Fact-checked by Grok 2 weeks ago

Chemical structure

Chemical structure refers to the connectivity of atoms within a chemical entity and their three-dimensional spatial arrangement, including the types and positions of chemical bonds that hold them together. This arrangement uniquely identifies a or and dictates its fundamental behavior in chemical reactions and interactions. Chemical structures are represented in diverse formats to communicate this information effectively, with two-dimensional diagrams being the most common for illustrating atomic connectivity and approximate geometry. Lewis structures depict all atoms, bonds (as lines or dots), and lone pairs explicitly, while skeletal or line-angle formulas simplify organic molecules by omitting hydrogen atoms and using lines to represent carbon-carbon bonds, assuming standard valences. For conveying and precise spatial details, three-dimensional models such as ball-and-stick (showing atoms as spheres and bonds as rods) or space-filling representations (illustrating atomic van der Waals radii) are employed, often in computational or experimental contexts like . International standards, such as those from IUPAC, ensure these diagrams are unambiguous, with guidelines for bond lengths, angles (e.g., 120° for sp²-hybridized carbons), and orientations to minimize overlap and enhance readability. Understanding chemical structure is essential across chemistry and related disciplines, as it directly influences a compound's physical properties (e.g., , ) and chemical reactivity. In organic synthesis, structural knowledge enables the design of molecules with targeted functions, while in structure-activity relationships (SAR), correlations between specific structural features and biological effects guide and assessments. For instance, subtle changes in bond arrangement or can alter pharmacological potency or toxicity, underscoring structure's role in advancing pharmaceuticals, , and .

Fundamentals

Definition and Scope

Chemical structure is defined as the arrangement of atoms and the bonds connecting them in a chemical entity, encompassing both the connectivity () of atoms and their spatial geometry. According to the International Union of Pure and Applied Chemistry (IUPAC), describes the identity of the atoms and their linkages, including bond multiplicities, while stereochemical refers to the fixed spatial arrangements that distinguish stereoisomers, such as those arising from double bonds or chiral centers. Conformation, another key aspect, involves variable spatial arrangements interconvertible by rotation around single bonds. In broader contexts, chemical structure may also include the distribution of electrons, particularly in quantum chemical descriptions, but it fundamentally focuses on atomic positions and interactions rather than dynamic behaviors like reactivity or spectroscopic properties. The scope of chemical structure extends beyond isolated molecules to include extended systems such as polymers, crystalline solids, and surfaces. For discrete molecules, it specifies the precise and three-dimensional layout that determine the compound's identity. In polymers, involves repeating units and their sequential , often termed primary structure, alongside higher-order folding. Crystalline materials feature periodic lattices of atoms or ions, where defines unit cells and symmetry, while surfaces encompass two-dimensional arrangements at interfaces, such as in or . This scope distinctly separates structural features from emergent physical properties; for instance, while structure dictates potential reactivity, it does not encompass measured reaction rates or spectral signatures, which arise from interactions with external factors. Central concepts in chemical structure include atomic connectivity () as the foundational element, which establishes the molecular and topology, and spatial arrangements— for fixed and conformation for variable forms—that refine the three-dimensional layout through angles, distances, and orientations. These elements collectively define the identity of a , ensuring that two entities with identical structures are considered the same compound, regardless of preparation method or isotopic variations unless specified. For example, the molecule (H₂O) exhibits a bent with a bond angle of approximately 104.5°, arising from the tetrahedral arrangement around oxygen, contrasting with the linear of (CO₂), where the O=C=O arrangement yields a 180° bond angle due to sp hybridization at carbon. Such structural distinctions underpin differences in properties like and , highlighting structure's role in chemical uniqueness.

Historical Background

The concept of chemical structure originated from early atomic theories that gradually incorporated ideas of bonding and spatial arrangement, transforming chemistry from empirical observations to a predictive science. In 1808, introduced his in A New System of Chemical Philosophy, proposing that elements consist of indivisible atoms combining in simple whole-number ratios to form compounds, though his model did not specify interatomic bonds. Building on this, developed the electrochemical theory in the 1810s, conceptualizing compounds as aggregates of electropositive and electronegative atoms and establishing the modern notation for chemical formulas, such as H₂O for water. A pivotal advancement occurred in 1858 when formulated his structural theory for organic compounds, positing that carbon atoms are tetravalent and can link to form chains or rings, allowing chemists to represent molecular skeletons and explain isomerism. This framework was revolutionized in 1874 by , who proposed the tetrahedral geometry of carbon atoms, introducing and accounting for the optical activity of chiral molecules. The late 19th century saw confirmation of structural ideas in through Alfred Werner's coordination theory of 1893, which differentiated primary ionic bonds from secondary coordination bonds in metal complexes, resolving discrepancies in isomerism and . The 1920s brought the integration of into chemical bonding, with and Fritz London's 1927 work explaining covalent bonds as shared electron pairs, providing an electronic basis for . Key milestones validated these concepts experimentally: in 1913, William Henry Bragg and William Lawrence Bragg determined the first crystal structure of sodium chloride using X-ray diffraction, confirming its ionic lattice arrangement. A landmark application came in 1953, when James Watson and Francis Crick elucidated the double-helical structure of DNA, demonstrating how structural principles underpin biological function. This evolution addressed fundamental challenges, such as the inadequacy of empirical formulas—obtained via , which yielded only elemental ratios without bonding details—for distinguishing isomers, prompting the adoption of structural formulas that depict .

Structural Representations

Two-Dimensional Notations

Two-dimensional notations provide simplified, planar representations of chemical structures, focusing primarily on and rather than spatial . These methods use symbols, lines, and abbreviations to depict molecules on flat surfaces like paper or digital screens, making them essential for quick communication in . They evolved from early systems to standardize the illustration of electrons and bonds, enabling chemists to infer molecular without exhaustive detail. Lewis structures, introduced by in , represent molecules by showing valence electrons as dots and covalent bonds as lines connecting atomic symbols. In these diagrams, lone pairs are depicted as pairs of dots, while shared electron pairs form single, double, or triple bonds indicated by one, two, or three lines, respectively. The guides construction, aiming for most atoms (except hydrogen, which follows the duet rule) to achieve eight valence electrons through bonding and lone pairs. For example, (H₂O) is shown with oxygen at the center bonded to two hydrogens via single lines, and two lone pairs on oxygen. To assess and alternative forms, is calculated for each atom using the formula: \text{[formal charge](/page/Formal_charge)} = V - N - \frac{1}{2}B where V is the number of valence electrons in a neutral atom, N is the number of nonbonding () electrons, and B is the number of bonding electrons. Structures with minimal formal charges and adherence to the are preferred, as in the carbonate ion (CO₃²⁻), where one resonance form places a -1 charge on an oxygen atom. This approach helps predict reactivity but assumes idealized electron distribution. Condensed structural formulas offer a more compact alternative by linearly arranging symbols and omitting explicit bond lines between carbons or hydrogens in chains. For instance, is written as CH₃CH₂OH, implying single bonds between the carbons and the oxygen, with hydrogens attached to satisfy valences. This notation groups atoms efficiently, such as (CH₃)₂CHOH for isopropanol, but requires familiarity to reconstruct the full connectivity. It is widely used in and reaction schemes for its brevity. Skeletal formulas, also known as line-angle or bond-line drawings, further simplify representations in by implying carbon atoms at line intersections and endpoints, with hydrogens omitted unless attached to heteroatoms. Bonds are shown as straight lines, and rings like are depicted as without explicit labels. For example, appears as a simple , assuming CH₂ groups at each vertex. This method prioritizes the carbon for rapid sketching and of functional groups. Despite their utility, two-dimensional notations have inherent limitations, as they neglect three-dimensional and , potentially misleading interpretations of molecular shape and interactions. Extensions like wedges and dashes can indicate depth but are not standard in pure 2D forms. These simplifications suit connectivity-focused tasks but require complementary methods for spatial insights.

Three-Dimensional Models

Three-dimensional models extend the representation of chemical structures by capturing the spatial relationships among atoms, including distances, angles, and orientations that influence molecular properties and interactions. These models bridge the planar depictions of bonding in two-dimensional notations, such as Lewis structures, to the actual geometries observed in molecules. By incorporating elements like bond lengths, angles, and torsional twists, they enable better prediction of reactivity, , and physical behavior. Physical models provide tangible ways to manipulate and study molecular architectures. Ball-and-stick models depict atoms as colored spheres connected by rods representing , emphasizing connectivity and approximate skeletal geometry while allowing rotation to explore conformations. Originating in the , these models remain valuable for educational purposes due to their simplicity and ability to illustrate and lengths visually. In contrast, space-filling models represent atoms as interlocking spheres sized according to their van der Waals radii, offering a realistic portrayal of molecular volume, surface area, and potential steric interactions. Developed in the mid-20th century by Robert Corey, , and Walter Koltun—known as models—these emphasize atomic overlap and packing density, which is crucial for understanding non-bonded repulsions in crowded molecules. Computational visualizations enhance these concepts through software that generates dynamic, high-resolution images of molecular structures. Tools like PyMOL allow rendering in wireframe mode to highlight bond frameworks, surface modes to display solvent-accessible areas based on van der Waals envelopes, and orbital representations to visualize distributions around atoms. These digital approaches facilitate analysis of complex systems, such as proteins, by enabling zooming, rotation, and overlay of multiple representations. Key geometric descriptors quantify the three-dimensional features of molecules. For instance, the standard carbon-carbon length in alkanes is approximately 1.54 , while angles in tetrahedral carbon environments, as in , measure 109.5°. Dihedral angles, which define torsional orientations, are 60° in the staggered conformation of , minimizing steric strain compared to the 0° eclipsed form. These parameters, derived from experimental and theoretical data, serve as benchmarks for model accuracy and molecular simulations. Stereochemical notations convey three-dimensional and conformations in simplified drawings. Solid wedges indicate projecting toward the viewer from a chiral center, while dashed lines denote those receding, allowing depiction of tetrahedral stereocenters without full models. Newman projections, viewed along a specific axis, illustrate relative positions of substituents; for , the staggered arrangement shows groups offset by 60° dihedrals, contrasting the higher-energy eclipsed state where they align. These conventions are essential for communicating asymmetry and rotational barriers. The provides a foundational framework for predicting three-dimensional shapes from arrangements around a central atom. Proposed by Ronald Gillespie and Ronald Nyholm, it posits that repel each other, arranging to maximize separation: lone pairs occupy more than bonding pairs. For (PCl₅), five bonding pairs result in a trigonal bipyramidal geometry, with equatorial angles of 120° and axial-equatorial angles of 90°. This model underpins the design of many 3D representations by forecasting ideal geometries for main-group compounds.

Determination Methods

Experimental Techniques

Experimental techniques for determining chemical structures rely on empirical measurements from spectroscopic and diffraction methods, providing direct evidence of atomic arrangements, bonding, and functional groups in molecules. These laboratory-based approaches capture physical interactions of matter with radiation or particles, yielding data that can be interpreted to reconstruct molecular architectures with high precision. Key methods include for solid-state structures, (NMR) spectroscopy for solution-phase connectivity, for molecular formulas and fragments, and (IR) or for vibrational signatures of functional groups. By integrating these techniques, chemists achieve comprehensive structural elucidation that accounts for both local and global features. X-ray crystallography determines the three-dimensional arrangement of atoms in crystalline solids by analyzing diffraction patterns produced when X-rays interact with the electron clouds of atoms. The fundamental principle is , which describes the constructive interference condition for diffraction: n\lambda = 2d \sin\theta, where n is an integer, \lambda is the X-ray wavelength, d is the interplanar spacing in the crystal , and \theta is the incidence angle. This law enables of lattice parameters and, through transformation of diffraction intensities, generation of maps that reveal atomic positions with typical resolutions of approximately 1 for small molecules. These maps display regions of high electron density corresponding to atomic nuclei, allowing refinement of bond lengths, angles, and in the crystal phase. NMR elucidates molecular and spatial arrangements in by exploiting the magnetic properties of atomic , such as ^1H and ^{13}C. Chemical shifts (\delta, measured in parts per million, ppm) indicate the electronic environment around each , while scalar coupling constants (J, in hertz, Hz) reveal through-bond interactions between neighboring atoms, providing evidence for skeletal frameworks. Two-dimensional techniques like COSY (correlation ) map J-coupled protons to establish adjacency, and NOESY ( ) detects through-space proximities via dipole-dipole interactions, aiding in stereochemical assignments. For example, NOESY cross-peaks between non-adjacent protons confirm spatial orientations in complex organic molecules. Mass spectrometry identifies molecular formulas and substructures by ionizing molecules and analyzing the mass-to-charge ratios (m/z) of resulting fragments. The molecular ion peak represents the intact ionized molecule, offering its exact mass and thus empirical formula when combined with isotopic patterns. Fragmentation patterns arise from bond cleavages, revealing connectivity; for instance, in carbonyl compounds like aldehydes or ketones, the McLafferty rearrangement produces a characteristic odd-electron at m/z = 58 for methyl ketones, involving \gamma-hydrogen transfer and elimination. This process, prominent in spectra, helps deduce positions without requiring pure crystals. IR and Raman spectroscopies probe vibrational modes of molecules, identifying functional groups through characteristic absorption or scattering frequencies. In , the carbonyl stretch (C=O) appears as a strong band around 1700 cm^{-1}, diagnostic of ketones, aldehydes, or esters depending on the exact position and intensity. Raman complements by detecting symmetric vibrations less active in IR, such as C=C stretches near 1650 cm^{-1}, with both techniques providing complementary fingerprints for bond types and in gaseous, liquid, or solid samples. Integrating these techniques enhances structural accuracy by leveraging complementary strengths; for example, provides precise solid-state atomic coordinates, while NMR offers dynamic solution-phase details, and their joint refinement resolves ambiguities in multidomain systems like proteins. Mass and vibrational data further corroborate functional groups and formulas, ensuring robust validation across phases. This multifaceted approach is essential for complex molecules where single methods may overlook conformational or environmental effects.

Computational Approaches

Computational approaches to chemical structure prediction and refinement rely on theoretical models and algorithms to simulate molecular geometries and energies from first principles or empirical parametrizations, enabling the exploration of structures inaccessible to direct experimentation. These methods range from quantum mechanical treatments that solve the approximately to classical simulations using force fields, often implemented in specialized software for tasks like geometry optimization and dynamic behavior analysis. In , the Hartree-Fock () method serves as a foundational for multi- systems by assuming each moves in an created by the others, leading to a set of self-consistent equations solved iteratively to obtain molecular orbitals and energies. This approach neglects electron correlation but provides a starting point for more accurate methods. (DFT), building on the framework, approximates the rather than the wavefunction, using exchange-correlation functionals to account for interactions; for optimization, DFT minimizes the total energy with respect to coordinates via algorithms, yielding equilibrium structures with high efficiency for medium-sized molecules. Molecular mechanics employs classical force fields to model structures by summing potential energy terms for bonded and non-bonded interactions, treating atoms as classical particles with predefined parameters. The MMFF94 force field, for instance, includes harmonic terms for bond stretching, such as E_{\text{bond}} = k (r - r_0)^2, where k is the force constant and r_0 the equilibrium distance, alongside angle bending, torsional rotations, and non-bonded van der Waals and electrostatic interactions, making it suitable for diverse organic and biomolecular systems. Software like Gaussian facilitates ab initio calculations, including HF and DFT, for precise static structure predictions, while molecular dynamics simulations—often using these force fields—model dynamic processes such as protein folding by propagating atomic trajectories over time under Newtonian mechanics. Accuracy varies by method: semi-empirical approaches like AM1 offer computational speed for large systems by parametrizing integrals empirically, though with higher errors in energies and geometries compared to post-HF methods such as , which include second-order correlation for improved precision in small molecules. DFT typically achieves errors of about 0.01 relative to experimental values, balancing accuracy and scalability. Hybrid applications, such as molecular in , combine these techniques to predict binding geometries by sampling poses and scoring interactions within protein pockets, aiding lead optimization.

Advanced Concepts

Stereochemistry

is the study of the three-dimensional spatial arrangement of atoms and substituents in , leading to distinct stereoisomers (including enantiomers and diastereomers) with potentially different physical, chemical, and biological properties. arises when a lacks an internal plane of and cannot be superimposed on its , resulting in analogous to left and right hands. A classic example is a tetrahedral carbon atom bonded to four different substituents, known as a or . Stereoisomers resulting from include enantiomers, which are nonsuperimposable mirror images of each other and exhibit identical physical properties except for their interaction with plane-polarized light. Diastereomers, in contrast, are stereoisomers that are not mirror images and thus possess different physical properties, such as melting points or solubilities, even though they share the same connectivity. Atropisomers represent a special class of stereoisomers arising from restricted rotation about a , typically due to steric hindrance from bulky substituents, allowing isolation of stable conformers at room temperature. To designate the at a chiral center, the Cahn-Ingold-Prelog () priority rules are employed, assigning priorities to s based on (higher receives higher priority), with ties resolved by comparing attached atoms in order of decreasing . The lowest-priority is oriented away from the viewer, and the configuration is labeled R (rectus, clockwise) or S (sinister, counterclockwise) based on the sequence of the remaining s. Chiral molecules exhibit optical activity, the ability to rotate the plane of polarized light, quantified by the [ \alpha ], defined as: [ \alpha ] = \frac{\alpha}{c \cdot l} where \alpha is the observed rotation in degrees, c is the concentration in g/mL, and l is the path length in decimeters. Enantiomers rotate plane-polarized light to an equal extent but in opposite directions, while racemic mixtures (1:1 enantiomer pairs) are optically inactive. Resolution of racemic mixtures into enantiomers can be achieved through chiral chromatography, which employs a chiral stationary phase to differentially interact with each , enabling their separation based on transient diastereomeric complexes. The biological implications of are profound, as enantiomers can elicit vastly different responses in chiral biological environments. For instance, the (R)- of exhibits sedative effects, while the (S)-enantiomer is teratogenic, causing severe birth defects; this tragedy in the 1950s–1960s underscored the need for enantiopure drugs.

Conformations and Isomerism

In chemical structures, conformations refer to the various spatial arrangements of atoms that result from rotation around single bonds, allowing interconversion without breaking covalent bonds. These arrangements occupy local energy minima on the molecule's , with barriers separating them that determine the rates of interconversion. For instance, in , the staggered conformation is the global minimum, separated from the eclipsed by a torsional barrier of approximately 12 kJ/mol. In more complex alkanes like , rotational isomers or rotamers include the anti (most stable) and gauche forms, with the gauche conformation higher in energy by about 3.8 kJ/mol due to steric interactions between the methyl groups. Isomerism encompasses molecules with identical molecular formulas but distinct arrangements of atoms, broadly classified into constitutional isomers and stereoisomers. Constitutional isomers, also known as structural isomers, differ in the connectivity of atoms, such as n-pentane (straight chain) and (branched), both C5H12. A subset of constitutional isomers are tautomers, which interconvert rapidly via proton transfer, like the keto-enol forms in β-dicarbonyl compounds. In , the tautomer predominates in solution ( Kenol/keto ≈ 7–8), stabilized by intramolecular hydrogen bonding, and the compound exhibits a pKa of approximately 9 for of the enol. Stereoisomers share the same but differ in the spatial orientation of atoms; they are subdivided into enantiomers, which are non-superimposable mirror images, and diastereomers, which are stereoisomers that are not mirror images. This classification tree highlights how constitutional differences precede spatial variations in defining molecular diversity. The dynamics of conformations involve energy barriers to rotation or interconversion, influencing observable populations. (NMR) provides evidence for these dynamics by resolving separate signals for rotamers at low temperatures when barriers exceed ~20–30 kJ/mol, or by line-shape analysis for faster exchanges. The difference between conformers relates to their populations via ΔG = −RT ln(K), where K is the ratio of concentrations; for at , Kanti/gauche ≈ 4.6 yields ΔG ≈ −3.8 kJ/mol, favoring the anti form. In cyclic systems, such as , the chair conformation interconverts to an equivalent chair conformation via a or twist-boat with an of about 45 kJ/mol, occurring rapidly at ambient temperatures despite the high barrier.

Specialized Tools and Formats

Digital Structure Formats

Digital structure formats enable the storage, exchange, and computational processing of chemical structures in machine-readable ways, facilitating cheminformatics applications such as database querying and molecular modeling. These formats encode atomic connectivity, coordinates, and optional stereochemical details, promoting interoperability across software tools and databases. The Simplified Molecular Input Line Entry System (SMILES) is a compact, linear string notation for representing molecular structures without coordinates. In SMILES, atoms are denoted by their elemental symbols, bonds are implied as single unless specified (e.g., = for double), branches are enclosed in parentheses, and rings are indicated by matching numbers after the connected atoms. For example, the SMILES string CC(O)CC represents 2-butanol, where the parentheses denote the hydroxyl branch on the second carbon. This format supports canonicalization for unique representations and is widely used for substructure searching due to its simplicity and compactness. The (InChI), developed by IUPAC, provides a hierarchical, non-proprietary string-based encoding that captures molecular in layers. The standard InChI begins with "InChI=1S/" followed by the molecular formula, then layers for connectivity (/c), hydrogens (/h), (/t for tetrahedral, /b for double bonds), and isotopes if applicable. For , the InChI is InChI=1S/[C6H6](/page/C6H6)/c1-2-4-6-5-3-1/h1-6H, where the /c layer describes the ring connectivity and the /h layer specifies positions. InChI ensures uniqueness and reversibility to the original structure, making it suitable for database indexing and cross-platform exchange. MOL and SD files, originating from MDL Information Systems (now part of ), store chemical structures as plain-text files with explicit 2D or atomic coordinates. A file consists of a header, a connection table with atom records (listing , x/y/z coordinates), records (atom indices and bond types), and optional properties; SD files extend this to multiple structures for batch processing. Atom blocks specify positions in Angstroms, enabling visualization and simulation, while these formats are prevalent in cheminformatics databases for storing conformer data. The (PDB) format is tailored for biomolecular structures, archiving atomic coordinates from experimental sources like . It uses fixed-width records, with ATOM lines detailing residue name, chain identifier, atom name, and x/y/z coordinates (e.g., for a carbon alpha atom: "ATOM 1 N MET A 1 26.760 13.920 22.950 1.00 21.87 N"). Designed for proteins, nucleic acids, and complexes, PDB supports multiple models and includes metadata like resolution. An extension, the macromolecular Crystallographic Information File (mmCIF), adopts a dictionary-based, self-documenting syntax for richer data representation, including hierarchical assemblies and experimental details, and has become the deposition standard since 2014. These formats enhance interoperability in cheminformatics workflows, as seen in toolkits like RDKit, which parses SMILES, InChI, , and PDB for operations such as fingerprint generation and . They also enable efficient structure-based searches in repositories like , where users query by SMILES or InChI to retrieve millions of compounds with associated bioactivity data.

Typesetting Chemical Structures

Typesetting chemical structures involves specialized techniques to visually represent molecular diagrams in documents, ensuring clarity, accuracy, and compatibility across print and . These methods have evolved from manual illustrations to automated tools that integrate seamlessly with text and equations, facilitating communication in scientific publications. Key approaches include markup-based systems for static rendering and interactive libraries for web environments, each addressing the need to depict bonds, atoms, and spatial arrangements without distortion. Historically, chemical structures were rendered by hand-drawing, a labor-intensive process prone to inconsistencies and errors in depicting complex molecules. This shifted in the late with the advent of automated software; for instance, ISIS/Draw, developed by MDL Information Systems in the , introduced digital sketching capabilities that allowed users to create and export 2D diagrams for inclusion in reports and journals, marking a transition to reproducible, editable representations. In LaTeX-based typesetting, packages like chemfig and XyMTeX enable the creation of vector-based chemical diagrams directly within documents, producing scalable output suitable for high-resolution printing. The chemfig package, for example, uses a simple syntax to draw structures, such as \chemfig{H-O-H} for , allowing bonds to be specified with angles and lengths for precise control. XyMTeX, an earlier system, focuses on structural formulas using mathematical notation extensions, generating vectors that integrate well with LaTeX equations but require more setup for organic molecules. Pros of these vector approaches include infinite scalability without and embedding in PDFs for searchable documents; however, they can be verbose for intricate structures and may demand familiarity with syntax, potentially slowing workflow compared to graphical editors. For digital and web-based rendering, and technologies support inline chemical diagrams through libraries like ChemDoodle Web Components, which generate interactive and visuals from molecular data. These libraries produce output that scales responsively in browsers and enhances accessibility by incorporating labels for screen readers to describe atomic connections and . Best practices for emphasize scalability to maintain line widths and angles across sizes, using formats like or to avoid raster artifacts in publications. Color should be applied judiciously for element differentiation—e.g., black for carbon, red for oxygen—while ensuring compatibility and avoiding overload that obscures details; integration with equations involves aligning baselines for subscripts and superscripts to match document fonts. Challenges persist in rendering complex stereochemistry, where wedges and dashes for chiral centers must convey depth without ambiguity, often requiring custom macros in or layered SVGs that may not render consistently across devices. Font-independent symbols, such as sans-serif glyphs for atoms per IUPAC guidelines, help mitigate variability, but ensuring compatibility in legacy systems or diverse outputs remains difficult.

References

  1. [1]
    structural formula (S06061) - IUPAC Gold Book
    A formula which gives information about the way the atoms in a molecule are connected and arranged in space.
  2. [2]
    Quantum Definition of Molecular Structure - ACS Publications
    Jan 10, 2024 · Molecular structure is a concept that is invariant under rotations. For example, in a diatomic molecule, the molecular structure is completely ...
  3. [3]
    [PDF] Graphical Representation Standards for Chemical Structure Diagrams
    For the broadest understanding, it is preferable to depict the phosphorus-con- taining fragments fully with explicit atoms and bonds. It is not acceptable to ...
  4. [4]
    16 The importance of accurate structure determination in organic ...
    Oct 31, 2023 · Structural information is essential in chemistry not only for determining the geometrical arrangement of atoms within a molecule or a crystal, ...
  5. [5]
    Practice of Structure Activity Relationships (SAR) in Toxicology
    A structure activity relationship relates features of a chemical structure to a property, effect, or biological activity associated with that chemical. In ...
  6. [6]
    The importance of molecular structure and conformation - PubMed
    Molecular structure holds a key to understanding Nature's intricate design mechanisms and blueprints. If we can understand her blueprints and basic ...
  7. [7]
    constitution (C01282) - IUPAC
    The description of the identity and connectivity (and corresponding bond multiplicities) of the atoms in a molecular entityMissing: definition | Show results with:definition
  8. [8]
    configuration (C01249) - IUPAC Gold Book
    The term is restricted to the arrangements of atoms of a molecular entity in space that distinguishes stereoisomers.
  9. [9]
    [PDF] NIST/TRC SOURCE Data Archival System: The Next-Generation ...
    Oct 28, 2011 · Chemical structure implies the combination of composition (inclusive of specific isotopes and charges, if present) and bonding (connectivity) ...
  10. [10]
    Polymer Structure - Nondestructive Evaluation Physics : Materials
    A polymer is composed of many simple molecules that are repeating structural units called monomers. A single polymer molecule may consist of hundreds to a ...
  11. [11]
  12. [12]
    John Dalton and the London atomists: William and Bryan Higgins ...
    Aug 20, 2014 · In 1808 John Dalton published A New System of Chemical Philosophy, which described principles such as the uniqueness of atoms of the same ...
  13. [13]
    Jöns Jakob Berzelius | Science History Institute
    An avid and methodical experimenter, Jöns Jakob Berzelius (1779–1848) conducted pioneering experiments in electrochemistry and established the law of constant ...
  14. [14]
    August Kekulé and Archibald Scott Couper - Science History Institute
    The discovery by these two scientists depended on Kekulé's theory, proposed in 1857, that carbon is tetravalent, valence being defined at the time as the ...
  15. [15]
    Jacobus Henricus van 't Hoff - Linda Hall Library
    Aug 30, 2024 · Van 't Hoff suggested that the four atoms that bond to carbon occupy the four points of a tetrahedron, with the carbon atom in the center (see ...
  16. [16]
    History of Chemistry - C&EN - American Chemical Society
    Jan 29, 2007 · ... quantum mechanics shifted the philosophical outlook of chemical bonding. Once quantum mechanics entered the picture with the work of Heitler ...
  17. [17]
  18. [18]
    A Structure for Deoxyribose Nucleic Acid - Nature
    The determination in 1953 of the structure of deoxyribonucleic acid (DNA), with its two entwined helices and paired organic bases, was a tour de force in ...
  19. [19]
    THE ATOM AND THE MOLECULE. - ACS Publications
    Gilbert N. Lewis. ACS Legacy Archive. Open PDF. Journal of the American Chemical Society. Cite this: J. Am. Chem. Soc. 1916, 38, 4, 762–785. Click to copy ...
  20. [20]
    Lewis Structures
    Formal Charge. It is sometimes useful to calculate the formal charge on each atom in a Lewis structure. The first step in this calculation involves dividing ...
  21. [21]
    Basic Lewis Structure "Rules" - Chemistry 301
    In a structure, each atom is assigned a formal charge based upon the number of valence electrons for that atom as well as the distribution of electrons in the ...
  22. [22]
  23. [23]
    Writing Lewis Structures: Obeying The Octet Rule
    Sep 13, 2009 · A Lewis structure consists of the electron distribution in a compound and the formal charge on each atom.
  24. [24]
    CHE 120 - Introduction to Organic Chemistry - Textbook: Chapter 1
    Aug 18, 2025 · Condensed chemical formulas show the hydrogen atoms (or other atoms or groups) right next to the carbon atoms to which they are attached. Line- ...
  25. [25]
    CH103 - Chapter 5: Covalent Bonds and Introduction to Organic ...
    A structural formula shows how the various atoms are bonded, and can be more useful that only writing the molecular formula for a compound. There are various ...
  26. [26]
    Structure and Bonding – Organic Chemistry: Fundamental Principles ...
    Line-Angle Formulas (sometimes referred to as skeletal drawings): Skeletal structures are commonly used to represent organic compounds, in particular for cyclic ...
  27. [27]
    "Organic Chemistry in Virtual Reality: Bridging Gaps between Two ...
    May 22, 2024 · The traditional two-dimensional representations in organic chemistry education highlighted the lack of depth and interactivity, ...
  28. [28]
    Compare-Contrast-Connect: Chemical Structures—Visualizing the ...
    Lewis dot structures are two-dimensional representations of molecules that illustrate each atom as its chemical symbol.Missing: definition | Show results with:definition<|control11|><|separator|>
  29. [29]
    A Cache of Chemistry Models - Caltech Magazine
    Feb 26, 2019 · Early models, dating back as far as the 1860s, featured balls representing atoms and sticks signifying chemical bonds.
  30. [30]
    The Chemical Components of a Cell - Molecular Biology of ... - NCBI
    (B) Molecules formed from these atoms have a precise three-dimensional structure, as shown here by ball and stick models (more...) Go to: There Are Different ...
  31. [31]
    Precision space‐filling atomic models - Koltun - Wiley Online Library
    First published: December 1965 ; Citations · 151 ; The Corey-Pauling-Koltun Models to construct macromolecules soon will be available for research and teaching.Missing: history | Show results with:history
  32. [32]
    Primary structure and the CPK model kit - BSCI 1510L Literature and ...
    Sep 26, 2024 · CPK space-filling models incorporate both the covalent and van der Waals radii of atoms into their design. They approximate the true shape of a molecule.
  33. [33]
  34. [34]
    Surface - PyMOLWiki
    Oct 24, 2016 · The surface representation of a protein, in PyMol, shows the "Connolly" surface or the surface that would be traced out by the surfaces of waters in contact ...Surface quality · Surface color · Surface mode · Surface type
  35. [35]
    Staggered vs Eclipsed Conformations of Ethane
    Feb 28, 2020 · In ethane's eclipsed conformation, hydrogens are directly in front of each other, while in the staggered, they are spaced 60 degrees apart.
  36. [36]
    Wedge and Dash Representation - Chemistry Steps
    Wedge and dash representation is used to indicate whether a group on a tetrahedral carbon is pointing towards us or away from us.
  37. [37]
    X-ray crystallography at subatomic resolution - Europhysics News
    Deviations from the spherical atom model appear as electron density peaks in the bonds on deformation electron den- sity maps (calculated by the difference ...
  38. [38]
    An Introduction to Biological NMR Spectroscopy - PMC
    Abstract. NMR spectroscopy is a powerful tool for biologists interested in the structure, dynamics, and interactions of biological macromolecules.
  39. [39]
    Development of a Teaching Approach for Structure Elucidation ...
    Jul 9, 2025 · 2D COSY and 2D NOESY can be used in a complementary fashion to fully assign the 1H NMR spectrum. The factors that affect chemical shifts should ...
  40. [40]
    McLafferty Rearrangement - an overview | ScienceDirect Topics
    The McLafferty rearrangement, discovered in 1952, is the only named reaction in mass spectrometry. For 2-pentanone, the reaction is represented by eqn [2].
  41. [41]
    Power of Infrared and Raman Spectroscopies to Characterize Metal ...
    Dec 14, 2020 · The C═O modes of the dimers of aliphatic carboxylic acids are usually observed between 1740 and 1700 cm–1 and the bands are very intense in the ...
  42. [42]
    IR Absorption Frequencies - NIU - Department of Chemistry and ...
    Typical IR Absorption Frequencies For Common Functional Groups ; C=O · Aldehyde, 1740–1720 (s) ; Ketone, 1725–1705 (s) ; Carboxylic Acid, 1730–1700 (s) ; Ester, 1750 ...
  43. [43]
    Combination of NMR spectroscopy and X-ray crystallography offers ...
    Feb 14, 2011 · Combining NMR and X-ray crystallography offers unique power for elucidating protein assembly structures. NMR's sensitivity helps obtain high- ...
  44. [44]
    Best‐Practice DFT Protocols for Basic Molecular Computational ...
    This work provides best‐practice guidance on the numerous methodological and technical aspects of DFT calculations in three parts.
  45. [45]
    A mathematical and computational review of Hartree–Fock SCF ...
    We present a review of the fundamental topics of Hartree–Fock theory in quantum chemistry. From the molecular Hamiltonian, using and discussing the Born– ...Missing: seminal | Show results with:seminal
  46. [46]
    Merck molecular force field. I. Basis, form, scope, parameterization ...
    This article introduces MMFF94, the initial published version of the Merck molecular force field (MMFF). It describes the objectives set for MMFF, the form ...
  47. [47]
    Gaussian.com | Expanding the limits of computational chemistry
    New Chemistry in Gaussian 16. Gaussian 16 expands the range of molecules and types of chemical problems that you can model.G09 · List of Gaussian Keywords · Contacting Gaussian, Inc. · GaussView 6
  48. [48]
    Molecular dynamics and protein function - PNAS
    In this Perspective we survey two areas, protein folding and enzymatic catalysis, in which simulations have contributed to a general understanding of mechanism.
  49. [49]
    Semiempirical Quantum Mechanical Methods for Noncovalent ...
    Apr 13, 2016 · Semiempirical (SE) methods can be derived from either Hartree–Fock or density functional theory by applying systematic approximations, ...
  50. [50]
    Critical Assessment of the Performance of Density Functional ... - NIH
    In this work we assess the ability of many DFT methods to accurately determine atomic and molecular properties for small molecules containing elements commonly ...
  51. [51]
    Molecular Docking: From Lock and Key to Combination Lock - NIH
    Accurate modeling of protein ligand binding is an important step in structure-based drug design, is a useful starting point for finding new lead compounds ...
  52. [52]
    Chirality and Stereoisomers - Chemistry LibreTexts
    Oct 26, 2025 · Chirality. Chirality essentially means 'mirror-image, non-superimposable molecules', and to say that a molecule is chiral is to say that its ...
  53. [53]
    6.1: Chiral Molecules - Chemistry LibreTexts
    Dec 30, 2019 · The term chiral, from the Greek work for 'hand', refers to anything which cannot be superimposed on its own mirror image. Certain organic ...
  54. [54]
    Enantiomers vs Diastereomers vs The Same? Two Methods For ...
    Mar 8, 2019 · Enantiomers are stereoisomers that are non-superimposable mirror images. Diastereomers are stereoisomers that are not non-superimposable mirror ...
  55. [55]
  56. [56]
    5.4: Optical Activity - Chemistry LibreTexts
    Dec 15, 2021 · Specific rotation is the characteristic property of an optical active compound. The literature specific rotation values of the authentic ...
  57. [57]
    14.3: Chiral Chromatography - Chemistry LibreTexts
    Aug 20, 2020 · The direct resolution of enantiomers can be achieved by interaction of a racemic mixture with a chiral selector, either a part of a chiral ...
  58. [58]
    Thalidomide - American Chemical Society
    Sep 2, 2014 · The (R)-enantiomer, shown in the figure, has sedative effects, whereas the (S)-isomer is teratogenic. Under biological conditions, the isomers ...
  59. [59]
    3.6: Conformations of Ethane - Chemistry LibreTexts
    Sep 23, 2024 · Experiments show that there is a small (12 kJ/mol; 2.9 kcal/mol) barrier to rotation and that some conformations are more stable than others.
  60. [60]
    18. Conformational Analysis of Alkanes - Maricopa Open Digital Press
    Called the gauche conformation, it lies 3.8 kJ/mol (0.9 kcal/mol) higher in energy than the anti conformation even though it has no eclipsing interactions.
  61. [61]
    Constitutional Isomers with Practice Problems - Chemistry Steps
    Constitutional (structural) isomers are compounds with the same formula but different connectivity. Below are a few more examples of constitutional isomers.
  62. [62]
    Determination of keto–enol equilibrium constants and the kinetic ...
    The keno–enol equilibrium constants of acetylacetone, ethyl acetoacetate and ethyl benzoylacetate in water at 25 °C are determined by studying the influence ...
  63. [63]
    5.10: A Review of Isomerism - Chemistry LibreTexts
    Jul 30, 2024 · Constitutional isomers (Section 3.3) are compounds whose atoms are connected differently. Among the kinds of constitutional isomers we've seen ...
  64. [64]
    Butane rotation - Oregon State University
    The equilibrium between anti and gauche is a function of the energy difference (0.9 kcal/mol) and comes out to be Keq = 4.57 at room temperature. This works ...
  65. [65]
    [PDF] Chapter 3
    Ring flip interchanges the axial and equatorial positions. The barrier to a chair-chair interconversion is 45 KJ/mol. 45 KJ/mol. 63.
  66. [66]
    Daylight Theory: SMILES
    3.2 SMILES Specification Rules. SMILES notation consists of a series of characters containing no spaces. Hydrogen atoms may be omitted (hydrogen-suppressed ...
  67. [67]
    About the InChI Standard - InChI Trust
    InChI is a structure-based chemical identifier, developed by IUPAC and the InChI Trust. It is a standard identifier for chemical databases.
  68. [68]
    The IUPAC International Chemical Identifier (InChI)
    Abstract : The central token of information in Chemistry is a chemical substance, an entity that can often be represented as a well-defined chemical structure.<|control11|><|separator|>
  69. [69]
    MDL MOL format (mdl, mol, sd, sdf) - Open Babel - Read the Docs
    Open Babel supports an extension to the MOL file standard that allows cis/trans and tetrahedral stereochemistry to be stored in 0D MOL files.
  70. [70]
    MDL MOL files - Chemaxon Docs
    Basic export options · Compression and Encoding · Document formats · Graphics Formats · Molecule file conversion with Molconvert ...
  71. [71]
    File Format Documentation - wwPDB
    Protein Data Bank: a computer-based archival file for macromolecular structures. J. Mol. Biol. 112, 535-542.
  72. [72]
    RCSB PDB: Homepage
    RCSB Protein Data Bank (RCSB PDB) enables breakthroughs in science and education by providing access and tools for exploration, visualization, and analysis.Guide to understanding pdb data · About RCSB PDB · PDB Statistics · PDB History
  73. [73]
    MMCIF USER GUIDE
    Jun 7, 2024 · PDBx/mmCIF provides the foundation for the deposition, annotation, and archiving of structural data across various experimental techniques.
  74. [74]
    RDKit
    RDKit is open-source cheminformatics software. It has Python and C++ APIs, and downloadable documentation.The RDKit Documentation · RDKit 2012 UGM · Python API Reference
  75. [75]
    PubChem structure search - NIH
    PubChem Structure Search allows the PubChem Compound Database to be queried by chemical structure or chemical structure pattern.
  76. [76]
    [PDF] ISIS/Draw
    With ISIS/Draw, scientists can graphically represent almost any kind of chemical or polymer structure for use in reports, publi- cations, and presentations. • ...
  77. [77]
    ISIS/Draw - An Introductory Guide
    Feb 8, 2019 · This is an introduction to ISIS/Draw, a program from MDL that is free for non-commercial use. You can use it to draw chemical structures, and export them for ...
  78. [78]
    [PDF] chemfig - CTAN
    Dec 1, 2015 · Anyway, I hope that this package will help all LATEX users wishing to draw molecules. Although chemfig has been thoroughly tested and ...
  79. [79]
    Package xymtex - Typesetting chemical structures - CTAN
    XyMTeX is a set of packages for drawing a wide variety of chemical structural formulas in a way that reflects their structure. The package provides three ...
  80. [80]
    [PDF] XΥMTEX: Reliable Tool for Drawing Chemical Structural Formulas
    The development of the XΥMTEX system highly reflects the personal history of my researches aiming at the integration of chemistry and mathematics.
  81. [81]
    Chemistry formulae - Overleaf, Online LaTeX Editor
    This article highlights two LaTeX packages designed for typesetting chemical content and documentation: chemfig: a package to draw structural formulae of ...
  82. [82]
    ChemDoodle Web Components | JavaScript HTML5 Chemistry
    ChemDoodle Web Components allow the wielder to present publication quality 2D and 3D graphics and animations for chemical structures, reactions and spectra.Demos > 2D Sketcher · IUPAC Naming · ChemDoodle Mobile · ChemDoodle 3D
  83. [83]
    ChemDoodle Web Components: HTML5 toolkit for chemical ... - NIH
    Jul 16, 2015 · ChemDoodle Web Components is a very useful toolkit for chemical graphics display, cheminformatics, and chemistry app development for the web.
  84. [84]
    Style guide for chemical structures | Nature
    In such cases we recommend that the structure is first drawn using the standard settings and then scaled down to fill the available width.
  85. [85]
    Graphics Guidelines for - Chemistry Europe
    The ChemDraw template provided in the Author Guidelines is set up for drawing chemical structures in the correct dimensions and font sizes. Chemical formulae ...
  86. [86]
    Considering best practices in color palettes for molecular ...
    Jun 22, 2022 · This paper identifies some of the rationale for the broad use of color in molecular visualizations, provides a set of contemporary color palette examples,Missing: typesetting | Show results with:typesetting
  87. [87]
    Graphical Representation Standards for Chemical Structure Diagrams
    May 2, 2025 · Recommendations are provided for the display of two-dimensional chemical structure diagrams in ways that avoid ambiguity and are likely to be understood ...
  88. [88]
    [PDF] On the use of italic and roman fonts for symbols in scientific text
    Scientific manuscripts frequently fail to follow the accepted conventions concerning the use of italic and roman fonts for symbols. An italic font is ...
  89. [89]
    [PDF] Accessible Chemical Structural Formulas through Interactive ...
    The task of identifying molecular structure images in documents is challenging due to various reasons, especially because of complexity of molecule structures.