This article provides a comprehensive overview of natural product-based drug design, with a specific focus on the strategy of scaffold hopping.
This article provides a comprehensive overview of natural product-based drug design, with a specific focus on the strategy of scaffold hopping. It explores the foundational role of natural products as biologically prevalidated starting points for drug discovery, detailing the principles of scaffold hopping as defined by its key objective: retaining biological activity while altering the core molecular structure. The content covers a spectrum of methodological approaches, from traditional bioisosteric replacements and pharmacophore-based searches to modern, AI-driven generative models. It further addresses common challenges in the field, such as balancing structural novelty with maintained activity and navigating intellectual property, and presents validation frameworks through case studies and comparative analyses of different techniques. Aimed at researchers, scientists, and drug development professionals, this review synthesizes historical context, current state-of-the-art technologies, and future directions, offering a practical guide for leveraging natural product-inspired design to discover novel therapeutic candidates with improved properties.
Natural products (NPs) are chemical compounds derived from natural sources such as plants, microorganisms, marine organisms, and fungi. These molecules have served as a major source of chemically novel, bioactive therapeutics throughout human history and continue to play a pivotal role in modern drug discovery [1] [2]. Their structural diversity and evolutionary refinement make them indispensable for tackling complex medical challenges, particularly for cancer and infectious diseases [3] [4].
The historical use of natural products dates back to ancient civilizations, with the earliest records depicted on clay tablets in cuneiform from Mesopotamia (2600 B.C.) documenting oils from Cupressus sempervirens (Cypress) and Commiphora species (myrrh) for treating coughs, colds, and inflammation [1]. The continued relevance of these natural compounds in modern medicine underscores their inherent bioactivity and therapeutic value, validated by both traditional use and contemporary scientific research [1] [5].
Traditional medicinal practices have formed the foundation of most early medicines, with subsequent clinical, pharmacological, and chemical studies validating their efficacy [1]. Ancient records including the Ebers Papyrus (2900 B.C.), Chinese Materia Medica (1100 B.C.), and the works of Greek physician Dioscorides (100 A.D.) documented hundreds of plant-based drugs that established the foundation for modern pharmacotherapy [1].
Probably the most famous example is the development of acetylsalicyclic acid (aspirin) derived from the natural product salicin isolated from the bark of the willow tree Salix alba L. [1]. Similarly, investigation of Papaver somniferum L. (opium poppy) resulted in the isolation of several alkaloids including morphine, first reported in 1803, which remains a commercially important analgesic drug [1].
Natural products and their structural analogues have historically made a major contribution to pharmacotherapy, especially for cancer and infectious diseases [4]. Despite a decline in their pursuit by the pharmaceutical industry from the 1990s onwards, recent technological developments have revitalized interest in natural product-based drug discovery [3] [4].
Table 1: Historically Significant Natural Product-Derived Drugs
| Natural Product | Source Organism | Therapeutic Application | Discovery Timeline |
|---|---|---|---|
| Salicin | Willow tree (Salix alba L.) | Anti-inflammatory (precursor to aspirin) | Ancient use, isolated 1828 |
| Morphine | Opium poppy (Papaver somniferum) | Analgesic | Isolated 1803 |
| Artemisinin | Sweet wormwood (Artemisia annua) | Antimalarial | Discovered 1972 |
| Paclitaxel | Pacific yew tree (Taxus brevifolia) | Anticancer | Discovered 1971 |
| Teixobactin | Bacterium (Eleftheria terrae) | Antibiotic | Discovered 2015 |
Natural products continue to provide unique structural diversity in comparison to standard combinatorial chemistry, which presents opportunities for discovering novel low molecular weight lead compounds [1]. With less than 10% of the world's biodiversity evaluated for potential biological activity, numerous useful natural lead compounds await discovery [1].
Natural products distinguish themselves from synthetic libraries through their elevated molecular complexity, including higher proportions of sp3-hybridized carbon atoms, increased oxygenation, and decreased halogen and nitrogen content [2]. This chemical richness is coupled with rigid molecular frameworks and lower lipophilicity, traits that facilitate favorable interactions with biological targets, particularly those elusive to synthetic small molecules [2].
What sets NPs apart most profoundly is their evolutionary purpose [6]. These molecules function as defense chemicals, signaling agents, and ecological mediators, fine-tuned for optimal interactions with living systems through millions of years of evolutionary refinement [6] [2]. This natural selection has endowed NPs with mechanisms of action that exploit biological vulnerabilities, particularly in pathogens and cancer cells [2].
Natural products possess several innate functions, including the ability to allosterically alter the catalytic activity of enzymes, promote or disrupt macromolecular interactions, act as chemical messengers between cells, participate in inter-kingdom signaling, and serve as toxins for defense [6]. They can even carry out some protein-like functions of their own [6].
The biochemical diversity of natural products presents both a challenge and source of inspiration for biologists and chemists across the globe [6]. This diversity enables them to target specific pathways implicated in disease processes, offering tailored therapeutic strategies [5]. Moreover, the synergy observed within natural extractsâwhere multiple bioactive compounds act collaborativelyâenhances their overall efficacy and broadens their therapeutic potential [5].
Table 2: Innate Functions of Natural Products in Producing Organisms
| Function | Mechanism | Example Natural Products |
|---|---|---|
| Defense | Deter herbivory through bitter or toxic compounds | Pyrrolizidine alkaloids, glucosinolates |
| Signaling | Act as chemical messengers between cells | Flavonoids, strigolactones |
| Pollination | Attract pollinators through chromo-pigments | Carotenoids, anthocyanins |
| Symbiosis | Facilitate ecological associations | Nod factors in rhizobia-legume symbiosis |
| Environmental Adaptation | Protect against biotic and abiotic stresses | Osmoprotectants, phytoalexins |
The field of natural product drug discovery is experiencing a paradigm shift due to advanced technologies that increase speed, accuracy, and sustainability [2]. Traditional discovery workflows are being enhanced by high-throughput screening, artificial intelligence, machine learning, and omics technologies, which collectively streamline compound identification and development [2].
Advanced analytical techniques including ultrasonic-assisted extraction, supercritical fluid extraction, and various chromatographic methods have revolutionized the isolation and purification of natural bioactive compounds [5]. Characterization techniques such as mass spectrometry, nuclear magnetic resonance spectroscopy, and high-performance liquid chromatography provide detailed insights into chemical composition and structural elucidation [5].
Natural Product Drug Discovery Workflow. The diagram outlines the standard pipeline for discovering bioactive natural products, from initial extraction to final compound optimization. Key steps include bioassay-guided fractionation and target identification to ensure therapeutic relevance. UAE: Ultrasound-Assisted Extraction; SFE: Supercritical Fluid Extraction; MS: Mass Spectrometry; NMR: Nuclear Magnetic Resonance; HPLC: High-Performance Liquid Chromatography.
The integration of genome mining and biosynthetic engineering has revolutionized natural product discovery, offering solutions to longstanding challenges in the field [2]. Advances in understanding NP biosynthetic pathways, coupled with sophisticated genomic analysis tools, have paved the way for systematic exploration of microbial genomes [2].
Tools such as CRISPR-Cas systems, artificial intelligence, and bioinformatics platforms are accelerating hit discovery, de-replication, and biosynthetic pathway engineering, overcoming long-standing barriers to NP research [2]. Genome mining tools like DeepBGC and AntiSMASH enable rapid prediction and characterization of biosynthetic gene clusters, facilitating the discovery of novel compounds [2].
Principle: This protocol outlines a standardized approach for the extraction, fractionation, and identification of bioactive compounds from plant material using bioactivity-guided fractionation to isolate natural products with therapeutic potential [7] [5].
Materials:
Procedure:
Troubleshooting:
Principle: This protocol utilizes bioinformatics tools to identify biosynthetic gene clusters in microbial genomes, followed by heterologous expression to discover novel natural products [4] [2].
Materials:
Procedure:
Troubleshooting:
Table 3: Essential Research Reagent Solutions for Natural Product Research
| Reagent/Resource | Function | Application Examples |
|---|---|---|
| AntiSMASH | Identifies biosynthetic gene clusters in genomic data | Genome mining for novel natural products [2] |
| GNPS (Global Natural Products Social Molecular Networking) | Facilitates mass spectrometry data sharing and annotation | Dereplication and compound identification [4] |
| Sephadex LH-20 | Size exclusion chromatography medium for natural product separation | Fractionation of crude extracts by molecular size [5] |
| CRISPR-Cas Systems | Genome editing for pathway engineering | Activation of silent biosynthetic gene clusters [2] |
| HPLC-MS Systems | High-performance liquid chromatography coupled with mass spectrometry | Compound separation, quantification, and identification [5] |
| Dihydroevocarpine | Dihydroevocarpine, CAS:15266-35-0, MF:C23H35NO, MW:341.5 g/mol | Chemical Reagent |
| Nisamycin | Nisamycin, CAS:150829-93-9, MF:C24H27NO6, MW:425.5 g/mol | Chemical Reagent |
Natural products serve as excellent starting points for scaffold hopping approaches in drug discovery [8]. Scaffold hopping involves modifications to the core structure of an existing bioactive molecule to create new patentable molecules with potentially improved properties [8].
The complex molecular architectures of natural products, honed by evolutionary selection for bioactivity, provide privileged scaffolds that can be optimized through various hopping strategies [8]. These include heterocycle replacement, ring opening or closure, and peptidomimetics to enhance drug-like properties while maintaining biological activity [8].
Scaffold Hopping Strategy for Natural Products. This diagram illustrates how natural product scaffolds can be modified through various hopping approaches to generate diverse analog libraries for lead optimization. Multiple structural modification strategies are employed to enhance drug-like properties while maintaining core bioactivity.
Table 4: Successful Natural Product-Derived Drugs Developed Through Scaffold Hopping
| Original Natural Product | Scaffold Hopping Approach | Resulting Drug/Candidate | Therapeutic Application |
|---|---|---|---|
| Roxadustat | Heterocycle replacement | Novel HIF-PHD inhibitors | Renal anemia treatment [8] |
| GLPG1837 | Ring closure and expansion | SBD-100 | Cystic fibrosis (enhanced potency) [8] |
| Imidazo[1,2-a]pyrazine TTK inhibitors | Iterative heterocycle replacement | CFI-402257 | Anticancer agent (improved exposure) [8] |
| Sorafenib | Ring opening and amide bond modification | Quinazoline-2-carboxamide analogs | Enhanced VEGFR2 inhibition [8] |
Natural products remain a cornerstone of pharmaceutical innovation due to their unparalleled structural diversity, evolutionary optimization, and proven therapeutic potential [3] [2]. Their historical success in treating various diseases, combined with inherent bioactivity honed through millions of years of evolution, positions them as invaluable resources for addressing current and future health challenges [6] [4].
The integration of modern technologiesâincluding genomics, artificial intelligence, and advanced analytical techniquesâwith traditional knowledge is revitalizing natural product research [4] [2]. These developments are overcoming historical limitations and creating new opportunities to harness nature's chemical ingenuity for drug discovery, particularly through strategies such as scaffold hopping that optimize natural scaffolds for clinical application [8]. As global health challenges continue to evolve, natural products will undoubtedly remain at the forefront of therapeutic development.
Scaffold hopping is a fundamental strategy in modern medicinal chemistry and drug discovery. It is defined as the process of modifying the central molecular core, or scaffold, of a known bioactive compound to generate a novel chemotype while preserving or improving its biological activity [9] [10]. The core objective is to identify isofunctional molecular structures with significantly different molecular backbones [11].
This approach is pivotal for overcoming limitations associated with existing lead compounds, such as poor pharmacokinetics, toxicity, or intellectual property restrictions [12] [10]. By generating structurally novel compounds that retain the desired biological function, scaffold hopping enables researchers to create "me-better" or "fast-follower" drugs, expand the intellectual property space, and explore uncharted regions of chemical space for bioactive molecules [12] [13]. Its application is particularly valuable in natural product-based drug design, where complex structures often require optimization for drug-like properties [12] [14].
The structural changes in scaffold hopping can be systematically categorized based on the degree and nature of the modification to the parent molecule. The following table outlines a widely used classification system.
Table 1: A Classification Framework for Scaffold Hopping
| Degree of Hop | Core Modification | Key Characteristics | Exemplar Case |
|---|---|---|---|
| 1° (Heterocyclic Replacement) | Replacement, addition, or swap of heteroatoms within a ring system [9] [10]. | Retains the core spatial pharmacophore arrangement; tunes physicochemical properties [10]. | Sildenafil to Vardenafil: Swap of carbon and nitrogen atoms in a fused ring system [9] [10]. |
| 2° (Ring Opening or Closure) | Breaking a ring to create an acyclic chain or forming a new ring to rigidify a structure [9]. | Can significantly alter molecular flexibility and entropic penalties for binding [9]. | Morphine to Tramadol: Ring opening of a rigid, fused system to a more flexible molecule [9]. |
| 3° (Peptidomimetics) | Replacement of a peptide backbone with non-peptide moieties [9]. | Mimics the topology of a peptide while improving metabolic stability and oral bioavailability [9]. | Conversion of a therapeutic peptide (AMP1) to a small, non-peptide synthetic mimetic [12]. |
| 4° (Topology-Based Hopping) | Identification of cores with similar shape and pharmacophore features but distinct atomic connectivity [9]. | Leads to the highest degree of structural novelty; often relies on 3D shape and electrostatic similarity [9] [11]. | Use of the FTrees method to find distant chemical relatives based on "fuzzy pharmacophores" [11]. |
Several computational methodologies have been developed to facilitate scaffold hopping. The protocols below detail two primary approaches: one leveraging deep learning and another based on virtual screening with pharmacophore constraints.
DeepHop formulates scaffold hopping as a supervised molecule-to-molecule translation task, conditioned on a specific protein target [13].
The workflow for this deep learning approach is standardized as follows:
This protocol uses structure-based virtual screening, enhanced with pharmacophore constraints, to identify scaffold hops from commercial or internal compound libraries [11].
The following diagram illustrates the logical decision process within the virtual screening workflow:
Computational predictions require rigorous experimental validation to confirm successful scaffold hopping. The following table outlines key biophysical and cellular assays used for this purpose.
Table 2: Key Assays for Experimental Validation of Scaffold Hops
| Assay Type | Measured Parameter | Protocol Summary | Application in Validation |
|---|---|---|---|
| Intact Mass Spectrometry | Direct detection of ligand-bound protein complex and binding stoichiometry [15]. | Protein-ligand complexes are buffer-exchanged into volatile ammonium acetate and analyzed by native mass spectrometry [15]. | Confirms stabilization of a protein-protein interaction (PPI) by a molecular glue, demonstrating cooperative binding [15]. |
| Time-Resolved FRET (TR-FRET) | Change in fluorescence resonance energy transfer between labeled binding partners [15]. | A fluorescent donor and acceptor are attached to the two interacting proteins. Ligand-induced proximity increases FRET efficiency, measured over time [15]. | Quantifies the potency (EC50) of a scaffold hop in stabilizing or inhibiting a PPI in a purified system [15]. |
| Surface Plasmon Resonance (SPR) | Binding kinetics (association rate kon, dissociation rate koff) and affinity (K_D) [15]. | One binding partner is immobilized on a sensor chip. Analyte containing the other partner flows over it, and binding-induced refractive index changes are monitored in real-time [15]. | Determines if the scaffold hop maintains or improves binding affinity and residence time compared to the original ligand. |
| NanoBRET | Bioluminescence resonance energy transfer in live cells [15]. | Proteins of interest are tagged with NanoLuc luciferase (donor) and HaloTag (acceptor). Ligand-induced interaction is measured via BRET signal in live cells [15]. | Validates target engagement and functional efficacy (e.g., PPI stabilization) of scaffold hops in a physiologically relevant, cellular context [15]. |
Successful execution of scaffold hopping campaigns relies on a suite of computational tools and chemical resources.
Table 3: Essential Research Reagents and Software for Scaffold Hopping
| Tool / Resource | Type | Primary Function in Scaffold Hopping |
|---|---|---|
| Cresset Blaze / Spark [12] | Software | Blaze: Virtual screening of vendor libraries for whole-molecule replacement. Spark: Fragment replacement to generate ideas for synthesis [12]. |
| AnchorQuery [15] | Software / Virtual Library | Pharmacophore-based screening of a >31 million compound library of readily synthesizable (via MCR chemistry) scaffolds [15]. |
| SeeSAR & ReCore [11] | Software | SeeSAR: Interactive structure-based design and docking. ReCore: Topological replacement of molecular fragments based on 3D vector geometry [11]. |
| FTrees / infiniSee [11] | Software | FTrees: Similarity searching based on "Feature Trees" (fuzzy pharmacophores) to find distant chemical relatives in large chemical spaces [11]. |
| Scaffold Hunter [16] | Software | A visual analytics framework for analyzing chemical compound data, featuring scaffold tree visualization, clustering, and dataset comparison [16]. |
| ChEMBL Database [13] [17] | Database | A manually curated database of bioactive molecules with drug-like properties, used for training models and extracting bioactivity data [13] [17]. |
| ZINC Database [11] | Database | A freely available database of commercially available compounds for virtual screening [11]. |
| Groebke-Blackburn-Bienaymé (GBB) Reaction [15] | Chemical Reaction | A multi-component reaction used to rapidly synthesize drug-like imidazo[1,2-a]pyridine scaffolds identified through computational design [15]. |
| Trovafloxacin mesylate | Trovafloxacin mesylate, MF:C21H19F3N4O6S, MW:512.5 g/mol | Chemical Reagent |
| Lethedoside A | Lethedoside A | Lethedoside A is a natural flavone for cancer research. This product is for Research Use Only (RUO), not for human or veterinary use. |
In the pursuit of novel therapeutics, the strategy of scaffold hopping has become a cornerstone of modern medicinal chemistry, particularly within natural product-based drug design. This approach, defined as the modification of a compound's central core structure to generate a novel chemotype while retaining or improving biological activity, serves as a powerful method to overcome limitations of original leads [18] [9]. The ultimate goal is to discover structurally novel compounds that maintain efficacy against a biological target while achieving superior pharmacological properties [9].
The conceptual foundation of scaffold hopping, introduced in 1999, emphasizes two key components: different core structures and similar biological activities relative to the parent compound [18] [9]. This strategy appears to challenge the traditional similarity-property principle but is instead enabled by a more sophisticated understanding of molecular recognition. Ligands that fit the same protein pocket often share essential three-dimensional featuresâsuch as shape and electrostatio potential surfaceâeven if their underlying two-dimensional architectures belong to different chemotypes [18].
This application note systematically classifies scaffold hopping approaches, provides detailed experimental protocols, and frames these methodologies within the context of advancing natural product-based drug discovery.
Scaffold hopping strategies are broadly categorized into four distinct classes based on the degree and nature of structural modification applied to the parent molecule [18] [19] [9]. These classes represent a spectrum of structural change, from minor atomic substitutions to complete topological reorganization.
Table 1: Fundamental Classification of Scaffold Hopping Approaches
| Hop Class | Degree of Structural Novelty | Core Methodology | Primary Application Context |
|---|---|---|---|
| 1° Hop: Heterocycle Replacement | Low | Swapping or replacing atoms (e.g., C, N, O, S) within a ring system. | Lead optimization, patent circumvention, improving physicochemical properties like solubility [18] [9]. |
| 2° Hop: Ring Opening or Closure | Low to Medium | Breaking bonds to open fused rings or forming new bonds to create ring systems and control molecular flexibility [18] [9]. | Modifying pharmacokinetic profiles, enhancing potency by reducing entropy loss upon binding [18]. |
| 3° Hop: Peptidomimetics | Medium | Replacing peptide backbones with non-peptide moieties to mimic the spatial arrangement of key pharmacophoric groups [18] [9]. | Developing drug-like molecules from bioactive but metabolically unstable peptides [18]. |
| 4° Hop: Topology-Based Hopping | High | Identifying or designing cores with different connectivity but similar shape and pharmacophore alignment in 3D space [18] [19] [9]. | Discovering truly novel chemotypes, high-risk lead hopping for challenging targets [18]. |
The following diagram illustrates the logical relationships and decision-making pathways connecting these four classes of scaffold hops:
Molecular glues, which stabilize protein-protein interactions (PPIs), represent a challenging and promising frontier. A 2025 study detailed a scaffold-hopping approach to develop molecular glues for the 14-3-3Ï/ERα complex, starting from a known covalent molecular glue [20].
Objective: To design a novel, non-covalent molecular glue scaffold with improved drug-like properties using a computational topology-based hopping approach followed by synthesis via multicomponent reactions (MCRs) [20].
Table 2: Key Research Reagents & Solutions for Molecular Glue Development
| Reagent/Solution | Function/Description | Application in Protocol |
|---|---|---|
| AnchorQuery Software | Pharmacophore-based screening tool for a 31M+ compound library synthesizable via one-step MCRs [20]. | Virtual screening to identify novel scaffolds based on anchor and pharmacophore points from a known ligand. |
| GBB-3CR Reaction Components | Groebke-Blackburn-Bienaymé multicomponent reaction using aldehydes, 2-aminopyridines, and isocyanides [20]. | Rapid synthesis of the proposed imidazo[1,2-a]pyridine scaffold, enabling generation of diverse analogs. |
| TR-FRET Assay Kit | Time-Resolved Förster Resonance Energy Transfer assay for measuring PPI stabilization in a biochemical setting. | Biophysical validation of molecular glue efficacy in stabilizing the 14-3-3Ï/ERα interaction. |
| NanoBRET Assay System | Bioluminescence Resonance Energy Transfer assay configured for live-cell PPI analysis. | Cellular confirmation of target engagement and PPI stabilization under physiological conditions. |
Methodology:
The workflow for this protocol is summarized in the following diagram:
A groundbreaking 2024 study demonstrated a hybrid enzymatic-chemical strategy for scaffold hopping in complex natural product synthesis, moving beyond purely computational designs [14].
Objective: To efficiently generate diverse terpenoid natural product scaffolds from a single, commercially available terpenoid precursor via enzymatic oxidation followed by selective chemical rearrangement [14].
Methodology:
This protocol challenges traditional retrosynthetic analysis by establishing a shared, enzyme-generated intermediate for multiple target scaffolds, significantly improving synthetic efficiency.
Modern scaffold hopping is increasingly powered by artificial intelligence (AI) and advanced molecular representation methods, which move beyond traditional fingerprint-based approaches [19].
These AI-driven strategies can be categorized based on their approach to structural modification, as shown in the table below.
Table 3: AI-Driven Models for Molecular Modification in Scaffold Hopping
| Model Name | Core Architecture | Modification Strategy | Key Application Note |
|---|---|---|---|
| DeepFrag [21] | 3D Deep Convolutional Neural Network (DCNN) | Fragment Splicing | Treats fragment replacement as a classification task based on the protein-ligand complex structure. |
| FREED/FREED++ [21] | Graph CNN + Reinforcement Learning (RL) | Fragment Splicing | Uses RL to efficiently explore chemical space and generate molecules with high docking scores. |
| FRAME [21] | SE(3)-Equivariant Neural Network | Fragment Splicing | Explicitly models 3D protein-ligand interactions (H-bonds, Ï-Ï) for dynamic fragment selection. |
| MolEdit3D [21] | 3D Graph Neural Network | Fragment Editing & Splicing | A 3D graph editing model allowing for precise atomic and fragment-level modifications. |
| TACOGFN [21] | GFlowNet + Graph Transformer | Fragment Splicing | Incorporates target pocket information into a generative flow network for guided fragment addition. |
The systematic classification of scaffold hops into heterocycle replacements, ring operations, peptidomimetics, and topology-based changes provides a robust conceptual framework for medicinal chemists. This is particularly valuable in natural product research, where the goal is often to translate complex, bioactive scaffolds into viable drug leads. The integration of computational protocolsâranging from pharmacophore-based MCR screening to FEP-guided designâwith innovative experimental strategies, such as enzyme-enabled diversification, is redefining the scaffold hopping landscape. As AI-driven molecular representation and generation models continue to mature, they promise to further accelerate the discovery of novel chemotypes, enhancing our ability to explore chemical space and address unmet medical needs through natural product-inspired design.
The strategic modification of natural product (NP) scaffolds is a cornerstone of modern drug discovery, enabling the optimization of bioactive compounds for enhanced efficacy and druggability. This approach leverages the inherent, evolutionarily refined biological activities of NPs while overcoming limitations such as poor bioavailability, low potency, or high toxicity. The following application notes detail a transformative protocol for the diversification of terpenoid scaffolds, a class of NPs renowned for their structural complexity and broad bioactivity.
A pioneering application of enzyme-enabled scaffold hopping is demonstrated in the work of Renata and colleagues, who developed a versatile chemoenzymatic strategy to generate diverse terpenoid frameworks from a common precursor [14]. This method challenges the conventional retrosynthetic paradigm of designing a custom synthesis for each distinct molecular target. Instead of viewing an enzymatic modification as a final step, the team treated a biocatalytically installed functional group as a handle for subsequent abiotic skeletal rearrangements [22] [23].
Core Application Workflow: The process begins with sclareolide, a commercially available sesquiterpene lactone with a drimane-type skeleton [14]. The key enabling step involves the highly selective oxidation of a single carbon atom (C-3) on this scaffold using engineered cytochrome P450 enzymes (CYP450s) [14]. This biocatalytic transformation, difficult to achieve with traditional chemical methods, produces an alcohol intermediate. This intermediate is not an end-product but a versatile platform for "scaffold hopping" â a process that intentionally alters the core connectivity of the molecule [23]. Through carefully designed chemical reactions (e.g., ring-opening, rearrangement, and cyclization sequences), this single oxidized intermediate can be diverted down multiple synthetic pathways to produce terpenoids with vastly different carbon skeletons [14] [22].
Outcomes and Significance: Using this strategy, the research team successfully synthesized four distinct terpenoid natural products from sclareolide: merosterolic acid B, cochlioquinone B, (+)-daucene, and dolasta-1(15),8-diene [14] [22]. This demonstrates the remarkable structural divergence achievable from a single starting point. The implications for drug discovery are profound, as this platform technology significantly enhances synthetic efficiency. It saves time and cost by providing a shared entry point with branching pathways, thereby accelerating the exploration of structure-activity relationships (SAR) around complex terpenoid scaffolds for medicinal chemistry programs [14].
| Natural Product | Class | Key Structural Features Achieved | Potential/Reported Bioactivity |
|---|---|---|---|
| Merosterolic Acid B | Meroterpenoid | Complex ring system integrated with non-terpenoid structural units | Not specified in search results |
| Cochlioquinone B | Sesquiterpene-quinone | Fused quinonoid moiety installed via oxidation and rearrangement [23] | Environmentally relevant terpenoid [23] |
| (+)-Daucene | Sesquiterpene hydrocarbon | Altered ring junctions from the original drimane core [23] | Serves as a critical biosynthetic intermediate [23] |
| Dolasta-1(15),8-diene | Diterpene hydrocarbon | Unique double bond placements and ring fusion patterns [23] | Not specified in search results |
This protocol outlines the detailed methodology for the oxidative diversification of sclareolide and subsequent abiotic rearrangements to access distinct terpenoid skeletons, as pioneered by Renata's group [14] [22] [23].
| Reagent/Material | Function/Explanation |
|---|---|
| Sclareolide | The starting material; a commercially available sesquiterpene lactone providing a defined drimane-type scaffold. |
| Engineered Cytochrome P450 Enzymes (CYP450) | Biocatalysts engineered for high regio- and stereoselective oxidation of inert C-H bonds, installing a handle (alcohol) for further functionalization. |
| Cofactor System (e.g., NADPH) | Required for the enzymatic activity of CYP450s to drive the oxidation reaction. |
| Appropriate Buffer | To maintain optimal pH and ionic strength for enzymatic stability and activity. |
| Abiotic Reagents | Chemical reagents (e.g., acids, bases, catalysts) for skeletal rearrangements post-oxidation, enabling scaffold hopping. |
Essential Equipment: Bioreactor or controlled environment shaker for enzymatic reactions, standard organic synthesis glassware, analytical instruments (HPLC, LC-MS, NMR) for reaction monitoring and purification.
Step 1: Enzymatic C-H Hydroxylation of Sclareolide
Step 2: Abiotic Skeletal Rearrangement (Scaffold Hopping) This step is target-dependent. The following are generalized pathways based on the synthesized products.
Step 3: Purification and Characterization
The success of the protocol is determined by the structural confirmation of all final products. The key data for comparison is summarized below.
| Natural Product | Number of Synthetic Steps from Sclareolide (Representative) | Overall Yield (Representative, Over Multiple Steps) | Key Analytical Data (e.g., Specific Rotation [α]D) |
|---|---|---|---|
| Merosterolic Acid B | Not specified in search results | Not specified in search results | Not specified in search results |
| Cochlioquinone B | Not specified in search results | Not specified in search results | Not specified in search results |
| (+)-Daucene | Not specified in search results | Not specified in search results | The "(+)" designation indicates the compound is dextrorotatory. |
| Dolasta-1(15),8-diene | Not specified in search results | Not specified in search results | Not specified in search results |
Note: The source articles announce the achievement of the synthesis but do not report detailed quantitative yield data or step counts. The primary analytical confirmation is based on NMR and MS comparison with authentic samples or literature data [14] [22].
The following diagram illustrates the logical workflow of the enzyme-enabled scaffold hopping strategy.
| Tool/Reagent | Category | Function in Research |
|---|---|---|
| Engineered Cytochrome P450s | Biocatalyst | Key to regioselective C-H activation; provides the critical "handle" for downstream diversification that is unattainable purely chemically [14]. |
| Sclareolide | Natural Product Scaffold | Commercially available, complex starting material that serves as a versatile and privileged platform for generating structural diversity [14] [23]. |
| Cofactor Regeneration System | Biochemical Reagent | Maintains the activity of oxidative enzymes like CYP450s over prolonged reactions, improving efficiency and atom economy. |
| Molecular Generative AI Models | Computational Tool | For target-unknown scenarios, these AIDD models can propose novel structural modifications and predict bioactivity, guiding scaffold hopping efforts [21]. |
| Fragment Hotspot Maps (FHMs) | Computational Tool | Used in silico to identify optimal sites on a protein target for fragment binding, informing the design of new scaffolds for specific targets [21]. |
| Aprindine | Aprindine, CAS:33237-74-0, MF:C22H30N2, MW:322.5 g/mol | Chemical Reagent |
| Dodecamethylpentasiloxane | Dodecamethylpentasiloxane, CAS:141-63-9, MF:C12H36O4Si5, MW:384.84 g/mol | Chemical Reagent |
In the field of natural product-based drug design, scaffold hoppingâthe identification of isofunctional molecular structures with significantly different molecular backbonesâis a central strategy for discovering novel lead compounds with improved properties and intellectual potential [9]. This process enables researchers to transition from complex natural product scaffolds to synthetically accessible mimetics while preserving biological activity [24]. Traditional computational methods for scaffold hopping primarily rely on 2D molecular fingerprints and 3D pharmacophore models, each offering distinct advantages for exploring the complex chemical space of natural products [25] [26]. This application note details standardized protocols for employing these methods within natural product research programs, complete with performance benchmarks and implementation workflows.
Molecular fingerprints are vector representations that encode molecular structures into binary, count, or categorical formats based on specific structural or chemical patterns. They are predominantly used for rapid similarity searching and quantitative structure-activity relationship (QSAR) modeling [26].
Natural products often possess distinct chemical characteristics compared to synthetic drug-like molecules, including higher molecular complexity, more stereocenters, and a greater fraction of sp³-hybridized carbons [26]. These differences can significantly impact the performance of various fingerprinting algorithms. The table below summarizes the performance of major fingerprint categories on natural product datasets.
Table 1: Performance Evaluation of Molecular Fingerprints on Natural Product Datasets
| Fingerprint Category | Representative Examples | Basis of Calculation | Performance on Natural Products |
|---|---|---|---|
| Circular Fingerprints | ECFP, FCFP [26] | Atom-centered radial substructures | Good performance, but can be outperformed by other fingerprints for specific NP bioactivity prediction tasks [26]. |
| Path-Based Fingerprints | Atom Pairs (AP), Depth First Search (DFS) [26] | Linear paths through molecular graph | Performance varies; requires benchmarking for specific NP datasets [26]. |
| Pharmacophore Fingerprints | Pharmacophore Pairs (PH2), Triplets (PH3) [26] | Topological distances between pharmacophoric points | Can provide an abstract, interaction-based representation less dependent on specific scaffold [26]. |
| Substructure-Based Fingerprints | MACCS, PubChem [26] | Presence of predefined structural keys | Limited by predefined fragment dictionary, potentially hampering scaffold hopping [27]. |
| String-Based Fingerprints | MHFP, MAP4 [26] | Fragmentation of SMILES strings | MAP4 has shown competitive or superior performance to ECFP in NP bioactivity prediction [26]. |
Principle: This ligand-based protocol uses a known active natural product as a query to identify structurally different compounds with similar bioactivity by calculating molecular similarity in a 2D descriptor space [24].
Materials & Software:
Procedure:
A pharmacophore is defined as "the ensemble of steric and electronic features that is necessary to ensure the optimal supramolecular interactions with a specific biological target" [25] [29]. Pharmacophore modeling abstracts a molecule from its specific chemical structure to a set of generalized features essential for binding, making it inherently powerful for scaffold hopping [25].
Table 2: Core Pharmacophore Features and Model Generation Approaches
| Feature Type | Geometric Representation | Interaction Type | Generation Method | Description & Utility |
|---|---|---|---|---|
| H-Bond Donor (HBD) | Vector or Sphere [25] | Hydrogen Bonding | Structure-Based [25] | Derived from protein-ligand complex crystal structure. Most reliable method. |
| H-Bond Acceptor (HBA) | Vector or Sphere [25] | Hydrogen Bonding | Ligand-Based [25] | Generated from a set of aligned active ligands. Requires known actives. |
| Positive/Negative Ionizable (PI/NI) | Sphere [25] | Ionic | Manual [25] | Built based on deep expert knowledge of target and ligands. |
| Aromatic (AR) | Plane or Sphere [25] | Ï-Stacking, Cation-Ï | - | - |
| Hydrophobic (H) | Sphere [25] | Hydrophobic Contact | - | - |
| Exclusion Volumes | Sphere [25] | Steric Clash Prevention | - | Define regions the ligand must not occupy; often from protein structure [25]. |
Principle: This protocol uses the 3D structure of a biological target (often with a bound natural product ligand) to derive a set of essential interaction features, which are then used as a query to screen for novel scaffolds [25].
Materials & Software:
Procedure:
A synergistic approach that leverages the speed of 2D fingerprints and the scaffold-hopping power of 3D pharmacophores is often most effective. A common strategy is to use a 2D similarity pre-filter to narrow down a large database, followed by a more computationally intensive 3D pharmacophore screen on the resulting subset [30]. Furthermore, holistic 3D descriptors like WHALES (Weighted Holistic Atom Localization and Entity Shape) have been developed specifically to bridge the gap between natural products and synthetic mimetics by simultaneously encoding pharmacophore, shape, and partial charge distribution [24] [27].
The table below provides a comparative overview of the scaffold-hopping performance of different molecular representations, highlighting the effectiveness of 3D methods.
Table 3: Benchmarking the Scaffold-Hopping Performance of Different Molecular Representations
| Molecular Representation | Dimension | Core Principle | Reported Scaffold-Hopping Performance (SDA%)* | Key Advantage for NPs |
|---|---|---|---|---|
| ECFPs | 2D | Circular substructures | 73 ± 12 [27] | Fast, widely used, good overall performance [26]. |
| MACCS Keys | 2D | Predefined fragments | 75 ± 12 [27] | Interpretable, but limited by predefined dictionary [27]. |
| CATS2 | 2D | Topological pharmacophore pairs | - | Abstract representation aids hopping [27]. |
| WHIM | 3D | Statistical projection of 3D coordinates | - | Captures overall molecular shape [27]. |
| WHALES | 3D | Holistic integration of shape & pharmacophore | Outperformed benchmarks in 89% of 182 targets [27] | Specifically designed for NP-to-synthetic hopping; high success rate [24] [27]. |
*SDA% (Scaffold Diversity of Actives): The ratio of unique Murcko scaffolds to the number of active compounds retrieved in the top 5% of a virtual screening rank, with higher values indicating better scaffold-hopping ability [27].
Table 4: Key Software Tools for Molecular Fingerprint and Pharmacophore Modeling
| Tool Name | Category | Primary Function | Application in Protocol |
|---|---|---|---|
| RDKit | Cheminformatics | Open-source toolkit for cheminformatics | Fingerprint calculation (ECFP, etc.), molecular standardization, and similarity searching [28] [26]. |
| FTrees/InfiniSee | Virtual Screening | Pharmacophore-based similarity search | Algorithm behind the "Scaffold Hopper Mode" to find compounds related by pharmacophore features [30]. |
| ReCore (SeeSAR) | Structure-Based Design | Structure-based core replacement | Suggests molecular motifs that replace a scaffold while maintaining binding interactions and side chains [30]. |
| FlexS | Molecular Alignment | 3D ligand alignment | Used to align candidate compounds to a reference pharmacophore for similarity assessment [30]. |
| MOE | Integrated Suite | Molecular modeling and simulation | Comprehensive environment for pharmacophore model creation (both structure- and ligand-based) and virtual screening [9]. |
| Cetefloxacin | Cetefloxacin|CAS 141725-88-4|RUO | Cetefloxacin is a synthetic fluoroquinolone antibiotic for research use only (RUO). Inhibits DNA gyrase. Not for human or veterinary use. | Bench Chemicals |
| Penethamate hydriodide | Penethamate hydriodide, CAS:808-71-9, MF:C22H32IN3O4S, MW:561.5 g/mol | Chemical Reagent | Bench Chemicals |
The Weighted Holistic Atom Localization and Entity Shape (WHALES) descriptors represent an advanced molecular representation technique designed to facilitate scaffold hopping in computer-assisted drug discovery. Unlike traditional fingerprints that focus on molecular connectivity or presence of specific fragments, WHALES descriptors simultaneously capture critical information about molecular shape and partial charge distribution in a holistic manner. Originally developed to translate structural and pharmacophore information from bioactive natural products to synthetically accessible isofunctional compounds, WHALES has demonstrated remarkable capability in identifying novel ligand chemotypes that populate uncharted regions of the chemical space while maintaining desired biological activity [31] [32].
The fundamental innovation of WHALES lies in its integration of geometric interatomic distances with atomic physicochemical properties, enabling the identification of structurally diverse compounds that share similar bioactivity profiles. This approach has proven particularly valuable in natural product-based drug design, where complex molecular architectures often serve as inspiration for developing synthetically tractable compounds with improved drug-like properties. By enabling efficient navigation through chemical space, WHALES addresses a key challenge in medicinal chemistry: the discovery of novel bioactive chemotypes through straightforward similarity searching [32].
WHALES descriptors encode molecular information through a sophisticated algorithm that transforms three-dimensional molecular structures and their electronic properties into a numerical representation. The methodology employs weighted locally-centred atom distances computed for each atom position in a three-dimensional molecular conformation, creating a comprehensive profile that captures molecular shape and charge distribution simultaneously. This holistic approach allows WHALES to identify structurally diverse compounds that share similar steric and electronic properties, making it particularly effective for scaffold hopping applications where traditional fragment-based methods often fail [32].
The theoretical foundation of WHALES rests on the calculation of an atom-centred covariance matrix for each non-hydrogen atom, which captures the distribution of atoms and their partial charges around each atomic center. This matrix incorporates both spatial atomic coordinates and their associated partial charges, creating a weighted representation of the molecular environment that forms the basis for subsequent distance calculations and descriptor generation [32].
For each non-hydrogen atom (j) in the molecule, compute the weighted covariance matrix using the formula:
$${{\bf{S}}}{w(j)}=\frac{{\sum }{i=1}^{n}\,|{\delta }{i}|\cdot ({{\bf{x}}}{i}-{{\bf{x}}}{j}){({{\bf{x}}}{i}-{{\bf{x}}}{j})}^{{\rm{T}}}}{{\sum }{i=1}^{n}\,|{\delta }_{i}|}$$
where:
Compute the ACM distance matrix with the equation:
$${\bf{A}}{\bf{C}}{\bf{M}}\,(i,j)={({{\bf{x}}}{i}-{{\bf{x}}}{j})}^{{\rm{T}}}\cdot {{\bf{S}}}{w(j)}^{-1}\cdot ({{\bf{x}}}{i}-{{\bf{x}}}_{j})$$
This matrix collects all pairwise normalized interatomic distances according to the atom-centred covariance matrix, where atoms located in directions of high variance have smaller distances from the atomic center than those in low-variance regions [32].
From the ACM matrix (excluding diagonal elements), calculate:
Calculate the distribution statistics of atomic remoteness, isolation degree, and their ratios to generate 33 WHALES descriptors comprising:
Table 1: WHALES Descriptor Variants Based on Partial Charge Calculation Methods
| Descriptor Version | Partial Charge Method | Complexity Level | Key Applications |
|---|---|---|---|
| WHALES-DFTB+ | DFTB+ accelerated quantum mechanical simulation | High chemical detail | High-precision scaffold hopping |
| WHALES-GM | Gasteiger-Marsili connectivity-based method | Medium chemical detail | Large-scale virtual screening |
| WHALES-shape | δᵢ = 1 for all atoms (no charge information) | Basic shape-based representation | Shape-focused similarity searching |
WHALES descriptors have been rigorously evaluated for their scaffold-hopping potential through comprehensive benchmarking studies. In a systematic analysis comparing WHALES with seven state-of-the-art molecular representations across 30,000 bioactive compounds and 182 biological targets, WHALES demonstrated superior performance in identifying structurally diverse active compounds [32].
The benchmark included molecular descriptors spanning different dimensionalities and chemical information domains:
Performance was quantified using the Scaffold Diversity of Actives (SDA%) metric, calculated as:
$$S{D}_{A} \% =\frac{ns}{na}\cdot 100$$
where ns represents the number of unique Murcko scaffolds identified in the top 5% of the ranked list, and na is the number of actives present in that same portion [32].
Table 2: Performance Comparison of Molecular Descriptors in Scaffold Hopping
| Molecular Representation | Dimensionality | Mean SDA% ± Standard Deviation | Key Strengths |
|---|---|---|---|
| WHALES-DFTB+ | 3D + Electronic | 89 ± 9 | Best overall scaffold hopping |
| WHALES-GM | 3D + Electronic | 87 ± 10 | Balance of speed and performance |
| WHALES-shape | 3D Shape | 85 ± 11 | Pure shape similarity |
| GETAWAY | 3D | 84 ± 11 | Atom-weighted 3D descriptors |
| WHIM | 3D | 82 ± 12 | Principal axes molecular properties |
| CATS2 | 2D | 80 ± 13 | Pharmacophore pairs |
| Matrix-based descriptors | 2D | 78 ± 13 | Graph theory-based |
| MACCS FP | 1D | 75 ± 12 | 166 predefined substructures |
| ECFPs | 1D | 73 ± 12 | Atom-centered radial fragments |
| Constitutional descriptors | 0D/1D | 76 ± 13 | Basic molecular properties |
The benchmark results demonstrated that all three WHALES versions outperformed state-of-the-art methods in 89% of the tested biological targets, with WHALES-DFTB+ showing the highest scaffold-hopping ability [32]. This superior performance highlights WHALES' capacity to identify structurally diverse compounds that maintain similar bioactivity, a crucial capability in natural product-inspired drug design where structural complexity often necessitates simplification while retaining activity.
The scaffold-hopping capability of WHALES was validated in a prospective application targeting the retinoid X receptor (RXR). Using known RXR modulators as queries, WHALES descriptors identified four novel agonists with innovative molecular scaffolds that populated previously uncharted regions of the chemical space [32].
Notably, one identified agonist possessed a rare non-acidic chemotype that exhibited:
This successful prospective application confirmed WHALES' ability to detect novel bioactive chemotypes through straightforward similarity searching, demonstrating its practical utility in hit identification and lead optimization campaigns [32].
A 2025 study further validated WHALES' utility in drug repurposing for the human dopamine transporter (hDAT). Researchers employed WHALES descriptors to identify novel atypical inhibitors that bind to hDAT's allosteric site, using four benztropine-like atypical inhibitors as templates [33].
The workflow encompassed:
This integrated approach successfully identified three compounds with significant hDAT inhibitory potency (ICâ â values of 0.753 μM, 0.542 μM, and 1.210 μM, respectively), demonstrating WHALES' effectiveness in prospective drug discovery applications [33].
Purpose: To identify novel chemotypes with similar biological activity to a query compound through WHALES-based similarity searching.
Materials and Software:
Procedure:
Database preparation:
WHALES descriptor calculation:
Similarity searching:
Result analysis:
Troubleshooting:
Purpose: To translate structural information from complex natural products to synthetically accessible compounds using WHALES descriptors.
Materials and Software:
Procedure:
Synthetic library screening:
Scaffold hopping analysis:
Hit validation:
Validation: In the proof-of-concept study, WHALES identified seven natural-product-inspired synthetic compounds that modulated the cannabinoid receptor, featuring innovative scaffolds compared to actives annotated in ChEMBL [32].
Table 3: Essential Research Reagents and Computational Tools for WHALES Applications
| Tool/Resource | Type | Function in WHALES Workflow | Availability |
|---|---|---|---|
| DFTB+ | Software | Calculates accurate partial charges for WHALES-DFTB+ | Freely available |
| Gasteiger-Marsili Method | Algorithm | Rapid partial charge calculation for large datasets | Implemented in RDKit, OpenBabel |
| MMFF94 Force Field | Parameter Set | Generates energy-minimized 3D conformations | Implemented in most cheminformatics packages |
| RDKit | Cheminformatics Library | Handles molecular I/O, conformation generation, and descriptor calculation | Open-source Python library |
| - WHALES Implementation | Code Repository | Complete code for WHALES calculation and screening | github.com/ETHmodlab/scaffoldhoppingwhales [31] |
| ChEMBL Database | Chemical Database | Source of bioactive compounds for benchmarking and queries | Publicly available |
| - Murcko Scaffold Analysis | Algorithm | Quantifies scaffold diversity in screening results | Implemented in RDKit |
| Ziconotide | Ziconotide (Prialt) | Ziconotide is a synthetic ω-conotoxin for research into chronic pain. It is a selective, non-opioid N-type calcium channel blocker. For Research Use Only. Not for human use. | Bench Chemicals |
| Chloranil | Chloranil, CAS:118-75-2, MF:C6Cl4O2, MW:245.9 g/mol | Chemical Reagent | Bench Chemicals |
WHALES descriptors represent a significant advancement in molecular representation for scaffold hopping and natural product-inspired drug design. By simultaneously capturing molecular shape and charge distribution, WHALES enables efficient identification of structurally diverse compounds with similar biological activity, addressing a critical challenge in medicinal chemistry. The robust performance of WHALES across retrospective benchmarks and prospective applications highlights its value in hit identification and lead optimization workflows. As computational drug discovery continues to evolve, holistic representations like WHALES will play an increasingly important role in bridging the gap between complex natural product architectures and synthetically accessible therapeutic compounds.
Natural products (NPs) are invaluable resources for drug discovery, characterized by their intricate scaffolds and diverse bioactivities. However, their structural complexity often leads to undesirable properties such as toxicity, metabolic instability, or poor pharmacokinetic profiles [21]. Scaffold hopping has emerged as a critical strategy to overcome these limitations by designing molecules with novel core structures (scaffolds) that retain the desired biological activity of the original natural product [19]. This approach not only helps optimize drug-like properties but also facilitates the creation of novel chemical entities with freedom-to-operate advantages [24].
The rapid evolution of artificial intelligence (AI), particularly deep learning, has revolutionized the computational approach to scaffold hopping. Traditional methods relied on searching predefined databases using molecular fingerprints or pharmacophore models, inherently limiting exploration to known chemical space [28]. AI-driven methods, especially transformer networks and graph-based models, now enable generative exploration of vast chemical spaces, facilitating the discovery of novel scaffolds absent from existing libraries [19] [34]. These approaches have positioned AI-assisted drug design as a prominent research area, significantly accelerating early screening and lead compound identification [19].
A prerequisite for effective AI-driven scaffold hopping is translating molecular structures into computer-readable formats, known as molecular representations [19]. These representations bridge the gap between chemical structures and their biological, chemical, or physical properties [19].
Traditional molecular representations include:
AI-driven molecular representations employ deep learning techniques to learn continuous, high-dimensional feature embeddings directly from large datasets [19]. These include:
Table 1: Comparison of Molecular Representation Methods for Scaffold Hopping
| Representation Type | Key Examples | Advantages | Limitations |
|---|---|---|---|
| String-Based | SMILES, SELFIES | Simple, compact, human-readable [19] | Limited in capturing molecular complexity [19] |
| Fingerprint-Based | ECFP, Morgan Fingerprints [28] | Computationally efficient, effective for similarity search [19] | Predefined features, may overlook subtle structural relationships [19] |
| 3D Holistic | WHALES [24] | Encodes shape and pharmacophore information | Requires 3D conformer generation |
| Graph-Based | Graph Neural Networks [19] | Naturally represents molecular structure | Requires specialized architectures |
| Language Model-Based | Molecular Transformers [19] | Leverages NLP advances, captures sequential patterns | SMILES syntax may not reflect structural similarity |
Transformer networks, originally developed for natural language processing, have been successfully adapted for molecular design tasks including scaffold hopping. Their self-attention mechanism enables modeling of long-range dependencies and complex structural relationships in molecular data [35].
The DeepHop model represents a significant advancement in scaffold hopping by reformulating the task as a supervised molecule-to-molecule translation problem [28]. This multimodal architecture integrates molecular 3D conformer information through a spatial graph neural network and protein sequence information through a transformer model [28].
Architecture and Training:
Performance Metrics:
Figure 1: DeepHop Multimodal Transformer Architecture for Scaffold Hopping
Data Preparation:
Model Application:
Graph-based models, particularly Graph Neural Networks (GNNs), provide a natural representation of molecular structure by treating atoms as nodes and bonds as edges in a graph [19]. These models have demonstrated remarkable capability in capturing both local and global molecular features essential for effective scaffold hopping [36].
Recent advancements in graph-based approaches have incorporated 3D structural information to enhance scaffold hopping performance:
FRAME utilizes SE(3)-equivariant neural networks to explicitly model protein-ligand interactions (hydrogen bonds, Ï-Ï stacking) and dynamically select optimal connection points and fragments from starting molecules [21].
D3FG employs a diffusion model with rigid functional group definitions to preserve complex fragment structures during generation, capturing spatial relationships and interactions between proteins and ligands via graph neural networks [21].
STRIFE dynamically guides starting molecule expansion by extracting fragment hotspot maps (FHMs) from protein targets, using a combination of GNN and CNN architectures [21].
Fragment Library Preparation:
Model Implementation:
Validation:
Table 2: Performance Comparison of AI-Driven Scaffold Hopping Methods
| Method | Architecture | 2D Similarity Range | 3D Similarity Range | Success Rate | Key Applications |
|---|---|---|---|---|---|
| DeepHop [28] | Multimodal Transformer | â¤0.6 (Tanimoto) | â¥0.6 (SC Score) | 70% (Improved bioactivity) | Kinase inhibitors |
| WHALES [24] | Holistic Descriptors | Variable | High shape similarity | 35% (Experimental active) | Cannabinoid receptor modulators |
| ChemBounce [37] | Fragment Replacement | Tunable threshold | Electron shape similarity | Varies by target | General synthetic mimetics |
| FREED [21] | GCN + Reinforcement Learning | Not specified | Not specified | High docking scores | Target-informed design |
Figure 2: Graph-Based Fragment Replacement Workflow for Scaffold Hopping
Table 3: Research Reagent Solutions for Scaffold Hopping Implementation
| Resource Category | Specific Tools/Libraries | Function | Application Context |
|---|---|---|---|
| Molecular Representation | RDKit, ODDT (with ElectroShape) [37] | Chemical informatics and descriptor calculation | Preprocessing, similarity assessment |
| Fragment Libraries | ChEMBL-derived scaffolds (3.2M+ unique) [37], VEHICLe database [28] | Source of replaceable molecular fragments | Fragment-based scaffold replacement |
| Deep Learning Frameworks | PyTorch, TensorFlow, PyTorch Geometric [35] | Implementation of GNNs and transformers | Model development and training |
| Benchmarking Platforms | OpenGT [35], GuacaMol, MOSES [34] | Standardized evaluation of generative models | Performance validation |
| Property Prediction | MTDNN [28], DMPNN [28], FP-ADMET [19] | Virtual profiling of generated molecules | Bioactivity and ADMET assessment |
AI-driven scaffold hopping, particularly through transformer networks and graph-based models, has fundamentally transformed natural product-based drug design. These approaches enable systematic exploration of chemical space beyond the limitations of traditional database searching, facilitating the discovery of novel scaffolds with maintained bioactivity and improved drug-like properties [19] [34].
The integration of 3D structural information through spatial graph networks and the application of multimodal learning represent significant advancements in the field [21] [28]. Models like DeepHop demonstrate that combining molecular structure with target protein information yields substantially higher success rates in generating viable scaffold hops compared to traditional methods [28].
Future developments will likely focus on improving synthetic accessibility of AI-generated molecules, enhancing model interpretability, and addressing data scarcity for novel targets [34] [38]. As these computational methods continue to evolve, they will further accelerate the transformation of natural product inspiration into viable therapeutic candidates, bridging the gap between nature's chemical diversity and modern drug discovery needs.
Tuberculosis (TB), caused by the bacterial pathogen Mycobacterium tuberculosis (Mtb), remains a leading cause of death worldwide, with the emergence of multi-drug resistant (MDR) and extensively drug-resistant (XDR) TB posing significant challenges to global eradication efforts [39] [40]. The complexity of Mtb's physiology, including its lipid-rich cell envelope and extensive metabolic adaptation capabilities, necessitates novel treatment strategies and therapeutic targets [40]. Protein kinases (PKs), key regulators of cellular processes across all life forms, have emerged as major targets in anti-TB drug discovery due to their essential roles in Mtb survival, metabolism, and pathogenesis [39] [40].
The Mtb genome encodes approximately 60 different protein kinases, including 11 two-component regulatory systems (2CRS) and 11 serine/threonine protein kinases (STPKs), along with numerous metabolic kinases essential for fundamental biosynthetic pathways [40]. This application note presents practical case studies and detailed protocols for targeting Mtb kinases, with a specific focus on scaffold hopping strategies to discover novel chemotypes with improved properties.
Thymidylate kinase of Mtb (TMPKmt) catalyzes the phosphorylation of dTMP to dTDP in the pyrimidine biosynthesis pathway, an essential step in DNA synthesis [39] [40]. TMPKmt is mechanistically and structurally unrelated to the human enzyme, making it an attractive target for developing selective antitubercular agents with minimal host toxicity [40]. Gene essentiality studies have confirmed that TMPKmt is indispensable for mycobacterial survival, further validating its therapeutic potential [39].
Scaffold hopping from known natural product inhibitors of TMPKmt has yielded promising synthetic analogs with retained activity and improved synthetic accessibility. The application of holistic molecular similarity approaches, such as Weighted Holistic Atom Localization and Entity Shape (WHALES) descriptors, has enabled successful scaffold hopping by simultaneously capturing pharmacophore features, partial charge distributions, and molecular shape [24]. These descriptors facilitate the identification of isofunctional synthetic compounds that maintain key interaction patterns while reducing structural complexity compared to natural product templates [24].
Table 1: Classification of Scaffold Hopping Approaches for Kinase Inhibitor Design
| Hop Category | Structural Change | Degree of Novelty | Example Application |
|---|---|---|---|
| 1° Hop | Heteroatom replacements or swaps in backbone rings | Low | Carbon-nitrogen swaps in aromatic systems [9] |
| 2° Hop | Ring opening or closure | Medium | Morphine to tramadol transformation [9] |
| 3° Hop | Peptidomimetics | Medium-High | Replacement of peptide backbones with non-peptidic moieties [9] |
| 4° Hop | Topology-based changes | High | Field-based scaffold hopping using electrostatic similarity [12] |
Host-directed therapy (HDT) represents a promising alternative strategy for combating intracellular bacteria like Mtb by targeting host cellular mechanisms that support bacterial survival [41]. A recent high-throughput screen of 827 ATP-competitive kinase inhibitors from the Published Kinase Inhibitor Sets (PKIS1 and PKIS2) identified multiple compounds effective against intracellular Mtb in human cell lines and primary macrophages [41].
The screening workflow employed flow cytometry-based infection models using HeLa and MelJuSo cell lines infected with DsRed-expressing Mtb. Two distinct populations of infected cells were observed: DsRed-dim (containing few bacteria) and DsRed-bright (containing replicating bacteria, with 142 times more viable bacteria than dim cells) [41]. Compounds reducing the DsRed-bright population without host cell cytotoxicity were prioritized for further development.
Protocol 1: Primary Screening of Kinase Inhibitor Libraries
Table 2: Promising Kinase Inhibitor Chemotypes Identified as HDT Candidates Against Mtb
| Compound Chemotype | Primary Kinase Targets | Mtb Inhibition (Rescreen Z-score) | Cytotoxicity | Therapeutic Potential |
|---|---|---|---|---|
| Morpholino-imidazo/triazolo-pyrimidinones | PIK3CB | < -2 | Non-cytotoxic | PI3K/AKT pathway modulation [41] |
| 2-Aminobenzimidazoles | ABL1 | < -2 | Non-cytotoxic | Lysosomal acidification restoration [41] |
| 2-Anilino-4-pyrrolidinopyrimidines | JAK2, AAK1 | < -2 | Non-cytotoxic | Intracellular trafficking modulation [41] |
| 4-Anilinoquinolines | MAP2K5, RIPK2, RSK4 | < -2 | Non-cytotoxic | Multiple signaling pathway disruption [41] |
Shikimate kinase (SK), the fifth enzyme in the shikimate pathway, catalyzes the ATP-dependent phosphorylation of shikimate to shikimate-3-phosphate [42]. This pathway is essential in bacteria, fungi, and plants but absent in mammals, making it an attractive target for selective antimicrobial development [42]. Mtb SK (MtSK) is encoded by the aroK gene and has been validated as essential for bacterial growth through gene knockout studies [42].
Protocol 2: LC-MS Screening for Shikimate Kinase Inhibitors
This target-based approach identified 14 novel MtSK inhibitors with ICââ values <5 μg/mL against Mtb H37Rv and favorable selectivity indices (SI >10), demonstrating the utility of combining phenotypic screening with target-based validation [42].
The WHALES (Weighted Holistic Atom Localization and Entity Shape) molecular representation enables scaffold hopping from complex natural products to synthetically accessible mimetics by simultaneously encoding pharmacophore features, atomic distributions, and molecular shape [24]. The descriptor calculation involves four key steps:
In prospective applications using natural cannabinoids as queries, this approach achieved 35% success rate in identifying novel synthetic modulators of human cannabinoid receptors, with identified modulators being structurally less complex than their natural product templates [24].
Field-based molecular similarity methods provide another powerful approach for scaffold hopping in kinase inhibitor design. These methods compare molecules based on their electrostatic, steric, and hydrophobic fields rather than structural frameworks, enabling identification of diverse chemotypes with conserved interaction potential [12]. Successful applications include:
Table 3: Essential Research Reagents for Kinase-Focused Anti-TB Drug Discovery
| Reagent/Material | Specifications | Experimental Function | Application Examples |
|---|---|---|---|
| PKIS Libraries | PKIS1 & PKIS2 (827 ATP-competitive kinase inhibitors) | Source of chemically diverse, well-characterized kinase inhibitors with known target profiles | Host-directed therapy screening against intracellular Mtb [41] |
| DsRed-Expressing Mtb | Recombinant Mtb H37Rv expressing DsRed fluorescent protein | Enables quantification of intracellular bacterial burden via flow cytometry | High-throughput screening in host cell infection models [41] |
| Recombinant Kinases | Purified Mtb kinases (TMPKmt, MtSK) expressed in E. coli | Target proteins for biochemical inhibition assays | Enzyme activity assays, inhibitor characterization [39] [42] |
| LC-MS System | HPLC coupled to mass spectrometer with C18 column | Quantification of reaction substrates and products in enzyme assays | Shikimate-3-phosphate detection in MtSK inhibition assays [42] |
| 2-Undecanone | 2-Undecanone, CAS:112-12-9, MF:C11H22O, MW:170.29 g/mol | Chemical Reagent | Bench Chemicals |
Kinase inhibitors represent promising candidates for direct antitubercular activity and host-directed therapy approaches. The case studies and protocols presented herein demonstrate practical applications of kinase-focused drug discovery, highlighting the utility of scaffold hopping strategies to generate novel chemotypes with maintained activity against Mtb targets. Integration of computational design methods with robust experimental screening protocols provides a powerful framework for advancing new therapeutic candidates against drug-resistant tuberculosis.
The pursuit of novel therapeutic agents requires researchers to navigate a fundamental tension: the need for structural innovation against the imperative to maintain biological activity. This challenge is particularly acute in natural product-based drug design, where the complex scaffolds of natural products offer rich starting points for drug discovery but often require optimization to improve drug-like properties [24]. Scaffold hopping, defined as the strategic modification of a molecule's core structure to generate novel chemotypes while preserving bioactivity, has emerged as a powerful methodology to address this challenge [9] [18].
The similarity property principle posits that structurally similar molecules tend to exhibit similar biological activities [9]. However, scaffold hopping operates at the boundaries of this principle, demonstrating that significant structural changes can retain or even enhance desired activity when key pharmacophore elements are conserved [24]. This application note provides a structured framework for researchers to balance structural novelty with biological activity through scaffold hopping approaches, with particular emphasis on natural product-inspired design.
Scaffold hopping strategies can be systematically categorized into distinct classes based on the degree and nature of structural modification [9] [18]. Understanding this classification helps researchers select appropriate strategies for their specific design goals. The following table summarizes the four primary scaffold hopping categories:
Table 1: Classification of Scaffold Hopping Approaches
| Hop Category | Structural Change | Novelty Level | Success Rate | Primary Application |
|---|---|---|---|---|
| 1° Heterocycle Replacements | Swapping atoms in aromatic rings or replacing carbon with heteroatoms | Low | High | Lead optimization, patent protection |
| 2° Ring Opening/Closure | Breaking or forming ring systems to control molecular flexibility | Medium | Medium | Modifying pharmacodynamic and pharmacokinetic properties |
| 3° Peptidomimetics | Replacing peptide backbones with non-peptidic moieties | Medium-High | Medium | Converting bioactive peptides into drug-like molecules |
| 4° Topology-Based Hopping | Identifying chemotypes with similar shape and pharmacophores but different atomic connectivity | High | Low | Discovering truly novel chemotypes from natural products |
The trade-off between structural novelty and success rate is evident across these categories [9]. Small-step hops (1° and some 2°) generally offer higher probabilities of maintaining bioactivity but yield more modest structural innovations. Large-step hops (4° and some 3°), while offering greater novelty, present higher risks of activity loss [18]. This relationship must be carefully considered when planning scaffold hopping campaigns.
The strategic implementation of scaffold hopping requires quantitative assessment of both structural novelty and biological activity. The following experimental workflow establishes a standardized approach for evaluating this balance:
Different molecular descriptors capture complementary aspects of structural similarity and novelty [24]:
WHALES (Weighted Holistic Atom Localization and Entity Shape) descriptors: Encode information on geometric interatomic distances, molecular shape, and atomic properties in a holistic way. These descriptors are particularly effective for scaffold hopping from natural products to synthetic mimetics [24].
ECFP (Extended-Connectivity Fingerprints): Fragment-based representations that capture local atomic environments. While intuitive and widely used, they may be less effective for detecting similarities between structurally diverse natural products and synthetic compounds [24].
Pharmacophore descriptors: Focus on the spatial arrangement of functional groups essential for biological activity.
Table 2: Quantitative Data Analysis in Natural Product Scaffold Hopping
| Analysis Type | Statistical Methods | Application Example | Key Outcome Metrics |
|---|---|---|---|
| Dose-Response Analysis | ANOVA, Regression analysis | Natural product trials in rat models for neuroinflammation and memory deficits [43] | IC50, EC50, Hill coefficient |
| Correlation Analysis | Pearson/Spearman correlation | Assessing relationship between compound concentration and inflammatory marker levels [43] | Correlation coefficient (r), p-value |
| Longitudinal Analysis | Repeated measures ANOVA, mixed models | Monitoring disease progression in chronic disease models [43] | Slope of progression, treatment effect over time |
| Survival Analysis | Kaplan-Meier curves, log-rank test | Evaluating anti-cancer properties in xenograft models [43] | Hazard ratio, median survival time |
| Multivariate Analysis | PCA, PLS-DA | Accounting for age, sex, and housing conditions in in vivo tests [43] | Variable importance, clustering patterns |
Purpose: To compute holistic molecular descriptors that facilitate scaffold hopping from natural products to synthetic mimetics [24].
Materials:
Procedure:
Sw(j) = Σ[|δi| · (xi - xj)(xi - xj)T] / Σ|δi|
where δi is the partial charge of atom i, and xi represents 3D coordinates [24].ACM(i,j) = (xi - xj)T · Sw(j)-1 · (xi - xj)Validation: In prospective application, this approach achieved 35% success rate in identifying novel synthetic modulators of human cannabinoid receptors using natural cannabinoids as queries [24].
Purpose: To experimentally validate the biological activity of scaffold-hopped compounds derived from natural products.
Materials:
Procedure:
Functional Activity Assays:
In Vivo Efficacy Studies:
ADMET Profiling:
The transformation from morphine to tramadol represents a classic example of ring opening scaffold hopping [9] [18]. Morphine's rigid 'T'-shaped structure with multiple fused rings was modified by breaking six ring bonds, resulting in tramadol's more flexible structure. Despite significantly different 2D structures, 3D superposition conserves key pharmacophore features: the positively charged tertiary amine, aromatic ring, and hydroxyl group. This scaffold hop reduced potency but improved oral bioavailability and created a safer analgesic profile with reduced addiction potential [9].
A prospective study demonstrated successful scaffold hopping from natural cannabinoids to synthetic modulators using WHALES descriptors [24]. This holistic molecular representation captured pharmacophore and shape patterns, enabling identification of synthetic compounds with low-micromolar potency at human cannabinoid receptors CB1 and CB2. Of the selected compounds, 35% were experimentally confirmed as active, with five representing novel scaffolds not found in existing cannabinoid ligand databases [24].
The evolution of antihistamines illustrates multiple scaffold hopping strategies [9] [18]:
Table 3: Essential Research Reagents for Scaffold Hopping Studies
| Reagent/Category | Function/Application | Example Use Cases |
|---|---|---|
| WHALES Descriptors | Holistic molecular representation for scaffold hopping | Identifying synthetic mimetics of natural products [24] |
| Molecular Operating Environment (MOE) | Flexible Alignment program for 3D molecular superposition | Pharmacophore comparison between morphine and tramadol [9] |
| Liposomal Nanocarriers | Improve bioavailability of natural compounds | Delivery of antiviral natural compounds in in vivo models [43] |
| High-Throughput Screening Assays | Rapid activity assessment of compound libraries | Identifying active compounds from large collections [44] |
| qPCR Reagents | Quantify gene expression changes in disease models | Measuring inflammation-related gene expression in rat models [43] |
| MMFF94 Force Field | Energy minimization for 3D structure preparation | Conformational analysis for WHALES descriptor calculation [24] |
Successfully navigating the trade-off between structural novelty and biological activity requires a methodical approach that integrates computational design with experimental validation. The scaffold hopping classification system provides a strategic framework for selecting appropriate modification strategies based on project goals. Natural products continue to offer rich structural templates for drug discovery, with computational methods like WHALES descriptors enabling more effective translation of their complex architectures into synthetically accessible compounds with improved drug-like properties.
The experimental protocols and case studies presented herein demonstrate that strategic structural modifications, when guided by pharmacophore conservation principles and validated through rigorous biological testing, can yield novel chemotypes with maintained or improved therapeutic potential. As drug discovery faces increasing challenges, these scaffold hopping approaches will remain essential tools for expanding the structural diversity of chemical probes and therapeutic agents.
Natural products (NPs) are invaluable resources for drug discovery, characterized by their intricate scaffolds and diverse bioactivities [21]. However, their clinical application often faces challenges due to inherent complexities, including unfavorable ADMET properties, violation of Lipinski's rule of five, and low oral bioavailability [21]. Structural modification of NPs through scaffold hopping has emerged as a critical strategy to overcome these limitations while preserving biological activity [37]. This Application Note provides detailed protocols for addressing NP complexity and synthetic accessibility through computational scaffold hopping, enabling researchers to generate novel, patentable drug candidates with improved drug-like properties.
Table 1 summarizes key molecular properties that contribute to the complexity of natural products and their implications for drug discovery.
Table 1: Molecular Complexity Metrics of Natural Products and Implications for Drug Discovery
| Molecular Property | Typical NP Profile | Synthetic Compound Profile | Impact on Drug Discovery |
|---|---|---|---|
| sp³-hybridized carbon atoms | Higher proportion | Lower proportion | Increased 3D structural complexity, improved target selectivity [2] |
| Oxygen content | Increased | Decreased | Enhanced hydrogen bonding capacity, improved solubility [2] |
| Nitrogen/halogen content | Decreased | Increased | Reduced toxicity potential [2] |
| Molecular rigidity | Increased | Variable | Improved binding affinity, pre-organized bioactive conformation [2] |
| Lipophilicity (cLogP) | Lower | Variable | Improved solubility, reduced metabolic clearance [2] |
| Molecular weight | Often >500 Da | Typically <500 Da | Potential challenges with oral bioavailability despite successful examples [21] |
| Structural complexity | High (multiple stereocenters, macrocycles) | Moderate to low | Synthetic challenges, requires specialized strategies [21] |
Table 2 compares the performance of various computational tools for scaffold hopping, highlighting their advantages in generating synthetically accessible compounds.
Table 2: Performance Comparison of Scaffold Hopping Tools for Natural Product Optimization
| Tool/Platform | Methodology | Scaffold Library Size | Key Advantages for NP Optimization | Synthetic Accessibility Metrics |
|---|---|---|---|---|
| ChemBounce [37] | Fragment replacement, shape similarity | 3.2 million (ChEMBL-derived) | Open-source, high synthetic accessibility, ElectroShape similarity | Lower SAscore, higher QED values compared to commercial tools |
| FREED/FREED++ [21] | Reinforcement learning + GCN | Predefined fragments | Target-aware generation, explores diverse chemical space | Pharmacochemically acceptable molecules with high docking scores |
| D3FG [21] | Diffusion + SE(3)-equivariant GNN | Rigid functional groups | Preserves complex fragment structures, captures protein-ligand interactions | Maintains structural integrity of complex NP fragments |
| TACOGFN [21] | GFlowNet + graph transformer | 72 fragment types | Target pocket information integration, gradual fragment addition | Controlled growth ensures synthetic feasibility |
| DeepFrag [21] | 3D DCNN classification | Predefined fragments | Considers receptor pocket and full ligand context | Fragment-based approach ensures chemical authenticity |
Input Preparation and Validation
Scaffold Identification
Scaffold Replacement and Filtering
Output Analysis
Protein-Ligand Complex Preparation
Fragment-Based Optimization
Interaction-Preserving Generation
Multi-parameter Optimization
Table 3: Essential Research Reagents and Computational Tools for NP Scaffold Optimization
| Tool/Reagent | Type | Function in NP Scaffold Optimization | Example Sources/Platforms |
|---|---|---|---|
| ChEMBL-derived Fragment Library | Chemical Library | 3.2 million synthesis-validated scaffolds for replacement [37] | Curated from ChEMBL database [37] |
| COCONUT | NP Database | Over 400,000 non-redundant natural products for reference and inspiration [45] | Zenodo (open access) [45] |
| ScaffoldGraph | Software Library | Implements HierS algorithm for systematic scaffold decomposition [37] | Python package [37] |
| Open Drug Discovery Toolkit (ODDT) | Software Library | Provides ElectroShape implementation for shape similarity calculations [37] | Python library [37] |
| Predefined Fragment Sets | Chemical Building Blocks | 72+ fragment types for target-aware structure generation [21] | Various commercial and open sources [21] |
| SE(3)-Equivariant Neural Networks | Algorithmic Framework | Captures 3D spatial relationships in protein-ligand complexes [21] | D3FG, FRAME, other structure-based tools [21] |
| Reinforcement Learning Frameworks | Algorithmic Framework | Explores chemical space while optimizing multiple objectives [21] | FREED, FREED++ implementations [21] |
When working with structurally complex natural products such as macrocycles, high molecular weight peptides, or compounds with multiple stereocenters, specific considerations apply:
The protocols can be customized for different natural product classes:
The integration of computational scaffold hopping approaches with natural product research provides a powerful framework for addressing the inherent complexity and synthetic challenges of NPs. The protocols outlined in this Application Note enable systematic exploration of chemical space while maintaining biological relevance and ensuring synthetic feasibility. By leveraging these methodologies, researchers can accelerate the transformation of complex natural products into viable drug candidates with optimized properties, bridging the gap between nature's chemical diversity and modern pharmaceutical requirements.
Natural products (NPs) and their derivatives represent a cornerstone of drug discovery, particularly in therapeutic areas such as oncology and anti-infectives, accounting for approximately 30% of FDA-approved drugs over recent decades [21] [46]. However, their inherent structural complexity often leads to suboptimal pharmacokinetic (PK) profiles and toxicity, which present significant barriers to clinical application [21] [46]. Unmodified NPs frequently exhibit unfavorable ADMET (Absorption, Distribution, Metabolism, Excretion, and Toxicity) properties, including poor oral bioavailability, low specificity, and high toxicity [21]. Consequently, strategic structural modification of the core scaffold is not merely beneficial but essential to transform these naturally occurring compounds into viable therapeutic agents.
The process of "scaffold hopping" â the purposeful alteration of a molecule's core structure â has emerged as a powerful strategy to overcome these limitations while preserving or enhancing desired pharmacodynamic effects [46] [15]. This approach allows medicinal chemists to navigate the complex chemical space of natural products, systematically improving drug-like characteristics. The ultimate goal is to achieve an optimal balance within the "P3 properties model", which integrates Pharmacodynamic (efficacy), Physicochemical, and Pharmacokinetic/ADMET profiles through deliberate molecular design [46]. This Application Note provides detailed protocols and frameworks for employing core modifications to optimize PK and reduce toxicity within the context of modern natural product-based drug design.
Critically ill patients, such as those with sepsis or cancer, undergo significant physiological changes that profoundly impact drug PK. Understanding these alterations is critical for effective drug design and dosing. The major factors are summarized below [47]:
Natural products often contain structural motifs that, while biologically active, can cause off-target toxic effects. Key strategies to mitigate these include [46]:
The following tables consolidate key quantitative relationships and structural modifications that impact pharmacokinetics and toxicity.
Table 1: Impact of Physiological Alterations on Pharmacokinetic Parameters of Selected Drug Classes [47]
| Physiological Alteration | Impacted PK Parameter | Effect on Hydrophilic Drugs (e.g., β-lactams, Aminoglycosides) | Effect on Lipophilic Drugs (e.g., Azithromycin, Tigecycline) |
|---|---|---|---|
| Fluid Overload / Third-Spacing | Volume of Distribution (Vd) | Significantly Increased | Minimal Change |
| Hypoalbuminemia | Free Drug Concentration (Unbound Vd) | Increased for acidic, highly protein-bound drugs (e.g., Ceftriaxone) | Variable |
| Augmented Renal Clearance (ARC) | Clearance (CL) | Significantly Increased | Minimal Change |
| Acute Kidney Injury (AKI) | Clearance (CL) | Significantly Decreased | Minimal Change |
| Hepatic Dysfunction | Clearance (CL) | Minimal Change (for renally cleared) | Significantly Decreased |
Table 2: Representative Structural Modifications and Their Effects on Efficacy and Toxicity [46] [48]
| Natural Product / Compound | Core Modification | Impact on Efficacy | Impact on Toxicity / PK |
|---|---|---|---|
| Apratoxin A | Saturation of α,β-unsaturated carbonyl (to Apratoxin S4) | Retained potent cytotoxic activity | Reduced in vivo toxicity; eliminated nonspecific Michael addition [46] |
| Podophyllotoxin | Introduction of N-heterocyclic rings (e.g., β-Carboline) | Converted to Topo II inhibitor; retained anticancer activity | Reduced toxicity compared to original tubulin inhibitor [46] |
| Platinum Nanodrug (DACHPt/HANP) | Nanoformulation + Intralipid pre-treatment | Improved tumor targeting via EPR effect | Reduced liver/spleen/kidney accumulation; reduced organ toxicity [48] |
| Tanshinone I | Hybridization with piperidine scaffold | Enhanced anti-inflammatory activity (NLRP3 inhibition) | Improved drug-like properties [46] |
This protocol uses target-interaction-driven molecular generative models to optimize a natural product lead when the protein target is known [21] [49].
I. Research Reagent Solutions
II. Step-by-Step Methodology
This protocol describes a hybrid enzymatic-chemical approach for scaffold hopping in terpenoid natural products, moving beyond traditional total synthesis [14].
I. Research Reagent Solutions
II. Step-by-Step Methodology
This protocol assesses the in vivo performance of a novel nanodrug, using a platinum-based anti-cancer nanodrug as an example, and a strategy to reduce its RES-mediated toxicity [48].
I. Research Reagent Solutions
II. Step-by-Step Methodology
The following diagrams illustrate the core experimental and conceptual frameworks described in this note.
AI-Guided Optimization Workflow
Enzyme-Enabled Scaffold Hopping
Table 3: Key Research Reagent Solutions for Core Modification Studies
| Reagent / Resource | Function / Application | Example(s) |
|---|---|---|
| Engineered Cytochrome P450s | Enables site-selective oxidation of inert C-H bonds in complex NPs, creating handles for diversification. | Engineered P450 from B. megaterium for sclareolide C3 oxidation [14]. |
| Open-Source AI Models | Computational guidance for rational structure-based design via fragment replacement or scaffold hopping. | DeepFrag (group modification), ScaffoldGVAE (scaffold hopping) [21] [49]. |
| Intralipid 20% | An FDA-approved fat emulsion used to transiently blunt RES function, reducing nanodrug accumulation in off-target organs. | Pre-treatment to reduce liver/spleen/kidney toxicity of Pt-nanodrugs [48]. |
| Groebke-Blackburn-Bienaymé (GBB) Reaction | A multi-component reaction to rapidly generate complex, drug-like scaffolds (e.g., imidazo[1,2-a]pyridines) for scaffold hopping. | Used to develop novel molecular glues for the 14-3-3/ERα complex [15]. |
| Molecular Docking Software | Validates the predicted binding mode of modified compounds to the target protein. | AutoDock Vina, Glide, GOLD. |
| In-silico ADMET Platforms | Predicts pharmacokinetic and toxicity profiles of designed compounds prior to synthesis. | SwissADME, pkCSM, ProTox-II. |
Scaffold hopping, the practice of designing compounds with novel core structures that retain the biological activity of a parent molecule, is a crucial strategy in natural product-based drug design for overcoming limitations such as poor pharmacokinetics, toxicity, or intellectual property constraints [9] [18]. The core challenge lies in maintaining essential interactions with the biological target while significantly altering the molecular framework [50]. This application note details protocols for employing advanced computational models that strategically integrate multimodal dataâincluding 2D molecular graphs, 3D structural information, and protein target dataâto guide successful scaffold hops. By framing these methodologies within the context of natural product optimization, we provide researchers with a structured framework for discovering novel chemotypes with improved properties.
The following models represent the state-of-the-art in data-driven scaffold hopping, each utilizing different types of data and architectural principles.
Table 1: Performance Comparison of Deep Learning Models for Scaffold Hopping
| Model Name | Core Architecture | Primary Data Modalities | Reported Key Performance Metric | Applicable Scenario in NP Optimization |
|---|---|---|---|---|
| DeepHop [28] | Multimodal Transformer | Molecular Sequence, 2D Graph, 3D Conformer, Protein Sequence | ~70% of generated molecules had improved bioactivity, high 3D, but low 2D similarity (1.9x higher than other methods) | Target-informed hopping for NPs with known protein targets |
| DeLinker [51] | 3D-aware Graph Neural Network (GGNN) | 2D Molecular Graph, 3D Distance & Orientation | Designed 60% more molecules with high 3D similarity vs. baseline; 200% outperformance for linkers â¥5 atoms | Fragment linking in NP-derived partial structures |
| ScaffoldGVAE [17] | Graph Variational Autoencoder (GVAE) | 2D Molecular Graph (scaffold & side-chain views) | Generated validated novel inhibitors for LRRK2; superior scores on 4 scaffold-hopping specific metrics | Generating novel scaffolds while preserving NP-derived side chains |
| WHALES [32] | 3D Molecular Descriptors | 3D Molecular Conformation, Partial Atomic Charges | Outperformed 7 state-of-the-art descriptors in scaffold-hopping ability (SDA%) for 89% of 182 biological targets | Ligand-based hopping when 3D structure of NP is available |
This protocol is designed for scaffold hopping and linker design in natural product-derived fragments [51].
Research Reagent Solutions:
Procedure:
This protocol uses a supervised translation approach to generate scaffold-hopped molecules with improved bioactivity for a specific protein target [28].
Research Reagent Solutions:
Procedure:
X, Y) where molecule Y has significantly improved activity (e.g., pChEMBL value ⥠1) over X against the same target (Z), while having low 2D scaffold similarity (Tanimoto score ⤠0.6) but high 3D similarity (SC score ⥠0.6) [28].X.X.Z is encoded via a Protein Transformer.X into the output molecule Y conditioned on the target and structural information.X and the target protein Z into the trained model.Y predicted to have improved bioactivity against Z, a novel 2D scaffold, and a similar 3D profile to X.This protocol is for ligand-based scaffold hopping using advanced 3D molecular descriptors, useful when the target structure is unknown but a 3D conformation of the active natural product is available [32].
Research Reagent Solutions:
Procedure:
The following diagram illustrates the integrated workflow for multimodal scaffold hopping, synthesizing the key elements from the protocols above.
Diagram 1: Integrated Multimodal Scaffold Hopping Workflow. This diagram outlines the core process where a natural product input is decomposed into multiple data modalities that inform a generative model to produce novel, validated compound designs. The process is iterative, leveraging validation results to refine subsequent design cycles.
Table 2: Essential Research Reagents and Computational Tools
| Item/Tool Name | Category | Function in Scaffold Hopping | Example/Note |
|---|---|---|---|
| RDKit | Cheminformatics Library | Handles molecule I/O, SMILES processing, fingerprint generation, scaffold analysis, and 3D conformer generation. | Open-source; essential for preprocessing and feature extraction [28] [17]. |
| ChEMBL Database | Bioactivity Data | Provides curated bioactivity data for constructing training pairs and benchmarking. | Used in DeepHop, ScaffoldGVAE, and WHALES studies [28] [32] [17]. |
| PyTorch / TensorFlow | Deep Learning Framework | Provides the environment for building, training, and deploying complex generative models (GNNs, Transformers, VAEs). | Standard frameworks for implementing models like DeLinker and DeepHop [51] [28]. |
| ETKDG / OMEGA | 3D Conformer Generator | Generates realistic 3D molecular conformations required for 3D-aware models and descriptor calculations. | RDKit's ETKDG method; critical for DeLinker and WHALES inputs [51] [32]. |
| MMFF94 | Force Field | Used for energy minimization and geometry optimization of generated 3D structures. | Ensures generated conformers are energetically reasonable [28] [32]. |
| DFTB+ / Gasteiger-Marsili | Partial Charge Method | Calculates atomic partial charges, a key input for detailed 3D descriptors like WHALES. | DFTB+ for accuracy; Gasteiger for speed [32]. |
| ScaffoldGraph | Scaffold Analysis | Systematically decomposes molecules to extract hierarchical scaffolds for model training and analysis. | Used in ChemBounce and ScaffoldGVAE for scaffold definition [37] [17]. |
The drug discovery landscape is undergoing a profound transformation, moving from traditional, labor-intensive processes to modern, technology-driven approaches. Within natural product-based drug design and scaffold hopping research, understanding the quantitative performance differences between these paradigms is crucial for strategic research planning. Traditional drug discovery remains an arduous endeavor, typically requiring 10-15 years and exceeding $1-2 billion per approved therapy, with a dismally low success rate of fewer than 10% of candidates entering Phase I trials ultimately gaining approval [52]. Modern approaches, particularly those leveraging artificial intelligence (AI) and structured scaffold hopping, demonstrate potential to compress timelines to 12-18 months for early discovery phases and significantly improve efficiency [52] [53]. This application note provides a quantitative benchmark and detailed experimental protocols to guide researchers in adopting these advanced methodologies.
Table 1: Performance Metrics of Traditional vs. Modern Drug Discovery Methods
| Performance Metric | Traditional Methods | Modern AI-Driven Methods | Scaffold Hopping Approaches |
|---|---|---|---|
| Typical Discovery Timeline | 3-6 years (to preclinical candidate) [52] | 12-18 months (to preclinical candidate) [52] [53] | Varies (builds on existing molecules) [8] |
| Clinical Success Rate | <10% (Phase I to approval) [52] | Under evaluation; early data shows improved efficiency [52] [54] | Often higher (starting from validated starting points) [8] |
| Key Supporting Technologies | High-Throughput Screening (HTS), Structure-Based Design [55] | Machine Learning (ML), Deep Learning (DL), Multi-modal Data Fusion [52] [53] [56] | Computational Design (e.g., MORPH software), Advanced Synthetic Chemistry [8] |
| Representative Case Study | Conventional HTS for lead identification | Insilico Medicine's AI-generated anti-fibrotic (18 months to PCC) [52] | Roxadustat analogs developed via heterocycle replacement [8] |
Table 2: Analysis of AI Methodologies in Modern Drug Discovery (Analysis of 173 Studies, 2015-2025) [52]
| AI Methodology | Adoption Rate | Primary Application in Discovery |
|---|---|---|
| Machine Learning (ML) | 40.9% | Target identification, compound potency/ADMET prediction [52] [56] |
| Molecular Modeling & Simulation (MMS) | 20.7% | Molecular docking, binding affinity prediction, protein-ligand interactions [52] |
| Deep Learning (DL) | 10.3% | De novo molecular design, analysis of complex data (e.g., histopathology images) [52] [56] |
| Reinforcement Learning (RL) | Not quantified | Multi-objective optimization in generative chemistry [53] |
| Natural Language Processing (NLP) | Not quantified | Mining scientific literature and patents for target identification [53] [56] |
This protocol describes the classical approach to isolating active compounds from natural sources, which serves as a foundational starting point for many scaffold-hopping campaigns.
3.1.1 Materials and Reagents
3.1.2 Step-by-Step Procedure
This modern protocol uses AI to hypothesize and prioritize molecular targets for a natural product with a known phenotypic effect but an unknown mechanism of action (MoA).
3.2.1 Materials and Software
3.2.2 Step-by-Step Procedure
This protocol outlines a systematic approach, enhanced by computational chemistry, to generate novel chemical entities based on the core structure of an active natural product.
3.3.1 Materials and Software
3.3.2 Step-by-Step Procedure
Scaffold Hopping from a Natural Product Lead
AI-Driven Target Deconvolution for Natural Products
Table 3: Essential Research Reagents and Platforms for Modern Discovery
| Tool / Reagent | Function / Application | Key Characteristic / Benefit |
|---|---|---|
| PandaOmics (Insilico Medicine) [53] | AI-powered target discovery; identifies and prioritizes novel therapeutic targets from multi-omics data. | Leverages knowledge graphs from 1.9 trillion data points and NLP on 40M+ documents for holistic target assessment. |
| Chemistry42 (Insilico Medicine) [53] | Generative chemistry AI platform for de novo molecular design and lead optimization. | Uses generative adversarial networks (GANs) and reinforcement learning to balance potency, selectivity, and ADMET. |
| Recursion OS Platform [53] | Maps trillions of biological relationships using proprietary data and AI models (e.g., Phenom-2, MolPhenix). | Integrates automated wet-lab biology with AI computational models for phenotypic drug discovery. |
| Organ-on-a-Chip Models [57] | Microphysiological systems for human-relevant ADMET and efficacy testing (a key New Approach Methodology - NAM). | Provides more predictive human toxicity and PK data than animal models, reducing late-stage attrition. |
| MORPH Software [8] | Computational tool for systematic scaffold hopping and aromatic ring modification in 3D ligand models. | Enables complex scaffold-hopping strategies to generate novel, patentable chemical space from known leads. |
| CONVERGE (Verge Genomics) [53] | Closed-loop ML platform for target discovery, prioritizing targets using large-scale human-derived biological data. | Focuses on human clinical genomics data to improve translational relevance, particularly in neurology. |
In natural product-based drug design, the process of scaffold hoppingâthe identification of structurally novel compounds with similar biological activity to a parent natural productâis a crucial strategy for overcoming challenges associated with natural products, such as structural complexity, poor solubility, or limited availability [9] [24]. The success of these campaigns depends critically on the rigorous application of key validation metrics to ensure that newly designed compounds maintain the desired biological activity while achieving improved synthetic accessibility and drug-like properties. This document details the essential metricsâ2D/3D similarity, potency, and selectivityâframed within the context of natural product-inspired scaffold hopping, providing researchers with structured protocols for their application in early drug discovery.
Molecular similarity analysis forms the foundational pillar of scaffold hopping, operating on the principle that structurally similar molecules are likely to exhibit similar biological activities [58]. For natural products, this involves comparing novel synthetic mimetics against their complex natural templates.
2D similarity methods evaluate molecular structure based on topological descriptors, without considering three-dimensional conformation.
Fingerprint-Based Methods: These encode molecular structures as bit strings representing the presence or absence of specific structural features.
Protocol: Performing a 2D Similarity-Based Virtual Screen
3D similarity methods compare molecules based on their shape and the spatial arrangement of pharmacophoric features, which is critical for scaffold hopping where core structures differ significantly in 2D topology [61] [24].
Table 1: Key Metrics for 2D and 3D Molecular Similarity Methods
| Method | Descriptor Type | Key Metric | Typical Use Case in Scaffold Hopping | Advantages |
|---|---|---|---|---|
| 2D Similarity | ECFP Fingerprints | Tanimoto Coefficient | Rapid pre-filtering of large libraries; hit expansion [59]. | Computationally fast; intuitive for chemists [59]. |
| 3D Shape-Based | Molecular Shape/Features | TanimotoCombo Score | Identifying mimetics of complex natural products with different 2D structures [61]. | Captures functional similarity beyond topology [61]. |
| 3D Holistic | WHALES Descriptors | Euclidean Distance in WHALES space | Scaffold hopping from complex natural products to synthetically accessible leads [24]. | Integrates shape and pharmacophore features; robust to conformational change [24]. |
Combining 2D and 3D methods in a sequential or parallel workflow increases the success rate and novelty of identified hits [59].
Diagram 1: Integrated 2D/3D similarity screening workflow for identifying natural product mimetics, combining the speed of 2D methods with the functional insight of 3D approaches [59] [24].
Beyond mere structural similarity, confirming a compound's functional profile through potency and selectivity is paramount in transitioning from a initial hit to a viable lead.
Potency quantifies the concentration of a compound required to produce a given biological effect. Common metrics include ICâ â (half-maximal inhibitory concentration), ECâ â (half-maximal effective concentration), and Káµ¢ (inhibition constant). In scaffold hopping, the goal is to achieve low micromolar to nanomolar potency against the primary target, comparable to or better than the original natural product [62].
Selectivity ensures that a compound exerts its primary effect on the intended target without causing significant off-target effects, which can lead to adverse reactions. Traditional metrics like the Gini coefficient or selectivity entropy quantify the overall narrowness of a compound's bioactivity spectrum but do not focus on a specific target of interest [62].
For scaffold hopping, a target-specific selectivity score is more appropriate. This approach frames selectivity as a multi-objective optimization problem, seeking compounds that simultaneously demonstrate [62]:
Table 2: Key Metrics for Assessing Potency and Selectivity in Scaffold Hopping
| Metric | Definition | Interpretation in Scaffold Hopping |
|---|---|---|
| ICâ â / Káµ¢ | Concentration inhibiting 50% of target activity / Inhibition constant. | Primary measure of compound potency against the intended target. Aim for nM to low µM range. |
| Global Relative Potency (G) | ( G{ci,tj} = K{ci,tj} - \text{mean}(B{ci} \setminus {K{ci,t_j}}) ) [62] | Measures a compound's ((ci)) potency for a target ((tj)) relative to its mean potency against all other targets. A high value indicates high selectivity. |
| Local Relative Potency (L) | ( L{ci,tj} = K{ci,tj} - \text{mean}(B{ci,hNN(t_j)}) ) [62] | Measures a compound's potency for a target relative to its potency against the h-most similar off-targets. More sensitive to activity cliffs. |
| Partition Index | Fraction of total binding strength directed toward a reference target [62]. | Quantifies the fraction of a compound's overall binding energy devoted to the primary target. |
Table 3: Experimental Profiling Techniques for Validation
| Assay Type | Information Provided | Application in Scaffold Hopping |
|---|---|---|
| In vitro Binding Assay | Direct measurement of binding affinity (Kd, Ki) to a purified target protein. | Confirm direct target engagement and primary potency. |
| Cell-Based Functional Assay | Measures functional consequences (e.g., inhibition of cell growth, second messenger production) in a live cell system (ICâ â, ECâ â). | Validates activity in a more physiologically relevant context. |
| Kinase/GPCR Panel Screening | Broad profiling of activity against dozens to hundreds of related targets in a protein family. | Essential for quantifying polypharmacology and identifying potential off-target effects early. |
Diagram 2: The iterative validation cycle for a scaffold-hopped compound, moving from initial design to a confirmed potent and selective lead candidate through rigorous biochemical profiling [62].
The following table details key computational and experimental resources essential for conducting research in natural product-based scaffold hopping and validation.
Table 4: Essential Research Tools and Reagents for Scaffold Hopping and Validation
| Tool/Reagent | Function | Application Note |
|---|---|---|
| ChEMBL Database | Public repository of bioactive molecules with drug-like properties and curated bioactivity data [60] [61]. | Serves as a primary source for building reference datasets of known active compounds for similarity-based target prediction [60]. |
| ZINC20/COCONUT | Freely available databases of commercial (ZINC20) and natural (COCONUT) compounds for virtual screening [60]. | Used as source libraries for finding potential scaffold hops or for building a knowledge base of non-complex small molecules for target prediction [60] [61]. |
| ROCS (OpenEye) | A leading 3D shape-based molecular superposition tool and screening engine [61]. | Used for aligning query natural products to potential mimetics based on 3D shape and chemical features, quantified by the TanimotoCombo score [61]. |
| CTAPred | An open-source, command-line tool for predicting protein targets of natural products using similarity-based searches [60]. | Addresses the challenge of limited bioactivity data for NPs by using a focused compound-target activity reference dataset [60]. |
| Kinase/GPCR Profiling Services | Commercial services (e.g., Eurofins, Reaction Biology) that offer high-throughput screening against panels of kinase or GPCR targets. | Critical for experimentally determining the selectivity profile of a scaffold-hopped compound against a therapeutically relevant target family [62]. |
| WHALES Descriptors | A holistic molecular representation capturing pharmacophore, shape, and partial charge patterns [24]. | Enables scaffold hopping from complex natural products to isofunctional synthetic compounds by capturing functionally relevant molecular features [24]. |
Scaffold hopping is a foundational strategy in modern medicinal chemistry, defined as the structural modification of the core molecular framework of a known bioactive compound to generate novel, patentable molecules with potentially improved properties [8] [10]. Within the context of natural product-based drug discovery, this approach is invaluable for optimizing the complex scaffolds of natural leads, addressing common issues such as poor solubility, metabolic instability, toxicity, and limited intellectual property (IP) space [24] [21] [63]. By systematically exploring alternative chemotypes that retain the desired biological activity, scaffold hopping bridges the gap between biologically validated natural product starting points and drug-like candidates better suited for clinical development. This analysis provides a comparative examination of contemporary scaffold hopping strategies, software tools, and their practical application in a research setting.
Scaffold hopping encompasses a spectrum of structural modifications, which can be systematically classified into distinct degrees based on the type of alteration performed on the parent molecule's core [8] [10]. This classification aids medicinal chemists in rational design and communication.
Table 1: Degrees of Scaffold Hopping and Their Characteristics
| Degree | Type of Modification | Description | Key Applications |
|---|---|---|---|
| 1° (Heterocycle Replacement) | Substitution, addition, or removal of heteroatoms in a core ring; replacement of one heterocycle with a similar one [8] [10]. | The simplest form; retains the spatial arrangement of the pharmacophore and adjacent groups [10]. | Tuning physicochemical properties, optimizing PK profile, identifying key ligand-target interactions [10]. |
| 2° (Ring Opening or Closure) | Converting a cyclic moiety into an acyclic chain, or vice versa [8]. | Alters molecular flexibility and conformational entropy without drastically changing the core topology. | Modulating metabolic stability and conformational freedom [8]. |
| 3° (Peptidomimetics) | Replacing a peptide backbone with a non-peptide scaffold [8]. | Aims to mimic the spatial orientation of key pharmacophoric elements of a peptide. | Improving oral bioavailability and metabolic stability of peptide leads [8]. |
| 4° (Topology-based Alterations) | Introducing global molecular shape changes, such as fusing or dissociating rings [8]. | The most sophisticated degree; involves significant redesign of the core scaffold topology. | Exploring novel IP space, addressing complex property issues [8]. |
The logical relationship and strategic progression between these degrees and their objectives can be visualized in the following workflow.
The implementation of scaffold hopping strategies is powerfully enabled by a suite of specialized software tools. These platforms use various algorithms to propose novel molecular structures that meet specific design criteria.
Table 2: Comparative Analysis of Key Scaffold Hopping Software Tools
| Software Tool | Core Methodology | Key Features | Applicability in Natural Product Optimization |
|---|---|---|---|
| Spark [64] | Electrostatic and shape similarity for bioisosteric replacement. | User-friendly wizards for lead discovery/optimization; multi-parametric optimization (LogP, TPSA, MW) [64]. | Replacing complex, synthetically challenging NP fragments with simpler, isofunctional motifs. |
| ChemBounce [65] | Fragment replacement using a curated library of >3 million fragments from ChEMBL. | Evaluates Tanimoto and electron shape similarities; focuses on high synthetic accessibility [65]. | Systematic exploration of synthetic mimetics for natural product scaffolds. |
| SeeSAR [66] | 3D structure-based visualization and design with HYDE affinity estimation and FastGrow algorithm. | Intuitive visual interface; on-the-fly affinity estimation; ReCore tool for 3D-driven re-scaffolding [66]. | Visual, structure-guided optimization of NP analogs within a protein binding site. |
| WHALES [24] | Holistic molecular representation (pharmacophore, shape, partial charge). | Ligand-based; does not require target structure; effective for complex NP mimetic design [24]. | Scaffold hopping from complex NPs to synthetically accessible, isofunctional compounds when structural data is lacking. |
| AnchorQuery [15] | Pharmacophore-based screening of a vast, synthesizable MCR chemistry library. | Links to readily synthesizable scaffolds (e.g., GBB-3CR); prioritizes synthetic feasibility [15]. | Rapid generation of novel, drug-like, and synthetically accessible scaffolds inspired by an NP starting point. |
Aurones, natural "golden flavonoids," possess promising bioactivities but are hampered by poor solubility and metabolic stability [63]. A scaffold-hopping approach replaced the benzofuranone core's oxygen with nitrogen, yielding azaaurones (indolin-3-ones).
A scaffold-hopping approach was used to develop non-covalent molecular glues stabilizing the 14-3-3/ERα protein-protein interaction (PPI), a target in breast cancer [15].
Scaffold hopping has been pivotal in addressing drug-resistant tuberculosis (TB), by creating new chemotypes that circumvent existing resistance mechanisms [10].
The following diagram illustrates the general decision-making workflow that integrates these computational and experimental methods.
Successful scaffold hopping campaigns rely on a combination of computational and experimental resources.
Table 3: Key Research Reagent Solutions for Scaffold Hopping
| Reagent / Resource | Category | Function in Scaffold Hopping |
|---|---|---|
| Fragment Libraries (e.g., SeeSAR's MedChem Set, Hinge Binder Set) [66] | Computational/Chemical | Pre-curated sets of molecular fragments used by software like SeeSAR and Spark for bioisosteric replacement, growing, and linking to generate novel scaffold ideas. |
| Multi-Component Reaction (MCR) Libraries (e.g., in AnchorQuery) [15] | Chemical/Synthetic | Virtual libraries of readily synthesizable scaffolds from MCR chemistry (e.g., GBB-3CR), enabling rapid identification of synthetically feasible novel cores. |
| ChEMBL Database [65] | Data | A manually curated database of bioactive molecules with drug-like properties. Serves as a key source for fragment libraries (e.g., in ChemBounce) and for validating the potential bioactivity of novel scaffolds. |
| Protein Data Bank (PDB) [10] | Data | A repository of 3D structural data of proteins and protein-ligand complexes. Essential for structure-based scaffold hopping, providing coordinates for docking and structure analysis. |
| Groebke-Blackburn-Bienaymé (GBB) Reaction Components [15] | Synthetic Chemistry | Aldehydes, 2-aminopyridines, and isocyanides used to synthesize the imidazo[1,2-a]pyridine scaffoldâa privileged, drug-like core identified via scaffold hopping for PPIs. |
| Sonogashira Coupling Reagents [63] | Synthetic Chemistry | Palladium catalysts (e.g., Pd(PPhâ)â) and co-catalysts used in one-pot syntheses of scaffold-hopped cores like azaaurones, enabling efficient construction of the novel heterocycle. |
Within the framework of natural product-based drug design, the dual challenges of enhancing therapeutic efficacy and establishing robust intellectual property (IP) positions are paramount. Scaffold hopping, a medicinal chemistry strategy that modifies the core molecular structure of a known bioactive compound, has emerged as a powerful approach to address both challenges simultaneously [10]. This method leverages the validated bioactivity of existing molecules, often derived from natural products, while generating novel chemical entities with improved properties [67]. The ensuing application note details a recent, impactful case study where a scaffold-hopping strategy, powered by computational design and multi-component reaction chemistry, successfully led to a new class of molecular glues with enhanced efficacy and a strong basis for patentability.
The stabilization of protein-protein interactions (PPIs) represents a promising therapeutic strategy, particularly for targets considered "undruggable" by conventional means. This case focuses on the native interaction between the scaffolding protein 14-3-3 and the transcription factor Estrogen Receptor α (ERα). Stabilizing this PPI can inhibit ERα's transcriptional activity, offering a novel approach to treating ERα-positive breast cancer, especially in cases of acquired endocrine resistance [15]. The disordered C-terminus of ERα, which is recognized by 14-3-3, presents a challenging but druggable composite surface.
The discovery campaign began with a previously developed molecular glue, compound 127 (PDB: 8ALW), which was known to bind cooperatively at the 14-3-3Ï/ERα interface [15]. While effective, this initial compound presented opportunities for optimization. A scaffold-hopping approach was employed to design a novel, more rigid, and drug-like core structure.
The computational workflow utilized AnchorQuery software, which performs pharmacophore-based screening of a vast virtual library of over 31 million compounds synthesizable via one-step multi-component reactions (MCRs) [15]. The process was guided by the crystallographic binding pose of compound 127:
This in silico screen identified the Groebke-Blackburn-Bienaymé (GBB) three-component reaction as the optimal pathway to generate a novel imidazo[1,2-a]pyridine scaffold. Docking poses revealed that the proposed GBB compounds exhibited significant shape complementarity and an nearly identical three-dimensional conformation to the original compound 127, while being inherently more rigid and drug-like [15].
The following table summarizes the biophysical and cellular efficacy data for the original lead compound and the optimized analogs derived from the scaffold-hopping campaign. The data demonstrates the success of the approach in generating compounds with potent stabilization activity.
Table 1: Efficacy Data of Original and Scaffold-Hopped Molecular Glues
| Compound | Core Scaffold | SPR Affinity (KD, μM) | TR-FRET Stabilization (EC50, μM) | Cellular NanoBRET (EC50, μM) | Key Improvements |
|---|---|---|---|---|---|
| Compound 127 (Original) | Flexible, non-MCR | Not Specified | Not Specified | Not Specified | Original covalent binder |
| GBB Analog 1 | Imidazo[1,2-a]pyridine | < 10 | Low micromolar | Low micromolar | Non-covalent, rigid, drug-like |
| GBB Analog 2 (Most Potent) | Imidazo[1,2-a]pyridine | < 10 | Low micromolar | Low micromolar | Improved synthetic accessibility, optimal rigidity |
Data synthesized from [15]. SPR: Surface Plasmon Resonance; TR-FRET: Time-Resolved Förster Resonance Energy Transfer; NanoBRET: Bioluminescence Resonance Energy Transfer assay in live cells.
The scaffold-hopping strategy directly enhanced the patentability of the resulting compounds. The generation of a novel imidazo[1,2-a]pyridine core, distinct from the original chemical matter, satisfies the key requirement of novelty [10] [68]. Furthermore, the application of a specific and less conventional synthetic routeâthe GBB multi-component reactionâstrengthens the inventive step (non-obviousness) of the approach.
The new scaffold also provides a robust platform for rapid Structure-Activity Relationship (SAR) expansion through multiple points of variation, enabling the generation of a broad patent estate that protects not only the specific lead compounds but also a wide range of analogs [15] [68]. This "patent thicket" strategy is a critical lifecycle management tool, extending market exclusivity and protecting the significant investment required for drug development [68]. The imidazo[1,2-a]pyridine scaffold is a recognized privileged structure in medicinal chemistry, found in several clinical candidates and marketed drugs (e.g., zolpidem), which de-risks the development pathway and enhances the commercial attractiveness of the program [15].
Purpose: To computationally identify novel, synthetically accessible scaffolds that mimic the binding mode of a known active compound.
Materials:
Methodology:
Purpose: To experimentally validate and quantify the stabilization of the 14-3-3/ERα complex by novel compounds.
Materials:
Methodology:
Table 2: Essential Reagents and Tools for Scaffold Hopping and PPI Stabilization Research
| Reagent / Tool | Function / Description | Application in Case Study |
|---|---|---|
| AnchorQuery Software | Pharmacophore-based virtual screening platform for MCR chemistry. | Identified the novel GBB-based imidazo[1,2-a]pyridine scaffold. |
| Groebke-Blackburn-Bienaymé (GBB) Chemistry | A three-component reaction between an aldehyde, 2-aminopyridine, and an isocyanide. | Enabled the rapid, divergent synthesis of the novel scaffold with multiple points for variation. |
| TR-FRET Assay Kits | Homogeneous assay for detecting biomolecular interactions in a microplate format. | Quantified the stabilization of the 14-3-3/ERα complex in a biochemical setting. |
| NanoBRET Assay System | Cell-based assay to monitor PPIs in live cells using bioluminescence energy transfer. | Confirmed target engagement and PPI stabilization in a physiologically relevant, cellular environment. |
| Crystallography Platform | Determines the 3D atomic structure of protein-ligand complexes. | Provided the critical binding pose of the original lead (compound 127) to guide computational design and validated the binding mode of optimized analogs. |
Scaffold hopping, powered by the rich structural diversity of natural products, has evolved from a concept reliant on medicinal chemistry intuition to a data-driven discipline supercharged by AI. The integration of holistic molecular representations and deep learning models enables a more efficient exploration of chemical space, leading to the discovery of novel, isofunctional compounds with improved pharmacological profiles. Successful outcomes depend on a nuanced balanceâmaintaining critical pharmacophoric elements and 3D shape while achieving significant 2D structural novelty. As these computational methodologies continue to mature, their ability to generalize across new targets and generate synthetically accessible leads will be crucial. The future of natural product-based drug design lies in the seamless fusion of these advanced in silico techniques with experimental validation, accelerating the delivery of new therapeutics for complex diseases like cancer and drug-resistant infections.