Fact-checked by Grok 2 weeks ago

Structural formula

A structural formula is a diagrammatic representation in chemistry that illustrates the arrangement of atoms within a and the chemical bonds connecting them, providing a visual depiction of molecular connectivity. Unlike a molecular formula, which simply lists the types and numbers of atoms (e.g., C₂H₆O for ethanol), or an empirical formula, which gives the simplest whole-number ratio of atoms (e.g., CH₃O), a structural formula explicitly shows how atoms are linked, enabling distinction between isomers like ethanol (CH₃CH₂OH) and dimethyl ether (CH₃OCH₃). Structural formulas exist in several formats to balance detail and clarity, particularly in . Expanded structural formulas display every and individually, such as H–O–H for , highlighting all covalent connections. Condensed structural formulas simplify this by grouping atoms and omitting some bonds, as in CH₃CH₂OH for , while skeletal or line-angle formulas—common for complex organic molecules—represent carbon atoms implicitly at line intersections or ends, with hydrogens assumed to fill valences, focusing on the carbon and functional groups. structures, another variant, additionally include lone pairs, multiple bonds (single, double, triple), and formal charges to convey electron distribution, as seen in (CH₄) with four single bonds. These representations are fundamental in chemistry, especially , where a molecular alone cannot uniquely identify a due to possible structural isomers differing in atom and thus in and reactivity. Structural formulas embody the theoretical core of the discipline by serving as both descriptive notations and predictive models for molecular behavior, facilitating the design of syntheses, analysis of reaction mechanisms, and comprehension of . For instance, they reveal how double bonds or rings influence counts in stable carbon-based compounds, underscoring their role in advancing chemical understanding and applications.

Fundamentals of Structural Formulas

Definition and Purpose

A structural formula is a diagrammatic representation that depicts the arrangement of atoms within a , explicitly showing the connections between atoms via chemical bonds, in contrast to a molecular formula, which only specifies the types and numbers of atoms present without indicating their linkages. For instance, the molecular formula C₆H₆ represents but does not reveal its cyclic structure, whereas a structural formula illustrates the hexagonal ring of carbon atoms with alternating double bonds. This distinction is crucial because molecular formulas alone cannot differentiate between compounds with the same atomic composition but different connectivities, such as the various isomers of C₄H₁₀O. The concept of structural formulas originated in the , pioneered by chemists seeking to explain the and in molecules. Friedrich played a pivotal role in 1858 by proposing that carbon atoms could form chains, establishing tetravalency as a key principle, and in 1865, he introduced the cyclic structural formula for , depicting it as a ring of six carbon atoms to account for its stability and reactivity. This innovation marked a shift from empirical observations to visual models of molecular architecture, enabling chemists to represent bonding patterns systematically. Structural formulas serve to elucidate isomerism by highlighting variations in atomic connectivity, which lead to distinct molecular behaviors despite identical molecular formulas; for example, they distinguish n-butane from , both C₄H₁₀, by showing linear versus branched chains. They also facilitate the depiction of reaction mechanisms, allowing chemists to trace breaking and forming in stepwise processes, as seen in illustrations of nucleophilic substitutions where explicit arrangements clarify structures. Furthermore, by revealing patterns, structural formulas help predict physical influenced by molecular , such as points differing between straight-chain and branched alkanes due to variations in intermolecular forces.

Representation of Bonds

In structural formulas, covalent bonds are visually represented by lines connecting symbols, with the type of indicated by the number and style of lines used. A , involving the sharing of one pair of s between two atoms, is depicted as a single solid straight line; for instance, the C-H bonds in are shown this way. Double bonds, which share two pairs, are represented by two parallel solid lines, as seen in the C=O of (H₂C=O). Triple bonds, sharing three pairs, use three parallel lines, such as in the N≡N of nitrogen gas. Non-covalent interactions, like hydrogen bonds or van der Waals forces, are conventionally shown with dashed lines to distinguish them from covalent linkages, emphasizing their weaker nature without implying electron sharing. These representations adhere to fundamental valence rules, ensuring that atoms achieve stable electron configurations as per the for elements in the second period of the periodic table, where eight electrons are typically sought. In depictions, this means carbon, for example, forms four bonds to complete its octet, as illustrated in (CH₄), where the central carbon connects to four s via single bonds, each satisfying its duet rule with one bond. Similar conventions apply to other atoms: oxygen often forms two bonds and has two s (implied in basic structural formulas), while forms three bonds and one . Violations of the occur in certain cases, such as with elements beyond the second period, but standard organic structural formulas prioritize octet compliance for main-group elements unless exceptions like compounds (e.g., BF₃ with six electrons around ) are specified. For aromatic compounds, bond representation departs from simple alternating single and double bonds to convey delocalization. In (C₆H₆), the classic Kekulé structure uses three alternating double bonds in a hexagonal to suggest , but a more accurate notation employs a inscribed within the hexagon to symbolize the uniform, delocalized π-electron across all six C-C bonds, each of equal (approximately 1.39 ). This convention highlights the stability of the aromatic system without implying localized double bonds. Although two-dimensional structural formulas do not explicitly depict , they implicitly convey standard molecular geometries based on hybridization and valence electron pairs. For sp³-hybridized carbons with four single bonds, a tetrahedral with of about 109.5° is assumed, as in the carbon framework of alkanes; deviations, such as in cyclopropane's strained 60° , are noted separately but not shown in basic line drawings. This implicit geometry aids in understanding reactivity and shape without requiring three-dimensional projections.

Representation of Electrons and Charges

In structural formulas, non-bonding electrons, known as lone pairs, are represented by pairs of dots placed adjacent to the to which they belong, illustrating the complete configuration without overlapping with representations. For example, in the (H₂O), the oxygen is depicted with two lone pairs as four dots (two pairs) positioned above and below the , alongside its two single bonds to atoms. This notation emphasizes the fulfillment for second-period elements, where lone pairs contribute to the 's stability. Formal charges on atoms within a structural formula are indicated by superscript numbers placed next to the atom, such as +1 or -1, to denote deviations from neutrality due to distribution in bonds and lone pairs. The is calculated using the formula: \text{Formal charge} = (\text{valence electrons}) - (\text{non-bonding electrons}) - \frac{1}{2} (\text{bonding electrons}) For instance, in a like the methyl cation (CH₃⁺), the central carbon bears a +1 formal charge as a superscript, reflecting its six valence electrons (three from bonds and none non-bonding) compared to its group 4A valence of four. This convention helps identify reactive sites and validate resonance structures in molecules. Unpaired electrons, characteristic of , are shown as a single dot adjacent to the atom or group, distinguishing them from paired lone pairs or shared bonding electrons. In the methyl (CH₃•), the dot follows the formula to signify the carbon's , highlighting its high reactivity and odd-electron nature. This simple dot notation is essential for depicting involved in chain reactions and ./01%3A_Structure_and_Bonding/1.04%3A_Lewis_Structures_Continued) For polyatomic ions, the entire structural formula is enclosed in square brackets, with the net ionic charge indicated as a superscript outside the brackets to convey the overall electron imbalance distributed across the . The sulfate ion, for example, is represented as [O₃S(=O)₂]²⁻ (or in dot notation with lone pairs), where the 2- charge outside the brackets accounts for the two extra electrons beyond neutral sulfur and oxygen valences. This bracketing ensures clarity in ionic compounds and distinguishes the ion from molecules.

Planar Structural Representations

Lewis Structures

Lewis structures, also known as Lewis dot diagrams or electron-dot structures, represent the arrangement of valence electrons in atoms, ions, and molecules, illustrating covalent bonds as shared electron pairs and lone pairs on atoms. Developed by in his 1916 paper "The Atom and the Molecule," these diagrams emphasize the role of valence electrons in forming stable octet configurations, where most atoms achieve eight electrons in their outer shells to mimic stability. This approach provides a visual tool for predicting , reactivity, and bonding without delving into . The construction of a Lewis structure involves a step-by-step process to ensure accurate electron distribution:
  1. Calculate the total number of valence electrons by summing the valence electrons from all atoms in the formula, adding electrons for negative charges or subtracting for positive charges.
  2. Sketch the skeletal framework by arranging atoms, placing the central atom (typically the least electronegative, excluding ) and connecting others with single bonds represented as lines or pairs of dots.
  3. Distribute the remaining valence electrons as lone pairs to peripheral atoms first, aiming to complete their octets (or duets for ).
  4. Assign any leftover electrons to the central atom; if its octet is incomplete, form multiple bonds by converting lone pairs into shared double or triple bonds as needed.
For (NH3), the features as the central atom with three single bonds to atoms (each bond as a pair of shared electrons) and one on , resulting in eight electrons around and two around each . This satisfies the for while highlighting its potential as a Lewis base due to the . Certain molecules require resonance structures to depict delocalized electrons that cannot be confined to a single diagram; the actual bonding is a hybrid of these equivalent forms. In (O3), two resonance structures show a alternating between the central oxygen and each terminal oxygen, with the real exhibiting partial double-bond character in both O-O links and 1.5 . Exceptions to the arise in cases where stable structures deviate from eight valence electrons around an atom. Odd-electron molecules, like the (NO2), contain an , preventing all atoms from achieving octets and often resulting in reactive free radicals. Expanded octets occur in hypervalent compounds such as (SF6), where sulfur bonds to six fluorines using 12 electrons, possible for main-group elements beyond the second period, which can exceed the octet through hypervalent bonding mechanisms. Electron-deficient species, exemplified by (BF3), have with only six electrons in three bonds, making it a prone to accepting additional electron pairs. These exceptions underscore the octet rule's utility as a guideline rather than an absolute law, guiding structure prediction while accommodating diverse bonding scenarios.

Condensed Formulas

Condensed structural formulas represent the connectivity of atoms in a using a linear, text-based notation that omits explicit lines while grouping atoms and substituents to convey structure efficiently. This format is particularly useful in for depicting carbon chains and branches without the visual complexity of full diagrams. The notation relies on implied single bonds between adjacent atoms, with carbon atoms assumed at intersections or chains unless specified otherwise, and atoms grouped directly with their attached carbons using subscripts for multiples. Parentheses are employed to indicate branches or substituents attached to the same atom, ensuring clarity in non-linear arrangements; for example, (2-methylpropane) is written as CH₃CH(CH₃)CH₃, where the parentheses denote the branching from the second carbon. In straight-chain alkanes, repeating units are abbreviated, as in n-octane represented as CH₃(CH₂)₆CH₃, which compactly shows the six methylene groups between terminal methyls. One key advantage of condensed formulas is their compactness, making them ideal for describing long or repetitive carbon chains without drawing extensive lines, thus facilitating quick communication in chemical literature and calculations. This brevity retains essential connectivity information while reducing the space required compared to expanded structural representations. However, condensed formulas can introduce limitations in clarity, particularly with complex branching, where arises if parentheses are omitted or misinterpreted, potentially leading to confusion between isomers. Proper use of grouping symbols is essential to avoid such issues, though they do not depict three-dimensional aspects or multiple bonds as explicitly as other notations. To convert a to a condensed , remove all explicit lines and electron dots, then group atoms with their respective carbons, using subscripts for identical groups and parentheses for branches to preserve the skeletal arrangement. This process simplifies the visual while maintaining the molecular topology. Skeletal formulas serve as further simplifications by omitting indications entirely.

Skeletal Formulas

Skeletal formulas, also known as line-angle or bond-line notations, provide a streamlined visual of molecules by emphasizing the of the carbon framework. In this system, each or terminus of a line denotes a carbon atom, while the lines themselves signify covalent bonds, with single bonds as solid lines and double bonds as paired lines or indicated explicitly. Hydrogen atoms attached to carbon are routinely omitted, as they are inferred to satisfy the four-valence requirement of carbon, thereby reducing visual complexity without sacrificing essential structural information. To reflect the tetrahedral arrangement around carbon atoms, linear chains in skeletal formulas are conventionally drawn in a zig-zag , where each represents a C-C angled to mimic near 109.5°. This approach facilitates quick recognition of chain length and branching, as intersections indicate branch points or ring closures. For cyclic structures, bonds form closed polygons, with commonly illustrated as a regular featuring three alternating double bonds to denote its , or equivalently, a enclosing a circle to symbolize uniform electron delocalization across the ring./15%3A_Benzene_and_Aromatic_Compounds/15.02%3A_The_Structure_of_Benzene) In , skeletal formulas serve as the standard for illustrating intricate molecules like steroids, which possess a characteristic tetracyclic core of three six-membered s fused to one five-membered , allowing chemists to focus on placements and stereocenters amid dense connectivity. Heteroatoms, such as oxygen or , are explicitly labeled with their atomic symbols at bond junctions or ends, and electrons may be depicted as dots when clarification of reactivity or charge is required, though they are frequently implied based on standard valences. Basic skeletal formulas assume a two-dimensional planar , though wedge and dash notations can be incorporated briefly to denote at chiral centers.

Stereochemical and Perspective Representations

Basic Stereochemistry Notation

Basic stereochemistry notation in structural formulas provides a two-dimensional to indicate the spatial of atoms around chiral centers and double bonds, allowing chemists to depict stereoisomers without requiring full three-dimensional models. These notations are essential for representing molecules where the arrangement of substituents affects properties such as reactivity and . Commonly applied to planar representations like skeletal formulas, they use simple line conventions to convey depth or relative positioning. For chiral centers, typically tetrahedral carbon atoms with four different substituents, wedge and dash notations specify whether a bond projects toward or away from the viewer. A solid represents a substituent coming out of the toward the observer, while a dashed line or hashed indicates a substituent receding behind the . These conventions, with the narrow end of the attached to the stereogenic , unambiguously define the in a drawing. For example, in (2R,3R)-, solid wedges are used for the hydroxyl groups projecting forward. Hashed lines, consisting of parallel short dashes, are an alternative to plain dashed lines for enhanced clarity in printed media. In alkenes and other compounds with restricted rotation around s, cis-trans isomerism is denoted by the relative positioning of substituents on adjacent carbons. When the double bond is represented linearly, forward slashes (/) and backslashes () on the connecting bonds indicate whether substituents are on the same () or opposite () sides; for instance, in (E)-2-butene, the methyl groups are shown with opposing slashes to denote trans configuration. This slash notation is particularly useful in condensed or skeletal structural formulas to avoid ambiguity in chain depictions. For more complex cases where substituents differ, the E/Z system supersedes cis/trans, but the slash method remains a visual aid in drawings. When the at a center or is unspecified or represents a , wavy lines are employed to connect substituents, signaling an unknown or racemic without implying a specific . The "rac" may accompany such drawings to explicitly denote a of enantiomers. According to IUPAC guidelines, wavy bonds should be used judiciously, often with accompanying text for clarity, and are preferred over plain bonds when is intentionally ambiguous. To assign at chiral centers in these notations, the priority rules are applied: substituents are ranked by (or further by attached atoms), the lowest-priority group is oriented away, and the is designated (clockwise) or (counterclockwise) when viewing the remaining groups in decreasing priority. This system integrates seamlessly with wedge-dash drawings by determining the order after establishing the perspective. The rules, formalized in 1966, ensure consistent across structural representations.

Fischer Projections

Fischer projections are a two-dimensional convention for depicting the three-dimensional arrangement of with multiple chiral centers, particularly useful for linear representations of biomolecules. Developed by German chemist in 1891 to elucidate the of carbohydrates, this method simplifies the visualization of tetrahedral geometries by projecting the molecule onto a plane while adhering to specific orientation rules. The standard drawing rules for Fischer projections position the main carbon chain vertically, with the most oxidized carbon (such as a ) placed at the top and the least oxidized at the bottom. Horizontal bonds are understood to project forward out of the plane of the paper toward the viewer, while vertical bonds extend backward behind the plane. This eclipsed conformation assumes all bonds are in the same plane for simplicity, though actual molecules adopt staggered arrangements; the projection prioritizes stereochemical clarity over conformational accuracy. A classic example is the of D-glucose, an aldohexose, which illustrates the D configuration through the positioning of hydroxyl groups. The open-chain form is drawn as follows, with the at the top and the CH₂OH at the bottom:
  CHO
  |
H-C-OH
  |
HO-C-H
  |
H-C-OH
  |
H-C-OH
  |
 CH₂OH
Here, the hydroxyl groups on carbons 2, 4, and 5 point to the right (horizontal bonds forward), while the one on carbon 3 points to the left, defining the specific stereoisomer. This arrangement corresponds to the (2R,3S,4R,5R) at the chiral centers. Interconversions of projections must preserve ; a 180° in the plane of the paper yields an equivalent representation, as it maintains the relative positions of substituents. However, a 90° or 270° inverts the , effectively generating the , so such manipulations are invalid without adjustment. To assign R/S designations, one can mentally swap two horizontal substituents to reorient the lowest-priority group vertically (backward), then apply the Cahn-Ingold-Prelog priority rules while viewing the projection as a ; an odd number of swaps inverts the . Fischer projections find primary applications in representing the of , where the / designation is based on the configuration at the penultimate carbon (OH on the right for D-series), and in , with natural L- showing the amino group on the left when the is at the top. These projections facilitated resolution of sugar enantiomers and remain standard in biochemical for open-chain forms.

Newman and Sawhorse Projections

Newman projections are a type of structural representation used in to visualize the three-dimensional conformation of molecules by looking straight down a specific , typically a carbon-carbon ./Chapter_03:_Structure_of_Alkanes/3.4:_Structure_and_Conformations_of_Alkanes/3.4.1:_Newman_Projections) This method was introduced in 1952 by Melvin S. Newman to facilitate the analysis of rotational isomers and torsional strain in alkanes. In a , the front carbon atom is represented as a dot at the center of a circle, with its three substituents depicted as lines extending outward at 120-degree angles; the rear carbon is shown as a larger circle, with its substituents as lines projecting from the circle's perimeter./Chapter_03:_Structure_of_Alkanes/3.4:_Structure_and_Conformations_of_Alkanes/3.4.1:_Newman_Projections) For (C₂H₆), the staggered conformation—where the front and rear atoms are offset by 60 degrees—represents the lowest-energy arrangement due to minimized steric repulsion, while the eclipsed conformation, with substituents aligned at 0 degrees, is a high-energy ./Chapter_03:_Structure_of_Alkanes/3.4:_Structure_and_Conformations_of_Alkanes/3.4.1:_Newman_Projections) Central to interpreting Newman projections is the concept of the , which measures the torsion between bonds on adjacent atoms and is defined as the angle between two planes formed by four atoms connected to the two central atoms in question. In these projections, dihedral angles determine the relative positions of substituents: an (or antiperiplanar) arrangement occurs at 180 degrees, maximizing separation; a gauche position is at 60 degrees, introducing some steric interactions; and a synperiplanar (or eclipsed) alignment at 0 degrees leads to significant torsional strain. These terms describe the conformational preferences driven by van der Waals repulsions and effects, allowing chemists to predict molecular stability without computational modeling. Sawhorse projections complement Newman projections by providing a view of molecular conformations from an oblique angle, resembling the shape of a carpenter's sawhorse, with the carbon-carbon bond drawn as a zigzag line and substituents extending at tetrahedral angles. This representation emphasizes the three-dimensional torsion around the bond, making it easier to visualize angles and substituent orientations in a way that bridges two-dimensional drawings and physical models. Unlike the end-on of Newman projections, sawhorse drawings highlight the spatial relationships in acyclic chains, aiding in the qualitative assessment of steric hindrance. A key application of these projections is in analyzing n-butane (C₄H₁₀), where rotation around the central C2-C3 bond yields distinct conformers. The anti conformation, with the terminal methyl groups at a 180-degree , is the global energy minimum due to optimal spacing that avoids steric clashes between the methyl groups./Chapters/Chapter_04:_Alkanes/3.09:_Conformations_of_Butane) In contrast, the gauche conformer at 60 degrees is slightly higher in energy by approximately 0.9 kcal/mol (3.8 kJ/mol), reflecting the partial overlap of the methyl groups, while eclipsed forms represent barriers to rotation./Chapters/Chapter_04:_Alkanes/3.09:_Conformations_of_Butane) This energy profile, observable in both Newman and sawhorse views, underscores the preference for staggered arrangements in alkanes and informs reactivity in larger hydrocarbons.

Haworth Projections

Haworth projections provide a simplified two-dimensional representation of the cyclic forms of monosaccharides, particularly (five-membered) and (six-membered) rings, by depicting the ring as a flat polygon with substituents oriented above or below the plane to convey stereochemical information. This convention, introduced by British chemist Walter Norman Haworth during his structural elucidations of carbohydrates in the late , facilitates the visualization of anomeric configurations and other chiral centers relative to the ring plane. The construction of a begins with the of the open-chain , where cyclization occurs via formation, typically involving the carbonyl at C1 and the hydroxyl at C4 or C5. The is drawn as a regular for forms, with the ring oxygen positioned at the upper right corner and the anomeric carbon (C1) at the rightmost vertex; for forms, a pentagon is used with similar orientation. Substituents that appear on the right side of the are placed below the ring plane (dashed lines or downward bonds), while those on the left are placed above (wedged lines or upward bonds); in D-series sugars, the CH₂OH group at C5 projects upward. This method preserves the relative while approximating the envelope-like pucker of the actual . A representative example is the Haworth projection of β-D-glucopyranose, the predominant cyclic form of D-glucose. The hexagon ring has the anomeric hydroxyl (at C1) oriented upward (indicating the β configuration), the hydroxyl at C2 downward, the hydroxyl at C3 upward, the hydroxyl at C4 downward, and the CH₂OH at C5 upward; all these positions correspond to equatorial orientations in the more accurate conformation. This depiction highlights the all-equatorial arrangement that contributes to the stability of β-D-glucopyranose in . The influences the at the anomeric carbon in projections, favoring the axial (downward in standard depiction) orientation of electronegative substituents like the hydroxyl group in α-anomers due to hyperconjugative stabilization involving the ring oxygen's , which overrides typical steric preferences for equatorial positions. In glucose, however, the β-anomer (equatorial, upward) predominates because of additional anomeric stabilization from solvent interactions, but the effect is pronounced in other sugars like or in glycosides where axial configurations enhance reactivity in enzymatic contexts. Despite their utility, projections have limitations in accurately conveying three-dimensional structure, as they portray the as fully planar rather than puckered, omitting details of bond angles and torsional strains present in real or conformations that affect stability and biological function.

, a six-membered carbon , adopts various conformations that are represented in structural formulas to depict its three-dimensional arrangement and energy preferences. The most stable conformation is the form, where the puckers such that adjacent carbon atoms alternate above and below the average plane of the . In this structure, the twelve C-H bonds are divided into six axial bonds, which point vertically parallel to the 's axis of symmetry, and six equatorial bonds, which extend outward at an angle approximately 109.5 degrees from the plane. This arrangement minimizes both angle strain and torsional strain, making the the predominant conformation at , comprising over 99.9% of the population. Less stable conformations include the and twist-boat forms, which arise during dynamic interconversions of the . The conformation features four carbon atoms in a planar arrangement with the remaining two lifted out of the plane, resulting in eclipsed C-C bonds that introduce significant torsional , estimated at about 6.9 kcal/ above the . Additionally, steric repulsion occurs between the "flagpole" hydrogens on the lifted carbons (positions 1 and 4), further elevating its energy to approximately 7.2 kcal/ relative to the . The twist-boat, a slightly distorted version of the , alleviates some of this by twisting the , lowering its energy to about 5.5 kcal/ above the while still being a local minimum. In structural representations, the conformation of is typically drawn as a zigzag with alternating up and down bonds to indicate the puckered shape, while substituents are shown using solid wedges for bonds coming out of the plane (equatorial or axial as appropriate) and dashed wedges for those receding behind the plane. For example, in , the equatorial position for the methyl group is preferred due to reduced steric interactions with axial hydrogens, stabilizing that conformer by about 1.7 kcal/ compared to the axial form. This preference is crucial for predicting reactivity and physical properties in substituted cyclohexanes. The interconversion between chair conformations, known as a , occurs rapidly at through a involving half-chair or twist-boat intermediates, with an energy barrier of approximately 10.8 kcal/mol (45 kJ/mol). This process inverts the axial and equatorial positions but preserves the relative up/down orientation of substituents, allowing dynamic equilibration without breaking bonds. The barrier height was experimentally determined using , confirming the low that enables millions of flips per second per molecule.

Limitations and Extensions

Inherent Limitations

Structural formulas, being two-dimensional depictions, inherently struggle to convey the three-dimensional of molecules, including precise bond angles, torsional strains, and dynamic conformations that influence reactivity and . For instance, the tetrahedral around a carbon atom with 109.5° bond angles in is flattened into a planar , requiring chemists to rely on mental or physical models to infer spatial relationships. This disconnect can lead to incomplete understanding of molecular behavior, as 2D representations omit volumetric aspects and electrostatic distributions that are crucial for interactions like enzyme-substrate binding. Ambiguities arise particularly in simplified forms like skeletal or condensed formulas, where hydrogen atoms are often implied rather than explicitly shown, potentially causing errors in valence counting or isomer identification for those unfamiliar with conventions. Stereochemical information, such as chirality at asymmetric centers, is likewise absent without supplementary notations like wedges or dashes, allowing multiple stereoisomers to appear identical and risking misinterpretation in synthesis or analysis. Even bond representations can be unclear; for example, a linearly aligned might ambiguously suggest cis-trans isomerism if not specified. For large and complex molecules, such as proteins or , structural formulas become overcrowded with intersecting lines and symbols, rendering them impractical for detailed scrutiny and prone to visual clutter that obscures functional groups or . This challenge is exacerbated in biomolecules, where thousands of atoms lead to unwieldy diagrams that prioritize brevity over clarity, often necessitating drastic abbreviations that sacrifice precision. Historically, early structural formulas developed in the by chemists like emphasized connectivity but ignored spatial orientations and quantum mechanical phenomena, such as electron delocalization in structures. These classical depictions treated bonds as fixed and localized, failing to account for the nature of systems like , where quantum effects demand multiple contributing forms beyond a single 2D drawing. This limitation persisted until the integration of in the 1930s, which revealed how traditional formulas oversimplified electron distribution.

Computational and 3D Extensions

Molecular modeling software extends traditional structural formulas by enabling the creation and manipulation of interactive representations of molecules. Tools such as , developed by Signals, allow users to draw 2D structures and generate models with depth-shading and conformational analysis, facilitating of and for chemists and biologists. Similarly, Avogadro, an open-source molecular editor, supports cross-platform and editing, including optimization of molecular structures using force fields, making it widely used in and . These software packages address the static limitations of 2D formulas by providing rotatable, scalable views that reveal spatial arrangements not apparent in planar depictions. SMILES (Simplified Molecular Input Line Entry System) notation serves as a text-based extension for computational handling of structural formulas, allowing compact representation of molecules for input into modeling software and databases. For instance, the SMILES string "CC(O)C" denotes isopropanol, where "C" represents carbon atoms, "O" oxygen, and parentheses indicate branching. This linear format enables automated parsing and generation of coordinates, bridging manual drawing with algorithmic processing in cheminformatics applications. In , visualizations derived from (DFT) calculations produce advanced 3D models that incorporate electronic structure data beyond empirical formulas. Ball-and-stick models, where atoms are depicted as spheres connected by rods representing bonds, and space-filling models, showing atomic van der Waals radii, are commonly generated from DFT outputs to illustrate optimized geometries and electron densities. Software like (VMD) facilitates these representations by loading DFT results from packages such as Gaussian, enabling interactive analysis of molecular orbitals and vibrational modes. As of 2025, AI-driven methods have further integrated structural formulas with predictive , particularly for complex biomolecules. , developed by , predicts protein structures from sequences with high accuracy, outputting 3D coordinates that align with traditional structural notations while revealing folding patterns; recent database updates synchronized with UniProt release 2025_03 have expanded coverage to over 200 million structures. This approach combines sequence-based "formulas" with computed 3D ensembles, accelerating and by providing verifiable models validated against experimental data.

References

  1. [1]
    Chemical Formulas - FSU Chemistry & Biochemistry
    Another way to represent water would be its structural formula: This formula indicates that both of the hydrogens are attached to the oxygen but not each other ...
  2. [2]
    Chemical Formulas - Chembook
    What is shown is a structural formula which means that the way the atoms are connected is shown. Those two formulas can even be "shown" inline by knowing a ...
  3. [3]
    Molecular Structure & Bonding - MSU chemistry
    The number of hydrogen atoms in stable compounds of carbon, hydrogen & oxygen reflects the number of double bonds and rings in their structural formulas.
  4. [4]
    [PDF] LEWIS FORMULAS, STRUCTURAL ISOMERISM, AND ...
    Lewis formulas show atom connectivity, single, double, or triple bonds, formal charges, and unshared electrons. Structural isomers have the same molecular  ...
  5. [5]
    How do Structural Formulas Embody the Theory of Organic ...
    The structural formulas of organic chemistry embody the theory of this discipline because they are both descriptive names and models, the production and ...
  6. [6]
    2.4 Chemical Formulas - Chemistry 2e | OpenStax
    Feb 14, 2019 · The structural formula for a compound gives the same information as its molecular formula (the types and numbers of atoms in the molecule) but ...
  7. [7]
    August Kekulé and Archibald Scott Couper - Science History Institute
    Kekulé famously “saw” carbon atoms joining in a “giddy dance” in a daydream. Couper invented a symbolic language to represent carbon linkage.
  8. [8]
    the 150th anniversary of the Kekulé benzene structure - PubMed
    Jan 2, 2015 · In January 1865, August Kekulé published his theory of the structure of benzene, which he later reported had come to him in a daydream about a snake biting its ...
  9. [9]
    Reaction Mechanisms – Chemistry - UH Pressbooks
    In this figure, structural formulas are used to illustrate a chemical reaction, including an. In a sample of C4H8, a few of the rapidly moving C4H8 molecules ...
  10. [10]
    Chemical Bonding
    A single bond consists of one pair of shared electrons between two atoms. {e.g. H2}, H-H ; A double bond consists of two pairs of shared electrons between two ...
  11. [11]
    4.1 Bonding in Elements
    Representations of Molecular Structures ; single bondA chemical bond formed when two atoms share a single pair of electrons ; double bondA chemical bond formed ...
  12. [12]
    Molecular Interactions (Noncovalent Interactions) - Loren Williams
    Jun 10, 2024 · The dashed lines represent hydrogen bonds. The electrostatic force between two point charges is given by: Force = k q1 q2 / ε r2Missing: formulas | Show results with:formulas
  13. [13]
    Unit 1 - Elaborations - Structural Formulas
    In a skeletal structural formula, chemical symbols are only used to label the non-carbon atoms; if an atom is not labeled it is assumed to be a carbon atom.
  14. [14]
    [PDF] 2. COVALENT BONDING, OCTET RULE, POLARITY, AND BASIC ...
    The acetylene molecule provides an example of a triple bond. This terminology (single, double, or triple bond) is very loose and informal. The formulas shown ...
  15. [15]
    Benzene and Other Aromatic Compounds - MSU chemistry
    We know that benzene has a planar hexagonal structure in which all the carbon atoms are sp2 hybridized, and all the carbon-carbon bonds are equal in length. As ...
  16. [16]
    [PDF] Atomic Structure & Chemical Bonding - Projects at Harvard
    The bond angles found in real molecules can deviate from their ideal values when some bonds or lone pairs are more electron-rich, and therefore more repulsive, ...Missing: implicit | Show results with:implicit
  17. [17]
  18. [18]
    THE ATOM AND THE MOLECULE. - ACS Publications
    Teaching Formal Charges of Lewis Electron Dot Structures by Counting Attachments. ... A Convenient Approach for the Construction of Lewis Structures by Focusing ...
  19. [19]
    Lab Background - FSU | Department of Chemistry & Biochemistry
    Drawing Lewis Structures can be summarized into a series of steps: Count all the valence electrons for each atom. (Add or subtract electrons if it is an ...
  20. [20]
    Lewis Structure for NH 3 (Ammonia)
    Ammonia (NH3) has 8 valence electrons and is commonly tested, with a trigonal pyramidal geometry. Try drawing it before watching the video.
  21. [21]
    Lewis structures: Octet rule violations
    Oct 10, 2023 · Expanded octets. Deficient octets and coordinate covalent bonds. Odd electron species. As a generalization, the octet rule is a simple and ...
  22. [22]
    [PDF] Chem 51A S. King Chapter 1 Structure and Bonding Organic ...
    Condensed structures are a shorthand way of drawing molecules where some or all of the covalent bonds are left out. They are drawn in a way that gives ...
  23. [23]
    Lewis Structures: The Language of Organic Chemistry
    Generally organic chemists depict molecular structures using a convention known as bond line notation. According to this approach, a bond between two atoms is ...
  24. [24]
    [PDF] Chapter 1,2 Carbon Compounds, Chemical Bonds, Lewis Structures ...
    Bonding pairs of electrons are most often represented as straight lines: single bonds as a single line, double bonds as two parallel lines, and triple bonds as ...
  25. [25]
  26. [26]
  27. [27]
    3.4: Chirality and stereoisomers - Chemistry LibreTexts
    Jul 20, 2022 · Conversely, wedges may be used on carbons that are not chiral centers – look, for example, at the drawings of glycine and citrate in the figure ...Missing: notation | Show results with:notation
  28. [28]
    7.5: Cis-Trans Isomerism in Alkenes - Chemistry LibreTexts
    Apr 3, 2024 · This leads to a special kind of isomerism in double bonds. Consider the alkene with the condensed structural formula CH3CH=CHCH3.
  29. [29]
    Stereoisomers - MSU chemistry
    In the isomers illustrated above, for which cis-trans notation was adequate, Z is equivalent to cis and E is equivalent to trans. To see how complex ...
  30. [30]
    Absolute Configuration - R-S Sequence Rules - Chemistry LibreTexts
    Jul 30, 2023 · The method of unambiguously assigning the handedness of molecules was originated by three chemists: R.S. Cahn, C. Ingold, and V. Prelog and is ...Missing: 2D formulas
  31. [31]
    Fischer and Haworth Projections of Carbohydrates - News-Medical
    Oct 22, 2019 · The Fischer projection was devised by German chemist Emil Fischer in 1891, who was the winner of the Nobel Prize in Chemistry in 1902. In a ...
  32. [32]
    5.5: Fisher Projection - Chemistry LibreTexts
    Dec 15, 2021 · 180º rotation of A leads to B, A and B are identical. 3. Rotate the Fisher projection 90º get the configuration inverted. S-configuration ...Missing: interconversion | Show results with:interconversion
  33. [33]
    Ch 7: Fischer Projections - University of Calgary
    ... horizontal lines represent bonds coming out of the plane of the paper and vertical lines represent bonds going behind the plane of the paper. (This is the ...
  34. [34]
    12.2: Fischer Projections - Chemistry LibreTexts
    Dec 27, 2021 · Below are three representations of the open chain form of D-glucose: in the conventional Fischer projection, a wedge/dash version of a Fischer ...
  35. [35]
    How To Determine R and S Configurations On A Fischer Projection
    May 21, 2019 · Recall how the Fischer projection works. The longest carbon chain is arranged vertically and the substituents are drawn out to the side.
  36. [36]
    Fischer Projection: Definition, Illustration, and Examples
    Jul 4, 2020 · Fischer Projection Rules; Converting ... The Fischer projections were designed by German chemist and Nobel laureate Emil Fischer in 1891.
  37. [37]
    12.2: Structures of Amino Acids - Chemistry LibreTexts
    Apr 30, 2025 · In Fischer projections, naturally occurring amino acids are represented by placing the –CO2– group at the top and pointing the side chain ...Missing: applications | Show results with:applications
  38. [38]
    [PDF] Melvin Newman - National Academy of Sciences
    His very first publication from Ohio State University (1937) was entitled. “The Synthesis of 1,2-Benzanthracene Derivatives Related to 3,4-Benzpyrene.” One of ...
  39. [39]
    3.7 Conformations of Other Alkanes - Organic Chemistry | OpenStax
    Sep 20, 2023 · Figure 3.9 Newman projections of propane showing staggered and eclipsed conformations. The staggered conformer is lower in energy by 14 kJ/mol.
  40. [40]
    Norman Haworth – Facts - NobelPrize.org
    Norman Haworth studied the composition of carbohydrates, and around 1928 he mapped the composition and structure of various forms of sugar, starch, and ...Missing: projection origin publication
  41. [41]
    2.6: Cyclic Structures of Monosaccharides - Chemistry LibreTexts
    May 8, 2021 · In a Haworth projection or formula, the molecules are drawn as planar hexagons with a darkened edge representing the side facing toward the ...
  42. [42]
    4.3: Conformation Analysis of Cyclohexane - Chemistry LibreTexts
    Dec 15, 2021 · The most stable conformation of cyclohexane is called “chair” conformation, since it somewhat resembles a chair.Missing: twist- sources
  43. [43]
    The Cyclohexane Chair Flip - Master Organic Chemistry
    May 30, 2014 · Cyclohexane's ground state conformation is the chair, and it can undergo a ring 'flip', where axial substituents become equatorial substituents.Missing: representation | Show results with:representation
  44. [44]
    The Cyclohexane Chair Flip - Energy Diagram
    Jun 6, 2014 · The cyclohexane chair flip "costs" 10 kcal/mol, which is the barrier to go from the chair to the half chair. We map out the energy diagram, ...
  45. [45]
    4.1: 3D and 2D Representations
    ### Summary of Limitations of 2D Representations of Molecules
  46. [46]
    2.1 Drawing and Interpreting Organic Formulas
    Drawing structural formulas is a good starting point for a novice organic chemist, and works when dealing with small, simple structures.
  47. [47]
    The Chemical Bond - ScienceDirect
    Turning to the first set of issues, Kekulé, Crum Brown and Frankland all seem to agree on the limitations of structural formulae, which do not display the ...
  48. [48]
    A Brief History of the Theory of Resonance and of its Interpretation
    Aug 9, 2025 · Pauling was aware of the difficulties chemists faced in understanding such unfamiliar concepts as quantum mechanical resonance and the ...Missing: formulas | Show results with:formulas
  49. [49]
    The Evolution of ChemDraw | Revvity Signals
    Sep 17, 2025 · 3D Conformational Representation: These tools allow chemists to explore stereochemistry, optimize molecular geometries, and generate shareable ...
  50. [50]
    Avogadro - Free cross-platform molecular editor - Avogadro
    Avogadro is an advanced molecule editor and visualizer designed for cross-platform use in computational chemistry, molecular modeling, bioinformatics, materials ...InstallManual
  51. [51]
    Avogadro: an advanced semantic chemical editor, visualization, and ...
    Aug 13, 2012 · The Avogadro project has developed an advanced molecule editor and visualizer designed for cross-platform use in computational chemistry, molecular modeling, ...<|separator|>
  52. [52]
    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 ...
  53. [53]
    OpenSMILES specification
    May 15, 2016 · This document formally defines an open specification version of the SMILES language, a typographical line notation for specifying chemical structure.
  54. [54]
    [PDF] VMD Quantum Chemistry Visualization Tutorial
    With QM data loaded, VMD can display molecular orbitals, as well as access the calculated energy levels and various other data present in the loaded files [4, 3] ...
  55. [55]
    VMD - Visual Molecular Dynamics
    VMD is a molecular visualization program for displaying, animating, and analyzing large biomolecular systems using 3-D graphics and built-in scripting.Download VMD · VMD 1.9.3 · VMD 1.9.3 Documentation · What is VMD?
  56. [56]
    AlphaFold - Google DeepMind
    So far, AlphaFold has predicted over 200 million protein structures – nearly all catalogued proteins known to science. The AlphaFold Protein Structure Database ...
  57. [57]
    Highly accurate protein structure prediction with AlphaFold - Nature
    Jul 15, 2021 · AlphaFold greatly improves the accuracy of structure prediction by incorporating novel neural network architectures and training procedures ...<|control11|><|separator|>
  58. [58]
    EMBL-EBI and Google DeepMind renew partnership and release ...
    Oct 7, 2025 · EMBL-EBI and Google DeepMind announce partnership renewal, and major update to bring the AlphaFold Database in line with UniProt ...