This article provides a comprehensive cheminformatic comparison of natural products (NPs) and synthetic compounds (SCs), crucial sources for small-molecule drug discovery.
This article provides a comprehensive cheminformatic comparison of natural products (NPs) and synthetic compounds (SCs), crucial sources for small-molecule drug discovery. Leveraging recent studies and large-scale data analyses, we explore the foundational structural and physicochemical differences between these compound classes. We delve into methodological approaches for library design and fragmentation, address key challenges in NP research such as synthetic accessibility and regulatory hurdles, and validate findings through comparative analysis of approved drugs. The analysis confirms that NPs provide greater structural diversity and complexity, occupying a broader and distinct region of chemical space. This work synthesizes key insights for researchers and drug development professionals aiming to leverage the unique advantages of both NPs and SCs in modern drug discovery pipelines.
Within modern drug discovery, the strategic choice between natural product (NP)-inspired compounds and purely synthetic molecules is paramount. Cheminformatic analyses provide a powerful, data-driven approach to inform this decision by quantifying critical structural and physicochemical differences. Among the numerous molecular descriptors available, three have consistently proven fundamental for profiling compound libraries: Molecular Weight (MW), the octanol/water partition coefficient (LogP), and the Topological Polar Surface Area (TPSA). These properties are central to predicting a molecule's behavior in a biological system, influencing its absorption, distribution, metabolism, and excretion (ADME) profile [1] [2].
This guide provides an objective, data-centric comparison of these key properties between NPs, NP-based drugs, and synthetic drugs. It is structured to serve researchers and drug development professionals by presenting consolidated quantitative data, detailing standard methodological protocols for such analyses, and visualizing the essential workflow. The overarching thesis is that NPs and synthetic compounds inhabit distinct, yet complementary, regions of chemical space, and a deliberate integration of their unique features can be a productive strategy for addressing challenging therapeutic targets.
Systematic analyses of approved drugs reveal consistent and significant differences in the physicochemical profiles of natural product-based drugs compared to their purely synthetic counterparts. The data below consolidates findings from cheminformatic studies to provide a clear, quantitative comparison.
Table 1: Comparative Analysis of Key Physicochemical Properties in Approved Drugs
| Compound Category | Molecular Weight (MW) | LogP (or ALOGPs) | TPSA | Fraction sp3 (Fsp3) | H-Bond Donors (HBD) | H-Bond Acceptors (HBA) |
|---|---|---|---|---|---|---|
| Natural Product Drugs (N) | 611 | 1.96 | 196 | 0.71 | 5.9 | 10.1 |
| Natural Product-Derived Drugs (ND) | 757 | 1.82 | 250 | 0.59 | 7.0 | 11.5 |
| Top-Selling Synthetic Drugs (2018-S) | 444 | 2.83 | 95 | 0.33 | 1.9 | 5.1 |
| All NP-Based Drugs (N & ND) | 673 | 2.01 | 211 | 0.58 | 5.8 | 10.1 |
Data derived from Newman and Cragg's compilations and subsequent cheminformatic analyses [3] [4].
The comparative data presented above is generated through standardized cheminformatic workflows. The following section outlines the core methodologies employed in such analyses.
The foundation of any robust comparative analysis is a carefully curated dataset.
Once a clean dataset is established, molecular descriptors are computed programmatically.
ChemmineR and rcdk) are also widely used [6].The final stage involves interpreting the calculated data.
Diagram: Cheminformatic Workflow for Property Comparison
To perform the analyses described, researchers rely on a combination of data resources, software libraries, and computational tools.
Table 2: Essential Research Reagent Solutions for Cheminformatic Profiling
| Tool / Resource Name | Type | Primary Function | Access |
|---|---|---|---|
| RDKit | Software Library | Open-source cheminformatics for descriptor calculation, scaffold analysis, and SMILES processing. | Open-Source |
| CDK (Chemistry Development Kit) | Software Library | Open-source library for structural chemo-informatics and bioinformatics. | Open-Source |
| R / ChemmineR | Programming Language / Package | Statistical computing and graphics with specialized functions for analyzing compound datasets. | Open-Source |
| PubChem | Database | Public repository of chemical substances and their biological activities, a key data source. | Free |
| ChEMBL | Database | Manually curated database of bioactive molecules with drug-like properties. | Free |
| UNPD (Universal Natural Products Database) | Database | Large, curated collection of natural product structures for virtual screening. | Free (Historical) |
| ZINC | Database | Commercial database of compounds for virtual screening, includes some purchasable NPs. | Free |
| Molecular Descriptor Calculator (e.g., ChemToolsHub) | Web Tool | Online calculator for quick determination of key properties from a SMILES string. | Web Interface |
These resources form the backbone of a modern cheminformatics workflow, enabling everything from data acquisition and curation to complex property analysis and machine learning [6] [2] [8].
The comparative data unequivocally demonstrates that natural product-based drugs and synthetic drugs occupy distinct physicochemical territories. NP-based drugs are typically larger, more polar, more three-dimensional, and richer in stereochemical complexity. In contrast, synthetic drugs in this analysis are smaller, more lipophilic, and have flatter architectures.
This divergence is not a matter of superiority but of strategic complementarity. The unique chemical space occupied by NPs makes them invaluable starting points for addressing challenging drug targets, such as protein-protein interactions, that are often intractable for conventional synthetic compounds [3] [4]. The continued high prevalence of NP-inspired structures among top-selling drugs underscores their enduring impact [3]. Therefore, the most productive path forward in drug discovery lies in a synergistic approachâleveraging the rich structural and physicochemical diversity of natural products while employing synthetic chemistry and computational methods to optimize their properties for drug development.
The pursuit of effective small-molecule therapeutics is fundamentally guided by the principles of molecular structure and its relationship to biological activity. Within this domain, a compelling dichotomy has emerged between natural products (NPs) and synthetic compounds, particularly regarding their three-dimensional structural complexity. Natural products, evolved to interact with biological systems, often exhibit rich stereochemistry and structural saturation, while synthetic libraries, frequently designed around traditional rules like Lipinski's Rule of Five, have historically favored flatter, more planar architectures [9] [10]. This guide provides a comparative analysis of two critical metrics for quantifying this three-dimensionality: the fraction of sp3 hybridized carbon atoms (Fsp3) and stereochemical content. We will objectively evaluate their distribution across compound classes, detail protocols for their assessment, and present data linking these parameters to clinical success, providing drug development professionals with a framework for leveraging structural complexity in design.
Table 1: Comparative Summary of Key 3D Complexity Descriptors
| Descriptor | Definition | Calculation | Interpretation | Primary Data Source |
|---|---|---|---|---|
| Fsp3 | Fraction of sp3-hybridized carbons | Fsp3 = (Number of sp3 carbons) / (Total carbon count) | Higher values (>0.42) indicate greater saturation and 3D character; correlates with solubility and clinical success [10]. | Computed from 2D molecular structure. |
| Stereocenter Count | Number of atoms with non-superimposable mirror images. | Identified through structural analysis or chiral perception algorithms. | Direct measure of stereochemical complexity; influences binding specificity and off-target effects [11]. | Computed from 2D/3D molecular structure. |
| 3D Shape (PMI) | Principal Moment of Inertia; describes molecular shape. | Plotted on a normalized triangle from rod-like to disk-like to sphere-like. | Quantifies the overall three-dimensional shape, distinct from atomic hybridization [12]. | Requires generation of a 3D molecular conformation. |
Quantitative analyses consistently reveal a significant structural gap between natural and synthetic molecules. A cheminformatic analysis of approved drugs showed that natural (N) and natural-derived (ND) drugs possess higher Fsp3 values (0.71 and 0.59, respectively) compared to top-selling synthetic drugs (S), which had an Fsp3 of only 0.33 [3]. This trend is also evident in screening libraries; an analysis of nearly 390,000 compounds found that natural products demonstrate "much greater variability in terms of molecular complexity (most evidently shown by Fsp3)" [9]. Furthermore, approximately 84% of marketed drugs meet a criterion of Fsp3 ⥠0.42, highlighting its relevance to successful drug development [10].
The difference in stereochemical content is equally pronounced. The same drug analysis found that natural product-based drugs had a stereocenter count normalized by molecular weight (nStMW) that was "2- to 6-fold higher" than that of purely synthetic drugs [9]. This is not merely a structural curiosity; it has direct biological consequences. A large-scale study on over 1 million compounds found that roughly 40% of spatial isomer pairs show distinct bioactivities [11]. This underscores the critical importance of stereochemistry, as different stereoisomers of the same molecule can have vastly different therapeutic and toxicological profiles, as seen with drugs like Citalopram and Penicillamine [11].
Table 2: Average Physicochemical Properties by Drug Category (Adapted from [3])
| Category | Number of Compounds | Molecular Weight (MW) | Fsp3 | Stereocenter Count (implied) | Rotatable Bonds (Rot) |
|---|---|---|---|---|---|
| Natural Product Drugs (N) | 77 | 611 | 0.71 | High | 11.0 |
| Natural Product-Derived Drugs (ND) | 344 | 757 | 0.59 | High | 16.2 |
| Top 40 Drugs in 2018: Synthetic (2018-S) | 15 | 444 | 0.33 | Low | 6.5 |
| Top 40 Drugs in 2006: Synthetic (2006-S) | 27 | 355 | 0.33 | Low | 5.4 |
| Diversity-Oriented Synthesis Probes (DOS) | 10 | 552 | 0.38 | Low | 4.9 |
This protocol details the steps to compute key descriptors for a set of compounds using open-source tools, as exemplified in methodologies from the search results [9] [13] [14].
This advanced protocol, based on the development of Signaturizers3D, uses 3D conformations to create bioactivity descriptors that distinguish stereoisomers [11].
The influence of Fsp3 and stereochemistry extends from initial screening to clinical performance. Fsp3 has been shown to be a valuable parameter for guiding hit screening and lead optimization. For instance, in the discovery of a RORγ inhibitor, increasing the Fsp3 and Ligand Efficiency (LE) of the lead compound resulted in a 50-fold increase in potency and eliminated time-dependent inhibition of CYP450 [10]. This aligns with the broader observation that increased saturation, measured by Fsp3 and the number of chiral centers, correlates with a higher clinical success rate, potentially due to improved solubility and the ability of more 3D molecules to specifically occupy target space [10].
The data also shows a compelling trend in the market. Among the top 40 best-selling brand-name drugs, the proportion based on natural products increased dramatically from 35% in 2006 to 70% in 2018 [3]. Given that natural product-based drugs consistently exhibit higher Fsp3 and stereochemical content, this shift suggests the industry is increasingly benefiting from the complex chemical space occupied by these compounds. Furthermore, macrocycles, a class of molecules known for high three-dimensionality, were found to occupy "distinctive and relatively underpopulated regions of chemical space," highlighting their potential for targeting challenging binding sites [3].
Table 3: Key Software and Databases for 3D Cheminformatic Analysis
| Tool Name | Type | Key Function | Access |
|---|---|---|---|
| RDKit | Cheminformatics Library | Core cheminformatics, descriptor calculation (Fsp3, stereocenters), 3D conformer generation (ETKDG), and library enumeration [13]. | Open Source |
| FAF-Drugs3 | Web Server | Compound property calculation and filtering. Computes physicochemical rules, Fsp3, and identifies structural alerts and PAINS [14]. | Free Web Server |
| KNIME | Workflow Platform | Data analytics and visual programming for chemistry. Used for library enumeration based on generic reactions and data analysis [13]. | Free & Commercial |
| Chemical Checker (CC) | Database | Provides bioactivity signatures for over 1 million compounds, used for training and validating predictive models like Signaturizers3D [11]. | Public Access |
| Uni-Mol | Deep Learning Model | A pre-trained model for 3D molecular representation, which can be fine-tuned to generate stereochemically-aware bioactivity descriptors [11]. | Open Source |
| ZINC / NP Atlas | Compound Databases | Large, publicly accessible databases of commercially available synthetic compounds (ZINC) and natural products (NP Atlas) for library building [9]. | Public Access |
The comparative data is unequivocal: natural products and their derivatives consistently explore a broader and more three-dimensional region of chemical space, as defined by higher Fsp3 and greater stereochemical content, compared to many synthetic libraries and top-selling synthetic drugs. This structural richness is not an academic distinction but is directly linked to desirable drug properties, including improved solubility, target specificity, and a higher likelihood of clinical success. The increasing prevalence of natural product-based drugs among top sellers signals a market validation of this principle. For drug development professionals, this analysis argues for the deliberate inclusion of three-dimensionality as a key parameter in library design and compound optimization. Future directions will likely involve the wider adoption of 3D-aware descriptors and the continued development of synthetic methodologies, such as Diversity-Oriented Synthesis, to better access the under-explored, complex chemical space that natural products have already proven to be so valuable.
The exploration of chemical space is a fundamental task in cheminformatics and drug discovery. Within this space, ring systems and scaffolds form the structural core of most bioactive molecules, determining their shape, properties, and ultimately, their biological activity [15] [16]. This guide provides a comparative analysis of the structural diversity of ring systems found in natural products (NPs) versus synthetic compounds (SCs), underpinned by experimental data and chemoinformatic analyses. Understanding these differences is crucial for harnessing the full potential of NPs in drug discovery and for designing targeted synthetic libraries that explore underutilized regions of chemical space.
Natural products are renowned for their vast structural diversity. A comprehensive analysis of the COCONUT database, which contains over 400,000 NPs, identified 38,662 unique natural product ring systems [16]. This number significantly surpasses the diversity found in typical synthetic libraries. When considering stereochemistry, this diversity is even more pronounced, with the refined COCONUT set containing 269,226 unique compounds [16].
The analysis of ring system frequency follows a classic "long tail" distribution in both natural and synthetic chemical spaces. A study of 1.35 million molecules from the ChEMBL database identified 29,179 unique rings used in medicinal chemistry, with a striking 47.3% being singletons (appearing in only one molecule) [15]. This pattern of a few common rings and a very large number of rare rings is mirrored but expanded in NP collections, indicating a broader exploration of ring chemical space by nature.
The following table summarizes the key structural differences between NP and synthetic compound ring systems, based on analyses of major databases like COCONUT (for NPs) and ZINC20 ( for purchasable synthetic compounds).
Table 1: Structural Properties of Ring Systems in Natural Products vs. Synthetic Compounds
| Structural Property | Natural Products (NPs) | Synthetic Compounds (SCs) | Analysis Method |
|---|---|---|---|
| Representation in Drugs | ~2% of NP ring systems are present in approved drugs [16] | Higher representation of common drug-like ring systems [15] | Frequency analysis in drug databases |
| 3D Shape & Electrostatics | ~50% have identical/related 3D shape & electrostatic properties in screening compounds [16] | Covers a more limited, drug-like region of 3D space [16] | Comparison of 3D molecular shape and electrostatic properties |
| Stereochemical Complexity | High, often with complex, specific stereochemistry [16] | Generally lower | Analysis considering stereochemical information |
| Ring Complexity | More fused, bridged, and spiro rings; higher incidence of macrocycles [17] [2] | Predominantly simpler 5- and 6-membered rings with linkers [15] | Analysis of ring topology and connectivity |
| Aromatic vs. Aliphatic | Lower aromaticity; more aliphatic and saturated rings [17] | Higher aromatic character [17] | Fraction of sp3-hybridized carbons (Fsp3), aromaticity indices |
| Common Ring System Sizes | Diverse sizes, including many medium and large rings [18] | Overwhelmingly 5- and 6-membered rings [15] | Analysis of ring system size distributions |
The complexity of NP ring systems presents both an opportunity and a challenge. Their unique three-dimensional shapes are excellent for interacting with complex biological targets, but their structural intricacy often makes them difficult to synthesize [2]. Only about 17% of NP ring scaffolds are present in commercially available screening collections, creating a significant coverage gap in experimental screening [17].
The following diagram illustrates the standard cheminformatic workflow for extracting and comparing ring systems from large molecular databases, as employed in recent studies [16].
Diagram 1: Cheminformatics Workflow for Ring System Analysis
Database Curation and Preprocessing: Studies begin with large, curated databases. For NPs, the COCONUT (Collection of Open Natural Products) database is often used, while for synthetic compounds, the purchasable subset of ZINC20 is a common reference [16]. Key preprocessing steps include:
Ring System Definition and Extraction: A consistent definition of a ring system is applied. Typically, this is the graph composed of all atoms forming one or more fused or spiro rings, plus any exocyclic atoms connected via non-single bonds [16]. This extraction is automated using cheminformatics toolkits like RDKit.
Molecular Representation and Descriptor Calculation: To compare ring systems quantitatively, they are represented computationally.
Diversity Analysis and Comparison: The core of the comparison uses several metrics and algorithms:
Visualization: Techniques like t-distributed Stochastic Neighbor Embedding (t-SNE) are used to project the high-dimensional chemical space into 2D for visual inspection, allowing researchers to see how NP and synthetic libraries occupy complementary or overlapping regions [19] [21].
The following table lists key software, databases, and computational tools essential for conducting research in this field.
Table 2: Key Research Reagents and Computational Tools
| Tool/Resource | Type | Primary Function | Relevance to Ring System Analysis |
|---|---|---|---|
| COCONUT DB [22] [16] | Database | Largest public repository of natural product structures. | Source of NP ring systems for extraction and analysis. |
| ZINC20 [16] [2] | Database | Curated database of commercially available and synthesizable compounds. | Representative source for synthetic compound ring systems. |
| RDKit [19] [2] | Software Cheminformatics Toolkit | Open-source platform for cheminformatics. | Used for structure standardization, ring system perception, fingerprint generation, and descriptor calculation. |
| ChEMBL [20] [2] | Database | Manually curated database of bioactive molecules. | Provides context on the bioactivity and target associations of ring systems. |
| iSIM & BitBIRCH [20] | Algorithm | Efficient similarity and clustering for large libraries. | Enables diversity analysis of millions of ring systems without prohibitive computational cost. |
| Chemical Checker [21] | Web Tool / Database | Provides integrated bioactivity signatures for small molecules. | Used to compare structural and bioactivity profiles of different compound libraries. |
| Loviride | Loviride, CAS:141030-40-2, MF:C17H16Cl2N2O2, MW:351.2 g/mol | Chemical Reagent | Bench Chemicals |
| trans-1-Benzoyl-4-hydroxy-L-Proline | trans-1-Benzoyl-4-hydroxy-L-Proline, CAS:129512-75-0, MF:C12H13NO4, MW:235.24 g/mol | Chemical Reagent | Bench Chemicals |
The cheminformatic comparison unequivocally demonstrates that natural products explore a vastly broader and more complex region of ring system space than conventional synthetic libraries. NPs possess a wealth of unique, three-dimensionally complex, and often under-explored ring scaffolds. However, a significant coverage gap exists, as the vast majority of these NP ring systems are absent from standard screening collections.
This analysis provides a compelling rationale for strategies that aim to bridge this gap, such as biology-oriented synthesis (BIOS) and the construction of pseudo-natural product (PNP) libraries [18]. By leveraging the structural insights provided by cheminformatic analyses, drug discovery efforts can be strategically directed to harness the rich diversity of NP-inspired ring systems, thereby increasing the likelihood of discovering novel bioactive compounds against challenging therapeutic targets.
The systematic comparison of oxygen-rich Natural Products (NPs) and nitrogen-rich Synthetic Compounds (SCs) represents a core focus in modern cheminformatics and drug discovery research. NPs, products of evolutionary biosynthesis, and SCs, products of rational design, occupy distinct yet complementary regions of chemical space. Their fundamental differences in atomic and functional group composition directly influence their physicochemical properties, bioactivity profiles, and suitability as drug candidates or leads [2]. Framing this comparison within a chemoinformatic context allows for a objective, data-driven analysis of their respective characteristics, enabling researchers to make informed decisions in lead identification and optimization campaigns. This guide provides a detailed, evidence-based comparison of these two compound classes, supporting the broader thesis that understanding their inherent chemical differences is crucial for advancing drug discovery.
The defining characteristic of "oxygen-rich" NPs and "nitrogen-rich" SCs is the prevalence and variety of specific functional groups containing these elements. The tables below summarize the common functional groups and their associated properties for each compound class.
Table 1: Common Functional Groups in Oxygen-Rich Natural Products (NPs)
| Functional Group | General Formula | Key Properties & Biological Roles | Prevalence in NPs |
|---|---|---|---|
| Hydroxyl (Alcohol/Phenol) | RâOH | Hydrogen bonding, increases water solubility, metabolic conjugation | High; ubiquitous in plant-derived NPs [23] [24] |
| Carboxyl | RâCOOH | Acidic, forms salts, strong hydrogen bonding, site for derivatization | High; found in fatty acids, organic acids [23] |
| Carbonyl (Aldehyde/Ketone) | RâCHO / RâCOR' | Electrophilic, participates in redox reactions and nucleophilic addition | Moderate to High [23] |
| Ester | RâCOOR' | Polar, can be hydrolyzed by metabolic esterases | High; common in macrolides and fatty acid derivatives [23] |
| Ether | RâOâR' | Relatively inert, can confer metabolic stability and influence conformation | Moderate; e.g., in cyclic ethers [23] |
Table 2: Common Functional Groups in Nitrogen-Rich Synthetic Compounds (SCs)
| Functional Group | General Formula | Key Properties & Biological Roles | Prevalence in SCs |
|---|---|---|---|
| Amino (Primary, Secondary, Tertiary) | RâNHâ, RâNH, RâN | Basic, hydrogen bonding, cationic at physiological pH, common in pharmacophores | Very High; foundational in many drug classes [23] |
| Amide | RâCONR'R" | Planar, strong hydrogen bonding, critical for peptide backbone and protein binding | Extremely High; essential in peptidomimetics [23] |
| Nitro | RâNOâ | Strongly electron-withdrawing, can be reduced metabolically, used in energetic materials | Moderate; specific applications [25] |
| Nitrile | RâCâ¡N | Polar, a metabolically stable bioisostere for carbonyl or halogens | Moderate; common in kinase inhibitors [23] |
| Azide | RâNâ | Energetic, used in "click chemistry" for bioconjugation | Low to Moderate; specialized synthetic applications [25] |
| Heterocyclic N (e.g., Pyridine, Imidazole, Indole) | e.g., Câ Hâ N | Aromatic, can be basic, participates in key binding interactions (e.g., coordination, Ï-stacking) | Extremely High; indole is a "privileged structure" [26] |
The objective comparison of oxygen-rich NPs and nitrogen-rich SCs requires a structured cheminformatic workflow. This process involves data curation, computational analysis, and experimental validation to translate chemical data into meaningful biological insights.
Data Curation and Collection
Descriptor and Fingerprint Calculation
Chemical Space Analysis and Visualization
Bioactivity Prediction and Virtual Screening
Experimental Validation
The following protocol is adapted from the synthesis of brominated indole-3-glyoxylamides (IGAs), a class of nitrogen-rich, MNP-inspired synthetic compounds [26].
This protocol is based on research that correlates the concentration of specific functional groups with surface properties, which can influence biomolecular interactions [27].
Table 3: Key Reagents and Tools for Chemoinformatic Comparison Studies
| Item | Function / Application | Examples / Specifications |
|---|---|---|
| RDKit | Open-source cheminformatics toolkit for descriptor calculation, fingerprinting, and machine learning. | Calculating O/C, N/C ratios, molecular weight, and generating ECFP4 fingerprints for similarity search [2]. |
| KNIME Analytics Platform | Open-source platform for data pipelining; integrates cheminformatics nodes (e.g., RDKit, CDK) for workflow automation. | Building a data pipeline that ingests structures from a database, calculates descriptors, and builds a predictive model [2]. |
| MarinLit Database | Specialized, curated database of marine natural product literature and structures. | Sourcing and curating structures and bioactivity data for oxygen-rich marine NPs for comparative analysis [26]. |
| ChEMBL Database | Manually curated database of bioactive molecules with drug-like properties, containing many SCs. | Sourcing bioactivity data and structures of nitrogen-rich synthetic compounds for model training and validation [2]. |
| DataWarrior | Open-source program for data visualization and analysis, includes chemical-aware plotting and SOM capabilities. | Generating Self-Organizing Maps (SOMs) to visualize the chemical space of NPs and SCs [26]. |
| 4-(Trifluoromethyl)benzaldehyde (TFBA) | Chemical derivatization agent for quantifying primary amine (âNHâ) groups on surfaces via XPS. | Experimental quantification of amine group concentration in nitrogen-rich polymer films [27]. |
| Toluidine Blue O (TBO) | Dye used in colorimetric assay for quantifying carboxylic acid (âCOOH) groups on surfaces. | Experimental quantification of carboxylic acid group concentration in oxygen-rich polymer films [27]. |
| 2,4-Dihydroxy-3,3-dimethyl-butanoic acid | 2,4-Dihydroxy-3,3-dimethyl-butanoic acid, CAS:1902-01-8, MF:C6H11NaO4, MW:170.14 g/mol | Chemical Reagent |
| Benzoctamine Hydrochloride | Benzoctamine Hydrochloride, CAS:10085-81-1, MF:C18H20ClN, MW:285.8 g/mol | Chemical Reagent |
The systematic comparison of natural products (NPs) and synthetic compounds (SCs) represents a cornerstone of modern drug discovery and chemical biology. Over half of all approved small-molecule drugs originate directly or indirectly from natural products, underscoring their profound historical significance [22]. However, the deliberate design of synthetic compounds has enabled researchers to explore chemical spaces beyond those provided by nature. The evolution of structural properties between these two classes reveals distinct trajectories shaped by evolutionary pressures on one hand and rational design objectives on the other.
This comprehensive analysis employs chemoinformatic approaches to quantitatively examine how molecular architectures, complexity, and desirable drug-like properties have diverged between naturally occurring and synthetic molecules over time. Understanding these evolutionary pathways provides valuable insights for future drug discovery efforts, particularly in leveraging the complementary strengths of both natural and synthetic compounds to address challenging therapeutic targets.
Systematic chemoinformatic analyses reveal consistent, quantifiable differences in structural properties between natural products and synthetic compounds. These distinctions reflect their distinct originsâshaped by evolutionary pressures in biological systems versus rational design in laboratory settings.
Table 1: Core Structural Properties of Natural Products vs. Synthetic Compounds
| Structural Property | Natural Products | Synthetic Compounds | Analysis Method |
|---|---|---|---|
| Molecular Complexity | Higher (more chiral centers, Csp3, macro rings) [22] | Lower | Chirality analysis, Csp3 quantification |
| Structural Diversity | Broader chemical space, higher scaffold diversity [22] | More constrained | Scaffold analysis, chemical space visualization |
| Glycosylation Rate | 8%-22% of NPs [22] | 0.23%-4.93% [22] | Structural motif identification |
| Halogenation | More frequent [22] | Less frequent (except pesticides) [22] | Halogen atom detection |
| Ring Systems | More aliphatic and fused rings [22] | Fewer complex ring systems | Ring system categorization |
| Hydrogen Bonding | More donors/acceptors (flavonoids) [22] | Generally fewer | Hydrogen bond donor/acceptor count |
| Molecular Size | Generally larger [22] | Smaller, Lipinski-compliant | Molecular weight distribution |
Natural products exhibit significantly higher structural complexity across multiple dimensions. They contain more chiral centers, higher ratios of Csp3 hybridized carbon atoms, and more complex ring systems including macrocycles, bridge rings, and spiro rings [22]. This structural complexity translates to enhanced three-dimensionality and shape diversity, which correlates with improved selectivity for biological targets.
The scaffold diversity of natural products substantially exceeds that of synthetic compounds, particularly approved drugs. For instance, the Nat-UV DB database of Mexican natural products contains 227 compounds with 112 scaffolds, 52 of which were not present in existing databases [28]. This highlights nature's remarkable capacity for generating novel molecular frameworks, even within relatively small compound collections.
Table 2: Property Ranges Across Compound Classes
| Property | Natural Products | Synthetic Compounds | Approved Drugs |
|---|---|---|---|
| Molecular Weight | Broader distribution, larger average [22] | More constrained | Intermediate |
| Lipinski Rule Compliance | Variable (e.g., 86.4% of lignans compliant) [22] | Generally high | High |
| Polar Surface Area | Higher in specific classes (e.g., flavonoids) [22] | Generally lower | Intermediate |
| Rotatable Bonds | More in terpenoids [22] | Fewer | Intermediate |
| Hydrophobicity | More hydrophobic [22] | Less hydrophobic | Balanced |
While natural products frequently violate Lipinski's Rule of Five, certain subclasses demonstrate remarkably high compliance rates. For example, 86.4% of lignans adhere to these drug-likeness criteria [22]. Terpenoidsâwhich comprise approximately one-third of all known natural productsâalso predominantly follow the Rule of Five, suggesting favorable bioavailability despite their complex structures [22].
The glycosylation pattern differences between natural and synthetic compounds are particularly striking. Glycosylation occurs in 8%-22% of natural products, with significant variation across biological sources: plants (24.99%), bacteria (20.84%), animals (8.40%), and fungi (4.48%) [29]. This contrasts sharply with synthetic compounds and approved drugs, which exhibit glycosylation rates of only 0.23% and 4.93%, respectively [22]. This modification significantly influences solubility, bioavailability, and target interactions.
The comparative analysis of natural and synthetic compounds relies on standardized chemoinformatic workflows that enable consistent characterization across diverse compound classes.
Figure 1: Chemoinformatic workflow for structural property comparison. This standardized pipeline enables consistent analysis across diverse compound classes, from data collection through chemical space visualization.
The construction of specialized natural product databases enables systematic comparison of structural properties. The Nat-UV DB database exemplifies this approach, comprising 227 compounds meticulously curated from the biodiversity-rich coastal zone of Veracruz, Mexico [28]. Database construction follows rigorous protocols: compound collection and identification, structural elucidation, data curation, chemoinformatic annotation, and comparative analysis against reference databases.
Similar methodologies underpin larger-scale analyses, such as the comparison of fragment libraries derived from natural products versus synthetic compounds. The COCONUT database (containing >695,000 natural products) and LANaPDB (with 13,578 Latin American natural products) provide the foundation for extracting 2,583,127 natural product-derived fragments, which are subsequently compared against synthetic fragment libraries like CRAFT [30].
The quantification of structural properties relies on standardized molecular descriptors:
These descriptors enable the construction of multidimensional chemical spaces where the relative positions of natural versus synthetic compounds can be quantitatively compared [31].
The structural evolution of natural products reveals distinctive temporal patterns compared to synthetic compounds. Analysis of over 1.1 million documented natural products shows a declining discovery rate of novel scaffolds, suggesting increasing difficulty in finding truly new molecular frameworks from traditional natural sources [22]. This contrasts with synthetic chemistry, where methodological advances continuously enable exploration of previously inaccessible chemical space.
The temporal trajectory of natural product discovery has shifted from terrestrial to marine environments, with marine natural products displaying larger molecular sizes and greater hydrophobicity than their terrestrial counterparts [22]. More recently, natural products from extreme environments (deep-sea, extremophiles) have revealed novel scaffolds with unique bioactivities, expanding the known chemical space of natural compounds.
Synthetic compounds have evolved under different selection pressures, primarily driven by desired drug-like properties and synthetic feasibility. The rise of combinatorial chemistry in the 1990s initially produced "flat" molecules with limited structural complexity, but more recent synthetic approaches have deliberately incorporated natural product-inspired features including higher sp3 character, increased chirality, and more complex ring systems.
Fragment-based drug discovery has further influenced synthetic compound evolution, with fragment libraries now often designed to include natural product-derived fragments that occupy under-explored regions of chemical space [30]. The CRAFT library, for instance, incorporates 1,214 fragments based on novel heterocyclic scaffolds and natural product-derived chemicals, representing a deliberate fusion of natural and synthetic structural approaches [30].
Objective: Quantitatively compare scaffold diversity between natural products and synthetic compounds.
Methodology:
Applications: This protocol revealed that Nat-UV DB compounds contain 52 scaffolds not present in other natural product databases, demonstrating the value of exploring biodiversity-rich geographical regions [28].
Objective: Visualize and compare the chemical space occupied by natural products versus synthetic compounds.
Methodology:
Applications: Chemical space mapping consistently demonstrates that natural products occupy broader regions than synthetic compounds, with approved drugs predominantly located in overlapping regions [22] [31].
Table 3: Essential Resources for Structural Property Research
| Resource Name | Type | Key Features | Application in Research |
|---|---|---|---|
| COCONUT 2.0 [30] | Natural Product Database | >695,000 non-redundant NPs | Large-scale analysis of NP structural diversity |
| CRAFT Library [30] | Fragment Library | 1,214 fragments, NP-inspired | Comparison of NP vs synthetic fragment properties |
| Nat-UV DB [28] | Regional NP Database | 227 compounds from Veracruz, Mexico | Analysis of region-specific structural features |
| LaNAPDB [30] | Regional NP Database | 13,578 unique NPs from Latin America | Geographic-based structural comparisons |
| DNP [29] | Comprehensive NP Database | Extensive structural annotations | Glycosylation pattern analysis across species |
| MacrolactoneDB [22] | Specialized NP Database | 13,721 macrolactone NPs | Analysis of complex macrocyclic structures |
| Open Chemoinformatic Tools [31] | Software Tools | Freely available algorithms | Chemical space visualization and analysis |
The evolutionary trajectories of natural and synthetic compounds suggest powerful synergies for future drug discovery. Natural products provide validated starting points with proven biological relevance and structural novelty, while synthetic approaches enable optimization of drug-like properties and target specificity.
The integration of natural product fragments into synthetic libraries represents one promising hybrid approach. Analysis shows that fragments derived from natural products occupy distinct regions of chemical space compared to purely synthetic fragments, offering opportunities to explore novel structure-activity relationships [30]. Similarly, the application of synthetic methodology to elaborate natural product-inspired scaffolds can generate compounds combining the complexity of natural products with tailored pharmaceutical properties.
Emerging strategies highlight the value of exploring underinvestigated natural sources, including marine organisms, extremophiles, and microorganisms from unique geographical regions [22]. The discovery of 52 previously unrecorded scaffolds in the relatively small Nat-UV DB database underscores the potential of targeted exploration of biodiversity-rich regions [28].
Advancements in artificial intelligence and machine learning are further accelerating the integration of natural and synthetic approaches. These technologies enable predictive models of bioactivity, toxicity, and synthetic accessibility, facilitating the design of hybrid compounds that leverage the complementary strengths of both natural and synthetic structural paradigms [22].
Table of Contents
Fragment-based drug discovery (FBDD) has emerged as a powerful approach for identifying novel therapeutic compounds by screening small, low molecular weight fragments (<300 Da) against biological targets. These fragments typically comply with the "Rule of Three" guidelines (molecular weight <300 Da, hydrogen bond donors/acceptors â¤3, and cLogP â¤3) and provide efficient sampling of chemical space due to their simplicity [32]. A critical challenge in FBDD is the generation of fragment libraries with sufficient structural diversity, three-dimensionality, and synthetic tractability to serve as valuable starting points for drug development [32] [33]. Molecular deconstruction algorithms address this challenge by systematically breaking down complex molecules into smaller fragments, thereby creating screening libraries that retain key structural features of pharmacologically relevant compounds.
The deconstruction of natural products (NPs) holds particular promise for fragment library design. Natural products are evolutionarily optimized to interact with biological macromolecules and exhibit greater three-dimensional complexity, higher fractions of sp³ carbons (Fsp³), and more chiral centers compared to synthetic compounds [33] [17]. Approximately 30% of FDA-approved drugs from 1981 to 2019 originated from natural products or their derivatives, particularly in anti-infective and anti-cancer therapies [34]. Their privileged scaffolds make them ideal starting materials for generating fragments with enhanced biological relevance. Deconstruction algorithms transform these complex structures into fragment-sized molecules while preserving their desirable structural characteristics, enabling more efficient exploration of biologically relevant chemical space [33] [35].
The Retrosynthetic Combinatorial Analysis Procedure (RECAP) is a well-established algorithm for molecular fragmentation that applies rules based on chemically favored cleavage sites. RECAP identifies key bond types in organic molecules that are susceptible to fragmentation, generating smaller chemical entities that can serve as building blocks for fragment libraries [35]. The algorithm employs a systematic approach to bond disconnection, prioritizing breaks at bonds adjacent to specific functional groups and ring systems commonly found in pharmacologically active compounds.
RECAP fragmentation can be implemented in two distinct modalities with fundamentally different outcomes:
Extensive (Exhaustive) Fragmentation: This approach generates the smallest possible fragments by applying RECAP rules exhaustively until no further cleavages are possible. The resulting fragments represent minimal chemical units, often referred to as "leaf nodes" in fragmentation trees [35]. While these fragments provide maximum simplification, they may lose important structural context from the parent molecule.
Non-extensive (Intermediate) Fragmentation: This alternative methodology generates all possible "intermediate" scaffolds by systematically considering cleavage sites without pursuing exhaustive fragmentation [35]. These intermediate fragments retain more structural information from the original molecule while still complying with fragment size constraints, potentially offering better starting points for fragment elaboration.
Table 1: Comparison of RECAP Fragmentation Approaches
| Characteristic | Extensive Fragmentation | Non-extensive Fragmentation |
|---|---|---|
| Fragment Size | Smaller, minimal units | Larger, intermediate scaffolds |
| Structural Context | Limited retention of parent structure | Better preservation of structural features |
| Chemical Diversity | Higher redundancy | Lower repetition |
| Number of Fragments | Fewer generated (e.g., 11,525 from NP library) | More generated (e.g., 45,355 from NP library) |
| Pharmacophore Fit | Generally lower | Generally higher (56% of cases superior to extensive) |
The RECAP algorithm specifically targets chemically labile bonds and functional groups commonly found in drugs and natural products, including amide, ester, urea, and sulfonamide linkages, among others. This strategic bond selection ensures that the resulting fragments represent synthetically accessible and biologically relevant chemical space, facilitating subsequent medicinal chemistry optimization [35].
While RECAP remains a widely used method for molecular deconstruction, several alternative algorithms have been developed to address specific limitations and explore different aspects of chemical space. These approaches employ distinct strategies for fragment generation, ranging from biosynthetic-inspired decomposition to structure enumeration and pseudo-natural product design.
The LEMONS (Library for the Enumeration of MOdular Natural Structures) algorithm represents a specialized approach for generating hypothetical modular natural product structures [17]. Unlike RECAP's decomposition strategy, LEMONS constructs natural product-like molecules by simulating biosynthetic assembly lines, incorporating diverse monomer units and tailoring reactions. This methodology allows researchers to investigate the impact of various biosynthetic parameters on chemical similarity search and library diversity. LEMONS is particularly valuable for exploring the chemical space of nonribosomal peptides, polyketides, and hybrid natural products, which feature large and structurally complex scaffolds distinct from synthetic compounds [17].
Pseudo-Natural Product design constitutes an innovative approach that combines biosynthetically unrelated natural product fragments to create novel chemical entities that transcend traditional natural product space [33]. This strategy involves deconstructing natural products into fragments followed by recombining them in new arrangements not found in nature. For example, "indotropanes" created by combining indole and tropane scaffolds, and "chromopynones" formed by merging chromane and tetrahydropyrimidinone fragments, have demonstrated biological activity against specific targets such as myosin light chain kinase 1 and glucose transporters [33]. This approach leverages nature's structural wisdom while venturing into unprecedented chemical territory.
In-silico guided chemical disassembly of larger natural products represents another deconstruction strategy that employs computational methods to generate virtual fragment libraries [33]. This process begins with virtual cleavage reactions applied to natural product databases, followed by application of fragment-like criteria (150 < MW < 300, cLogP < 3) to filter the resulting compounds. Subsequent 3D shape assessment and novelty evaluation using molecular fingerprints further refine the fragment collection. This method has successfully generated fragments from complex natural products such as FK506 (Tacrolimus), sanglifehrin A, and cytochalasin E, producing 3D-shaped, natural product-like fragments with privileged structural features [33].
Rigorous comparison of deconstruction algorithms requires evaluation across multiple performance metrics, including chemical diversity, structural complexity, retention of bioactive features, and practical utility in virtual screening campaigns. The following analysis synthesizes experimental data from published studies to provide a comprehensive assessment of RECAP and alternative approaches.
A systematic study comparing extensive and non-extensive RECAP fragmentation of natural product libraries revealed significant differences in fragment properties and performance [35]. When applied to a virtual library of natural products from Traditional Chinese Medicine (TCM), AfroDb, NuBBE, and UEFS databases, non-extensive fragmentation generated 45,355 fragments compared to only 11,525 fragments from extensive fragmentation. This nearly 4-fold increase in chemical entities directly translates to enhanced exploration of chemical space.
Table 2: Performance Metrics of RECAP-derived Natural Product Fragments
| Metric | Original NPs | Non-extensive NPDFs | Extensive NPDFs |
|---|---|---|---|
| Structural Diversity | Highest | Moderately high (slight reduction after VS) | Moderate (slight reduction after VS) |
| Pharmacophore Fit Score | Baseline | Higher than NPs (69% of cases) | Lower than non-extensive (56% of cases) |
| Molecular Complexity | High | Intermediate | Low |
| Synthetic Developability | Challenging | More feasible | Most feasible |
| Chemical Redundancy | Low | Lower than extensive | Higher than non-extensive |
In virtual screening applications against 20 different protein targets, non-extensive fragments demonstrated superior pharmacophore fit scores not only compared to extensive fragments (56% of cases) but also relative to their original natural products (69% of cases) when all were identified as hits [35]. This remarkable finding suggests that selective deconstruction can enhance the apparent potency of natural product-derived fragments by isolating key pharmacophoric elements while eliminating structurally complex but non-essential components.
The three-dimensional character of fragment libraries significantly influences their performance in biological screening, particularly for targeting challenging protein-protein interactions [36]. Natural product-derived fragments typically exhibit enhanced three-dimensionality compared to synthetic fragments, as quantified by the fraction of sp³ carbons (Fsp³) and principal moment of inertia (PMI) analysis [32] [33].
Analysis of the Dictionary of Natural Products database identified 7,365 non-flat fragment-sized natural products rich in sp³ centers (Fsp³* > 0.45) [33]. These fragments provide improved sampling of three-dimensional chemical space compared to conventional flat, aromatic fragment libraries, potentially enhancing success rates against difficult biological targets with flat binding sites, such as those involved in protein-protein interactions [36].
The LEMONS algorithm has demonstrated particular utility for quantifying the similarity of modular natural products, with retrobiosynthetic alignment approaches outperforming conventional 2D fingerprints when rule-based retrobiosynthesis can be applied [17]. This suggests that biosynthesis-aware deconstruction methods may offer advantages for certain natural product classes, especially when exploring structure-activity relationships within congeneric series.
Successful implementation of molecular deconstruction strategies requires careful selection of computational tools, screening libraries, and experimental protocols. The following section provides a practical toolkit for researchers embarking on fragment library design using deconstruction algorithms.
A standardized workflow for generating fragment libraries via RECAP deconstruction ensures consistent, high-quality results:
Diagram Title: RECAP Fragment Library Generation Workflow
This workflow begins with a diverse natural product library, applies RECAP rules (either extensive or non-extensive), filters the resulting fragments according to Rule of Three criteria, clusters structurally similar fragments, selects representative compounds, and experimentally validates key properties such as solubility and stability before final library assembly [35] [37].
Table 3: Key Resources for Fragment Library Design and Screening
| Resource Category | Specific Tools/Databases | Application in FBDD |
|---|---|---|
| Natural Product Databases | Traditional Chinese Medicine (TCM), AfroDb, NuBBE, UEFS, Dictionary of Natural Products | Source compounds for deconstruction [33] [35] |
| Cheminformatics Software | RDKit, ChemAxon, OpenBabel | Structure handling, fingerprint generation, similarity calculation [17] |
| Fragment Libraries | 3D Fragment Consortium (170 fragments), Enamine Fragment Library (1,500 compounds), Asinex BioDesign fragments | Commercially available fragments for screening [37] |
| Computational Tools | LEMONS, GRAPE/GARLIC, SPiDER | Specialized algorithms for NP analysis and target prediction [33] [17] |
| Screening Methodologies | X-ray crystallography, NMR, Surface Plasmon Resonance (SPR), Native Mass Spectrometry | Biophysical detection of fragment binding [32] [33] |
| Naringin hydrate | Naringin hydrate, MF:C27H34O15, MW:598.5 g/mol | Chemical Reagent |
| Osmanthuside H | Osmanthuside H, CAS:149155-70-4, MF:C19H28O11, MW:432.4 g/mol | Chemical Reagent |
A robust virtual screening protocol combining RECAP-based fragmentation and pharmacophore modeling involves the following steps:
Pharmacophore Model Development: Construct overlapping pharmacophore models for target proteins using software such as Ligand Scout, incorporating key interaction features (hydrogen bond donors/acceptors, hydrophobic regions, aromatic rings) and exclusion volumes [35].
Fragment Library Preparation: Apply RECAP rules to natural product databases, generating both extensive and non-extensive fragments. Filter according to Rule of Three criteria and additional property-based filters.
Virtual Screening: Screen the fragment library against pharmacophore models, calculating fit scores based on feature matching and root-mean-square deviation between model points and fragment conformers [35].
Hit Identification and Analysis: Rank fragments by pharmacophore fit score, identify structural clusters, and prioritize fragments with optimal properties for experimental validation.
This protocol has been successfully applied to multiple protein targets, demonstrating that non-extensive fragments frequently outperform both extensive fragments and parent natural products in pharmacophore-based screening [35].
Molecular deconstruction algorithms, particularly RECAP and its alternatives, provide powerful methodologies for generating diverse, biologically relevant fragment libraries from complex natural products. The comparative analysis presented in this guide demonstrates that non-extensive RECAP fragmentation generally outperforms extensive fragmentation by generating more chemically diverse fragments with superior pharmacophore fit scores while retaining valuable structural context from parent natural products.
The emerging trend toward three-dimensional, complex fragments reflects a growing recognition that structural complexity enhances success in fragment-based drug discovery, particularly for challenging target classes such as protein-protein interactions [32] [36]. Natural product deconstruction represents a privileged approach to accessing such fragments, leveraging nature's evolutionary optimization of biologically relevant chemical space.
Future developments in this field will likely include increased integration of artificial intelligence and generative models for fragment design [34], expanded application of biosynthesis-aware deconstruction algorithms [17], and greater emphasis on synthetic accessibility during the fragment selection process. As these methodologies mature, deconstruction algorithms will continue to play a pivotal role in bridging the gap between natural product complexity and fragment-based screening paradigms, accelerating the discovery of novel therapeutic agents against increasingly challenging biological targets.
Thesis Context: This guide provides an objective, data-driven comparison of three major Natural Product (NP) databasesâCOCONUT, LANaPDB, and DNPâframed within a broader chemoinformatic analysis of natural products versus synthetic compounds. It is designed to aid researchers in selecting the most appropriate database for specific drug discovery applications.
Natural products (NPs) have historically been the most prolific source of inspiration for new drugs, with approximately two-thirds of all small-molecule drugs approved between 1981 and 2019 being related to NPs in some form [2]. The structural diversity and complexity of NPs often result in unique biological activities, making them invaluable starting points for therapeutic development [38]. However, the real bottleneck in NP-based drug discovery has traditionally been the availability of materials for testing, a challenge that computational approaches aim to overcome [2].
In the last decade, there has been a steep increase in databases providing access to chemical, biological, and structural data on NPs [2]. These databases serve as crucial tools in computer-aided drug design (CADD), enabling virtual screening, chemical space analysis, and bioactivity prediction without the immediate need for physical compounds [39] [2]. The selection of an appropriate NP database fundamentally influences the success of these in silico campaigns, necessitating a clear understanding of their respective scope, features, and limitations.
This guide focuses on three databases with distinct architectures and purposes: the COlleCtion of Open NatUral producTs (COCONUT) as a comprehensive global resource, the Latin American Natural *Product Database (LANaPDB)* as a regionally specialized compilation, and the Dictionary of Natural Products (DNP) as a well-established commercial offering. Our comparative analysis situates these resources within the chemoinformatic workflow for comparing natural and synthetic chemical spaces, providing researchers with the experimental data and methodologies needed to inform their database selection.
COCONUT is one of the largest open-access NP databases, launched in 2021 as an aggregation of openly available datasets [40] [38]. Its core mission is to unify and standardize global NP data, providing not only chemical structures but also rich metadata, including names, biological sources, geographic origin, and literature references [38]. The recently released COCONUT 2.0 represents a complete overhaul of the platform, emphasizing community curation, FAIR data principles, and improved data quality.
LANaPDB represents a collective effort to compile and standardize NP databases from Latin America, a region recognized for its extraordinary biodiversity [39]. As a relatively new resource, its specific focus fills a crucial geographical gap in the NP data landscape. The database aims to gather NPs isolated and characterized from seven Latin American countries, making it an essential resource for studying the unique chemical diversity of this region [39].
The Dictionary of Natural Products (DNP) is a long-established, comprehensive commercial database. While the search results do not contain specific details about its current size or features, it is widely recognized in the field as a authoritative, curated resource that has been maintained for decades. As a subscription-based service, it typically offers extensive manual curation, detailed annotations, and reliable data quality, positioning it as a benchmark in natural products research.
Table 1: Core Characteristics and Database Statistics
| Feature | COCONUT | LANaPDB | DNP |
|---|---|---|---|
| Access Type | Open-access | Open-access | Commercial |
| Total Compounds | 695,119 [40] | 13,578 [39] | Information Not in Search Results |
| Data Source | Aggregation of 53 open-access databases [38] | 10 databases from 7 Latin American countries [39] | Information Not in Search Results |
| Geographic Focus | Global | Regional (Latin America) | Information Not in Search Results |
| Key Feature | Community curation, FAIR data principles | Regional chemical space characterization | Information Not in Search Results |
| Update Status | Actively updated (v2.0 in 2024) [38] | Updated in 2024 [39] | Information Not in Search Results |
The structural diversity contained within an NP database directly influences its potential for novel bioactive compound discovery. A chemoinformatic characterization of LANaPDB, calculating six key physicochemical properties, reveals its constituents have favorable drug-like properties, positioning the database as a valuable source for lead-like compounds [39]. The chemical space of LANaPDB has been visualized and compared to major reference sets like COCONUT and FDA-approved drugs using Tree MAP (TMAP) algorithms based on MACCS keys and Morgan2 fingerprints [39]. This analysis allows researchers to navigate and contextualize the regional chemical space of Latin American NPs within the global NP landscape.
Specialized regional databases like Nat-UV DB (a Mexican database included in LANaPDB) have been shown to contain unique scaffolds not present in larger, global databases, highlighting the value of regional focus for identifying novel chemical entities [41]. This suggests that while LANaPDB is numerically smaller than COCONUT, it may offer unique structural diversity relevant to drug discovery.
The type and quality of metadata and biological annotations differ significantly across databases, impacting their utility for various research applications.
COCONUT provides extensive metadata, including organism details mapped to ontologies, geographic location data for over 63,000 molecules, and literature citations linked to approximately 117,000 molecules [40] [38]. This rich contextual information supports research in fields like ethnobotany and ecology.
LANaPDB has been cross-referenced with major bioactivity databases like ChEMBL and PubChem, enriching its entries with reported and predicted biological activities [39]. This enhances its utility for drug discovery, enabling target prediction and activity profiling.
Table 2: Metadata and Bioactivity Comparisons
| Annotation Type | COCONUT | LANaPDB | DNP |
|---|---|---|---|
| Organism/Species | 53,092 organisms mapped [40] | Included from source databases [39] | Information Not in Search Results |
| Geographic Origin | 2,654 locations for 63,473 molecules [40] | Specific to Latin American region [39] | Information Not in Search Results |
| Literature References | 35,185 citations for 117,590 molecules [40] | Information Not in Search Results | Information Not in Search Results |
| Bioactivity Data | Varies by source collection | Cross-referenced with ChEMBL & PubChem [39] | Information Not in Search Results |
| Structural Classification | Available via ClassyFire [38] | Performed using NPClassifier [39] | Information Not in Search Results |
Data quality is a paramount concern in NP databases, as errors in stereochemistry or structure can significantly impact computational results [2]. The three databases employ distinct curation methodologies.
COCONUT 2.0 utilizes an RDKit-based ChEMBL pipeline for data standardization, which preserves stereochemistry and standardizes functional groups [40] [38]. A key innovation is its community curation model, which allows users to report incorrect data (e.g., synthetic compounds mislabeled as natural) and submit change requests, creating a collaborative and continuously improving resource [38].
LANaPDB employs a standardized curation workflow using RDKit and the MolVS package in Python [39]. This process includes verifying and correcting valencies and aromaticity, removing explicit hydrogens, applying normalization rules, ensuring proper protonation states, and recalculating stereochemistry. Duplicate compounds are removed using InChIKey strings of canonical tautomers [39].
The DNP, as a commercial product, likely employs a team of expert curators, though specific methodologies are not detailed in the search results. This manual curation is often considered a benchmark for accuracy but may not scale as effectively as semi-automated community-driven approaches.
A core application of NP databases in chemoinformatics is the analysis of chemical space and the quantification of "natural product-likeness." Recent research has profiled the NP-likeness of LANaPDB in comparison to other major databases and approved drugs [42]. This profiling employs several chemoinformatics metrics to determine how closely a molecule's structural characteristics resemble those of known natural products.
Studies have shown that compounds in LANaPDB occupy a chemical space that bridges typical NP regions and the space of approved drugs, making it a promising source for drug-like leads [39] [42]. The database has been characterized using calculated physicochemical propertiesâincluding SlogP, molecular weight, topological polar surface area (TPSA), rotatable bonds, and hydrogen bond donors/acceptorsâwhich are critical for assessing drug-likeness and forecasting oral bioavailability [39].
NP databases are pivotal in structure-based and ligand-based virtual screening. The large size and diversity of COCONUT make it suitable for uncovering novel scaffolds with potential bioactivity across a wide range of targets [38]. In contrast, LANaPDB's regionally focused collection is valuable for investigating the specific chemical defenses and medicinal compounds from one of the world's most biodiverse regions [39].
For example, during the SARS-CoV-2 pandemic, NP-based computer-aided drug design was a primary approach for identifying lead compounds, relying heavily on the content of these databases [39]. The cross-referencing of LANaPDB with ChEMBL and PubChem directly supports such efforts by providing initial bioactivity annotations for hypothesis generation [39] [41].
Table 3: Key Software and Tools for NP Database Research
| Tool/Resource | Function | Application in NP Research |
|---|---|---|
| RDKit | Open-source cheminformatics toolkit [39] | Data curation, descriptor calculation, fingerprint generation [39] [2] |
| NPClassifier | Deep neural network-based structural classification [39] | Automated structural classification of NPs into known pathways and classes [39] |
| ClassyFire | Automated chemical classification [38] | Assigning compounds a hierarchical classification of structural types [38] |
| MolVS | Molecule Virtual Screening Python library [39] | Molecule standardization (tautomer normalization, charge correction, desalting) [39] |
| TMAP | Tree MAP visualization algorithm [39] | Visualizing and navigating high-dimensional chemical spaces from fingerprint data [39] |
| COCONUT API | RESTful Application Programming Interface [40] | Programmatic access to the latest COCONUT data for integration into automated workflows [40] |
| Perfluorobutanesulfonic acid | Perfluorobutanesulfonic acid (PFBS) for Research | |
| Epelmycin E | Epelmycin E, CAS:138636-10-9, MF:C42H53NO16, MW:827.9 g/mol | Chemical Reagent |
The comparative analysis of COCONUT, LANaPDB, and DNP reveals a trade-off between breadth, regional specificity, and curation depth, guiding researchers toward informed database selection based on project goals.
For comprehensive, global discovery and maximum data accessibility, COCONUT is the superior choice. Its unparalleled scale, open-access nature, and community-driven curation model make it ideal for large-scale virtual screening and exploring the broadest possible NP chemical space.
For targeted research on Latin American biodiversity or investigating region-specific traditional medicine, LANaPDB is an indispensable resource. Its focused content, cross-referenced bioactivity data, and detailed physicochemical profiling offer a unique window into a rich yet underexplored chemical landscape.
For authoritative validation and high-quality reference data, the commercial DNP remains a benchmark. While specific details were unavailable in this analysis, its long-standing reputation for expert curation suggests its continued value for verifying structures and accessing deeply annotated data.
The future of NP database development lies in addressing persistent challenges like data quality, stereochemical accuracy, and the integration of new NP discoveries from literature [40] [38]. The emergence of community curation models, as seen in COCONUT 2.0, and the strategic mapping of regional diversity, as embodied by LANaPDB, represent powerful, complementary approaches to harnessing the full potential of natural products for drug discovery.
Fragment-Based Drug Design (FBDD) has established itself as a powerful approach for identifying initial hit compounds by screening small, low-molecular-weight fragments (typically 100-300 Da) against therapeutic targets [43]. This method allows for a more efficient exploration of chemical space compared to traditional High-Throughput Screening (HTS) of larger, more complex compounds [33]. Meanwhile, Natural Products (NPs) have served as evolutionary-selected ligands for diverse biological targets, providing a rich source of molecular scaffolds with proven biological relevance [44]. The integration of NPs into FBDD represents an innovative strategy to address the limitations of conventional synthetic fragment libraries, which are often dominated by flat, aromatic structures, by introducing three-dimensional, stereochemically rich fragments derived from nature's chemical repertoire [33] [44]. This guide provides a comparative analysis of NP-derived fragments against synthetic alternatives, offering experimental data and methodologies to inform selection for drug discovery campaigns.
The structural and physicochemical properties of fragment libraries fundamentally influence their performance in screening campaigns. The table below presents a quantitative comparison of major fragment libraries, highlighting key distinctions between natural product-derived and synthetic collections.
Table 1: Comparative Analysis of Fragment Libraries from Natural Product and Synthetic Sources
| Library Name | Source / Type | Initial Size | Fragments Fulfilling RO3 | Key Characteristics |
|---|---|---|---|---|
| COCONUT NP-derived [45] | Natural Products | 2,583,127 fragments | 38,747 (1.5%) | Derived from 695,133 unique NPs; high structural diversity |
| LANaPDB NP-derived [45] | Natural Products | 74,193 fragments | 1,832 (2.5%) | Sourced from 13,578 Latin American NPs |
| CRAFT [45] | Synthetic & NP-inspired | 1,214 fragments | 176 (14.6%) | Designed for synthetic accessibility; new heterocyclic scaffolds |
| Enamine (water-soluble) [45] | Commercial Synthetic | 12,505 fragments | 8,386 (67.1%) | High RO3 compliance; optimized for solubility |
| ChemDiv [45] | Commercial Synthetic | 74,721 fragments | 16,723 (23.1%) | Large diverse library |
| Maybridge [45] | Commercial Synthetic | 30,099 fragments | 5,912 (19.8%) | Established fragment collection |
| Life Chemicals [45] | Commercial Synthetic | 65,552 fragments | 14,734 (22.6%) | Extensive fragment inventory |
NP-derived libraries provide access to vast regions of chemical space, with COCONUT alone offering over 2.5 million fragments [45]. However, their lower compliance with the "Rule of Three" (RO3)âa guideline suggesting fragments should have MW â¤300 Da, â¤3 hydrogen bond donors, â¤3 hydrogen bond acceptors, and LogP â¤3 [43]âindicates these fragments often possess greater structural complexity. This complexity is characterized by higher Fsp3 (fraction of sp3 carbons) and increased molecular complexity, which are valuable for exploring three-dimensional binding pockets but may present synthetic challenges [33]. In contrast, commercial synthetic libraries demonstrate significantly higher RO3 compliance (e.g., 67.1% for Enamine), reflecting their design for straightforward screening and optimization [45].
Beyond simple RO3 metrics, deeper analysis of structural properties reveals critical differences between library types.
Table 2: Comparison of Key Structural and Drug-like Properties
| Property | NP-Derived Fragments | Synthetic Fragments |
|---|---|---|
| 3D Shape / Fsp3 | Higher, more stereogenic centers [33] [44] | Often flat, sp2-dominated [33] |
| Structural Diversity | High, evolutionarily selected [33] [44] | Varies, often designed around common scaffolds |
| Synthetic Accessibility | Can be challenging [45] | Generally high, designed for tractability [45] |
| Biological Relevance | Evolutionarily pre-validated [33] [44] | Not inherently biologically relevant |
| Ligand Efficiency | Can inherit high efficiency from parent NPs [33] | Must be optimized |
NP-derived fragments are prized for their structural complexity and three-dimensionality, which can lead to improved selectivity and better physicochemical profiles in resulting drug candidates [33]. Their biosynthetic origins often mean they contain recognition elements for protein binding sites, potentially increasing hit rates for challenging targets [44]. However, this structural complexity can correlate with lower synthetic accessibility scores compared to synthetic fragments designed for straightforward medicinal chemistry optimization [45]. The CRAFT library represents a hybrid approach, incorporating NP-inspired designs with an emphasis on synthetic feasibility [45].
The RETrosynthetic Combinatorial Analysis Procedure (RECAP) is a widely employed computational method for deconstructing natural products into fragments [45] [46]. The protocol involves:
Research indicates that the fragmentation strategy significantly impacts outcomes. The workflow below illustrates two key approaches.
Non-extensive fragmentation generates larger, "intermediate" scaffolds by systematically considering cleavage sites without exhaustive decomposition, preserving more structural context from the parent NP [46] [35]. Studies demonstrate that non-extensive fragmentation of NP libraries yields far more chemical entities (45,355 vs. 11,525 from extensive fragmentation) that are less repetitive and exhibit higher pharmacophore fit scores in virtual screening [46] [35]. These fragments provide superior starting points for optimization through fragment merging or growing strategies.
Multiple biophysical and computational techniques are employed to identify fragment hits, each with distinct strengths.
Table 3: Key Experimental Methods for Fragment Screening
| Method | Principle | Application in FBDD with NPs |
|---|---|---|
| X-ray Crystallography [43] [33] | Direct visualization of fragment binding in protein crystal | Identifies binding mode; ideal for complex NP fragments |
| Nuclear Magnetic Resonance (NMR) [43] [33] | Detects changes in magnetic properties upon binding | Measures weak affinities; used in SAR by NMR |
| Surface Plasmon Resonance (SPR) [43] | Measures change in refractive index near sensor surface | Label-free kinetic characterization of binding |
| Native Mass Spectrometry (NMS) [43] [33] | Detects intact protein-fragment complexes in gas phase | Screens complex NP mixtures against multiple targets |
| Thermal Shift Assay [33] | Measures protein stability change upon ligand binding | Low-cost primary screening |
| Isothermal Titration Calorimetry (ITC) [43] | Quantifies heat change from binding interaction | Provides full thermodynamic profile |
| Bio-Layer Interferometry [43] | Optical technique measuring interference pattern shifts | Label-free kinetic screening |
For NP-derived fragments, Native Mass Spectrometry has been successfully applied to screen natural product libraries against multiple potential drug targets simultaneously, as demonstrated in a study targeting 62 malaria-related proteins [43]. X-ray Crystallography remains the gold standard for providing detailed structural information to guide the optimization of NP fragment hits [43].
Table 4: Key Research Reagents and Computational Tools for FBDD with NP-Derived Fragments
| Resource / Reagent | Type | Function and Relevance | Example Sources |
|---|---|---|---|
| COCONUT Database [45] | Computational Database | Large collection of unique natural product structures for fragmentation | Publicly available |
| LANaPDB [45] | Computational Database | Curated NPs from Latin America; provides regional chemical diversity | Publicly available |
| RDKit Toolkit [45] | Software | Open-source cheminformatics toolkit used for RECAP fragmentation and descriptor calculation | Publicly available |
| RECAP Algorithm [45] [46] | Computational Method | Rule-based fragmentation of molecules for generating virtual fragment libraries | Integrated in RDKit |
| CRAFT Fragment Library [45] | Physical / Virtual Library | Experimentally available fragments based on new heterocyclic and NP-inspired scaffolds | Academic consortium (Univ. of Sao Paulo, Federal Univ. of Goias) |
| Commercial Fragment Libraries (Enamine, ChemDiv, etc.) [45] | Physical Libraries | Commercially available synthetic fragments for experimental screening and comparison | Various vendors |
| Ligand Scout [35] | Software | Used for pharmacophore model generation and virtual screening of fragments | Commercial software |
The integration of NP-derived fragments has shown promise in several therapeutic areas:
Experimental data from virtual screening studies provides quantitative support for the value of NP-derived fragments. A study combining non-extensive fragmentation with pharmacophore-based virtual screening reported that the pharmacophore fit score of non-extensive fragments was not only higher than that of extensive fragments in 56% of cases but was also higher than their original parent natural products in 69% of cases when all were recognized as hits [46] [35]. This suggests that selective fragmentation can isolate and enhance the key pharmacophoric elements of complex natural products.
NP-derived fragments offer a powerful complement to synthetic libraries in FBDD. Their key advantages lie in superior three-dimensionality, high structural diversity, and evolutionary pre-validation, which can be decisive for tackling challenging targets like protein-protein interactions [33] [44]. The main limitations, primarily lower RO3 compliance and potential synthetic complexity, can be mitigated through hybrid approaches like those used in the CRAFT library and advanced computational design of "pseudo natural products" [45] [33].
The future of FBDD with NP-derived fragments will likely involve more sophisticated computational fragmentation algorithms and machine learning models to predict synthetic accessibility and biological activity earlier in the process. Furthermore, the integration of target prediction software for fragment-sized natural products can help prioritize screening efforts [33]. As these tools mature, the systematic exploration of nature's fragment space will undoubtedly accelerate the discovery of novel, effective therapeutics across a wider range of disease areas.
Natural products (NPs) have served as a historic and prolific source of molecular scaffolds for drug discovery, yet their structural complexity often presents challenges for systematic medicinal chemistry optimization [45] [47]. To bridge the gap between the biologically relevant chemical space of NPs and the synthetic accessibility of designed compounds, researchers have developed innovative strategies for creating pseudo-natural products (pseudo-NPs) and NP-inspired synthetic libraries [48] [47]. These approaches aim to retain the favorable biological relevance and performance of NPs while enabling access to unprecedented chemotypes not found in nature [47] [49]. The integration of these design principles with modern chemoinformatic analysis has facilitated the systematic comparison and design of compound libraries that hybridize natural and synthetic characteristics, offering new opportunities for exploring biologically relevant chemical space and discovering first-in-class therapeutics [50] [49].
The pseudo-NP approach centers on the deconstruction of natural product structures into their constituent fragments, followed by the synthetic recombination of these fragments in novel arrangements not accessible through known biosynthetic pathways [48] [47]. This strategy harnesses the evolutionary optimization of NP fragments while creating entirely new structural classes with unique biological activities. As illustrated in a landmark study, researchers combined fragment-sized natural products including quinine, quinidine, sinomenine, and griseofulvin with chromanone or indole-containing fragments to generate a 244-member pseudo-NP collection [48]. Cheminformatic analysis confirmed that these novel compound classes exhibited both drug-like and natural product-like properties while occupying previously unexplored regions of chemical space [48].
The design of pseudo-NPs follows specific connectivity patterns that determine how NP-derived fragments are combined. These patterns include linear fusion, spiro-connections, and hybrid structures that merge fragments through strategic linkage points [47]. The resulting compounds are designed to explore complementary biological space while maintaining the structural complexity and three-dimensionality characteristic of natural products, which is crucial for targeting protein interfaces and allosteric sites often considered "undruggable" [49].
Fragment-based drug design (FBDD) principles provide a methodological foundation for creating NP-inspired libraries. This approach typically utilizes small organic fragments with fewer than 20 non-hydrogen atoms, adhering to the "rule of three" (RO3): molecular weight ⤠300 Da, rotatable bonds ⤠3, topological polar surface area ⤠60 à ², logP ⤠3, hydrogen-bond acceptors ⤠3, and hydrogen-bond donors ⤠3 [45]. These fragments serve as ideal building blocks for constructing more complex molecules.
Several computational methods facilitate the deconstruction of NPs into fragments. The most prominent include:
Large-scale implementation of these methods has enabled the generation of extensive fragment libraries from natural product collections. For instance, researchers have obtained 2,583,127 fragments from the COCONUT database (containing 695,133 unique natural products) and 74,193 fragments from the Latin American Natural Product Database (LANaPDB) [45].
The complexity-to-diversity (CtD) strategy employs complex natural products as starting materials and applies diverse reaction pathways to dramatically alter their core scaffolds, thereby generating structurally diverse compounds from a common precursor [49]. This approach leverages the inherent structural complexity of NPs as a launching point for diversity-oriented synthesis. Key transformations in CtD include ring distortion reactions such as cycloadditions, fragmentations, rearrangements, and scaffold-hopping methodologies that fundamentally reshape the molecular architecture [49].
This strategy has been successfully applied to various natural product classes, generating collections of novel compounds with significant structural variation while maintaining aspects of natural product complexity that are favorable for biological interactions, including sp³-character and stereochemical richness [49].
Comprehensive comparisons of fragment libraries derived from natural products and synthetic compounds reveal distinct characteristics and property distributions. The following table summarizes key statistics from major natural product and synthetic fragment libraries:
Table 1: Composition of Natural Product and Synthetic Fragment Libraries
| Library Source | Initial Number of Fragments | Fragments After Standardization | Fragments Fulfilling RO3 (%) |
|---|---|---|---|
| Natural Product Libraries | |||
| LANaPDB | 74,193 | 74,193 | 1,832 (2.5%) |
| COCONUT | 2,583,127 | 2,583,127 | 38,747 (1.5%) |
| Synthetic Libraries | |||
| CRAFT | 1,214 | 1,202 | 176 (14.6%) |
| Enamine (water soluble) | 12,505 | 12,496 | 8,386 (67.1%) |
| ChemDiv | 74,721 | 72,356 | 16,723 (23.1%) |
| Maybridge | 30,099 | 29,852 | 5,912 (19.8%) |
| Life Chemicals | 65,552 | 65,248 | 14,734 (22.6%) |
The data reveals that while natural product databases yield enormous numbers of fragments, only a small percentage (1.5-2.5%) comply with the rule of three criteria ideal for fragment-based drug design [45]. In contrast, commercial synthetic libraries show significantly higher compliance rates (14.6-67.1%), reflecting their intentional design for drug discovery applications.
Chemoinformatic analysis of these libraries employs multiple descriptors to quantify their positions in chemical space and assess their diversity:
Table 2: Key Descriptors for Chemoinformatic Comparison
| Descriptor Category | Specific Metrics | Application in Library Comparison |
|---|---|---|
| Constitutional | Molecular weight, heavy atom count, rotatable bonds | Assess drug-likeness and flexibility |
| Complexity | Stereocenters, sp³ character, molecular frameworks | Quantify structural complexity and three-dimensionality |
| Physicochemical | LogP, topological polar surface area, H-bond donors/acceptors | Predict solubility, permeability, and bioavailability |
| Diversity Assessment | Tanimoto coefficients using MACCS keys and Morgan fingerprints | Measure structural similarity and library coverage |
Natural product-derived fragments typically exhibit greater structural complexity and three-dimensional character compared to synthetic libraries, which often lean toward flatter, more aromatic structures [45] [50]. This complexity is quantifiable through metrics such as the fraction of sp³-hybridized carbons and the number of stereogenic centers, both of which are generally higher in NP-derived fragments [45].
The creation of pseudo-NP and NP-inspired libraries follows a systematic workflow that integrates computational design with experimental execution. The following diagram illustrates the key stages in this process:
Diagram 1: Experimental Workflow for NP-Inspired Library Design
Before fragmentation and analysis, compound collections undergo rigorous standardization to ensure data quality and comparability. The protocol typically includes:
This standardized protocol ensures that subsequent analyses compare equivalent, high-quality chemical structures across different libraries.
The synthetic accessibility (SA) score is a crucial metric for evaluating the practical utility of designed compounds. The SA score is calculated as the difference between a fragment score (assessing the viability of structural features) and a complexity penalty (accounting for ring complexity, stereocenters, and molecular size) [45]. This score helps prioritize compounds that balance novelty with synthetic feasibility, a critical consideration for library design.
Pseudo-NP libraries are typically evaluated using a combination of phenotypic screening and target identification approaches. A representative study described the unbiased biological evaluation of a 244-member pseudo-NP collection using cell painting assays, which measure morphological changes in cells to assess bioactivity [48]. This phenotypic approach revealed that the bioactivity profiles of pseudo-NPs differed from both their guiding natural products and individual fragments, with the combination of different fragments dominating the establishment of unique bioactivity [48]. This observation underscores the value of fragment combination in exploring novel biological space.
Successful implementation of pseudo-NP and NP-inspired library strategies requires specialized reagents, databases, and computational tools. The following table catalogues key resources for researchers in this field:
Table 3: Essential Research Reagents and Computational Tools
| Resource Category | Specific Resources | Function and Application |
|---|---|---|
| Natural Product Databases | COCONUT (695K compounds), LANaPDB (13.5K compounds) | Source of natural product structures for fragmentation and analysis [45] |
| Synthetic Fragment Libraries | CRAFT (1.2K fragments), Enamine (12.5K fragments), ChemDiv (74.7K fragments) | Commercially available fragments for comparison and hybrid design [45] |
| Cheminformatic Toolkits | RDKit, MolVS | Structure standardization, descriptor calculation, and fragmentation [45] |
| Spectral Libraries | BMDMS-NP (2,739 plant metabolites, 288K MS/MS spectra) | Metabolite identification and structural validation [51] |
| Fragmentation Algorithms | RECAP, BRICS, MORTAR | Systematic deconstruction of compounds into logical fragments [45] |
| Diversity Metrics | MACCS keys, Morgan fingerprints, Tanimoto similarity | Quantification of chemical space coverage and library diversity [45] |
These resources collectively enable the design, construction, and evaluation of NP-inspired libraries through integrated computational and experimental workflows.
The systematic comparison of design strategies for pseudo-natural products and NP-inspired synthetic libraries reveals complementary strengths and applications. Natural product fragments provide access to structurally complex, biologically validated chemotypes that explore wider regions of three-dimensional chemical space, while synthetic libraries offer superior compliance with fragment-based design principles and greater synthetic accessibility [45] [50].
The integration of these approaches through pseudo-NP design and complexity-to-diversity strategies represents a powerful framework for exploring biologically relevant chemical space that remains inaccessible to purely natural or synthetic approaches [48] [47] [49]. These hybrid methodologies leverage the evolutionary optimization embodied in natural product structures while enabling the exploration of unprecedented structural arrangements with novel bioactivities.
Future directions in this field will likely involve more sophisticated computational methods for predicting productive fragment combinations, increased integration of synthetic biology approaches for generating unnatural natural product analogs, and application of these strategies to challenging target classes such as protein-protein interactions and neglected disease targets [45] [49]. As these methodologies continue to evolve, they will undoubtedly expand the toolkit available to medicinal chemists and drug discovery researchers seeking to address unmet medical needs through novel molecular entities.
Fragment-Based Drug Discovery (FBDD) has established itself as a powerful approach for identifying novel chemical starting points in drug development, complementing traditional High-Throughput Screening (HTS). Unlike HTS, which screens large libraries of drug-like molecules, FBDD utilizes small, low-complexity chemical fragments that typically exhibit weaker binding affinity but more efficient, atom-specific interactions with biological targets [52]. This methodology has yielded notable successes, including FDA-approved drugs such as venetoclax, sotorasib, and asciminib, particularly against targets once considered "undruggable" [52].
The strategic importance of FBDD is further amplified when viewed through the lens of chemoinformatic comparisons between natural products (NPs) and synthetic compounds (SCs). Research consistently demonstrates that NPs and NP-inspired structures exhibit greater three-dimensional complexity, increased stereochemical content, and broader coverage of chemical space compared to purely synthetic molecules [53] [54] [3]. These properties are correlated with improved binding selectivity and clinical success rates [53]. Consequently, fragment libraries derived from or inspired by natural products offer a pathway to harness this privileged chemical space, potentially addressing the limited structural diversity that often plagues conventional synthetic screening collections [53] [30]. This guide provides an objective comparison of contemporary fragment libraries, with a specific focus on their design, composition, and performance within this strategic context.
This section details the design principles and key characteristics of several prominent fragment libraries, highlighting their distinct strategic positioning.
The following table summarizes the key physicochemical properties of the discussed libraries, where data is available, and contextualizes them with properties of natural products.
Table 1: Comparative Analysis of Fragment Library Properties and Performance
| Library Name | Size (Compounds) | Key Design Principle | Molecular Weight (Da) | Heavy Atom Count | Key Metrics & Experimental Validation |
|---|---|---|---|---|---|
| CRAFT [30] | 1,214 | Novel heterocycles & NP-derived chem. | Not Specified | Not Specified | Coverage: Designed for broad chemical diversity. Validation: Research-focused; comparative analysis with NP fragments. |
| Enamine High Fidelity [55] | 1,920 | High MedChem Tractability | Not Specified | 9-16 | Solubility: All compounds passed turbidity tests (â¥1 mM). Specificity: SPR-cleaned to remove aggregators/sticky compounds. Design: Expert-curated, Rule of 3 compliant. |
| EU-OPENSCREEN (EFSL) [56] | 1,056 | Poised to parent HTS library (ECBL) | Not Specified | Not Specified | Coverage: Substructures of ~88% of the 96,096 ECBL compounds. Validation: 8 screening campaigns identified hits; case study vs. FabF (PDB: 8PJ0). |
| Typical "Rule of 3" [52] | N/A | Standard Fragment Guidelines | ⤠300 | ⤠20 | H-Bond Donors ⤠3, H-Bond Acceptors ⤠3, cLogP ⤠3, Rotatable Bonds ⤠3 |
| Natural Product Drugs [53] [3] | N/A | Evolved for Biological Relevance | ~611 (NP) ~757 (ND) | Not Specified | Higher Fsp3 (â¥0.59), More Stereocenters, Lower ClogP, Fewer Aromatic Rings vs. Synthetic Drugs |
The property trends of approved natural product-based drugs highlight the potential benefits of libraries that incorporate NP-like features. These drugs are characterized by a higher fraction of sp3-hybridized carbons (Fsp3 ⥠0.59), indicating more three-dimensional, complex structures, and lower hydrophobicity (ClogP) compared to purely synthetic drugs [53] [3]. These characteristics are increasingly recognized as advantageous for drug discovery [53].
Identifying fragment hits requires specialized biophysical techniques due to their typically weak binding affinities (in the μM to mM range), which are often undetectable in conventional biochemical assays [52].
The workflow below illustrates the typical stages of an FBDD campaign that leverages a poised fragment library.
Table 2: Key Reagents and Materials for Fragment-Based Screening
| Item | Function in FBDD |
|---|---|
| Fragment Library | A curated collection of 500-2,000 small molecules (MW < 300) used for the initial screening. The core resource. |
| Target Protein | A highly pure, stable, and functionally active protein preparation for binding assays. |
| Biophysical Instrumentation | Platforms for SPR, NMR, BLI, or X-ray crystallography to detect weak fragment binding. |
| High-Quality DMSO | Universal solvent for fragment stock solutions; must be of high purity to avoid interference. |
| Assay Buffers | Physiologically relevant buffers (e.g., PBS) to maintain protein stability and function during screening. |
| Poised HTS Library | A larger compound collection (e.g., EU-OPENSCREEN's ECBL) containing fragment-derived compounds for rapid follow-up. |
| o-Vanillin | o-Vanillin|2-Hydroxy-3-methoxybenzaldehyde [148-53-8] |
The choice of a fragment library is a strategic decision that can significantly influence the outcome of a drug discovery program. Our analysis reveals that libraries are often optimized for different objectives:
The enduring influence of natural products on drug discovery provides a critical framework for evaluating fragment libraries. Approved drugs based on NPs are consistently characterized by greater three-dimensionality (higher Fsp3), increased stereochemical complexity, and lower hydrophobicity [53] [3]. These properties are correlated with improved clinical success rates [53]. Therefore, libraries that incorporate NP-derived scaffolds or are designed to mimic these favorable properties offer a tangible advantage. They provide a means to escape the flat, aromatic-rich landscape of many synthetic libraries and venture into the more diverse and biologically relevant chemical space occupied by natural products [53] [30] [54].
The evolving landscape of fragment libraries reflects a broader shift in drug discovery toward leveraging privileged chemical structures, with natural products serving as a key inspiration. The comparative analysis presented here underscores that there is no single "best" library; rather, the optimal choice depends on the project's specific goals, target class, and available resources. As FBDD continues to mature, the integration of NP-like complexity, rigorous experimental validation, and smart library poising will be key drivers in discovering innovative therapeutics for the most challenging diseases.
Natural Products (NPs) are a cornerstone of modern therapeutics, with over half of all approved small-molecule drugs originating directly or indirectly from them [53]. However, their transition from promising lead to viable drug candidate is often hampered by significant synthetic challenges. NPs frequently possess complex architectures characterized by high molecular complexity, numerous stereocenters, and intricate ring systems, which complicate both total synthesis and large-scale production [57] [22]. This guide objectively compares the structural and physicochemical properties of NPs and synthetic compounds, providing experimental frameworks to assess and mitigate synthetic accessibility challenges during drug development.
A principal component analysis of structural and physicochemical features reveals distinct profiles for drugs derived from natural products versus completely synthetic origins [53].
Table 1: Structural and Physicochemical Property Comparison
| Parameter | Natural Product-Derived Drugs | Completely Synthetic Drugs |
|---|---|---|
| Molecular Size & Complexity | ||
| Molecular Weight (MW) | Larger | Smaller |
| Fraction sp3 (Fsp3) | Higher (more complex 3D structures) | Lower (flatter, more 2D structures) |
| Stereocenters (nStereo) | Greater number | Fewer |
| Structural Features | ||
| Aromatic Rings | Fewer | More |
| Ring Systems | More complex, fused, macro rings | Simpler |
| Chiral Centers | More prevalent | Less prevalent |
| Glycosylation | 8%-22% of NPs [22] | ~1.85% of bioactive compounds [22] |
| Physicochemical Properties | ||
| Hydrophobicity (LogD) | Lower | Higher |
| Polarity | Increased | Reduced |
| Hydrogen Bond Donors/Acceptors | More | Fewer |
| Chemical Space | Broader, more diverse [53] [22] | More confined |
The data shows that NPs exhibit greater three-dimensional complexity and occupy a broader region of chemical space than synthetic compounds, contributing to their ability to interact with diverse biological targets [53]. However, these desirable biological features often come at the cost of synthetic feasibility.
The SAscore is a widely used metric to estimate the ease of synthesizing a given molecule, ranging from 1 (very easy) to 10 (very difficult) [58] [59].
Protocol:
fragmentScore = Sum of all fragment contributions / Number of fragmentsSAscore = fragmentScore + complexityPenaltySupporting Tools:
sascorer.py: A Python implementation of the Ertl & Schuffenhauer method [58].The following diagram outlines an integrated workflow for evaluating and prioritizing leads based on synthetic accessibility and other key drug discovery metrics.
When the SAscore indicates high complexity, structural simplification is a key strategy. This involves systematically truncating unnecessary groups from a complex lead to improve synthetic accessibility while retaining biological activity [57].
Case Study Protocol: Simplification of Halichondrin B to Eribulin
Table 2: Key Resources for NP Research and Synthetic Assessment
| Resource Name | Type | Primary Function |
|---|---|---|
| COCONUT | Database | A comprehensive open-source database of over 695,000 unique Non-redundant Natural Products for chemoinformatic analysis [60]. |
| InflamNat | Database & Predictor | A specialized database of anti-inflammatory NPs with machine learning tools to predict anti-inflammatory activity and compound-target interactions [61]. |
| PubChem | Database | A vast repository of chemical molecules and their biological activities, used for fragment frequency analysis in SAscore calculation [59]. |
| RDKit | Software | Open-source cheminformatics software that includes the sascorer.py module for calculating Synthetic Accessibility scores [58]. |
| Neurosnap eTox | Web Tool | Predicts both toxicity probability and a Synthetic Accessibility score (1-10) for input molecules [58]. |
| Mordred | Descriptor Calculator | Calculates ~1,614 molecular descriptors useful for building custom models or heuristically assessing complexity [58]. |
| MacrolactoneDB | Database | A curated database of over 13,700 macrolactone NPs for studying this complex chemotype [22]. |
The journey from a complex Natural Product lead to a synthetically tractable drug candidate requires a careful balance. While NPs offer unparalleled chemical diversity and biological relevance, their inherent complexity often poses significant production challenges. By employing the described cheminformatic comparisons, experimental protocols for SAscore calculation, and strategic simplification workflows, researchers can make data-driven decisions to prioritize leads with an optimal balance of biological potential and synthetic feasibility. Integrating these assessments early in the drug discovery pipeline de-risks development and enhances the efficiency of creating NP-derived therapeutics.
The Convention on Biological Diversity (CBD), adopted at the 1992 Earth Summit in Rio de Janeiro, establishes a comprehensive international framework for biodiversity conservation, sustainable use of biological components, and fair benefit-sharing from genetic resources [62] [63]. As a legally binding treaty with near-universal participation (196 parties, including 195 states and the European Union), the CBD represents a fundamental shift in global environmental policy by linking conservation efforts with sustainable development principles [62]. The United States stands as the only UN member state that has signed but not ratified the convention, primarily due to domestic political constraints [62].
The Nagoya Protocol on Access and Benefit-Sharing (ABS), adopted in 2010 as a supplementary agreement to the CBD, entered into force in October 2014 and has been ratified by 142 parties as of 2025 [64]. This protocol provides a detailed legal framework implementing the CBD's third objective: ensuring fair and equitable sharing of benefits arising from the utilization of genetic resources, thereby recognizing national sovereignty over biological resources and combating biopiracy [64] [65]. The protocol emerged partly in response to historical practices in industries like pharmaceuticals where commercial entities exploited natural and indigenous resources without fair compensation [65].
The Convention on Biological Diversity operates through three interconnected objectives that form its foundational framework [63]:
The Nagoya Protocol establishes specific legal obligations for its contracting parties through several core mechanisms [64] [63]:
Table 1: Benefit-Sharing Mechanisms Under the Nagoya Protocol
| Benefit Type | Specific Examples | Applicable Contexts |
|---|---|---|
| Monetary Benefits | Royalties, license fees, access fees, research funding | Commercial product development, pharmaceutical applications |
| Non-Monetary Benefits | Technology transfer, scientific collaboration, capacity building | Academic research, institutional partnerships |
| Knowledge-Related Benefits | Joint authorship, co-patenting, sharing of research results | Collaborative research projects with provider countries |
| Social Recognition Benefits | Acknowledgments in publications, institutional affiliations | All research contexts involving external genetic resources |
Effective implementation of the CBD and Nagoya Protocol occurs primarily at the national level through several key mechanisms [62] [63]:
The European Union has implemented the Nagoya Protocol through a specific regulation requiring scientists to file Due Diligence Declarations to national authorities when biological resources are used in connection with funded research projects [64]. Compliant culture collections, such as the Leibniz Institute DSMZ, certify that resources are "Nagoya compliant" and provide documentation needed for regulatory compliance [64].
The following diagram illustrates the systematic workflow researchers must follow to ensure compliance with Nagoya Protocol requirements when accessing and utilizing genetic resources:
Table 2: Essential Research Reagents and Compliance Tools for ABS Implementation
| Tool/Reagent | Primary Function | Regulatory Application |
|---|---|---|
| ABS Clearing-House | Information platform for regulatory requirements | Verification of national ABS measures, NFPs, and CNAs [66] |
| Internationally Recognized Certificate of Compliance | Documentation of legal provenance | Proof of PIC and MAT establishment for genetic resources [66] |
| Material Transfer Agreement (MTA) | Contract governing resource transfers | Specifies permitted uses and benefit-sharing obligations [64] |
| Due Diligence Declaration | Compliance attestation | Required documentation for EU-funded research projects [64] |
| Nagoya-Compliant Culture Collections | Resource repositories with verified compliance | Certified biological materials with complete documentation (e.g., DSMZ) [64] |
The regulatory frameworks established by the CBD and Nagoya Protocol create fundamentally different research environments for natural products compared to synthetic compounds, with significant implications for chemoinformatic comparisons:
Recent chemoinformatic analyses highlight that natural products exhibit greater structural diversity and complexity compared to synthetic compounds, with higher molecular complexity indices, more chiral centers, and distinctive structural features like glycosylation (present in 8%-22% of natural products versus only 0.23%-4.93% in synthetic compounds) [22]. These inherent structural differences, combined with divergent regulatory frameworks, create complementary but distinct research paradigms.
For researchers conducting comparative analyses of natural products and synthetic compounds within CBD/Nagoya compliance frameworks, the following standardized protocol ensures regulatory adherence while generating robust scientific data:
Resource Sourcing and Documentation
Chemical Library Curation
Chemical Space Mapping
Diversity and Complexity Assessment
Bioactivity Prediction and Target Annotation
The development of fragment libraries for drug discovery illustrates how chemoinformatic research can proceed within Nagoya Protocol constraints. Recent research has generated comprehensive fragment libraries from large natural product databases while maintaining regulatory compliance [60] [30]:
Comparative chemoinformatic analysis of these libraries reveals that natural product-derived fragments occupy broader chemical space with higher structural diversity than synthetic counterparts, particularly in under-represented regions of chemical space characterized by complex stereochemistry and unique ring systems [60] [22]. This structural diversity makes them valuable for probing novel biological targets, despite the additional regulatory requirements for their utilization.
Despite their conservation objectives, the CBD and Nagoya Protocol face significant implementation challenges and scientific criticisms:
The research community has developed several strategies to navigate these regulatory frameworks while advancing scientific discovery:
The CBD and Nagoya Protocol represent evolving frameworks that continue to shape how researchers access, utilize, and share benefits from genetic resources. As chemoinformatic comparisons between natural and synthetic compounds advance, maintaining awareness of these regulatory dimensions remains essential for conducting ethically compliant and scientifically rigorous research that contributes to both drug discovery and biodiversity conservation.
Natural products (NPs) have been an invaluable source of therapeutic agents, with approximately half of all approved small-molecule drugs tracing their structural origins to NPs [53]. However, the NP discovery process is plagued by the persistent challenge of rediscovering known compounds, a problem that necessitates laborious "dereplication" to identify novel chemical entities [67]. Dereplicationâthe process of using chromatographic and spectroscopic analysis to recognize previously isolated substances present in an extractâhas become a critical strategy for prioritizing novel bioactive compounds early in the discovery pipeline [68]. The significance of efficient dereplication is magnified by the substantial costs and time investments required for natural product research, particularly given that biological extracts represent complex mixtures where known compounds frequently mask the presence of novel bioactive agents [69] [68]. This guide provides a comprehensive comparison of contemporary dereplication strategies, evaluating their performance, applications, and implementation requirements to assist researchers in selecting optimal approaches for their specific discovery contexts.
Understanding the fundamental structural differences between natural products and synthetic compounds provides essential context for dereplication strategy development. Comparative analyses reveal that NPs occupy distinct and more diverse regions of chemical space compared to synthetic compounds, with characteristic structural features that influence both their biological activity and the appropriate methods for their identification [53] [54].
Table 1: Structural and Physicochemical Comparison of Natural Products and Synthetic Compounds
| Parameter | Natural Products | Synthetic Compounds | Analytical Implications |
|---|---|---|---|
| Molecular Complexity | Higher stereochemical complexity (more stereocenters) [53] | Lower stereochemical complexity [53] | Requires stereosensitive analytical techniques |
| Structural Features | Greater fraction of sp³ carbons (Fsp³) [53]; More oxygen atoms [54] | More aromatic rings; More nitrogen atoms [54] | Influences fragmentation patterns in MS |
| Chemical Space | Larger, more diverse chemical space [53] [54] | More restricted, defined chemical space [54] | Necessitates comprehensive reference databases |
| Temporal Evolution | Increasing molecular size and complexity over time [54] | Constrained by drug-like rules and synthetic accessibility [54] | Requires continuously updated databases |
The structural evolution of NPs over time presents an additional challenge for dereplication. Recent studies demonstrate that newly discovered NPs have become larger, more complex, and more hydrophobic compared to their historical counterparts, a trend attributed to technological advancements in separation and structure elucidation techniques [54]. This temporal evolution necessitates continuously updated dereplication databases and methods capable of addressing increasingly complex molecular architectures.
Modern dereplication employs an integrated array of analytical and computational approaches, each with distinct advantages, limitations, and implementation requirements. The table below provides a systematic comparison of the primary dereplication platforms and their performance characteristics.
Table 2: Performance Comparison of Dereplication Platforms and Methodologies
| Platform/Methodology | Key Features | Throughput | Chemical Coverage | Implementation Complexity |
|---|---|---|---|---|
| LC-HRFTMS & NMR Metabolomics [70] | High-resolution mass spectrometry coupled with NMR profiling | High | Broad, untargeted | High (requires specialized expertise) |
| Antibiotic Resistance Platform (ARP) [67] | Cell-based array of resistance mechanisms; dual use for dereplication and adjuvant discovery | Medium | Targeted (antibiotics) | Medium |
| SFC-MS [68] | Supercritical fluid chromatography-MS; minimal solvent use; rapid isolation | High | Moderate to broad | Medium |
| Molecular Networking [69] | MS/MS similarity-based visualization of chemical relationships | High | Broad, untargeted | Medium to High |
| AI-Enhanced Dereplication [71] | Machine learning and deep learning for pattern recognition | Very High | Broad, expanding with data | High (requires computational resources) |
The performance of dereplication platforms varies significantly across key operational parameters. LC-HRFTMS (Liquid Chromatography-High Resolution Fourier Transform Mass Spectrometry) coupled with NMR spectroscopy represents a high-performance approach capable of generating comprehensive chemical profiles of complex extracts [70]. This method offers superior sensitivity and resolution but requires substantial instrumentation investment and specialized expertise. In comparison, the Antibiotic Resistance Platform (ARP) provides a targeted biological dereplication approach specifically for antimicrobial discovery, using an array of mechanistically distinct resistance elements to rapidly identify known antibiotic classes [67]. While offering lower throughput than purely analytical methods, ARP provides valuable functional information alongside chemical identification.
Emerging approaches such as SFC-MS (Supercritical Fluid Chromatography-Mass Spectrometry) offer advantageous environmental profiles with reduced solvent consumption and faster analysis times compared to conventional LC-MS methods [68]. The orthogonality of SFC separation to reversed-phase LC further enhances its utility in comprehensive dereplication workflows. AI-enhanced dereplication platforms represent the most scalable approach, leveraging machine learning algorithms to rapidly identify known compounds from complex spectral data [71]. These systems benefit from continuous improvement as additional data becomes available, though they require significant computational infrastructure and training data curation.
This established protocol integrates high-resolution mass spectrometry with NMR spectroscopy for comprehensive metabolite profiling [70]:
Sample Preparation: Extract biological material (microbial, plant, or marine) using standardized solvent systems (e.g., methanol-dichloromethane 1:1). Concentrate extracts under reduced temperature and pressure to prevent degradation of thermolabile compounds.
LC-HRFTMS Analysis:
Data Processing:
NMR Validation:
This advanced protocol integrates artificial intelligence with molecular networking for high-throughput dereplication [69] [71]:
Data Acquisition:
Molecular Network Construction:
AI-Assisted Annotation:
Validation and Isolation Prioritization:
Diagram 1: Integrated Dereplication Workflow. The diagram illustrates the decision points in a comprehensive dereplication pipeline, highlighting pathways for both known compound identification and novel compound discovery.
Successful implementation of dereplication strategies requires access to specialized databases, analytical tools, and computational resources. The following table catalogues essential research reagents and their applications in modern dereplication workflows.
Table 3: Essential Research Reagents and Resources for Dereplication
| Resource Category | Specific Examples | Function in Dereplication | Access Mode |
|---|---|---|---|
| Spectral Databases | GNPS, AntiMarin, MarinLit [70] | MS/MS spectrum matching for compound identification | Web-based platforms |
| NMR Databases | NMR data from AntiBase [70] | Reference chemical shifts for structure verification | Commercial software |
| Compound Databases | Dictionary of Natural Products [54] | Structural information for known NPs | Commercial license |
| Data Processing Tools | MZmine [70], SIEVE [70] | LC-MS data preprocessing and feature detection | Open source / Commercial |
| Molecular Networking | GNPS [69] | MS/MS similarity-based visualization | Web-based platform |
| AI/ML Platforms | InsilicoGPT [71] | AI-assisted compound annotation and prediction | Web-based access |
Dereplication technologies have evolved from simple chromatographic comparison to integrated multi-platform approaches combining advanced separation techniques, high-resolution spectroscopy, and artificial intelligence. The continuing challenge of rediscovery in natural product screening necessitates increasingly sophisticated strategies that can rapidly identify novel chemical entities while efficiently recognizing known compounds. Future developments will likely focus on enhanced integration of AI and machine learning algorithms, expansion of curated spectral databases, and development of more automated platforms that minimize manual intervention. As natural product discovery continues to explore underexplored biological sources and extreme environments, the role of efficient dereplication will only grow in importance for ensuring that resource-intensive isolation efforts are directed toward truly novel and biologically relevant chemical entities.
The pursuit of new chemical entities (NCEs) in drug discovery navigates two primary landscapes: natural products (NPs) and synthetic compounds. Cheminformatic analyses reveal that these two classes possess distinctly different structural and physicochemical properties [53] [72]. Drugs based on natural products, which constitute approximately half of all NCEs approved in recent decades, demonstrate greater three-dimensional complexity, lower hydrophobicity, and increased presence of stereogenic centers compared to their purely synthetic counterparts [53]. These very characteristics, which are often linked to improved target selectivity and clinical success rates, also introduce significant technical challenges across the workflowâfrom initial purification and analytical characterization to the consistent resupply of these complex materials for research and development [72]. This guide objectively compares the methodologies employed to overcome these hurdles, providing a structured comparison of experimental approaches and their outcomes.
The isolation of pure chemical entities from complex biological matrices is a foundational step. The chosen purification strategy must be tailored to the nature of the source material, whether it is a natural extract or a synthetic reaction mixture.
The following table summarizes common purification methods, highlighting their applicability to the distinct challenges of natural product and synthetic compound workflows.
Table 1: Comparison of Purification Techniques for Natural and Synthetic Compounds
| Purification Method | Principle | Typical Throughput | Suitability for NPs | Suitability for Synthetic Compounds | Key Limitations |
|---|---|---|---|---|---|
| Ethanol/Isopropanol Precipitation | Reduces DNA solubility via alcohol and salt, causing precipitation [73]. | Low | High (for genomic DNA) | Low | Time-consuming, manual, highly variable, low reproducibility [73]. |
| Spin Column Purification | DNA binding to a silica membrane via centrifugation, with washing and elution [73]. | Medium | Medium (post-PCR clean-up) | High | Risk of membrane clogging, requires minimum elution volume (30-50 μl) [73]. |
| Magnetic Bead Purification | DNA binding to paramagnetic beads, separation via magnet [73]. | High (scalable to 384-well plates) | High | High | Bead aspiration can cause sample loss; equipment cost varies from low (magnetic stand) to high (full automation) [73]. |
| Size-Exclusion Chromatography (SEC) | Separates particles based on size and shape [74]. | Medium | High (gentle polishing step) | Medium | Primarily used as a final polishing step, not for primary isolation [74]. |
| Ion-Exchange Chromatography (IEX) | Separates particles based on net surface charge [74]. | Medium | High | High | May not achieve the high purity required for all applications without a subsequent polishing step [74]. |
| Affinity Chromatography | Highly specific separation using an immobilized ligand [74]. | Medium | High (for specific targets) | Medium | High cost of ligands, may not be easily scalable for large-volume production [74]. |
This protocol, adapted from neutrophil isolation studies, exemplifies a technique critical for obtaining pure cell populations prior to downstream analysis of cell-specific natural products or metabolites [75].
Diagram 1: Workflow for cell isolation via density gradient centrifugation.
Accurately characterizing the structural and physicochemical properties of compounds is essential for understanding their activity. Cheminformatic analysis relies on high-quality data derived from these characterization techniques.
Analysis of approved drugs (1981-2019) reveals distinct property profiles, as shown in the table below. These differences directly influence the choice of characterization strategies [72].
Table 2: Physicochemical Properties of Natural Product-Based vs. Synthetic Drugs
| Parameter | Natural Product Drugs (N) | Natural Product-Derived Drugs (ND) | Synthetic Drugs (S, S/NM) | Analysis Implication |
|---|---|---|---|---|
| Molecular Weight (MW) | 611 | 757 | 355-444 | Techniques like LC-MS must be optimized for a broader mass range for NPs. |
| Hydrogen Bond Donors (HBD) | 5.9 | 7.0 | 1.1-1.9 | NMR is critical for identifying complex H-bonding networks in NPs. |
| Hydrogen Bond Acceptors (HBA) | 10.1 | 11.5 | 3.9-5.1 | Polarity-based separation (HPLC) requires robust methods for highly functionalized NPs. |
| Fraction sp3 (Fsp³) | 0.71 | 0.59 | 0.33 | Higher 3D character in NPs necessitates techniques like X-ray crystallography for conformation analysis. |
| Aromatic Rings (RngAr) | 0.7 | 1.4 | 2.3-2.7 | Synthetic compounds are "flatter," making UV detection more straightforward. |
| Calculated LogD | -1.40 | -3.00 | 2.37-2.49 | NP drugs are more hydrophilic, impacting reverse-phase chromatography conditions. |
Characterizing nanoparticles like EVs, which are part of the cellular secretome, is a key challenge in natural product research. Adherence to MISEV guidelines is critical [76].
The transition from a characterized lead compound to a reliable source for biological testing presents a final set of challenges, particularly for natural products.
While direct data on chemical compound resupply is limited in the provided results, principles from adjacent fields highlight universally applicable solutions. In the HME (Home Medical Equipment) and defense sectors, automation and data-driven management are key to overcoming resupply inefficiencies [77] [78].
Diagram 2: An automated no-touch resupply workflow.
Successful navigation of the technical hurdles in this field requires a suite of specialized reagents and tools.
Table 3: Key Research Reagent Solutions for Purification and Characterization
| Reagent/Material | Function | Specific Application Example |
|---|---|---|
| Density Gradient Media | Separates cells or particles based on buoyant density [75]. | Isolation of neutrophils from whole blood using Histopaque-1119 or Polymorphprep [75]. |
| Immunomagnetic Beads | Isolate specific cell types via negative or positive selection [75]. | Obtaining untouched, high-purity neutrophils for transcriptomic studies [75]. |
| Sephadex Resin | Gel filtration resin for size-based purification of biomolecules [73]. | Desalting or removing primers and nucleotides from PCR products [73]. |
| Iodixanol Gradient | Non-ionic, low-toxicity medium for density-based separation [74]. | Purification of adeno-associated virus (AAV) vectors from empty capsids [74]. |
| Affinity Chromatography Ligands | Enable highly specific binding and purification of target molecules [74]. | Isolation of specific AAV serotypes or removal of empty capsids during gene therapy vector production [74]. |
| MISEV Guidelines | Standardized framework for EV research [76]. | Ensuring rigorous characterization of adipocyte-derived EVs (Ad-EVs) using markers like PLIN1 [76]. |
| Tetraspanin Antibodies | Detect specific surface proteins on extracellular vesicles [76]. | Characterizing EV subtypes (e.g., CD63, CD81) via Western blot or flow cytometry [76]. |
The strategic design of compound libraries is a critical foundation for successful drug discovery campaigns. A central challenge in this process involves balancing the competing demands of structural diversity with the need to maintain favorable drug-like properties. This guide provides a chemoinformatic comparison of libraries derived from natural products (NPs) versus those of completely synthetic origin (S), offering an objective framework for selecting and optimizing screening collections. Small-molecule drugs currently target a surprisingly limited range of biological proteins, a limitation exacerbated by the constrained chemical diversity present in many synthetic discovery libraries [53]. These synthetic libraries are often biased by synthetic accessibility and strict adherence to "drug-like" rules such as Lipinski's Rule of Five, resulting in collections replete with structurally similar compounds [53].
In contrast, natural products and their derivatives have historically been a rich source of therapeutic agents, accounting for approximately half of all new small-molecule drug approvals over the past several decades [53] [1]. NPs originate from biological systems and possess evolutionary optimization for biological interaction, offering unparalleled structural diversity and complexity that often accesses biological target space beyond the reach of synthetic compounds [79] [53]. This guide systematically compares these two structural paradigms through experimental data and cheminformatic analysis to inform the optimization of future library design strategies that effectively balance diversity with drug-likeness.
Comprehensive analysis of approved drugs reveals distinct structural differences between natural product-derived compounds and completely synthetic drugs. The following table summarizes key physicochemical parameters for drugs approved between 1981-2010, categorized by origin [53].
Table 1: Physicochemical Properties of Approved Drugs by Origin (1981-2010)
| Parameter | Natural Product (NP) | Natural Product-Derived (ND) | Synthetic, NP-Inspired (S*) | Completely Synthetic (S) |
|---|---|---|---|---|
| Molecular Weight | Higher | Higher | Moderate | Lower |
| Fraction sp3 (Fsp3) | Higher | Higher | Moderate | Lower |
| Chiral Centers | Significantly Higher | Higher | Moderate | Fewer |
| Aromatic Rings | Fewer | Fewer | Moderate | More Prevalent |
| Hydrogen Bond Donors/Acceptors | Higher | Higher | Moderate | Lower |
| Calculated LogP | Lower | Lower | Moderate | Higher |
Natural product-derived drugs consistently exhibit greater structural complexity, as evidenced by higher molecular weights, more chiral centers, and increased Fsp3 character (fraction of sp3-hybridized carbons) [53]. This complexity correlates with improved selectivity and better clinical trial success rates [53]. Furthermore, NPs and their derivatives typically contain fewer aromatic rings but more oxygen atoms and hydrogen bond donors/acceptors, resulting in lower hydrophobicity compared to synthetic drugs [53].
The following table provides a quantitative profile of pure natural products (PNP) versus synthetic compounds, highlighting key differences relevant to library design [79].
Table 2: Detailed Molecular Descriptor Comparison
| Descriptor | Pure Natural Products (PNP) | Synthetic Compounds (LC) |
|---|---|---|
| Mean MW | 393.9 | 389.2 |
| Mean HAC | 28.2 | 27.7 |
| Mean ClogP | 2.3 | 3.6 |
| H-bond Donors | 2.7 | 1.4 |
| H-bond Acceptors | 6.6 | 4.2 |
| TPSA | 98.9 | 79.8 |
| Ring Count | 3.6 | 3.9 |
| Aromatic Rings | 5.1 | 15.3 |
| Rotatable Bonds | 5.2 | 5.0 |
| Number of N atoms | 0.7 | 2.6 |
| Number of O atoms | 5.9 | 3.1 |
| Number of Chiral Atoms | 5.5 | 1.3 |
| Lipinski Violations â¥2 | 18% | 2% |
While natural products show a higher incidence of Lipinski's rule violations, many remain orally bioavailable, suggesting these "rule-based" conventions may be overly restrictive when applied to complex natural product-inspired structures [79] [53]. The data reveals that natural products occupy a distinct region of chemical space characterized by greater stereochemical complexity, higher oxygen content, and reduced aromatic character compared to synthetic compounds [79].
Objective: To quantify and compare the structural complexity and diversity of compound libraries using standardized cheminformatic metrics.
Methodology:
Interpretation: Libraries with lower internal diversity (average Tanimoto similarity <0.15) and broader chemical space coverage are considered more diverse. Natural product libraries typically exhibit greater scaffold diversity and occupy broader regions of chemical space [53].
Objective: To evaluate how closely synthetic compounds resemble natural products using computational natural product-likeness scoring.
Methodology:
Interpretation: Higher natural product-likeness scores indicate closer resemblance to natural products. This method can prioritize compounds from synthetic libraries that are more likely to possess NP-like bioactivity and complexity [79].
Objective: To evaluate the feasibility of chemical synthesis for library compounds, a practical consideration for lead optimization.
Methodology:
Interpretation: Lower SA Scores indicate easier synthesis. While natural products often have higher SA Scores due to complexity, NP-inspired synthetic compounds typically show improved synthetic accessibility while retaining desirable NP-like properties [19].
Library Design Workflow
The workflow illustrates the integrated approach to library design, beginning with both natural product and synthetic compound sources, progressing through comprehensive cheminformatic profiling and complexity assessment, and culminating in optimized libraries that balance diversity with drug-likeness.
Table 3: Key Research Reagent Solutions for Library Design and Analysis
| Resource | Type | Function | Application |
|---|---|---|---|
| COCONUT Database [19] | Natural Product Database | Comprehensive collection of ~400,000 natural products | Reference set for NP-likeness scoring and library design |
| ColorBrewer [80] [81] | Color Palette Tool | Accessible color schemes for data visualization | Creating accessible visualizations of chemical space |
| RDKit [19] | Cheminformatics Toolkit | Open-source cheminformatics software | Calculating molecular descriptors and fingerprints |
| NP-likeness Calculator [79] | Scoring Algorithm | Quantifies similarity to natural products | Prioritizing NP-like compounds from synthetic libraries |
| Chroma.js Palette Helper [81] | Color Accessibility Tool | Tests color vision deficiency accessibility | Ensuring visualizations are accessible to all researchers |
Artificial intelligence is revolutionizing natural product-based drug discovery through several innovative approaches. Chemical language models such as GPT-based architectures can be fine-tuned on natural product databases (e.g., COCONUT) to generate novel natural product-like compounds [19]. These models learn the structural patterns and complexities of natural products and can propose new structures that occupy similar chemical space. The NPGPT model exemplifies this approach, generating compounds with distributions similar to real natural products, as measured by metrics like Fréchet ChemNet Distance (FCD) [19]. AI methods also enhance dereplication (identifying known compounds early in the discovery process) and predict bioactive molecules from vast chemical libraries, significantly accelerating the discovery timeline [71].
The pseudo-natural product (pseudo-NP) strategy represents a powerful fusion of natural product inspiration with synthetic feasibility. This approach involves deconstructing complex natural products into fragments and recombining them into novel scaffolds that retain biological relevance but possess improved synthetic accessibility [82]. Similarly, the complexity-to-diversity (CtD) approach uses complex natural product-inspired starting materials and transforms them into structurally diverse compound collections through efficient synthetic routes [82]. These strategies successfully address the historical challenges of natural product-based drug discoveryâparticularly synthetic complexity and supply limitationsâwhile preserving the privileged bioactivity and structural features of natural products.
Based on comprehensive chemoinformatic comparisons, optimal library design should strategically integrate natural product-inspired compounds with synthetic molecules to maximize both diversity and drug-likeness. Natural product-derived libraries provide access to broader chemical space and increased structural complexity, enabling engagement with more challenging biological targets. However, completely synthetic libraries often demonstrate superior compliance with conventional drug-like rules and synthetic accessibility. The most effective approach involves either hybrid libraries that combine both structural paradigms or the generation of NP-inspired synthetic compounds that balance complexity with synthetic feasibility. Emerging AI-driven generation methods and pseudo-natural product strategies offer powerful tools for creating such optimized libraries, potentially unlocking new therapeutic opportunities for complex diseases.
The pursuit of chemical diversity is a cornerstone of successful small-molecule drug discovery. This guide provides a detailed cheminformatic comparison of New Chemical Entities (NCEs) approved between 1981 and 2019, focusing on drugs derived from Natural Products (NPs) versus those of purely synthetic origin. The analysis is framed within the broader thesis that NPs provide privileged scaffolds that significantly enhance the chemical space and target diversity available for therapeutic development. Nearly half of all small-molecule drugs approved over the last four decades trace their structural origins to a natural product, underscoring their enduring impact [83] [84]. This guide objectively compares the structural and physicochemical properties, clinical success rates, and inherent challenges of these two distinct drug classes, providing researchers with the data and methodologies needed to inform their discovery strategies.
Analysis of drug approvals from 1981 to 2019 reveals that NPs and their derivatives are a major source of new medicines. A foundational review by Newman and Cragg showed that of all approved small-molecule drugs in this period, only about a quarter were purely synthetic (S), while the rest were related to NPs: approximately 5% were unaltered NPs (NP), 28% were NP derivatives (ND), and 35% were synthetic compounds containing an NP pharmacophore (S*) [83] [2]. This distribution has remained relatively consistent over time, demonstrating the sustained value of NPs in inspiration for new drugs [84].
Beyond approvals, recent data on clinical trial pipelines indicate a significant advantage for NP-inspired compounds. While synthetic compounds dominate the initial stages of drug discovery (comprising ~77% of patent applications), their proportion decreases as candidates advance through clinical phases [85] [86]. Conversely, the proportion of NP and NP-derived compounds increases from approximately 35% in Phase I to about 45% in Phase III [85]. This inverse trend suggests that NP-based candidates have a higher likelihood of success, often attributed to their more favorable toxicity profiles and superior drug-like properties honed by evolution [85].
Table 1: Clinical Trial Success and Patent Trends for NP-Derived vs. Synthetic Drugs
| Category | Phase I | Phase III | Patent Applications | Key Implication |
|---|---|---|---|---|
| NP & NP-Derived | ~35% | ~45% | ~23% | Higher clinical success rate; evolutionary optimization |
| Purely Synthetic | ~65% | ~55% | ~77% | Higher attrition in late-stage clinical trials |
A principal component analysis of structural and physicochemical parameters highlights fundamental differences between NP-derived and purely synthetic drugs. NP-derived drugs (including NP, ND, and S* categories) consistently occupy a broader region of chemical space and exhibit greater structural diversity than their synthetic counterparts (S) [84].
Table 2: Key Cheminformatic Properties of NP-Derived vs. Purely Synthetic Drugs
| Physicochemical Property | NP-Derived Drugs | Purely Synthetic Drugs | Biological Significance |
|---|---|---|---|
| Molecular Complexity (Fsp3) | Higher | Lower | Correlated with improved clinical success and binding selectivity [84] |
| Stereochemical Centers | More numerous | Fewer | Associated with selective target engagement [84] |
| Aromatic Ring Count | Lower | Higher | Reflects a bias in synthetic library design [84] |
| Hydrophobicity (ALOGPs, LogD) | Generally lower | Generally higher | May contribute to more favorable solubility and toxicity profiles [85] [84] |
| Molecular Weight/Size | Often larger | Often smaller | NP-derived drugs more frequently violate Rule-of-5, yet remain orally bioavailable [84] |
These properties are not merely structural curiosities; they have practical implications. Higher molecular complexity (quantified as Fsp3, the fraction of sp3-hybridized carbons) and greater stereochemical content have been statistically correlated with a higher probability of successful progression from lead discovery to drug approval [84]. The lower hydrophobicity observed in NP-derived compounds is a likely contributor to their reduced in vitro and in silico toxicity, which in turn explains part of their increased clinical success rate [85].
For researchers seeking to replicate or extend this type of analysis, the following methodology provides a robust framework.
The first step is to categorize approved drugs based on their origin. The established classification system is [84]:
Data for this analysis can be sourced from public databases. The ChEMBL database is an essential resource for bioactivity data [2]. For structural information, NP-specific databases such as Super Natural II, UNPD, and the Natural Product Atlas are invaluable [2]. The study by Newman and Cragg serves as a definitive reference for categorizing approved drugs up to 2019 [83].
Once the dataset is curated, a standard set of 20+ structural and physicochemical descriptors should be calculated for each molecule. Essential parameters include [84] [87]:
These calculations can be performed using open-source toolkits like RDKit or the Chemistry Development Kit (CDK) [2]. Subsequent multivariate statistical analysis, such as Principal Component Analysis (PCA), is then used to visualize and quantify the differences in chemical space occupied by NP-derived and synthetic drugs [84].
Success in this field relies on leveraging a combination of public data resources and specialized software tools.
Table 3: Essential Resources for NP and Cheminformatics Research
| Resource Name | Type | Primary Function | Key Feature |
|---|---|---|---|
| ChEMBL | Database | Bioactivity data for drug-like molecules | Manually curated bioactivity data from scientific literature [2] |
| Super Natural II | Database | Encyclopedic NP information | Contains >325,000 NP entries; queryable via chemistry-aware web interface [2] |
| Natural Product Atlas | Database | Specialized NP resource | Focused collection of >25,000 NPs from bacteria and fungi [2] |
| RDKit | Software | Cheminformatics toolkit | Open-source platform for descriptor calculation, fingerprinting, and machine learning [2] |
| KNIME | Software | Data analytics platform | Graphical workflow for data blending, preprocessing, and model execution [2] |
| DataWarrior | Software | Data visualization and analysis | Integrated tool for generating Self-Organizing Maps (SOMs) to visualize chemical space [26] |
This cheminformatic comparison unequivocally demonstrates that natural products and their inspired synthetic analogs are not merely historical artifacts but remain indispensable to modern drug discovery. They provide a critical source of chemical diversity, occupying regions of chemical space that are under-represented by purely synthetic compounds. The structural hallmarks of NP-derived drugsâgreater three-dimensionality, increased stereochemical complexity, and lower hydrophobicityâare statistically linked to their higher rates of clinical success and more favorable toxicity profiles. For researchers aiming to broaden the scope of addressable biological targets and improve the efficiency of drug development, prioritizing NP-derived scaffolds and leveraging the computational tools and databases outlined in this guide presents a powerful and empirically validated strategy.
Natural products (NPs) and their structural analogs have been a cornerstone of drug discovery for decades. Current analyses reveal that over half of all approved small-molecule drugs are directly or indirectly derived from natural products [22]. This trend is not merely historical but continues to shape the modern pharmaceutical landscape, particularly among top-selling therapeutic agents. Chemoinformatic analyses demonstrate that drugs based on natural product structures interrogate broader regions of chemical space and exhibit greater structural diversity compared to their completely synthetic counterparts [84]. This guide provides a comparative analysis of the performance of natural product-based drugs against synthetic alternatives, supported by experimental cheminformatic data and structural property comparisons relevant to researchers and drug development professionals.
Drugs originating from natural product templates exhibit distinct structural and physicochemical profiles that differentiate them from completely synthetic drugs. These differences have significant implications for target selection, binding specificity, and overall drug performance.
Table 1: Property Comparison of Drug Classes [84] [3]
| Property | Natural Product Drugs (N) | Natural Product-Derived Drugs (ND) | Completely Synthetic Drugs (S) |
|---|---|---|---|
| Molecular Weight (MW) | 611 | 757 | 355-444 |
| Hydrogen Bond Donors (HBD) | 5.9 | 7.0 | 1.1-2.4 |
| Hydrogen Bond Acceptors (HBA) | 10.1 | 11.5 | 3.9-6.0 |
| ALOGPs | 1.96 | 1.82 | 2.08-3.15 |
| LogD | -1.40 | -3.00 | 0.40-2.49 |
| Rotatable Bonds (Rot) | 11.0 | 16.2 | 5.4-7.6 |
| Topological Polar Surface Area (tPSA) | 196 | 250 | 61-111 |
| Fraction sp3 (Fsp3) | 0.71 | 0.59 | 0.33-0.54 |
| Aromatic Rings (RngAr) | 0.7 | 1.4 | 2.0-2.7 |
The data reveal that natural product-based drugs typically possess higher molecular complexity (as measured by Fsp3), increased stereochemical content, and lower hydrophobicity (evidenced by lower LogD values) compared to completely synthetic drugs [84]. These properties correlate with improved binding selectivity and decreased preclinical toxicity profiles [3].
The influence of natural product structures extends significantly into the commercial pharmaceutical market. Analysis of top-selling drugs reveals a striking increase in the prevalence of natural product-based structures over time.
Table 2: NP-Based Drugs Among Top-Selling Pharmaceuticals [3]
| Year | Total Top 40 Drugs | NP-Based Drugs (Count) | NP-Based Drugs (%) |
|---|---|---|---|
| 2006 | 41 (unique structures) | 14 | 34% |
| 2018 | 49 (unique structures) | 34 | 69% |
This significant increase in natural product-based drugs among top sellers underscores their growing commercial importance and therapeutic value. Notably, this trend coincides with the industry's challenge to address a wider range of biological targets, as natural products exhibit broader chemical diversity and can engage more challenging protein targets [84] [3].
Experimental Objective: To systematically categorize drug compounds based on their structural origins for comparative cheminformatic analysis.
Classification Protocol (adapted from Newman and Cragg [84] [3]):
Data Collection Parameters:
Experimental Objective: To quantify and compare structural and physicochemical properties across drug categories.
Computational Methodology:
Molecular Descriptor Calculation [84]:
Key Parameters Measured [84]:
Statistical Analysis:
Figure 1: Experimental workflow for cheminformatic comparison of drug classes
Table 3: Essential Research Resources for NP-Based Drug Discovery [60] [22]
| Resource Type | Specific Tools/Databases | Application in Research |
|---|---|---|
| NP Databases | COCONUT, LANaPDB, Dictionary of Natural Products (DNP) | Source structures for fragment library generation and chemical space analysis |
| Fragment Libraries | CRAFT, NP-derived fragment libraries | Access to privileged scaffolds for hit identification and optimization |
| Cheminformatics Software | Design Hub, RDKit, ChemAxon | Calculation of molecular descriptors, property prediction, and chemical space visualization |
| Structural Analysis Tools | PCA algorithms, clustering methods, diversity indices | Comparative analysis of chemical space occupancy and scaffold diversity |
Curated NP Fragment Libraries [60]:
NP-Specific Cheminformatic Scripts [22]:
Specialized Property Prediction Tools [84] [3]:
The cheminformatic data reveal that natural product-based drugs occupy distinct regions of chemical space compared to completely synthetic drugs, characterized by several structurally advantageous features:
Enhanced Three-Dimensionality: NP-based drugs exhibit significantly higher Fsp3 values (0.59-0.71 vs. 0.33-0.54 for synthetic drugs), contributing to improved target selectivity and clinical success rates [84] [3].
Balanced Hydrophobicity Profiles: Lower measured LogD values (-3.00 to -1.40 vs. 0.40-2.49 for synthetic drugs) correlate with improved solubility and reduced metabolic clearance [84].
Structural Complexity: Increased stereochemical content (nStereo) and reduced aromatic ring count (RngAr) enable engagement with more challenging target classes, including protein-protein interactions [3].
The growing prevalence of NP-based structures among top-selling drugs suggests a strategic reorientation in successful drug discovery approaches:
Library Design: Incorporation of NP-inspired scaffolds can significantly increase the chemical diversity of screening collections, expanding the range of addressable biological targets [84].
Lead Optimization: Embracing NP-like properties (higher Fsp3, balanced lipophilicity) may improve compound quality and clinical success rates, particularly for challenging targets [3].
Chemical Biology: NP-inspired synthetic compounds (S* category) represent a powerful strategy to access NP-like chemical space while overcoming supply and optimization challenges associated with complex natural products [84].
The integration of NP-informed design principles with modern synthetic and analytical technologies represents a promising trajectory for addressing current challenges in small-molecule drug discovery, particularly for target classes that have historically proven difficult with conventional synthetic approaches.
The systematic exploration of chemical spaceâthe multidimensional universe of all possible organic compoundsâis a fundamental objective in modern drug discovery. Within this vast space, natural products (NPs) and synthetic compounds represent two major continents, each with distinct topological features. This guide provides a chemoinformatic comparison of these domains, demonstrating how the unique structural properties of NP-based drugs significantly expand the diversity of accessible biological targets. Over half of all approved small-molecule drugs originate directly or indirectly from natural products, underscoring their pivotal role in addressing complex disease mechanisms [22]. Their structural evolution through millennia of biological optimization provides a rich source of chemical diversity that often surpasses designed synthetic libraries in complexity and novelty.
Natural products exhibit distinct structural characteristics that differentiate them from synthetic compounds and commercial fragment libraries. These differences directly influence their ability to interact with diverse biological targets.
Table 1: Physicochemical Property Comparison Between Natural Products and Synthetic Compounds
| Property Category | Natural Products | Synthetic Compounds | Significance for Target Diversity |
|---|---|---|---|
| Molecular Size | Larger molecular size [22] | Smaller, more uniform size | Enables interaction with larger protein surfaces |
| Structural Complexity | More chiral centers, Csp3 atoms, rotatable bonds [22] | Fewer stereocenters, higher aromaticity | Facilitates specific binding to complex binding pockets |
| Hydrophobicity | Higher hydrophobicity [22] | Generally lower Log P | Improves membrane penetration for intracellular targets |
| Ring Systems | More aliphatic and fused rings [22] | Simpler ring systems | Provides structural rigidity and defined 3D geometry |
| Heteroatom Content | Higher oxygen content, fewer nitrogen/sulfur atoms [22] | More diverse heteroatom distribution | Influces hydrogen bonding patterns with targets |
Table 2: Fragment Library Diversity Metrics Across Natural and Synthetic Sources
| Library Source | Total Fragments | RO3-Compliant Fragments | Percentage RO3 | Structural Diversity Index |
|---|---|---|---|---|
| LANaPDB NPs | 74,193 | 1,832 | 2.5% | 0.89 |
| COCONUT NPs | 2,583,127 | 38,747 | 1.5% | 0.92 |
| CRAFT Synthetic | 1,202 | 176 | 14.6% | 0.76 |
| Enamine Synthetic | 12,496 | 8,386 | 67.1% | 0.71 |
The data reveals that while NP libraries generate a much larger absolute number of fragments, a smaller percentage comply with the strict Rule of Three (RO3) for fragment-based drug design compared to commercial synthetic libraries [45]. However, NP-derived fragments explore a broader chemical space as indicated by higher diversity indices, making them valuable for targeting unconventional biological interfaces.
Robust experimental protocols and standardized methodologies are essential for meaningful comparison of chemical space coverage between natural and synthetic compounds.
Figure 1: Experimental workflow for chemical space analysis of natural products and synthetic compounds
The unique structural properties of natural products directly translate to enhanced capabilities for addressing diverse biological targets through multiple mechanisms.
Advanced machine learning approaches leverage the structural diversity of NPs for novel target identification. The eXplainable Graph-based Drug response Prediction (XGDP) framework represents drugs as molecular graphs, incorporates gene expression data from cancer cell lines, and uses Graph Neural Networks with attention mechanisms to identify salient functional groups and their interactions with significant genes [89]. This approach demonstrates that NP-derived fragments often contain structural motifs that correlate with activity against under-explored biological targets.
Table 3: Target Class Affinity Distribution Across Compound Types
| Target Class | Natural Product Affinity | Synthetic Compound Affinity | Representative NP Scaffolds |
|---|---|---|---|
| Kinases | Moderate | High | Flavonoids, indolocarbazoles |
| GPCRs | High | High | Alkaloids, terpenoids |
| Nuclear Receptors | High | Moderate | Steroids, diterpenoids |
| Ion Channels | High | Moderate | Peptide toxins, macrolides |
| Protein-Protein Interactions | Very High | Low | Cyclic peptides, complex polyketides |
| Epigenetic Regulators | Emerging | Moderate | Chromomycin, trapoxin |
Table 4: Essential Research Reagents and Computational Tools for Chemical Space Analysis
| Resource Category | Specific Tools/Databases | Primary Function | Access Information |
|---|---|---|---|
| Natural Product Databases | COCONUT, LANaPDB, Dictionary of Natural Products (DNP) | Source of curated NP structures and annotations | Publicly available (COCONUT, LANaPDB); Commercial (DNP) |
| Synthetic Compound Libraries | CRAFT, Enamine, ChemDiv, Life Chemicals | Source of synthetic compounds and fragments | CRAFT: GitHub; Others: Commercial vendors |
| Cheminformatics Toolkits | RDKit, MolVS, DeepChem | Molecular standardization, descriptor calculation, fingerprint generation | Open-source Python packages |
| Chemical Space Visualization | MolCompass, Chemical Space Networks (CSNs), TMAP | Dimensionality reduction and interactive visualization | MolCompass: GitHub; CSNs: RDKit/NetworkX |
| Fragment Analysis | RECAP, BRICS, MORTAR | Deconstruction of molecules into logical fragments | Implemented in RDKit and other cheminformatics platforms |
| Machine Learning Frameworks | XGDP, Graph Neural Networks, Parametric t-SNE | Predictive modeling and interpretable AI for drug response | Custom implementations (e.g., XGDP) and open-source libraries |
The systematic chemoinformatic comparison of natural products and synthetic compounds reveals a compelling narrative: while synthetic libraries provide excellent coverage of "drug-like" chemical space with high synthetic accessibility, natural products explore broader structural territories with superior complexity and uniqueness. This expanded chemical space coverage directly translates to the ability to address a more diverse target landscape, particularly for challenging target classes like protein-protein interactions, allosteric sites, and macromolecular assemblies. The integration of advanced machine learning methods with high-quality NP libraries will further enhance our ability to navigate this valuable chemical territory, accelerating the discovery of innovative therapeutics for complex diseases. As chemoinformatic methodologies continue to evolve, the strategic integration of NP-derived fragments with synthetic libraries represents the most promising path forward for comprehensive chemical space exploration and target diversity expansion in drug discovery.
The pursuit of new therapeutic agents consistently navigates the intricate balance between molecular complexity, synthetic feasibility, and biological activity. Within this landscape, Natural Products (NPs) and synthetic compounds represent two foundational pillars of drug discovery. NPs, defined as compounds produced by living organisms, have a long and successful history as sources of therapeutic agents, with over half of all approved small-molecule drugs originating directly or indirectly from them [22]. In contrast, synthetic compounds originate entirely from chemical synthesis, while semi-synthetic compounds incorporate both natural and synthetic components in their molecular structure [91].
This guide provides a chemoinformatic comparison of NPs and synthetic compounds, focusing on performance metrics that correlate with clinical success. By examining quantitative structural properties, diversity measures, and adherence to drug-likeness guidelines, we aim to objectively evaluate how NP-like features influence the drug discovery pipeline and ultimate clinical outcomes.
The fundamental structural differences between NPs and synthetic compounds significantly influence their performance in drug discovery. Chemoinformatic analyses reveal that NPs generally exhibit greater structural complexity, higher sp3 carbon count, more chiral centers, and increased molecular rigidity compared to their synthetic counterparts [22]. These characteristics contribute to distinct profiles in terms of target engagement, selectivity, and developmental outcomes.
Table 1: Comparative Physicochemical Properties of Natural Products and Synthetic Compounds
| Property | Natural Products | Synthetic Compounds | Clinical Implications |
|---|---|---|---|
| Molecular Weight | Generally higher [22] | Generally lower [22] | Higher MW can complicate oral bioavailability but may improve target specificity |
| cLogP | Variable, marine NPs often more hydrophobic [22] | More controlled during design [92] | Lower logP generally correlates with reduced toxicity risks [92] |
| Csp3 Carbon Count | Higher [22] | Lower [22] | Higher Csp3 correlates with better solubility and clinical success [22] |
| Chiral Centers | More prevalent [22] | Fewer [22] | Increased stereochemical complexity impacts synthesis and specificity |
| Structural Rigidity | More macro rings, fused rings [22] | More flexible structures [22] | Rigidity can improve binding selectivity but reduce adaptability |
| Glycosylation Rate | 8-22% [22] | ~0.23% (purchasable compounds) [22] | Glycosylation significantly influences solubility and target recognition |
Fragment-Based Drug Design (FBDD) utilizes small organic molecules (<300 Da) adhering to the "Rule of Three" (RO3) to efficiently explore chemical space [45]. The application of NPs in FBDD involves deconstructing them into fragments using algorithms like RECAP (Retrosynthetic Combinatorial Analysis Procedure), which breaks specific chemical bonds to generate useful building blocks [45].
Table 2: Fragment Library Performance Metrics [45]
| Library Source | Total Fragments | RO3-Compliant Fragments | RO3 Compliance Rate | Key Characteristics |
|---|---|---|---|---|
| COCONUT (NP) | 2,583,127 | 38,747 | 1.5% | High structural diversity, complexity |
| LANaPDB (NP) | 74,193 | 1,832 | 2.5% | Latin American NP sources, novel scaffolds |
| CRAFT (Synthetic) | 1,202 | 176 | 14.6% | Designed for synthetic accessibility |
| Enamine (Commercial) | 12,496 | 8,386 | 67.1% | Optimized for solubility, drug-likeness |
| ChemDiv (Commercial) | 72,356 | 16,723 | 23.1% | Diverse heterocyclic scaffolds |
| Maybridge (Commercial) | 29,852 | 5,912 | 19.8% | Established drug-like properties |
| Life Chemicals | 65,248 | 14,734 | 22.6% | Focused libraries for screening |
The data reveals a crucial trade-off: while NP-derived fragments offer exceptional structural diversity and complexity, they exhibit significantly lower RO3 compliance rates compared to synthetically-designed libraries [45]. This suggests that synthetic libraries are intentionally curated for drug-like properties from the outset, whereas NP fragments prioritize structural novelty, requiring more optimization to become drug-like.
To ensure reproducible and objective comparisons between NPs and synthetic compounds, researchers employ standardized computational workflows. These methodologies enable quantitative assessment of chemical properties, diversity, and drug-likeness.
Purpose: To generate standardized, comparable molecular representations from diverse data sources by removing inconsistencies and errors that could bias analysis [45].
Workflow:
Tools: RDKit (2024.03.5) and MolVS (0.1.1) toolkits are commonly employed for this protocol [45].
Purpose: To deconstruct molecules into meaningful fragments for chemical space analysis and diversity quantification [45].
Workflow:
Purpose: To quantitatively estimate the feasibility of synthesizing a molecule, which is crucial for assessing development potential [45].
Workflow:
Diagram 1: Chemoinformatic Workflow for NP vs. Synthetic Compound Comparison. This workflow outlines the standardized process for comparing natural products and synthetic compounds, from initial data curation to final analysis.
The high clinical success rate of NPs and NP-derived compoundsâaccounting for over 50% of approved small-molecule drugs [22]âsuggests that specific NP-like structural features correlate favorably with therapeutic outcomes. This section examines the quantitative relationships between these features and key development metrics.
Beyond simple structural metrics, ligand efficiency measures provide crucial insights into binding quality. These metrics help explain why structurally complex NPs often achieve successful clinical outcomes despite sometimes suboptimal physicochemical properties.
Table 3: Ligand Efficiency Metrics and NP Correlations [92]
| Efficiency Metric | Calculation | NP Correlation | Clinical Relevance |
|---|---|---|---|
| Ligand Efficiency (LE) | ÎG per heavy atom | Often higher in NP-derived drugs | Identifies compounds maximizing binding per atomic investment |
| Ligand Lipophilic Efficiency (LLE) | pIC50 - cLogP | Favorable in optimized NPs | Higher LLE correlates with reduced toxicity and better selectivity |
| LELP | cLogP/LE | Lower values in successful NPs | Combines size and lipophilicity corrections; discriminates compounds with acceptable ADMET profiles |
The high structural complexity of NPsâevidenced by increased chirality, stereochemical complexity, and ring fusionâdirectly contributes to their clinical success through enhanced target selectivity [22]. Complex three-dimensional structures are better able to distinguish between closely related biological targets, reducing off-target effects and associated toxicity in clinical trials.
Terpenoid NPs provide an excellent case study, as their high structural complexity (e.g., more chiral centers, Csp3, bridge rings, and spiro rings) contributes to enhanced selectivity toward specific targets [22]. This structural sophistication, while challenging synthetically, provides a natural advantage in clinical development where specificity is paramount.
Synthetic compounds frequently undergo "molecular obesity" during optimizationâincreases in molecular weight and lipophilicity that negatively impact clinical success [92]. In contrast, NPs often serve as optimized starting points from an evolutionary perspective, having been pre-validated through biological interactions.
This fundamental difference creates a divergence in development trajectories: NP-based programs often focus on simplifying complex structures while maintaining efficacy, whereas synthetic programs frequently struggle with adding complexity without introducing pharmacokinetic or toxicity issues [92].
Diagram 2: Divergent Optimization Paths for Natural Products vs. Synthetic Compounds. This diagram illustrates how natural product and synthetic compound optimization follow different trajectories, with NPs often requiring simplification while synthetic compounds risk molecular obesity.
Successful comparison of NPs and synthetic compounds requires specialized computational tools and databases. This toolkit outlines essential resources for conducting comprehensive chemoinformatic analyses.
Table 4: Essential Research Resources for Chemoinformatic Analysis
| Resource Category | Specific Tools/Databases | Function | Access |
|---|---|---|---|
| NP Databases | COCONUT, LANaPDB, Dictionary of Natural Products (DNP) | Provide curated structural and source information for natural products [45] [22] | Publicly available |
| Synthetic Compound Databases | CRAFT, Enamine, ChemDiv, ChEMBL | Offer synthetic compound libraries with drug-like properties [45] [93] | Commercial & academic |
| Cheminformatics Toolkits | RDKit, MolVS, MOE | Enable molecular standardization, descriptor calculation, and structural analysis [45] [94] | Open source & commercial |
| Fragmentation Algorithms | RECAP, BRICS, MORTAR | Deconstruct molecules into fragments for FBDD and diversity analysis [45] | Integrated in toolkits |
| Visualization Platforms | DataWarrior, KNIME, Python libraries | Facilitate chemical space visualization and pattern recognition [94] | Open source & commercial |
| Predictive Modeling Tools | QSAR models, Deep-PK, HobPre | Predict ADMET properties, bioavailability, and synthetic accessibility [45] [94] | Various access models |
The chemoinformatic comparison of NPs and synthetic compounds reveals a nuanced relationship between structural features and clinical success. NPs offer exceptional structural diversity, complexity, and biomolecular recognitionâfeatures that correlate with their disproportionate contribution to approved drugs. However, these advantages come with challenges in synthesis, optimization, and RO3 compliance. Synthetic compounds provide superior synthetic accessibility, controlled physicochemical properties, and higher fragment screening efficiency, yet may lack the structural sophistication needed for challenging biological targets.
The most successful drug discovery strategies leverage the complementary strengths of both approaches: using NP-inspired complexity and privileged scaffolds as starting points, while applying synthetic chemistry and computational design to optimize drug-like properties. This integrated approach, guided by the performance metrics outlined in this review, offers the most promising path for addressing the high failure rates in drug development and delivering novel therapeutics to patients.
Natural products (NPs) and their derived pharmacophores have been pivotal in drug discovery, with over half of all approved small-molecule drugs originating directly or indirectly from these natural compounds [22]. NPs exhibit greater structural novelty, diversity, and complexity compared to synthetic compounds, making them invaluable reservoirs for new chemical entities [22]. The pharmacophore conceptâdefined as the ensemble of steric and electronic features necessary for optimal supramolecular interactions with a specific biological targetâprovides a powerful framework for translating these complex natural structures into effective drugs [95]. This review examines successful drugs derived from natural product pharmacophores through a chemoinformatics lens, comparing their performance against synthetic alternatives and highlighting the methodologies that have enabled these discoveries.
The Pharmacophore Concept in Natural Product Drug Discovery In medicinal chemistry, pharmacophore-based methods have become an indispensable component of modern computer-aided drug design workflows [95]. The official IUPAC definition states that a pharmacophore represents "the ensemble of steric and electronic features that is necessary to ensure the optimal supra-molecular interactions with a specific biological target structure and to trigger (or to block) its biological response" [95]. This abstract description allows researchers to identify structurally different molecules possessing similar pharmacophoric patterns that are recognized by the same binding site, enabling the transformation of natural product scaffolds into therapeutic agents with optimized properties [95].
Table 1: Structural Characteristics of Natural Products vs. Synthetic Compounds
| Structural Feature | Natural Products | Synthetic Compounds | Significance in Drug Discovery |
|---|---|---|---|
| Molecular Complexity | Higher (more chiral centers, Csp³ atoms) [22] | Lower | Enhanced selectivity for specific targets [22] |
| Structural Diversity | Broader chemical space [22] | More confined | Access to novel scaffolds [22] |
| Glycosylation Rate | 8%-22% [22] | 0.23%-4.93% [22] | Improved solubility and bioavailability |
| Adherence to Rule of 5 | Variable (class-dependent) [22] | Typically designed to comply | Natural products explore beyond traditional drug-like space |
| Scaffold Novelty | High, evolutionary optimized [96] | Lower, often based on known chemotypes | Potential for novel mechanisms of action |
Natural products occupy a broader chemical space than synthetic compounds and exhibit distinct structural characteristics that contribute to their success as drug starting points [22]. NPs tend to be more hydrophobic and possess larger molecular size, more macro rings, chiral centers, Csp³ atoms, and rotatable bonds [22]. Analysis of fragment libraries reveals that NP-derived fragments contain more aliphatic and fused rings, fewer heteroatoms (except oxygen), and exhibit higher structural diversity and complexity compared to synthetic fragment libraries [60]. These characteristics enable NPs to interact with diverse biological targets, explaining why more than half of approved small-molecule drugs between 1981 and 2019 are directly or indirectly derived from NPs [22].
Table 2: Drug Discovery Metrics - Natural Product-Derived vs. Synthetic Approaches
| Metric | Natural Product-Derived Drugs | Synthetic Compound Drugs | Data Source |
|---|---|---|---|
| Contribution to Approved Drugs (1981-2019) | >50% [22] | <50% | Newman & Cragg, 2020 [22] |
| Documented Compounds Available | >1.1 million NPs documented [22] | ~100 million synthesizable [22] | NP database analysis |
| Novel Drug-Productive Species (1991-2010) | 59 new species yielding drugs [96] | Not applicable | Zhu et al., 2012 [96] |
| Drugs from Untapped Species (1991-2010) | 7.1%-14.5% of new approvals [96] | Not applicable | Zhu et al., 2012 [96] |
| Scaffold Diversity | Higher diversity in fragment libraries [60] | Lower diversity in fragment libraries [60] | Comparative chemoinformatic analysis [60] |
The productivity of natural product-derived drugs remains substantial despite shifts in pharmaceutical screening strategies. Between 1991-2010, 46-126 nature-derived drugs were approved every five years, with 7.1%-14.5% originating from previously untapped species [96]. This trend suggests that untapped drug-productive species are not near extinction, and future bioprospecting efforts are expected to yield new drugs at comparable levels [96]. Notably, 55% of new drug-productive species emerging in 1991-2010 came from existing drug-productive species families, while another 37% came from new species families in existing drug-productive clusters, indicating a high probability of finding new drug-productive species from these sources [96].
Protocol 1: Structure-Based Pharmacophore Modeling
Protocol 2: Ligand-Based Pharmacophore Modeling
The experimental workflow for pharmacophore-based natural product drug discovery involves multiple stages, from model generation to virtual screening and experimental validation, as visualized in the following diagram:
Protocol 3: Non-Extensive Fragmentation of Natural Products
Non-extensive fragmentation has been shown to produce fragments with higher pharmacophore fit scores than both extensively fragmented compounds and their original parent natural products in the majority of cases (56% and 69% respectively) [35]. This approach yields a much higher number of chemical entities (45,355 vs. 11,525 compounds for extensive fragmentation) that are far less repetitive and cover broader chemical space [35].
Experimental Background: Eribulin, approved in 2009, originated from the natural products homohalichondrin B and halichondrin B isolated from previously untapped western Pacific sponges Halichondria and Axinella [96]. These compounds demonstrated strong cytotoxic and tubulin polymerization inhibitory activities but faced significant supply problems [96].
Pharmacophore Optimization Strategy:
Performance Data: Eribulin maintains the potent tubulin-binding pharmacophore of the parent natural product while achieving synthetic accessibility and improved drug-like properties. It demonstrates potent cytotoxic activity toward both paclitaxel-sensitive and paclitaxel-resistant cells [96].
Experimental Background: Ixabepilone, approved in 2009, was derived from epothilone B identified from the previously untapped myxobacterium Sorangium cellulosum [96]. Epothilones were discovered as tubulin-interacting anticancer agents with potent activity against taxane-resistant cells [96].
Pharmacophore Optimization Challenges:
Performance Data: Ixabepilone maintains the potent microtubule-stabilizing activity of the parent natural product while exhibiting improved metabolic stability and pharmacokinetic properties [96]. It demonstrates efficacy in taxane-resistant malignancies, highlighting the value of this natural product-derived pharmacophore in overcoming resistance mechanisms.
Experimental Background: The development of imatinib (Gleevec), approved in 2001, illustrates how natural product pharmacophores can inspire drugs for novel targets [96]. The identification of BCR-ABL as a key target in chronic myeloid leukemia prompted searches for ABL inhibitor drugs [96].
Pharmacophore Evolution:
Performance Data: Imatinib represents a milestone in targeted cancer therapy, demonstrating how natural product pharmacophores can be progressively optimized through structure-based design to achieve target specificity while maintaining potency [96].
The following diagram illustrates the conceptual workflow for translating natural product pharmacophores into optimized drugs, integrating computational and experimental approaches:
Table 3: Key Research Resources for Natural Product Pharmacophore Research
| Resource Type | Specific Examples | Function/Application | Access Information |
|---|---|---|---|
| Natural Product Databases | COCONUT, LANaPDB, Dictionary of Natural Products, UNPD, SuperNatural 3.0 [60] [22] | Source of natural product structures for virtual screening | Publicly available via GitHub repositories or institutional access [60] |
| Fragment Libraries | CRAFT library, Natural Product-Derived Fragments (NPDFs) [60] [35] | Fragment-based drug design starting points | Available through research publications and associated data repositories [60] |
| Pharmacophore Modeling Software | Ligand Scout, Catalyst [95] [97] | Generation and validation of 3D pharmacophore models | Commercial and academic software packages |
| Chemical Space Analysis Tools | Chemoinformatic workflows for diversity assessment [60] [22] | Comparison of NP vs synthetic chemical space | Custom implementations based on published methodologies |
| Virtual Screening Platforms | Molecular docking, pharmacophore screening [95] [97] | High-throughput in silico screening of compound libraries | Various commercial and open-source platforms |
Natural product pharmacophores continue to provide valuable starting points for drug discovery, offering structural diversity and complexity that often exceeds what is available in synthetic compound libraries [22]. The success stories of drugs like eribulin, ixabepilone, and imatinib demonstrate how natural product-derived pharmacophores can be optimized to address limitations of the original compounds while maintaining their therapeutic activity [96].
Future directions in natural product pharmacophore research include increased integration of artificial intelligence for target prediction and activity forecasting [22], exploration of untapped species and extreme environments for novel bioactive compounds [22] [96], and advanced computational methods for exploring the extensive chemical space occupied by natural products [60] [22]. As these technologies mature, natural product pharmacophores are poised to continue their critical role in delivering innovative therapeutic agents for diverse diseases.
The comparative analysis presented in this review demonstrates that natural products and their pharmacophores provide complementary advantages to synthetic compounds in drug discovery. While synthetic compounds often excel in drug-like properties and synthetic accessibility, natural products offer unparalleled scaffold diversity and evolutionary-optimized bioactivity. The most successful drug discovery strategies leverage both approaches, using natural product pharmacophores as inspiration and synthetic chemistry to optimize these starting points into developable drugs.
This cheminformatic comparison unequivocally demonstrates that natural products and their derived fragments offer unparalleled chemical diversity, structural complexity, and broader coverage of biologically relevant chemical space compared to synthetic compound libraries. Key differentiators of NPs, including higher Fsp3 character, greater stereochemical content, and unique ring systems, are increasingly recognized as valuable for addressing challenging drug targets. Despite practical hurdles, the successful track record of NP-based drugsâcomprising approximately half of all new small-molecule approvalsâvalidates their critical role. The future of drug discovery lies in hybrid strategies that integrate the rich structural inspiration of NPs with the scalability and tractability of synthetic chemistry. Leveraging AI, advanced database integration, and continued exploration of underutilized biological sources will be pivotal in harnessing the full potential of nature's chemical arsenal for developing next-generation therapeutics.