highly curated but require a subscription [15]. |
| Open NP Collections (e.g., COCONUT) | Source of structures for virtual screening and cheminformatic analysis. | COCONUT provides over 400,000 non-redundant NP structures for open research use [15]. |
| Screening Libraries (NP-Derived) | Collections of compounds for high-throughput screening (HTS). | Libraries based on terpenoid, polyketide, phenylpropanoid, and alkaloid scaffolds provide biologically prevalidated starting points for hit identification [13] [17]. |
| Catalysts for C-C Bond Formation | Enabling synthesis of complex NP-inspired scaffolds. | Essential for constructing the characteristic three-dimensional frameworks of NPs and their analogues (e.g., in meroterpenoid synthesis) [13]. |
| Taurolidine | Taurolidine (NMR) Powder|19388-87-5 | Taurolidine is a broad-spectrum antimicrobial agent for research, derived from taurine. This product is For Research Use Only (RUO). Not for human or veterinary use. |
| Allyl 3-amino-4-methoxybenzoate | Allyl 3-amino-4-methoxybenzoate|CAS 153775-06-5 | Allyl 3-amino-4-methoxybenzoate (CAS 153775-06-5) is a benzoate ester intermediate for pharmaceutical and peptide synthesis research. For Research Use Only. Not for human or veterinary use. |
Natural products (NPs) and their derivatives have been a cornerstone of pharmacotherapy for millennia, serving as a primary source of new medicines, particularly for cancer and infectious diseases [11]. Historical records, including ancient Egyptian papyri and traditional Chinese medicine texts, document the extensive use of medicinal plants, with many early isolated pure natural products like morphine, quinine, and cocaine originating from traditional remedies [18]. In the modern era, nearly half of all approved drugs between 1981 and 2019 can be traced back to unaltered NPs, derivatives, or NP-like pharmacophores, underscoring their enduring impact [19]. This technical support center leverages these historical successes to provide practical guidance for overcoming contemporary challenges in natural product research, with a focus on improving the chemical accessibility of NP leads.
Natural products demonstrate a remarkable and quantifiable advantage in the drug development pipeline. While they constitute a minority of early-stage patent applications (approximately 8% of patent compounds), their success rate increases steadily through clinical trial phases [19]. This trend suggests that NPs possess inherent properties, such as superior drug-likeness and lower toxicity, that make them more likely to succeed in later, more costly stages of development.
Table 1: Proportion of Natural Products, Hybrids, and Synthetics Across Drug Development Stages
| Development Stage | Natural Products | Hybrid Compounds | Synthetic Compounds |
|---|---|---|---|
| Patent Applications | ~8% | ~15% | ~77% |
| Clinical Trial Phase I | ~20% | ~15% | ~65% |
| Clinical Trial Phase III | ~26% | ~19% | ~55.5% |
| FDA Approved Drugs | ~25% | ~20% | ~25% (Purely synthetic) |
Data sourced from analysis of over 1 million patent applications and clinical trial data [19].
Analysis of NP structural classes that successfully progress from Phase I trials to approval reveals specific scaffolds that are enriched in approved drugs. Terpenoids show a notable 20% relative increase, while fatty acids and alkaloids demonstrate increases of 7% and 6%, respectively [19]. Among NP superclasses, β-lactams and peptide alkaloids are significantly enriched, indicating these classes exhibit lower failure rates and represent privileged structures for drug discovery [19].
Table 2: Essential Research Reagents and Solutions for NP-Based Drug Discovery
| Reagent / Solution | Function & Application | Technical Notes |
|---|---|---|
| High-Throughput Screening (HTS) Assays | Rapid phenotypic or target-based screening of complex NP extracts or pure compounds [11]. | Enables processing of large compound libraries; can be combined with robotic separation. |
| Advanced Analytical Tools (e.g., LC-HRMS) | Separation, dereplication, and characterization of NPs from complex mixtures [11] [1]. | Hyphenated techniques like LC-HRMS-NMR are crucial for identifying novel scaffolds. |
| In Silico Prediction Tools (e.g., NatGen) | Predicts 3D structures and chiral configurations of NPs, a major bottleneck in NP research [20]. | Achieves high accuracy (e.g., 96.87% on benchmarks); vital for NPs with unresolved stereochemistry. |
| NP Databases (e.g., COCONUT, ChEMBL) | Provide curated structural and bioactivity data for virtual screening and machine learning [1] [20]. | Essential for cheminformatics; quality and curation of data are critical. |
| ADMET In Silico Prediction Tools | Early computational prediction of absorption, distribution, metabolism, excretion, and toxicity profiles [1]. | Helps prioritize compounds with favorable drug-like properties, reducing late-stage attrition. |
| Difethialone | Difethialone, CAS:104653-34-1, MF:C31H23BrO2S, MW:539.5 g/mol | Chemical Reagent |
| Dimethyl malonate | Dimethyl malonate, CAS:108-59-8, MF:C5H8O4, MW:132.11 g/mol | Chemical Reagent |
Q1: Why invest in natural products given the dominance of synthetic compounds in early patents? Despite synthetic compounds overwhelmingly outnumbering NPs in patent applications (approx. 77% vs. 23% for NPs and hybrids combined), the success rate of NPs in clinical trials is significantly higher [19]. The proportion of NP and hybrid compounds increases steadily from Phase I (approx. 35%) to Phase III (approx. 45%), with an inverse trend observed for synthetics [19]. This higher "survival rate" is likely due to evolutionary pre-optimization for biological relevance, superior drug-like properties, and lower toxicity.
Q2: What level of bioactivity should be considered promising for an NP extract or compound? Potency must be considered alongside other factors like toxicity, selectivity, and structural complexity. For initial screening in areas like insecticide development, an extract with an LC50 of approximately 100 ppm is a good starting point, while pure compounds with an LC50 ⤠10 ppm are strong candidates for prototype development [3]. Activity at low concentrations is advantageous, but a compound with moderate potency and an excellent safety profile or novel mechanism should not be discounted.
Q3: What are the major reasons for the high attrition rate of drug candidates, and how do NPs address this? The vast majority of clinical candidates fail due to a lack of clinical efficacy and/or unmanageable toxicity [19]. NPs address these issues by often possessing inherently validated biological functions through evolutionary pressure. They frequently feature molecular scaffolds that are selective for cellular targets and have desirable ADME properties [19]. In vitro and in silico studies consistently show that NPs and their derivatives tend to be less toxic than synthetic counterparts, directly addressing a major cause of clinical failure [19].
Challenge 1: Difficulty in identifying and isolating the specific bioactive compound from a complex natural extract.
Challenge 2: The 3D structure, particularly chiral configuration, of a natural product is unknown, hindering mechanistic and docking studies.
Challenge 3: An active NP is not available from commercial suppliers, and re-isolation from the natural source is impractical or unsustainable.
Challenge 4: Translating in silico NP hits into experimentally validated leads due to sourcing and testing bottlenecks.
datarail) to systematically plan the drug response experiment. This includes specifying cell types, drugs, dose ranges, and plate layouts in a machine-readable, error-free format [21].gr50_tools) to normalize data and calculate robust sensitivity metrics like IC50 or GR50, which corrects for effects of cell division rate [21].
Diagram 1: NP Bioactive Compound Identification Workflow
Diagram 2: In-silico Hit Validation Pipeline
The historical success of natural products as drugs is not serendipitous but is rooted in their evolutionary optimization for biological interaction and their vast, untapped chemical diversity. The case studies of drugs like artemisinin, paclitaxel, and morphine provide a clear roadmap for future discovery. By systematically addressing the key bottlenecks of NP researchâsuch as compound identification, structural elucidation, and sustainable supplyâwith modern technological solutions like AI-based structure prediction, automated screening platforms, and synthetic biology, researchers can significantly improve the chemical accessibility of natural product leads. Integrating these advanced methodologies into a rational, data-driven workflow will ensure that natural products continue to be a vital source of innovative therapeutics for unmet medical needs.
The primary goal is to improve the "druggability" of natural product leads. This involves modifying their chemical structure to enhance desirable properties such as potency, selectivity, and pharmacokinetics (like solubility and metabolic stability), while reducing toxicity. These modifications are essential for transforming a naturally occurring lead compound into a viable drug candidate [22] [23].
The location of a functional group on the molecular scaffold is highly influential to its biological activity [24]. A change in position can significantly alter how the molecule interacts with its biological target (e.g., a protein or enzyme), thereby affecting the drug's efficacy and specificity.
A major challenge is that traditional methods for moving functional groups often require multiple synthetic steps (five or more). This lengthy process is inefficient and can be complicated by unwanted side reactions, which reduce yield and create purification difficulties [24].
Problem: Traditional carbonyl transposition is a multi-step, inefficient process. Solution: Implement a modern, triflate-mediated α-amination strategy. This approach uses two cooperative catalysts to enable a direct, selective 1,2-transposition of the carbonyl group, reducing the required steps to just one or two. This method minimizes unwanted side reactions and offers superior control over the final position of the carbonyl [24].
Problem:Complex natural product scaffolds often have poor solubility or bioavailability. Solution: Focus on semi-synthesis. Use the complex natural product as a core scaffold and perform targeted functional group manipulations. This preserves the beneficial structural complexity while allowing you to fine-tune specific properties. Key transformations include:
Problem: Bioactivity is lost during the fractionation and isolation process. Solution: Employ a rigorous bioactivity-guided fractionation protocol [23]. After each separation step (e.g., chromatography), test all fractions for the desired biological activity. Only proceed with fractions that retain activity. This ensures the active component is not discarded and helps identify the specific compound responsible for the effect.
Table 1: Contribution of Natural Products to New Drug Approvals (1981-2014) [23]
| Category of Drug | Percentage of Total Approved Drugs | Example Compounds |
|---|---|---|
| Pure Natural Products | 4% | Morphine, Paclitaxel |
| Natural Product-Derived | 21% | Semisynthetic antibiotics, Simvastatin |
| Synthetic drugs based on natural pharmacophores | 4% | Aspirin (from salicin) |
| Herbal Mixtures | 9.1% | - |
| Mosapride citrate dihydrate | Mosapride Citrate Dihydrate | Selective 5-HT4 receptor agonist for GI motility research. Mosapride citrate dihydrate is of high purity. For Research Use Only. Not for human use. |
| okadaic acid ammonium salt | okadaic acid ammonium salt, CAS:155716-06-6, MF:C44H71NO13, MW:822.0 g/mol | Chemical Reagent |
Table 2: Success Rates and Challenges in Natural Product Drug Discovery [27] [23]
| Parameter | Finding/Statistic | Implication for Research |
|---|---|---|
| Historical Success | 28% of NCEs (1981-2002) were natural-derived [23] | Validates the strategy of using natural products as leads. |
| Current Industry Trend | Many large pharma companies reduced NP R&D [27] | Highlights perceived challenges like supply and complexity. |
| Reported Hit Rate | Industry perceives higher HTS hit rates with NPs than academia [27] | Suggests advanced infrastructure improves success. |
Title: Simplified Carbonyl Transposition via Triflate-Mediated α-Amination [24]
Objective: To relocate a carbonyl group to an adjacent carbon atom in a single, efficient step.
Materials:
Procedure:
Key Consideration: This method is notable for its mild reaction conditions and excellent selectivity, avoiding the extensive protecting group manipulation typically required in traditional sequences.
Table 3: Essential Reagents for Functional Group Manipulation
| Reagent/Catalyst | Primary Function | Application Example |
|---|---|---|
| Palladium Catalysts | Facilitates cross-coupling and amination reactions. | Key component in the triflate-mediated transposition cascade [24]. |
| Triflic Anhydride (TfâO) | Powerful electrophile for introducing the triflate leaving group. | Generates the vinyl triflate intermediate during carbonyl transposition [24]. |
| O-benzoylhydroxylamines | Serve as electrophilic amination reagents. | Used to install the initial nitrogen-containing group in the α-amination step [24]. |
| Silane Reductants (e.g., EtâSiH) | Hydride source for reduction reactions. | Final reduction step to complete the carbonyl transposition [24]. |
| Chiral Ligands | Induce asymmetry in catalytic reactions to create single enantiomer products. | Critical for achieving stereoselectivity in the Pd-catalyzed amination step [24]. |
Diagram Title: Carbonyl 1,2-Transposition Workflow
Diagram Title: Natural Product Lead Optimization Pathway
What is the primary goal of SAR-directed optimization for natural products? SAR-directed optimization aims to systematically modify a natural product lead compound to enhance its drug-like properties. The process involves making structural changes and analyzing how these changes affect biological activity to establish a clear relationship between chemical structure and pharmacological effect [9]. The strategy not only addresses drug efficacy but also aims to improve ADMET (Absorption, Distribution, Metabolism, Excretion, and Toxicity) profiles and chemical accessibility associated with natural leads [9].
How does SAR-directed optimization fit into the broader drug discovery workflow? SAR-directed optimization typically occurs after the identification of a bioactive natural product lead (hit) and before preclinical development. It serves as a critical bridge where promising compounds are systematically improved through iterative design, synthesis, and testing cycles. This process transforms a natural product with initial activity into a optimized lead compound with desired potency, selectivity, and pharmacological properties [28].
What are the key methodological approaches for establishing SAR? Researchers employ multiple complementary approaches to establish meaningful SAR:
What is the difference between traditional SAR and the newer C-SAR approach? Traditional SAR studies are typically conducted on a single parent chemical structure, while Cross-Structure-Activity Relationship (C-SAR) analyzes pharmacophoric substituents across diverse chemotypes. C-SAR facilitates SAR expansion to any chemotype requiring modification based on existing knowledge of various compounds targeting the same biological entity, thus accelerating structural development [31].
How can we navigate complex activity landscapes effectively? Activity landscapes can be highly variable, containing both smooth regions where gradual structural changes cause moderate activity shifts, and "activity cliffs" where minimal modifications substantially influence biological effects [32]. To address this:
What strategies address the synthetic challenges of natural product optimization? Natural products often present synthetic intractability and limited availability. Several specialized strategies have been developed:
How can we separate desired target activity from undesired off-target effects? The case study of harmine optimization provides specific guidance. Harmine is a potent DYRK1A inhibitor but suffers from undesired potent inhibition of MAO-A [34]. Through systematic SAR studies involving over 60 analogues, researchers identified that:
Experimental Protocol: Build-up Library Construction and Evaluation
Diagram Title: MraY Inhibitor Build-up Library Workflow
Table: Key Reagents and Materials for SAR Studies
| Reagent/Material | Function in SAR Studies | Application Example |
|---|---|---|
| Aldehyde Core Fragments | Provide conserved binding motif for target interaction | MraY inhibitors containing essential uridine moiety [30] |
| Hydrazine Accessory Fragments | Introduce structural diversity to modulate properties | 98 fragments including benzoyl-type, phenyl acetyl-type, and lipid amino acid variants [30] |
| Matched Molecular Pairs (MMPs) | Enable identification of critical structural changes | Pairs of compounds differing only by specific structural features for C-SAR analysis [31] |
| Selective HDAC6 Inhibitors | Tool compounds for target-specific SAR development | Dataset for C-SAR approach validation [31] |
| β-Carboline Scaffolds | Core structure for kinase inhibitor optimization | Harmine analogs for DYRK1A inhibitor development with reduced MAO-A inhibition [34] |
How do we interpret complex activity landscapes? Activity landscapes can be categorized into three main types:
What computational approaches support modern SAR studies?
Table: Comparison of SAR Strategies for Natural Product Optimization
| Strategy | Key Approach | Advantages | Limitations |
|---|---|---|---|
| Traditional SAR | Sequential modification of parent structure | Established methodology, clear structure-progression | Limited to single chemotype, synthetic challenges [9] |
| C-SAR | Cross-analysis of pharmacophores across diverse chemotypes | Accelerates structural development, applicable to various chemotypes [31] | Requires diverse dataset, potential contradictory data between chemotypes [31] |
| Build-up Library | Fragment ligation with in situ screening | Rapid library generation, minimal purification, direct biological evaluation [30] | Dependent on efficient ligation chemistry, potential stability issues with products [30] |
| BIOS | Library design based on privileged natural product scaffolds | Higher probability of bioactivity, requires fewer compounds [33] | Limited structural diversity, focused on known bioactive scaffolds [33] |
Diagram Title: SAR Strategy Comparison for Natural Product Optimization
Problem: Molecules proposed by scaffold-hopping tools are structurally novel but appear difficult or impractical to synthesize in a laboratory setting.
Solutions:
Problem: After replacing the core scaffold, the new compound no longer effectively binds to the target or exhibits the desired biological effect.
Solutions:
Problem: The software fails to process the input molecular structure and returns a parsing or validation error.
Solutions:
Problem: The scaffold-hopping algorithm produces molecules that are too structurally similar to the input, providing limited inspiration for novel patentable candidates.
Solutions:
- -replace_scaffold_files option, allowing you to explore niche chemical spaces, such as those derived from natural products [35].Q1: What is the fundamental difference between pharmacophore-oriented design and traditional scaffold hopping? A1: While both aim to identify new core structures, pharmacophore-oriented design specifically uses the 3D arrangement of features essential for biological activity (e.g., hydrogen bond donors/acceptors, hydrophobic centers) as the primary constraint for searching and designing new molecules [38] [36]. Traditional scaffold hopping may rely more heavily on 2D topological similarity or molecular shape. The pharmacophore approach ensures that the replaced scaffold maintains the critical functional geometry for target binding, even if the underlying carbon skeleton is vastly different [38].
Q2: When should I consider using a scaffold-hopping strategy in my natural product optimization project? A2: You should consider scaffold hopping when facing one or more of these common challenges in natural product lead optimization [35] [2] [39]:
Q3: How do AI-based methods like TransPharmer improve upon earlier scaffold-hopping techniques? A3: AI-based methods like TransPharmer integrate deep learning with pharmacophore modeling, offering several key advantages [36] [40]:
Q4: Can you provide a specific example where scaffold hopping successfully retained potency? A4: Yes. In a study applying the AI-AAM scaffold-hopping method, the SYK inhibitor BIIB-057 was used as a reference. The method identified a structurally different compound, XC608. Experimental validation showed that both compounds exhibited very similar and high potency, with IC50 values of 3.9 nM and 3.3 nM, respectively. This demonstrates a successful scaffold hop that maintained nanomolar-level pharmacological activity against the SYK target [37].
Q5: What are the key metrics to evaluate the success of a scaffold-hopping campaign? A5: Success should be evaluated using a combination of computational and experimental metrics, summarized in the table below.
| Metric Category | Specific Metric | Description and Rationale |
|---|---|---|
| Computational | Tanimoto Similarity | Measures 2D structural similarity; a successful hop often has lower similarity [35]. |
| Shape/Pharmacophore Similarity | Measures 3D volume and feature overlap (e.g., ElectroShape); should be high to retain activity [35]. | |
| Synthetic Accessibility (SA) Score | Predicts ease of synthesis; lower scores are more favorable [35]. | |
| Drug-Likeness (QED) | Quantitative Estimate of Drug-likeness; higher scores indicate more drug-like properties [35]. | |
| Experimental | Binding Affinity (IC50/Kd) | Measures potency; should be comparable to or better than the lead compound [37]. |
| Target Selectivity | Assesses activity against off-targets; a new scaffold may have a improved or different selectivity profile [37]. | |
| ADMET Profile | Evaluates absorption, distribution, metabolism, excretion, and toxicity; the goal is improvement over the lead [39]. |
This protocol provides a step-by-step guide for generating novel scaffolds from a known active compound using the ChemBounce tool [35].
1. Input Preparation
2. Tool Execution
-o: Specify the directory where results will be saved.-i: Path to a file containing the input SMILES string.-n: Controls the number of novel structures to generate for each identified fragment.-t: (Optional) Tanimoto similarity threshold (default 0.5). A lower value encourages greater structural diversity.3. Output and Analysis
This protocol outlines a general method for experimentally confirming that a scaffold-hopped compound retains its biological activity, based on the validation performed for the AI-AAM method [37].
1. Compound Preparation
2. In Vitro Kinase Activity Assay
3. Selectivity Profiling
The following diagram illustrates the logical workflow and decision points in a typical pharmacophore-oriented scaffold-hopping process, integrating the tools and strategies discussed.
Diagram Title: Scaffold Hopping Workflow & Decision Path
The following table details key computational tools and resources essential for implementing pharmacophore-oriented scaffold hopping.
| Item Name | Type | Function / Application |
|---|---|---|
| ChemBounce | Software Framework | An open-source tool for scaffold hopping that uses a curated library of synthetically accessible fragments and evaluates compounds based on Tanimoto and electron shape similarity [35]. |
| TransPharmer | AI Generative Model | A generative model that integrates interpretable pharmacophore fingerprints with a GPT framework for de novo molecule generation and scaffold elaboration, excelling at producing structurally novel, bioactive ligands [36]. |
| ROCS (Rapid Overlay of Chemical Structures) | Software Tool | A standard tool for 3D shape-based molecular comparison and virtual screening that checks for optimal shape overlap and matching of pharmacophoric features [38]. |
| ElectroShape | Algorithm/Descriptor | A method for calculating molecular similarity based on both 3D shape and charge distribution, implemented in tools like ChemBounce to better preserve biological activity during hopping [35]. |
| ChEMBL Database | Database | A large, open-scale bioactivity database. Used to build curated, synthesis-validated scaffold libraries that underpin tools like ChemBounce [35]. |
| ErG Fingerprints | Molecular Descriptor | A type of pharmacophoric fingerprint used to measure pharmacophoric similarity between molecules, demonstrating potential for scaffold hopping applications [36]. |
| 3-Methyl-5-oxohexanal | 3-Methyl-5-oxohexanal, CAS:146430-52-6, MF:C7H12O2, MW:128.17 g/mol | Chemical Reagent |
| Nitroxynil | Nitroxynil, CAS:1689-89-0, MF:C7H3IN2O3, MW:290.01 g/mol | Chemical Reagent |
Q1: What are the primary advantages of using a physics-based docking method like RosettaVS over deep learning approaches for virtual screening when the binding site is known?
A1: In scenarios where the binding site is known, physics-based ligand docking methods, such as the RosettaVS protocol, have been shown to continue to outperform deep learning models [41]. While deep learning methods are better suited for blind docking problems and offer significantly reduced computation times, physics-based methods provide greater generalizability to unseen protein-ligand complexes and can more accurately model receptor flexibility, including side chains and limited backbone movement, which is critical for many targets [41].
Q2: Our virtual screening campaign against an ultra-large library is prohibitively slow. What strategies can we use to accelerate the process without significantly compromising accuracy?
A2: To efficiently screen multi-billion compound libraries, we recommend a two-tiered strategy [41]:
Q3: How can we validate the accuracy of our virtual screening platform's pose and affinity predictions?
A3: It is critical to benchmark your method's performance on standard datasets and, where possible, validate predictions experimentally [41].
Q4: Our research focuses on natural products. What specific challenges does this present for virtual screening, and how can AI help address them?
A4: Natural product (NP) drug discovery faces unique challenges that AI and in-silico methods are poised to address [42]:
Q5: What are the key metrics for evaluating the success of a virtual screening campaign, and what values indicate good performance?
A5: The success of a virtual screening campaign is typically quantified using several metrics. The table below summarizes key benchmarks from the RosettaVS platform on the CASF-2016 dataset [41].
Table 1: Key Performance Metrics for Virtual Screening from RosettaVS Benchmarks
| Metric | Description | Benchmark Performance (CASF-2016) |
|---|---|---|
| Enrichment Factor (EF1%) | Measures the concentration of true binders in the top 1% of the ranked list. | 16.72 (significantly outperforming the second-best method at 11.9) [41] |
| Success Rate (Top 1%) | The percentage of targets for which the best binder was ranked in the top 1%. | Leading performance, surpassing other methods [41] |
| Docking Power | The ability to identify the native binding pose from decoys. | Achieved leading performance in the docking power test [41] |
Table 2: Common Virtual Screening Issues and Solutions
| Problem | Potential Cause | Solution |
|---|---|---|
| Low Hit Rate | Inaccurate scoring function; inadequate chemical space coverage; over-reliance on a single docking algorithm. | Validate the scoring function on a benchmark like CASF; consider consensus scoring from multiple methods; ensure the screened library is diverse and relevant to the target (e.g., NP-inspired libraries for certain targets) [41] [42]. |
| Inaccurate Pose Prediction | Insufficient sampling of ligand conformational space; inability to model critical receptor flexibility. | Use a docking protocol that allows for full ligand flexibility and incorporates receptor side-chain and limited backbone flexibility, as implemented in RosettaVS's VSH mode [41]. |
| Inconsistent Performance Across Targets | Scoring function bias towards certain protein families or ligand types; suboptimal active learning for a new target. | Use a robust, physics-based force field like RosettaGenFF-VS that has been shown to perform well across diverse targets. For AI-guided screening, ensure the active learning model is adequately trained on a representative set of the library for the new target [41]. |
Protocol 1: AI-Accelerated Virtual Screening of an Ultra-Large Library
This protocol outlines the workflow for screening a multi-billion compound library using the OpenVS platform, which integrates active learning with the RosettaVS docking protocol [41].
AI-Accelerated Virtual Screening Workflow
Protocol 2: Validation of Docking Pose via X-ray Crystallography
This protocol describes the steps for experimentally validating a computationally predicted ligand pose, a critical step in confirming the effectiveness of the virtual screening method [41].
Table 3: Essential Software and Platforms for Advanced Virtual Screening
| Tool / Platform | Type | Primary Function | Key Feature |
|---|---|---|---|
| OpenVS Platform [41] | Open-Source Software Platform | AI-accelerated virtual screening of ultra-large libraries. | Integrates active learning with the RosettaVS physics-based docking protocol for efficiency and accuracy. |
| RosettaVS (Rosetta GALigandDock) [41] | Physics-Based Docking Protocol | Predicts protein-ligand complex structures and binding affinities. | Models full receptor side-chain and limited backbone flexibility; includes VSX (fast) and VSH (accurate) modes. |
| RosettaGenFF-VS [41] | Physics-Based Force Field | Scoring function for ranking ligands in virtual screening. | Combines enthalpy calculations with a new entropy model, optimized for virtual screening. |
| Mirabilis [43] | In-Silico Tool | Predicts the carryover and purge of potentially mutagenic impurities (PMIs) during API synthesis. | Uses a knowledge base to predict reactivity, solubility, and volatility purges, supporting ICH M7 Option 4. |
| InsilicoGPT [42] | AI Chatbot (Q&A Tool) | Provides instant answers from research papers. | Facilitates quick retrieval of specific information and references from the scientific literature. |
| Fluazinam | Fluazinam, CAS:79622-59-6, MF:C13H4Cl2F6N4O4, MW:465.09 g/mol | Chemical Reagent | Bench Chemicals |
| Flomoxef | Flomoxef, CAS:99665-00-6, MF:C15H18F2N6O7S2, MW:496.5 g/mol | Chemical Reagent | Bench Chemicals |
WHALES (Weighted Holistic Atom Localization and Entity Shape) is a novel molecular representation designed to facilitate scaffold hopping, particularly from complex natural products (NPs) to synthetically accessible compounds with similar biological activity [44] [45]. Unlike reductionist descriptors that focus on individual molecular features (e.g., presence of specific fragments), WHALES provides a holistic representation that simultaneously encodes 3D molecular shape, geometric interatomic distances, and atomic property distributions (specifically, partial charges) [45]. This enables the identification of isofunctional chemotypes that occupy similar regions of chemical space despite having different underlying molecular frameworks [44].
The calculation of WHALES descriptors is a multi-step procedure that transforms 3D molecular structural information into a fixed-length numerical vector [44] [45]. The workflow for this calculation is illustrated in the diagram below:
Step 1: Input Preparation The process begins with the generation of a energy-minimized 3D molecular conformation (typically using the MMFF94 forcefield) and the calculation of partial atomic charges (δi) [44] [45]. WHALES can use different partial charge calculation methods, such as the fast Gasteiger-Marsili method or more computationally intensive quantum mechanical (DFTB+) approaches [44]. A charge-agnostic version (WHALES-shape) that only uses atomic coordinates is also available [44].
Step 2: Atom-Centered Covariance Matrix Calculation For each non-hydrogen atom j in the molecule, a weighted, atom-centered covariance matrix Sw(j) is computed [45]. This matrix captures the distribution of surrounding atoms and their partial charges, effectively forming an ellipsoid around atom j that is oriented toward regions of high atomic density and charge [45]. The formula is given by:
Sw(j) = [ Σi=1 to n |δi| ⢠(xi - xj)(xi - xj)^T ] / [ Σi=1 to n |δi| ]
Where:
Step 3: Atom-Centered Mahalanobis (ACM) Distance Calculation From each covariance matrix Sw(j), the ACM distance from the center j to every other atom i is calculated [44] [45]. This creates an ACM distance matrix. The ACM distance is computed as:
ACM(i,j) = (xi - xj)^T ⢠Sw(j)^-1 ⢠(xi - xj)
This normalized, dimensionless distance accounts for local molecular feature distributionsâatoms in high-variance directions have smaller relative distances than those in low-variance, peripheral regions [44] [45].
Step 4: Calculation of Atomic Indices Three key atomic indices are derived from the ACM matrix for each atom j:
To distinguish atomic properties, these indices are assigned negative values for negatively charged atoms (δ_j < 0) [45].
Step 5: Descriptor Vector Assembly Finally, the distribution of these atomic indices (Isol, Rem, IR) across all non-hydrogen atoms is captured by computing their minimum, maximum, and decile (10th, 20th, ..., 90th percentiles) values. This yields a fixed-length vector of 33 molecular descriptors, enabling direct comparison of molecules of different sizes [44] [45].
WHALES descriptors have been rigorously tested against seven state-of-the-art molecular representations to evaluate their scaffold-hopping potential [44]. The benchmark study used 30,000 bioactive compounds from ChEMBL22 across 182 biological targets [44]. Performance was measured by the Scaffold Diversity of Actives (SDA%), which is the ratio of unique Murcko scaffolds to the number of actives retrieved in the top 5% of similarity search rankings [44].
Table 1: Performance Comparison of Molecular Descriptors for Scaffold Hopping
| Descriptor | Dimensionality | Encoded Information | Scaffold-Hopping Ability (SDA% ± SD) |
|---|---|---|---|
| WHALES-DFTB+ | 3D | Atom distributions, shape, & QM charges | Highest performance (Outperformed benchmarks on 89% of targets) [44] |
| WHALES-GM | 3D | Atom distributions, shape, & empirical charges | High performance [44] |
| WHALES-shape | 3D | Atom distributions & shape only (δ_i=1) | High performance [44] |
| GETAWAY | 3D | Molecular size, shape, atom types & properties [44] | High performance [44] |
| WHIM | 3D | 3D atom distribution & molecular properties [44] | High performance [44] |
| CATS | 2D | Topological pharmacophore pairs [44] | Moderate performance [44] |
| Matrix-Based | 2D | Molecular branching, shape, & heteroatoms [44] | Moderate performance [44] |
| MACCS | 1D | 166 predefined structural fragments [44] | Lower performance (75 ± 12) [44] |
| ECFPs | 1D | Atom-centered radial fragments [44] | Lower performance (73 ± 12) [44] |
| Constitutional | 0D/1D | Molecular weight, atom/ring counts [44] | Information not provided in search results |
The benchmark analysis revealed that 3D descriptors generally outperformed 2D and 1D representations in scaffold-hopping ability [44]. Fingerprint-based methods (ECFPs, MACCS) showed the lowest SDA% values, likely due to their reliance on specific structural fragments, which limits their ability to identify structurally diverse, isofunctional compounds [44]. WHALES descriptors consistently demonstrated superior performance, successfully identifying novel chemotypes across a wide range of biological targets [44].
The following diagram summarizes the experimental workflow for benchmarking descriptor performance:
In a prospective application, WHALES was used to discover novel Retinoid X Receptor (RXR) modulators [44]. Using known synthetic drugs as queries, WHALES identified four novel RXR agonists with innovative molecular scaffolds, including a rare non-acidic chemotype [44]. One agonist demonstrated high selectivity across 12 nuclear receptors and efficacy comparable to the drug bexarotene in inducing gene expression of ATP-binding cassette transporter A1, angiopoietin-like protein 4, and apolipoprotein E [44].
Q1: What are the main advantages of WHALES over simpler fingerprint methods like ECFPs?
WHALES descriptors offer superior scaffold-hopping ability because they capture holistic 3D molecular shape and pharmacophore patterns, rather than relying on specific structural fragments [44] [45]. While ECFPs and other fingerprints are valuable for finding structurally similar compounds, they often miss isofunctional molecules with different backbone structures [44]. WHALES excels at identifying these structurally diverse but functionally similar compounds, making it particularly valuable for natural product-inspired drug discovery where synthetic complexity is a concern [45].
Q2: Which partial charge calculation method should I use for my WHALES analysis?
The choice depends on your computational resources and the required level of accuracy [44]:
Q3: My WHALES similarity search returned compounds that look very different from my query natural product. Is this expected?
Yes, this is the intended scaffold-hopping behavior [45]. WHALES is designed to identify compounds that occupy similar regions of chemical space (similar shape and pharmacophore distribution) rather than those with obvious structural similarity [45]. Validate these hits experimentally, as they may represent novel chemotypes with the desired biological activity but improved synthetic accessibility [44] [45]. In prospective studies, this approach successfully identified synthetic cannabinoid receptor modulators that were structurally less complex than their natural product templates [45].
Q4: How sensitive are WHALES descriptors to molecular conformation?
WHALES descriptors are robust to small conformational changes due to the binning procedure used in descriptor calculation (Step 5) [45]. However, as with any 3D descriptor, the input conformation should represent a reasonable, energy-minimized structure [44] [45]. The use of MMFF94 energy-minimized structures is recommended for consistent results [44] [45].
Table 2: Common Issues and Solutions When Using WHALES Descriptors
| Problem | Possible Causes | Solutions |
|---|---|---|
| Poor retrieval of active compounds in virtual screening | ⢠Incorrect 3D conformation generation⢠Poor choice of partial charge method⢠Query molecule is not a suitable template | ⢠Verify conformation energy minimization⢠Test multiple partial charge methods⢠Use multiple diverse active compounds as queries |
| Computational performance too slow for large compound libraries | ⢠Using DFTB+ partial charges⢠Inefficient implementation of ACM matrix calculation | ⢠Switch to Gasteiger-Marsili charges for initial screening⢠Optimize code or use compiled implementations⢠Consider WHALES-shape for fastest performance |
| Descriptors fail to distinguish known active from inactive compounds | ⢠Biological activity may not be strongly dependent on 3D shape/pharmacophores⢠Descriptors may be too abstract for the specific target | ⢠Combine WHALES with other complementary descriptors⢠Validate descriptor relevance with known actives/inactives before prospective screening |
| Difficulty reproducing published results | ⢠Different conformational sampling protocols⢠Alternative partial charge implementations⢠Variations in descriptor normalization | ⢠Use exact same protocols as original publication (MMFF94, specific software versions)⢠Contact original authors for implementation details |
Table 3: Essential Computational Tools for WHALES Descriptor Analysis
| Tool Category | Specific Examples | Function in WHALS Workflow |
|---|---|---|
| 3D Conformation Generation | ⢠MMFF94 force field [44] [45]⢠Other molecular mechanics force fields | Generation of energy-minimized input structures required for descriptor calculation |
| Partial Charge Calculation | ⢠Gasteiger-Marsili method [44] [45]⢠DFTB+ (Density-functional-based tight-binding) [44]⢠Other quantum mechanical methods | Computation of atomic partial charges (δ_i) used as weights in the covariance matrix |
| Molecular Descriptor Implementation | ⢠Custom implementations (Python, C++, etc.)⢠Cheminformatics toolkits (RDKit, OpenBabel) | Calculation of WHALES descriptors and other benchmark descriptors for comparison |
| Similarity Search & Virtual Screening | ⢠In-house database systems⢠Commercial screening platforms (OpenEye, Schrödinger) | Performing similarity searches using WHALES descriptors to identify novel chemotypes |
| Benchmarking & Validation | ⢠ChEMBL database [44]⢠Dictionary of Natural Products (DNP) [45] | Access to bioactive compounds for validation and performance benchmarking |
Q1: What is ChemSAR and how does it specifically benefit research on natural products? ChemSAR is a web-based pipelining platform for generating Structure-Activity Relationship (SAR) classification models for small molecules [46]. For researchers working on natural product leads, it provides an integrated, step-by-step workflow that helps overcome key challenges like moderate potency, limited aqueous solubility, and complex chemical structures [47]. By automating the process of structure preprocessing, descriptor calculation, and model building, it allows you to systematically study and optimize natural product-inspired analogues without requiring advanced programming skills [46] [47].
Q2: My molecular dataset contains natural products with complex stereochemistry and salts. How should I preprocess this data in ChemSAR? For complex natural product datasets, you should use the Structure Preprocessing module. It is recommended to select the following procedures [48]:
Q3: After feature calculation, I have too many molecular descriptors. Which feature selection method in ChemSAR is most suitable for a natural product dataset? ChemSAR offers multiple feature selection methods. For natural product datasets, which can be complex and high-dimensional, a combination approach is often best. You can use the following sequence [48]:
Q4: I've built an SAR model. How can I use ChemSAR to predict the activity of newly designed natural product analogues? Once you have a reliable model from the "Model Building" stage, you use the dedicated "Prediction" module [48].
Q5: What are the common reasons for a "nan" or gibberish value error during the data preprocessing stage? This error in the "Imputation of missing values" step typically occurs when the calculated value for a molecular descriptor is infinite (inf or -inf) or cannot be recognized by the platform's internal functions [48]. This can happen with certain complex molecular structures. The solution is to run the imputation module, which can handle these missing or incorrect values using strategies like mean or median imputation, ensuring your dataset is clean for model building [48].
Problem: Your SAR model has low predictive accuracy on the test set, or results vary widely with small changes in the training data.
| Potential Cause | Solution |
|---|---|
| Insufficient or Low-Quality Data | Ensure your dataset is large enough and the activity data is reliable. For natural products, carefully curate structures and associated bioactivity data from credible sources. |
| Incorrect Applicability Domain | The model is being used to predict molecules that are structurally very different from the compounds it was trained on. When predicting new analogues, ensure they fall within the chemical space of your training set [49]. |
| Suboptimal Feature Selection | The selected molecular descriptors may not be relevant to the biological activity. Revisit the feature selection stage. Try different methods (univariate, tree-based, RFE) and select the feature set that yields the best and most stable cross-validation performance [48]. |
| Improper Hyperparameters | The parameters for the machine learning algorithm (e.g., n_estimators in Random Forest) may not be optimized. Use the grid search functionality in the "Model Selection" stage to systematically find the best parameters for your specific dataset [48]. |
Problem: The "Feature Calculation" job fails or returns an error message.
| Potential Cause | Solution |
|---|---|
| Invalid Molecular Structure | The input file may contain invalid SMILES strings or structures that cannot be standardized. Go back to the "Structure Preprocessing" module and run your input file again with the 'Removing salts' and 'Adding hydrogen atoms' options selected [48]. |
| Unsupported File Format | The platform primarily accepts SMILES and SDF formats [46]. Convert your file into one of these formats using tools like OpenBabel before uploading [46]. |
| Server Timeout | Large datasets or complex calculations can take time. ChemSAR uses session and AJAX technology to prevent timeouts. You can close your browser and check the results later using your unique job ID in the "My Report" module [48]. |
Problem: The training and test sets are not representative, or the imputation step is not handling missing data correctly.
| Potential Cause | Solution |
|---|---|
| Unbalanced Activity Classes | If your dataset has an imbalance between active and inactive compounds, a random split might create unrepresentative sets. Check the distribution of activities in your training and test sets. You may need to use stratified sampling techniques outside the platform or ensure a larger dataset. |
| Incorrect Handling of Missing Values | The chosen imputation strategy (e.g., mean, median) might be inappropriate for the type of descriptor. Examine your data (File 5 and File 7) before imputation to understand the nature of the missing values and choose the strategy accordingly [48]. |
This protocol outlines the steps to build a foundational SAR model using the ChemSAR platform.
1. Structure Preprocessing:
data.csv) containing SMILES strings of your natural products and their analogues [48].2. Feature Calculation:
3. Data Preprocessing and Splitting:
y.csv) containing the true activity labels for each compound [48].File_X_train), a test set (File_X_test), and their corresponding activity labels.4. Model Building and Evaluation:
n_estimators=800, cv=10 for cross-validation) and initiate the grid search to find the best set of features and hyperparameters [48].This protocol is useful for understanding which structural features of your natural products are most critical for activity.
1. Univariate Feature Selection:
k=10) and select a score function (e.g., f_classif). Note that some score functions like chi2 require the data to contain only non-negative values [48].2. Tree-Based and Recursive Feature Elimination:
The following table details key computational "reagents" â the descriptors, fingerprints, and algorithms that are essential for constructingSAR models on the ChemSAR platform.
| Item Name & Function | Brief Explanation |
|---|---|
| Molecular DescriptorsQuantitative representations of molecular structure and properties. | ChemSAR can compute 783 1D/2D descriptors covering constitution, topology, charge, and molecular properties. These are the fundamental variables that the machine learning model uses to learn the relationship with biological activity [46]. |
| FingerprintsBinary vectors representing the presence or absence of specific substructures or paths in a molecule. | The platform calculates ten types of widely-used fingerprints. They are crucial for assessing molecular similarity and for models that rely on substructure patterns [46]. |
| Standardizer (e.g., ChemAxon Standardizer)Tool for molecular structure preprocessing. | Integrated into ChemSAR for tasks like salt removal, normalization, and tautomer standardization. This ensures all molecules are in a consistent representation before analysis [46]. |
| Scikit-learnA core machine learning library in Python. | ChemSAR integrates this library to provide algorithms for feature selection, model building (e.g., Random Forest, SVM), and cross-validation, making advanced ML accessible without programming [46]. |
| RDKit & ChemoPy PackagesOpen-source cheminformatics toolkits. | Used by ChemSAR for underlying molecular descriptor calculation and fingerprint generation [46]. |
The diagram below illustrates the complete workflow for using ChemSAR in natural product lead optimization, from data preparation to model deployment.
For researchers working to transform complex natural products into viable drug candidates, optimizing Absorption, Distribution, Metabolism, Excretion, and Toxicity (ADMET) properties presents unique challenges. Despite their therapeutic potential, natural products often face higher attrition rates due to poor pharmacokinetics and unforeseen toxicity issues, which account for approximately 70% of clinical failures [51]. This technical support center provides targeted troubleshooting guides and experimental protocols to help you identify and resolve common ADMET liabilities early in your optimization workflow, accelerating the development of safer, more effective therapeutics from natural product leads.
Natural products frequently possess complex molecular structures that differ significantly from synthetic compounds, leading to unpredictable pharmacokinetic behavior. Their structural complexity often results in:
Modern computational approaches have significantly improved prediction capabilities:
Solubility optimization requires a multi-faceted approach:
Addressing metabolic instability requires understanding degradation pathways:
Symptoms: Low dissolution rates, inconsistent exposure in animal models, formulation challenges.
Diagnostic Steps:
Resolution Strategies:
Prevention: Incorporate solubility prediction tools (e.g., ADMET predictor) during virtual screening and maintain LogD values between 1-3 when possible [55].
Symptoms: Steep exposure drop-off in PK studies, requirement for frequent dosing.
Diagnostic Steps:
Resolution Strategies:
Prevention: Incorporate human liver microsomal stability (HLM) and hepatocyte clearance assays early in lead optimization series [52].
Symptoms: In vitro safety flags, adverse findings in repeat-dose toxicology studies.
Diagnostic Steps:
Resolution Strategies:
Prevention: Implement routine safety pharmacology screening earlier in the discovery cascade and leverage predictive models like ADMET-AI [56] [54].
Symptoms: Significant discrepancies between predicted and observed human pharmacokinetics.
Diagnostic Steps:
Resolution Strategies:
Prevention: Use IVIVE approaches that account for free drug concentrations and incorporate transporter effects [53].
Purpose: Rapid screening of metabolic stability in liver microsomes to identify compounds with favorable clearance profiles.
Materials:
Procedure:
Data Interpretation: Compounds with human hepatic clearance >70% of liver blood flow are considered high-clearance; <30% are low-clearance [52].
Purpose: Evaluate cellular permeability and P-glycoprotein interaction potential.
Materials:
Procedure:
Data Interpretation: Efflux ratio >2 suggests potential P-gp substrate liability that may limit absorption or CNS penetration [52].
| Property | Primary Assay | Secondary Assay | Acceptance Criteria | Throughput |
|---|---|---|---|---|
| Solubility | Kinetic solubility (pH 7.4) | Thermodynamic solubility | >100 μM (early); >500 μM (candidate) | High |
| Permeability | PAMPA | MDR1-MDCKII | Papp >5 à 10â»â¶ cm/s (high) | Medium |
| Metabolic Stability | Liver microsomes (human/mouse) | Hepatocytes (suspended/plated) | CLhep <30% liver blood flow | Medium |
| CYP Inhibition | Fluorescent/LC-MS screening | IC50 determination | IC50 >10 μM (individual CYP) | Medium |
| hERG Inhibition | Patch-clamp | Binding assay | IC50 >30-fold over Cmax | Low |
| Plasma Protein Binding | Equilibrium dialysis | Ultracentrifugation | Fu >1% (preferred) | Medium |
| Tool/Platform | Prediction Capabilities | Strengths | Limitations |
|---|---|---|---|
| ADMET-AI [56] | hERG, CYP inhibition, permeability | Best-in-class results on Therapeutic Data Commons datasets | Requires SMILES or 3D structures as input |
| ADMET Predictor [55] | Metabolic stability, solubility, LogD | Useful for in silico screening in early drug discovery | Accuracy varies by chemical space |
| QSAR Models [51] | Toxicity, solubility, permeability | Interpretable features and relationships | Limited to chemical space of training data |
| Graph Neural Networks [53] | Multiple ADMET endpoints simultaneously | Learns directly from molecular structures | Black box nature; large training datasets needed |
ADMET Optimization Workflow
| Reagent/Assay | Vendor Examples | Primary Application | Key Considerations |
|---|---|---|---|
| Human Liver Microsomes | Corning, XenoTech, BioIVT | Metabolic stability, metabolite profiling | Consider donor pool size and demographics |
| Cryopreserved Hepatocytes | BioIVT, Lonza, CellzDirect | Intrinsic clearance, metabolite identification | Check viability and metabolic activity upon thawing |
| MDR1-MDCKII Cells | ATCC, academic sources | Permeability assessment, transporter studies | Monitor passage number and efflux ratio of controls |
| CYP450 Isozyme Kits | Promega, Thermo Fisher | Enzyme inhibition screening | Validate with known inhibitors for each CYP |
| hERG Expressing Cells | ChanTest, Eurofins | Cardiac safety assessment | Use reference compounds for assay validation |
| Equilibrium Dialysis Devices | HTDialysis, Thermo Fisher | Plasma protein binding measurement | Ensure equilibrium is reached for highly bound compounds |
| Accelerator Mass Spectrometry | Commercial providers | Human microdosing studies | Requires synthesis of radiolabeled compound (¹â´C) |
This technical support center provides practical solutions for researchers facing common supply chain and sustainability challenges in natural product (NP) sourcing for drug discovery.
1. How can we make our natural product supply chains more resilient to geopolitical and climate disruptions?
Modern supply chains must balance cost with resilience. The old model of single, globalized supply chains is being replaced by multiple regional sourcing networks and strategic redundancy [57].
2. What are the most critical raw material procurement challenges, and how can we address them?
Procurement of raw materials is the most affected part of the supply chain, with 94% of companies reporting disruptions [58]. Key challenges include rising costs, supplier reliability, and a lack of transparency [59].
1. Our consumers and stakeholders demand greater ethical sourcing. Where do we start?
Ethical sourcing is now a core business imperative, driven by consumer pressure, investor expectations, and emerging legislation [60] [61]. It spans social equity, ecological preservation, and geopolitical considerations [61].
2. How can we navigate the complex landscape of sustainability regulations?
Regulations like the EU's Deforestation Regulation and the Uyghur Forced Labor Prevention Act (UFLPA) in the U.S. are reshaping supply chain management, requiring greater transparency and corporate accountability [60].
1. How can we overcome low natural product titers or silent gene clusters in microbial fermentation?
A major challenge in NP research is activating biosynthetic gene clusters (BGCs) in native producers or heterologous hosts. This often involves manipulating complex regulatory networks [5].
The following workflow visualizes this protocol for activating and optimizing production:
2. What strategies exist for optimizing a natural product lead to improve its chemical accessibility?
Natural products often serve as leads rather than final drugs because they can be structurally complex and difficult to synthesize in large quantities [9]. Optimization is required to improve their chemical accessibility for further development.
The logical relationship between the optimization strategy and its purpose is outlined below:
Table 1: Survey data on supply chain disruption impacts across industry sectors (2025) [58].
| Area of Impact | Percentage of Respondents Affected | Key Sector-Specific Examples |
|---|---|---|
| Procurement of Raw Materials | 94% | Widespread across all sectors; lack of domestic availability for many key components. |
| Manufacturing & Production Capacity | 90% | Delayed projects and cost inflation are becoming commonplace. |
| Warehousing & Aftermarket Services | 76% | Impacts the entire logistics and service infrastructure. |
| Innovation & R&D | 80% (in advanced manufacturing) | Capital is being redirected from R&D and workforce development, threatening U.S. technological leadership. |
Table 2: Key megatrends shaping supply chains and strategic responses for research organizations [57].
| Megatrend | Impact on NP Research | Recommended Strategic Response |
|---|---|---|
| Rise of Economic Statecraft | Tariffs and trade policies increase the cost and complexity of global sourcing of raw materials and intermediates. | Diversify sourcing locations; leverage partnerships and joint ventures to pool investment risk. |
| Climate-Related Events as Strategic Risks | 8% of output from the world's top 50 manufacturing hubs is at risk. Consumer electronics and semiconductors are highly vulnerable. | Factor climate risk (e.g., extreme weather, sea-level rise) into site selection for sourcing and manufacturing partnerships. |
| Mounting Manufacturing Talent Bottlenecks | Shortages of blue-collar, white-collar, and digital talent hinder scale-up and operations in key regions. | Prioritize talent development and partner with institutions in regions with supportive immigration pathways for skilled workers. |
Table 3: Key tools and technologies for addressing supply chain and sourcing challenges in natural product research.
| Tool / Technology | Function | Application in NP Sourcing Research |
|---|---|---|
| Blockchain Platforms | Creates an immutable, transparent record of transactions and product journey. | Enables verification of ethical and sustainable sourcing claims for natural ingredients from origin to lab [61]. |
| Heterologous Expression Systems | Allows the transfer and expression of biosynthetic gene clusters in a tractable host. | Overcomes challenges of cultivating native producers or low titers; enables production of scarce NPs and their analogues [5]. |
| Supplier Management & Risk Software | Digital platforms for real-time monitoring of supplier performance and risk. | Provides transparency into supplier reliability, financial stability, and compliance status, mitigating procurement risks [59]. |
| AI-Powered Formulation Tools | Uses machine learning to analyze ingredient databases and predict formulation properties. | Can aid in the design of optimized NP formulations and in reverse-engineering (deformulation) for competitive analysis [61]. |
| Centralized Sustainability Data Platforms | Captures and standardizes ESG (Environmental, Social, Governance) data across complex global supply chains. | Helps researchers and companies conduct risk assessments, ensure regulatory compliance, and report on sustainability metrics [60]. |
1. My final compound yield is very low after isolation. What could be the cause? Low yields can stem from several issues in the extraction and purification workflow. Inefficient initial extraction from the raw material is a common culprit, where the solvent may not be effectively penetrating the solid matrix to dissolve the target solute [62]. During crystallization, using an excessive amount of solvent can lead to significant compound loss in the mother liquor, resulting in a poor final yield [63]. Furthermore, instability of the target compound can lead to its degradation during the process, especially if it is exposed to unfavorable conditions like high temperatures or extreme pH for prolonged periods [64] [65].
2. I suspect my target compound is degrading during the isolation process. How can I prevent this? Compound degradation is a major challenge, particularly for sensitive molecules. To mitigate this:
4. My compound will not crystallize. What can I do to induce crystallization? When a dissolved solution fails to crystallize, a hierarchical approach can be used [63]:
The table below summarizes frequent challenges, their potential causes, and recommended solutions.
| Problem | Potential Causes | Recommended Solutions |
|---|---|---|
| Low Yield [62] [63] | Inefficient extraction; excessive solvent in crystallization; compound degradation. | Optimize solvent polarity and extraction time; reduce solvent volume for crystallization; use low-temperature protocols [64] [65]. |
| Compound Degradation [64] [65] | Exposure to high temperature, improper pH, or prolonged processing times. | Lower process temperatures; use stability-indicating solvents/buffers; employ faster, modern extraction techniques (e.g., UAE, MAE). |
| Poor Separation Resolution [65] [66] | Inefficient chromatographic method; co-elution of impurities. | Re-develop method to maximize resolution of target peak; use column-switching or peak-trapping techniques. |
| Slow or No Crystallization [63] | Solution is supersaturated; lack of nucleation sites. | Scratch flask with glass rod; add a seed crystal; reduce solvent volume; change solvent system. |
| Excessive Peak Tailing/Broadening [65] | Sub-optimal chromatographic conditions (e.g., low temperature). | Adjust mobile phase pH/buffer; increase column temperature if compound stability allows. (Note: Sometimes broader peaks are accepted to prevent degradation at low temperatures). |
Protocol 1: Low-Temperature Chromatographic Isolation for Unstable Compounds This protocol is designed to isolate degradation-prone compounds by maintaining a cold chain throughout the process [65].
Protocol 2: Ultrasound-Assisted Extraction (UAE) for Heat-Sensitive Bioactives This protocol uses acoustic cavitation to efficiently extract compounds while minimizing thermal degradation [64] [62].
The following table lists essential materials and their functions for setting up isolation experiments.
| Research Reagent | Function / Application |
|---|---|
| Poly-Lysine Magnetic Beads [67] | Affinity-based purification of ribosomes and other RNA-protein complexes by binding to the negatively charged RNA backbone. |
| Trifluoroacetic Acid (TFA) [65] | A volatile ion-pairing agent used in reversed-phase HPLC mobile phases to improve peak shape for acidic and basic analytes. |
| Volatile Buffers (e.g., Ammonium Bicarbonate) [65] [66] | Buffers that can be easily removed by evaporation, facilitating the isolation of pure compounds after preparative HPLC. |
| Poly-D/L-Glutamic Acid [67] | Used as an elution agent to displace bound RNA or ribosomes from poly-lysine beads via competitive binding. |
| Enzyme Cocktails (e.g., Cellulase, Pectinase) [64] | Used in Enzyme-Assisted Extraction (EAE) to selectively break down plant cell walls and release intracellular compounds. |
The diagram below outlines a logical workflow for diagnosing and addressing common isolation challenges.
Troubleshooting Workflow for Pure Compound Isolation
The table below compares conventional and advanced extraction methods, highlighting how technique selection directly impacts the success of downstream isolation [64] [62].
| Extraction Technique | Key Principle | Advantages | Best for Compound Types |
|---|---|---|---|
| Maceration [62] | Soaking plant material in solvent at room temperature. | Simple, low equipment cost. | Stable, non-thermolabile compounds. |
| Soxhlet Extraction [64] | Continuous washing with hot solvent. | High throughput. | Non-polar, thermally stable compounds. |
| Ultrasound-Assisted (UAE) [64] | Uses acoustic cavitation to disrupt cells. | Higher yield, faster, lower temperature. | Heat-sensitive polyphenols, flavonoids. |
| Microwave-Assisted (MAE) [64] [62] | Uses microwave energy to heat solvent and cells. | Rapid, reduced solvent consumption. | A wide range of phytochemicals. |
| Enzyme-Assisted (EAE) [64] | Uses enzymes to break down cell walls. | High selectivity, mild conditions. | Glycosides, polysaccharides. |
Q1: What is a Synthetic Accessibility Score (SAscore)? A Synthetic Accessibility Score (SAscore) is a computational metric used to estimate how easy or difficult it is to synthesize a given molecule. It typically provides a numerical value where a lower score (e.g., closer to 1) indicates a molecule is easy to make, and a higher score (e.g., closer to 10) suggests significant synthetic challenges [68]. These scores help researchers triage and prioritize compounds in drug discovery projects.
Q2: Why is SAscore important in natural product lead optimization? Natural products often have complex structures that can be difficult and resource-intensive to synthesize. Using SAscores allows researchers to early on:
Q3: What are the main computational approaches for estimating synthetic accessibility? There are two primary computational approaches, each with different methodologies and resource requirements [69]:
| Approach Type | Description | Key Characteristics |
|---|---|---|
| Complexity-Based | Uses rules and fragment libraries to assess molecular complexity [68]. | Fast, suitable for high-throughput screening; relies on historical synthetic knowledge. |
| Retrosynthetic-Based | Uses AI and reaction databases to plan a complete synthetic route [70] [69]. | Resource-intensive, more realistic; provides a detailed route and step count. |
Q4: My generator produces molecules with good predicted activity but poor SAscores. How can I fix this? This is a common challenge. The solution is to integrate the SAscore directly into the molecular generation process itself, not just use it for post-generation filtering. You can:
Scenario 1: Inconsistent Scores Between Different SAscore Tools
| Symptom | Potential Cause | Solution |
|---|---|---|
| A molecule gets a "easy" score from one tool but a "hard" score from another. | Different tools use different underlying algorithms (fragment-based vs. retrosynthesis-based) and training data. | Standardize your toolset. Understand the basis of each score. Use a retrosynthesis-based score (e.g., RScore) for a more realistic assessment of synthetic steps, especially for novel or complex natural product-like structures [69]. |
Scenario 2: High SAscore on a Seemingly Simple Molecule
| Symptom | Potential Cause | Solution |
|---|---|---|
| A molecule without obvious complexity (e.g., large rings, many stereocenters) receives a high SAscore. | The molecule may contain rare or non-standard fragments that are underrepresented in historical synthetic data. It might also lack commercially available starting materials. | Perform a full retrosynthetic analysis using a tool like Spaya-API [69]. This can confirm if the high score is due to a lack of known synthetic pathways or available building blocks. |
Scenario 3: Handling Invalid Molecules in a Batch SAscore Request
| Symptom | Potential Cause | Solution |
|---|---|---|
| When submitting a batch of molecules, some return a null score or an error. | Input molecules may be hypervalent, have incomplete rings, or improper protonation [70]. | Pre-process and curate your chemical structures. Use a toolkit to standardize SMILES strings and validate structures before submitting them for SAscore calculation [70]. |
The table below summarizes several established SAscore tools and their characteristics to help you select the right one for your project [68] [70] [69].
| Score Name | Underlying Methodology | Score Range | Interpretation | Best Use Case |
|---|---|---|---|---|
| SAscore | Fragment contributions & molecular complexity penalty [68]. | 1 (easy) to 10 (hard) | Lower score = easier to synthesize. | High-throughput initial triage of large compound libraries (e.g., from virtual screening). |
| SYNTHIA SAS | Graph convolutional neural network (GCNN) trained on retrosynthetic data [70]. | 0 (easy) to 10 (hard) | Approximates the number of synthetic steps. Lower score = fewer steps. | Prioritizing leads with a more realistic step-count estimate. |
| RScore | Full retrosynthetic analysis via Spaya-API [69]. | 0.0 (no route) to 1.0 (one-step synthesis) | Higher score = more feasible route found. | In-depth analysis of final candidate molecules to assess synthetic viability. |
| SC Score | Neural network trained on reaction data [69]. | 1 to 5 | Lower score = less complex, more feasible. | Ranking molecules based on comparative complexity derived from reactions. |
This protocol outlines how to computationally assess and interpret the synthetic accessibility of a natural product lead or a derivative.
1. Objective To determine the synthetic accessibility of a natural product lead using multiple SAscore metrics and perform a basic retrosynthetic analysis to contextualize the score.
2. Research Reagent Solutions (Computational Tools)
| Item | Function |
|---|---|
| PubChem Database | Provides a vast repository of known chemical structures used to train fragment-based SAScores and establish historical synthetic knowledge [68]. |
| Spaya-API | A retrosynthesis software API used to perform a data-driven synthetic route planning and obtain the RScore [69]. |
| Commercial Compound Catalogs | Integrated into tools like Spaya, these databases of readily available starting materials are crucial for determining if a realistic synthesis can be launched [69]. |
3. Methodology
4. Workflow Diagram The diagram below illustrates the logical workflow for triaging molecules based on their Synthetic Accessibility Score.
For generative molecular design, using a fast, predictive model of synthetic accessibility is crucial. The diagram below outlines a pipeline where a predicted SAscore directly influences the generator to produce more synthesizable molecules [69].
FAQ 1: How can I systematically improve the ADMET profile of a complex natural product lead without compromising its potent bioactivity?
Answer: Employ a sequential, multi-task learning approach that explicitly models the pharmacokinetic (PK) hierarchy. Traditional methods treat Absorption, Distribution, Metabolism, and Excretion (ADME) as independent properties, leading to suboptimal predictions. The ADME-DL pipeline enhances molecular foundation models by pretraining them on 21 ADME endpoints in a sequential AâDâMâE order, which aligns with the established flow of a drug through the body [71]. This method encodes crucial PK information into the molecular embedding, allowing for a more accurate prediction of how structural changes will affect the overall drug-likeness and ADMET profile before synthesis. The resulting ADME-informed embeddings can then be used to classify molecules as drug-like or non-drug-like, significantly improving early-stage filtering [71].
FAQ 2: What computational strategies can I use to evaluate and plan the synthesis of a complex natural product-derived candidate?
Answer: A dual-path strategy is recommended for robust assessment.
FAQ 3: My natural product lead violates the Rule of 5 but appears to have good oral bioavailability. Should I reject it?
Answer: Not necessarily. Many natural product-based drugs successfully occupy 'beyond-rule-of-5' (bRo5) chemical space. Natural products often have higher molecular weight, more stereocenters, and greater structural complexity, which can be correlated with improved binding specificity and lower preclinical toxicity [75] [2]. Rather than relying solely on rigid rules, use property-based filters like Veber's rules (rotatable bonds ⤠10, TPSA ⤠140 à ²) or Egan's filter (TPSA ⤠131.6 à ², logP ⤠5.88) as additional benchmarks for oral bioavailability [76]. The key is to use these rules as guidelines, not absolute filters, and prioritize experimental data on permeability and bioavailability when available.
FAQ 4: How can I distinguish truly promising hits from compounds that are pan-assay interference compounds (PAINS)?
Answer: Always screen your virtual or physical library against a curated list of PAINS substructures. These are functional groups, such as rhodanines and certain quinones, known to cause false-positive results in high-throughput screens by engaging in non-specific interactions with biological targets [76]. Furthermore, apply filters for aggregators, which can be identified by a combination of high lipophilicity (e.g., SlogP < 3) and structural similarity to known aggregator databases [76]. Proactively filtering these compounds saves significant time and resources.
Problem: Your lead compound shows high potency in vitro but is rapidly metabolized (e.g., by CYP450 enzymes), leading to a short half-life.
Diagnosis and Solution Strategies:
| Step | Action | Protocol / Rationale |
|---|---|---|
| 1. Identify | Determine the site of metabolism. | Use in silico metabolism prediction tools (e.g., ADMETlab, admetSAR) to identify labile sites like aromatic hydroxylation or N-dealkylation [77] [74]. Validate these predictions with in vitro microsomal stability assays. |
| 2. Design | Implement strategic structural modifications. | Blocking: Introduce strategically placed deuterium (deuteriation) or fluorine atoms at the metabolically soft spot [9]. Bioisosteric Replacement: Replace a metabolically vulnerable group (e.g., methyl) with a bioisostere (e.g., cyclopropyl) [9]. |
| 3. Validate | Re-assess the optimized compound. | Use the sequential ADME MTL framework (ADME-DL) to predict the impact of your changes on the overall ADME profile, not just metabolism in isolation [71]. Follow up with experimental validation. |
Problem: Your candidate has excellent target binding but poor aqueous solubility or cell membrane permeability, limiting its efficacy.
Diagnosis and Solution Strategies:
| Step | Action | Protocol / Rationale |
|---|---|---|
| 1. Profile | Calculate key physicochemical properties. | Use RDKit or similar software to compute descriptors: Topological Polar Surface Area (TPSA), LogP, and the number of H-bond donors/acceptors [74] [76]. High TPSA (>140 à ²) and high rotatable bond count often correlate with poor permeability [76]. |
| 2. Optimize | Modify the structure to improve properties. | Increase Solubility: Introduce ionizable groups (e.g., amines) or reduce overall lipophilicity (clogP). Improve Permeability: Mask H-bond donors/acceptors through prodrug strategies or reduce molecular rigidity to fall within Veber's filter guidelines [9] [76]. |
| 3. Leverage NPs | Learn from natural products. | NPs often achieve good permeability despite high MW by having a high fraction of sp³-hybridized carbons (Fsp³), which confers 3D structure and reduces flatness. Consider increasing the Fsp³ of your lead [75] [2]. |
Use this data to benchmark your candidates against successful drugs. [75]
| Property | Natural Product (N) Drugs | Natural Product-Derived (ND) Drugs | Top-Selling Synthetic (2018-S) Drugs |
|---|---|---|---|
| Molecular Weight (MW) | 611 | 757 | 444 |
| Hydrogen Bond Donors (HBD) | 5.9 | 7.0 | 1.9 |
| Hydrogen Bond Acceptors (HBA) | 10.1 | 11.5 | 5.1 |
| Calculated LogP (ALOGPs) | 1.96 | 1.82 | 2.83 |
| Rotatable Bonds (Rot) | 11.0 | 16.2 | 6.5 |
| Topological Polar Surface Area (tPSA) | 196 | 250 | 95 |
| Fraction sp³ Carbons (Fsp³) | 0.71 | 0.59 | 0.33 |
| Aromatic Rings (RngAr) | 0.7 | 1.4 | 2.7 |
Based on analysis of approved anticancer drugs from 1981-2010. [9]
| Optimization Purpose | Key Strategies | Example Tactics |
|---|---|---|
| Enhance Drug Efficacy | Structure-Activity Relationship (SAR)-driven design; Direct functional group manipulation. | Systematic analogue synthesis; Bioisosteric replacement; Structure-based design if target is known. |
| Improve ADMET Profile | Structural modification to alter physicochemical properties. | Reduce logP for lower toxicity; Block metabolic soft spots; Introduce solubilizing groups. |
| Increase Chemical Accessibility | pharmacophore-oriented design; Simplification of core structure. | Identify & retain key pharmacophore; Synthesize simpler, more accessible analogs with core activity (Scaffold hopping). |
This protocol details the use of the ADME-DL pipeline for a more pharmacologically relevant assessment of drug-likeness [71].
Methodology:
z).z) from the pretrained model to train a simple classifier (e.g., a Multi-Layer Perceptron - MLP).
| Tool Name | Type | Primary Function | Relevance to Natural Product Optimization |
|---|---|---|---|
| ADME-DL [71] | AI Pipeline | Drug-likeness prediction via sequential ADME modeling. | Provides PK-aware evaluation of complex NPs, overcoming limitations of structure-only filters. |
| druglikeFilter [74] | Multi-dimensional Filter | Collective evaluation of physicochemical rules, toxicity, binding affinity, and synthesizability. | Offers a one-stop platform for comprehensive assessment, integrating retrosynthetic analysis (Retroâ) for complex molecules. |
| RDKit [77] [74] | Cheminformatics Library | Calculates molecular descriptors, fingerprints, and SAscore. | The foundational library for generating property profiles and rapid synthetic accessibility estimates. |
| SYLVIA [72] | Synthetic Accessibility Software | Predicts synthetic feasibility based on structural complexity and starting material information. | Useful for benchmarking the synthetic complexity of natural product scaffolds and their analogs. |
| ADMETlab / admetSAR [74] | Web Server / Database | Predicts ADMET-related parameters. | Used for initial profiling and troubleshooting of specific ADMET issues like metabolic stability or hERG inhibition. |
| Therapeutic Data Commons (TDC) [71] | Data Resource | Provides curated datasets for ADME endpoints. | Supplies the essential training and benchmarking data for building robust ADME prediction models. |
Q1: Why is validating optimized natural product leads particularly challenging? Natural products often possess complex chemical structures with multiple chiral centers and high molecular weight, which can lead to poor solubility, synthetic intractability, and unfavorable pharmacokinetic profiles [78] [9]. Validation must therefore address not just biological activity but also drug-like properties and chemical accessibility to ensure the simplified lead remains a viable drug candidate [9].
Q2: What is the primary goal of lead optimization in this context? The optimization aims to improve the chemical accessibility of complex natural leads through structural simplification while maintaining or improving their favorable biological activity [78]. This often involves reducing molecular complexity, such as the number of rings and chiral centers, to create more synthetically feasible drug-like molecules [78] [9].
Q3: How does the Design-Make-Test-Analyze (DMTA) cycle apply to lead optimization? The DMTA cycle is a fundamental, iterative strategy in lead optimization [79]. Researchers design new compound structures based on existing data, synthesize these compounds (make), evaluate their biological activity and properties (test), then analyze the results to inform the next design cycle. This process enables systematic improvement of lead compounds [79].
Q4: What key properties should be monitored during the validation process? Beyond potency, critical properties include selectivity, solubility, metabolic stability, permeability, and early toxicity indicators [9] [80]. Absorption, distribution, metabolism, excretion, and toxicity (ADMET) profiles are crucial for determining clinical translatability [9] [80].
Table: Common In Vitro Assay Issues and Solutions
| Problem | Possible Causes | Recommended Solutions |
|---|---|---|
| High variability in potency measurements | Low compound solubility, compound adhesion to plates, chemical instability | Use appropriate co-solvents (DMSO), include controls for non-specific binding, verify compound stability under assay conditions [9] |
| Poor correlation between binding and cellular activity | Poor cellular permeability, efflux by transporters, intracellular metabolism | Assess permeability in Caco-2 or PAMPA assays, check for P-glycoprotein substrate potential, measure intracellular concentration [80] |
| Cytotoxicity in absence of target engagement | Off-target effects, non-specific toxicity, reactive metabolites | Perform counter-screening against known toxicity targets, conduct reactive metabolite assays, check for chemical structural alerts [80] |
| Irreproducible results between assay runs | Compound precipitation, inconsistent cell passage number, assay protocol deviations | Standardize cell culture conditions, use fresh compound solutions, implement strict SOPs and quality controls [79] |
Table: Addressing Challenges in Natural Lead Simplification
| Challenge | Impact on Lead Validation | Mitigation Strategies |
|---|---|---|
| Loss of potency after simplification | Reduced target engagement and efficacy | Employ pharmacophore-based design to retain key interacting moieties; use structure-based simplification if target structure is available [78] |
| Unfavorable shift in ADMET profile | Poor pharmacokinetics or increased toxicity | Monitor key properties early (e.g., microsomal stability, CYP inhibition); use bioisosteric replacements to improve properties [9] |
| Introduction of structural instability | Compound degradation invalidates results | Assess chemical stability at various pH levels; identify and modify labile functional groups [9] |
| Increased chiral centers or synthetic complexity | Hindered chemical accessibility for scaling | Prioritize synthetic tractability in design; reduce chiral centers and complex ring systems where possible [78] |
Purpose: To systematically determine how structural modifications affect biological activity and selectivity during natural lead simplification.
Procedure:
Troubleshooting Tips:
Purpose: To identify potential pharmacokinetic and toxicity issues before advancing simplified leads.
Procedure:
Metabolic Stability:
Cellular Permeability:
Cytotoxicity Screening:
Key Interpretation Guidelines:
Diagram: Workflow for Validating Optimized Natural Product Leads
Diagram: In Silico to In Vitro Validation Bridge
Table: Essential Reagents and Tools for Lead Validation Studies
| Reagent/Resource | Primary Function | Application Notes |
|---|---|---|
| Liver Microsomes (Human) | Metabolic stability assessment | Lot-to-lat variability; use pooled donors for representative data; include positive controls [9] |
| Caco-2 Cell Line | Intestinal permeability prediction | Requires 21-day differentiation; standardized conditions critical for reproducibility [80] |
| Phospholipid Vesicles | Membrane binding studies | Relevant for natural products with lipophilic characteristics; can impact free concentration [9] |
| CYP450 Inhibition Kits | Drug-drug interaction potential | Screen against major CYP enzymes (3A4, 2D6, 2C9); key for safety assessment [80] |
| Plasma Protein Binding Assays | Free fraction determination | Use human plasma for most relevant data; equilibrium dialysis preferred method [80] |
| Structural Simplification Guides | Molecular design | Apply strategies like ring reduction, chiral center elimination, and functional group bioisosteres [78] |
The discovery of Bromodomain and Extra-Terminal (BET) family inhibitors, particularly those targeting BRD4, represents a promising frontier in epigenetic cancer therapy. While potent synthetic inhibitors like JQ1 have demonstrated significant anti-tumor efficacy in preclinical models, their clinical application has been hampered by challenges including poor pharmacokinetic profiles, low oral bioavailability, and dose-limiting toxicities [82]. Within the context of improving chemical accessibility of natural product leads research, this case study examines the systematic optimization of a hypothetical natural product template into a viable BRD4-targeting therapeutic agent. Natural products frequently offer privileged structural scaffolds with inherent bioactivity but often require substantial medicinal chemistry optimization to enhance their drug-like properties, target selectivity, and metabolic stability. This technical support document provides researchers with practical methodologies and troubleshooting guidance for navigating this complex optimization pathway, from initial virtual screening through experimental validation.
Q: Our virtual screening campaigns yield compounds with excellent predicted binding affinity but poor experimental activity. What could explain this discrepancy?
A: Several factors could contribute to this common issue:
Experimental Protocol: Pharmacophore-Based Virtual Screening
Q: How can we improve selectivity for specific BRD4 bromodomains (BD1 vs. BD2) to minimize off-target effects?
A: Achieving BD1/BD2 selectivity requires targeting non-conserved residues. Follow this structured approach:
Table: Key Residues for Selective BRD4 Inhibitor Design
| Residue Position | BRD4-BD1 | BRD4-BD2 | Selectivity Consideration |
|---|---|---|---|
| Residue 146/439 | Ile146 | Val439 | Smaller Val439 in BD2 allows bulkier substituents for BD2 selectivity. |
| Residue 81/374 | Trp81 | Trp374 | Highly conserved; key for acetyl-lysine mimic anchoring via water-mediated H-bonds. |
| Residue 83/375 | Phe83 | Phe375 | Conserved hydrophobic contact. |
| Residue 140/433 | Asn140 | Asn433 | Forms critical H-bond with inhibitor carbonyl group in both domains. |
Q: Our lead compound shows strong BRD4 inhibition in enzymatic assays but poor cellular potency. What strategies can improve cell permeability?
A: Poor cellular activity often stems from insufficient intracellular concentration. Consider these modifications:
Q: How can we develop dual-target inhibitors to enhance efficacy and overcome resistance?
A: Dual-targeting strategies can address pathway redundancy. For BRD4/STAT3 inhibition, follow this protocol:
Q: Our inhibitor effectively reduces cancer cell proliferation but induces senescence rather than cell death. How can we address this therapeutically?
A: Therapy-induced senescence can lead to tumor dormancy and relapse. Implement a combination strategy with senolytic agents:
Q: Our BRD4 inhibitor shows limited efficacy in solid tumor models. What combination strategies could be explored?
A: Limited single-agent efficacy in solid tumors is a known challenge. Consider these rational combinations:
Experimental Protocol: Evaluating Senescence Induction and Senolytic Combination
Table: Essential Reagents for BRD4 Inhibitor Discovery and Validation
| Reagent / Tool | Function/Application | Example/Specification |
|---|---|---|
| JQ1 | Pan-BET family inhibitor; positive control and tool compound | Useful for benchmarking new inhibitors in binding and functional assays. |
| OTX015 | Clinical-stage BET inhibitor; reference compound | For comparative in vitro and in vivo efficacy studies. |
| Recombinant BRD4-BD1/BD2 Proteins | In vitro binding and inhibition assays (TR-FRET, FP) | Ensure >95% purity; use for initial enzymatic activity screening. |
| Cell Lines | Cellular potency and mechanism studies | KYSE450 (esophageal), HepG2 (liver), CAKI-2 (renal), MOLM-13 (AML). |
| ABT737 | Senolytic agent for combination studies | BCL-2 inhibitor; use to clear senescent cells induced by BRD4 inhibition. |
| Erastin / RSL3 | Ferroptosis inducers for combination studies | Use to exploit BRD4i-induced ferroptosis sensitivity [86]. |
| Antibody Panel | Mechanistic validation via Western Blot | Anti-BRD4, anti-c-MYC, anti-p27, anti-Ki67, anti-TXNIP, anti-p-STAT3. |
| Crystal Structure (PDB: 4BJX) | Structure-based drug design | High-resolution (1.59 Ã ) structure for docking and pharmacophore modeling [83]. |
Q1: What are the primary chemical accessibility challenges associated with Natural Product (NP)-derived leads? NP-derived leads often face significant chemical accessibility challenges, including:
Q2: How do the molecular properties of NP-derived leads typically compare to those of purely synthetic compounds? NP-derived leads differ from purely synthetic compounds in several key aspects, which contribute to their high success rate as drugs despite their complexity [2]. The table below summarizes these comparative properties.
| Molecular Property | Natural Product-Derived Leads | Purely Synthetic Counterparts |
|---|---|---|
| Structural Complexity | High; more stereocenters, macrocyclic structures [2] | Typically lower and less structurally diverse [2] |
| Lipophilicity (cLogP) | Generally lower, leading to better solubility profiles [2] | Often higher [2] |
| sp3 Carbon Fraction | Higher, indicating more complex, 3D structures [2] | Lower, indicating flatter, more 2D structures [2] |
| Chemical Starting Point | Evolutionarily pre-validated bioactivity [2] [87] | Designed for specific target binding [10] |
| Synthetic Accessibility | Often low; complex total synthesis [2] [78] | Generally high; designed for efficient synthesis [10] |
Q3: What experimental strategies can improve the chemical accessibility and "drug-likeness" of a complex NP lead? A primary strategy is Structural Simplification, which aims to retain the core pharmacophore while removing unnecessary complexity [78]. Key approaches include:
Q4: What role do modern technologies play in overcoming NP accessibility hurdles? Advanced technologies are revolutionizing NP-based drug discovery:
Problem: Your NP-derived lead compound shows promising but weak activity, or it interacts with off-targets.
Solution: Implement a focused Structure-Activity Relationship (SAR) study.
| Step | Protocol Description | Key Reagents & Tools |
|---|---|---|
| 1. Analog Design | Design a library of analogues by systematically modifying different regions of the lead molecule. Focus on regions predicted to influence binding. | Cheminformatics Software (e.g., Schrodinger Suite, MOE) to model interactions. |
| 2. Synthesis & Purification | Synthesize the designed analogues. Use parallel synthesis techniques to increase efficiency. | Building Blocks (e.g., amino acids for peptides, heterocycles); Purification Systems (e.g., HPLC, flash chromatography). |
| 3. In Vitro Bioassay | Test the synthesized analogues in a target-specific bioassay (e.g., enzyme inhibition, cell-based phenotypic assay). | Target Protein/Cell Line; Assay Kits (e.g., fluorescence-based, ELISA); High-Throughput Screening (HTS) Systems. |
| 4. Data Analysis | Analyze the bioassay results to establish SAR trends. Identify which structural modifications enhance potency and selectivity. | Data Analysis Software (e.g., GraphPad Prism, StarDrop) for IC50/EC50 calculation and trend analysis. |
Problem: Your NP lead has good on-target potency but suffers from poor metabolic stability, low solubility, or high clearance.
Solution: Reshape the lead optimization cascade to focus on Absorption, Distribution, Metabolism, and Excretion (ADME) properties early in the process [88].
| Step | Protocol Description | Key Reagents & Tools |
|---|---|---|
| 1. In Vitro ADME Screening | Profile the lead and its analogues in a suite of in vitro assays. Key assays include: metabolic stability in liver microsomes, plasma stability, Caco-2 permeability, and solubility measurements. | Liver Microsomes (human/mouse); Caco-2 Cell Line; Plasma; LC-MS/MS for analyte quantification. |
| 2. Identify Metabolic Soft Spots | Use microsomal incubations and LC-HRMS to identify major metabolites and sites of rapid metabolism. | Human Liver Microsomes (HLM); High-Resolution Mass Spectrometer (HRMS). |
| 3. Medicinal Chemistry Intervention | Chemically modify the identified metabolic soft spots. Strategies include: blocking metabolically labile sites, introducing deuterium, or reducing lipophilicity. | Medicinal Chemistry Tools (e.g., peptide truncation, peptidomimetics, N-/C-terminal capping [88]). |
| 4. In Vivo PK Profiling | Administer the top 1-2 optimized leads to animal models (e.g., mice) to determine key in vivo parameters like half-life and bioavailability. | Animal Models (e.g., Sprague-Dawley rats); LC-MS/MS for bioanalysis. |
| Item | Function in NP Lead Research |
|---|---|
| Human Liver Microsomes (HLMs) | A critical reagent for in vitro assessment of metabolic stability and identification of metabolic soft spots in NP leads [78]. |
| Surface Plasmon Resonance (SPR) Chip | Used in biophysical assays (e.g., with a CM5 chip) to provide direct, label-free data on target engagement, binding affinity (KD), and binding kinetics (kon/koff) [88]. |
| Caco-2 Cell Line | A model of the human intestinal epithelium used to predict the oral absorption and permeability of NP-derived compounds [10]. |
| LC-MS/MS-SPE-NMR Platform | A hyphenated analytical system that combines separation, quantification, and structural elucidation to rapidly identify novel NPs from complex extracts, accelerating dereplication [11]. |
| Biosynthetic Gene Cluster (BGC) Prediction Tools | Bioinformatics platforms (e.g., AntiSMASH, DeepBGC) used to mine microbial genomes and identify clusters of genes responsible for producing specific NPs, enabling heterologous expression [2]. |
| Peptidomimetic Building Blocks | Synthetic chemical fragments used to replace peptide bonds in NP-derived peptides, improving metabolic stability and membrane permeability while maintaining biological activity [88]. |
This technical support center provides targeted guidance for researchers working to improve the chemical accessibility of natural product (NP) leads. Natural products are renowned for their potent biological activity and structural complexity, but this very complexity often renders them synthetically intractable, creating a significant bottleneck in drug discovery pipelines [89]. This resource addresses the core challenges of evaluating and optimizing NP-inspired compounds, focusing on the critical triad of potency, selectivity, and synthetic tractability. The following FAQs, troubleshooting guides, and standardized protocols are designed to help you navigate these challenges efficiently.
1. Why should we invest in fully synthesizing natural product analogs when semisynthesis is often faster? While semisynthetic modification is a major source of FDA-approved NP-derived drugs, fully synthetic approaches offer significant advantages [89]. De novo synthesis allows for more profound structural alterations through strategies like scaffold hopping, enabling you to discover novel chemotypes that maintain beneficial biological activity while improving synthetic accessibility and creating new intellectual property space [89] [90].
2. What does "synthetic accessibility" really mean, and how is it quantified? Synthetic Accessibility (SA) is a practical metric of how easy or difficult it is to synthesize a given molecule in the lab [91]. It is not a simple binary but a continuum. A commonly used scoring method is the Ertl & Schuffenhauer score, which assigns a value from 1 (very easy) to 10 (very difficult) based on:
sascorer.py and commercial platforms, can provide these scores to help prioritize compounds [91].3. Our high-throughput screening identified a potent natural product hit, but it has poor solubility. What strategies can we use? Poor aqueous solubility is a known issue with some complex, lipophilic natural products [92]. Several lead optimization techniques can address this:
Observed Issue: A natural product lead shows excellent on-target potency but also high cytotoxicity in mammalian cell assays, suggesting potential off-target effects or general toxicity.
Investigation & Resolution:
| Step | Action | Rationale & Details |
|---|---|---|
| 1 | Confirm Selectivity | Profile the lead against a panel of related and unrelated targets (e.g., kinase panels, GPCR panels) to assess selectivity. A promiscuous binding profile often underlies general cytotoxicity [90]. |
| 2 | Check for PAINS | Analyze the structure for Pan-Assay Interference Compounds (PAINS) motifs. These substructures can cause false positives or non-specific activity, leading to misleading toxicity readouts [92]. |
| 3 | Evaluate Physicochemical Properties | Calculate key properties like cLogP. Very high lipophilicity can lead to non-specific membrane disruption. Aim to lower cLogP through synthetic modification to reduce non-mechanistic toxicity [92]. |
| 4 | Scaffold Hop | If the above steps confirm non-selectivity, use scaffold hopping strategies. Identify the key pharmacophore and graft it onto a new, synthetically tractable core structure to retain potency while eliminating the toxicophore [90]. |
Observed Issue: Computational design or screening identifies a complex NP scaffold with ideal binding characteristics, but retrosynthetic analysis suggests the synthesis would be too long, low-yielding, or not scalable.
Investigation & Resolution:
| Step | Action | Rationale & Details |
|---|---|---|
| 1 | Obtain a SA Score | Use computational tools (e.g., RDKit, eTox) to calculate a Synthetic Accessibility score. This provides a quantitative baseline and helps identify the most problematic structural features [91]. |
| 2 | Simplify the Scaffold | Employ strategies from Function-Oriented Synthesis (FOS). Systematically reduce intrinsic complexity while aiming to retain the core biological function. This may involve simplifying ring systems or reducing stereocenters [89]. |
| 3 | Leverage a DOS Library | Screen a Diversity-Oriented Synthesis (DOS) library. These libraries are populated with compounds containing NP-like features (e.g., high sp3 content, stereogenicity) but are designed for synthetic feasibility, potentially providing a new, tractable lead [92]. |
| 4 | Plan a Modular Synthesis | If the full structure is essential, devise a synthesis using convergent coupling strategies. For example, the Myers group's synthesis of tetracycline analogs involved coupling separate D- and AB-ring precursors, enabling more efficient exploration of structure-activity relationships [89]. |
This workflow ensures consistent evaluation of NP analogs against the key metrics of potency, selectivity, and synthesizability.
Use this table to standardize the reporting and comparison of data for NP leads and their analogs. This ensures objective decision-making during lead optimization.
Table 1: Key Quantitative Metrics for Natural Product Lead Evaluation
| Metric Category | Specific Parameter | Target Range for Progression | Experimental Method |
|---|---|---|---|
| Synthetic Tractability | Synthetic Accessibility (SA) Score | ⤠6 (on a 1-10 scale) [91] | Computational Calculation (e.g., RDKit, eTox) [91] |
| Potency | IC50 / EC50 | < 100 nM (target-dependent) | In vitro biochemical or cell-based assay [90] |
| Selectivity | Selectivity Index (e.g., IC50 off-target / IC50 on-target) | > 100-fold [90] | Panel-based screening against related targets [90] |
| Drug-like Properties | cLogP | < 5 [92] | Computational Prediction |
| Polar Surface Area (TPSA) | 60-140 à ² [92] | Computational Prediction | |
| Solubility (PBS, pH 7.4) | > 50 µg/mL | Kinetic solubility assay (e.g., nephelometry) | |
| In vitro ADMET | Microsomal Stability (% remaining) | > 30% (human/rat liver microsomes) | In vitro metabolic stability assay [90] |
Table 2: Essential Tools and Resources for NP Lead Research
| Tool / Resource | Function & Utility | Example Application |
|---|---|---|
| DOS Libraries | Pre-made collections of compounds designed with NP-like complexity (high Fsp3, stereocenters) but for synthetic feasibility [92]. | Screening for novel, tractable chemical starting points when a NP lead is too complex to synthesize [92]. |
| Fragment Libraries | Collections of small, low molecular weight compounds (<300 Da) for Fragment-Based Drug Discovery [90]. | Identifying minimal binding motifs of a complex NP to guide the design of simplified, potent analogs [90]. |
| Retrosynthetic Software | AI-powered tools that propose plausible synthetic routes for a target molecule in seconds [93]. | Rapidly assessing the feasibility of synthesizing a computationally designed NP analog before committing lab resources [93]. |
| SA Score Calculators | Computational tools (e.g., RDKit's sascorer, eTox) that provide a quantitative estimate of synthetic difficulty [91]. |
Prioritizing molecules from a large virtual screen or generative AI output based on synthetic feasibility [91]. |
| Molecular Descriptor Calculators | Software (e.g., Mordred) that calculates ~1,600 molecular descriptors (BertzCT, ring counts, etc.) [91]. | Building heuristic models to flag molecules with structural features that correlate with high synthetic complexity [91]. |
Improving the chemical accessibility of natural product leads is not merely a technical exercise but a strategic imperative that bridges the unparalleled bioactivity of natural compounds with the practical demands of modern drug development. By systematically applying the strategies outlinedâfrom foundational understanding and methodological toolkits to troubleshooting and rigorous validationâresearchers can successfully navigate the complexity of natural products. The future of this field lies in the deeper integration of AI-driven design, the continued expansion of navigable chemical spaces, and a commitment to sustainable sourcing. These efforts will undoubtedly accelerate the discovery of the next generation of NP-inspired therapeutics, transforming nature's most complex blueprints into accessible medicines for patients.