Fact-checked by Grok 2 weeks ago

Substituent

In , a substituent is an atom or group of atoms other than that replaces one or more atoms in a or other parent structure, thereby altering the molecule's properties and reactivity. These groups are typically derived from alkanes, such as methyl (-CH₃) or ethyl (-C₂H₅), or may include functional groups like (-Cl), (-NO₂), or alkoxy (-OR). Substituents play a central in the systematic naming of compounds under IUPAC rules, where they are identified and prefixed to the parent chain name, with numbering chosen to give the lowest possible locants. The presence of substituents significantly influences molecular behavior through electronic effects, including the —where the substituent withdraws or donates electron density via sigma bonds—and the resonance effect, which involves delocalization of electrons through pi bonds, particularly in conjugated systems. In reactions, for instance, electron-donating substituents like alkyl or methoxy groups activate the ring and direct incoming electrophiles to and positions, while electron-withdrawing groups such as or carbonyl deactivate the ring and favor substitution. These effects extend to acidity modulation in carboxylic acids, where electron-withdrawing substituents enhance acidity by stabilizing the conjugate base. Beyond reactivity, substituents impact physical properties like boiling points, , and spectroscopic characteristics, making them essential in designing molecules for pharmaceuticals, materials, and synthetic applications. Common substituents are classified as activating or deactivating based on their influence on reaction rates, with quantitative measures like Hammett sigma constants describing their electronic contributions across diverse reaction types.

Fundamentals

Definition

In , a substituent is defined as an atom or group of atoms that replaces one or more atoms in the parent chain or ring, thereby becoming a part of the molecular structure, or one that attaches to a . This replacement alters the chemical and physical properties of the parent molecule without changing its carbon . Substituents can be simple atoms like (e.g., ) or more complex groups such as alkyl chains. The concept of a substituent applies both to the static architecture of molecules, where it denotes a fixed moiety attached to the parent structure, and to dynamic processes in substitution reactions, where one group displaces another (often a like a ) to form a new bond. In structural terms, substituents are integral to and property prediction, whereas in reactions such as , they represent the incoming or outgoing species that drive reactivity. The degree of substitution at a carbon atom is classified as primary, secondary, or based on the number of other carbon atoms attached to that carbon bearing the substituent. For illustration, unsubstituted (CH_4) has a central carbon bonded to four s; replacing one hydrogen with a substituent X yields CH_3X, a primary substitution where the carbon is attached to zero other carbons. In contrast, a secondary substitution occurs in structures like (CH_3)_2CHX, where the substituted carbon is bonded to two other carbons, and in (CH_3)_3CX, with three carbon attachments. This influences reactivity, with tertiary positions often more reactive in certain mechanisms due to stability. Substituents exert several key effects on molecular properties at a high level: the involves the transmission of electron density through sigma bonds, where electronegative substituents withdraw electrons (+I or -I designation), while the mesomeric () effect operates through pi systems or lone pairs, allowing delocalization that can donate or withdraw electrons depending on the group. Additionally, arise from the physical bulk of substituents, leading to spatial repulsion that impacts conformation, bond angles, and reaction rates without involving electronic changes. These effects collectively modulate acidity, basicity, and reactivity, with general notation like R- representing an alkyl substituent in structural formulas.

Historical Background

The concept of substituents in originated in the mid-19th century amid efforts to rationalize the composition and reactivity of organic compounds, building on the radical theory proposed by Justus Liebig and in 1832, which posited stable atomic groups (radicals) like the benzoyl radical as fundamental units in organic molecules. This theory, influenced by earlier work such as Joseph Louis Gay-Lussac's 1815 identification of the cyanide radical, viewed organic substances as assemblies of unchanging radicals, but it struggled to explain substitution reactions where atoms like were replaced by others. advanced this in the 1830s by demonstrating that could substitute for in hydrocarbons without altering equivalent weights, leading to the substitution theory that emphasized modifiable groups within molecular structures. A pivotal milestone occurred in 1844 when Charles Gerhardt introduced generic notation in his Précis de chimie organique, using the symbol "" to represent radicals or residues, enabling chemists to denote abstract substituting groups in formulas and bridging and ideas. This notation, possibly abbreviating "" (coined by Guyton de Morveau in 1786) or Gerhardt's own "residue" from 1839, facilitated the representation of atomic s in s. The term "substituent" itself emerged around the 1860s in the context of early structural , reflecting the growing recognition of groups replacing atoms in parent chains, as became central to understanding molecular diversity. In 1858, Archibald Scott Couper and independently developed the structural theory, emphasizing carbon's tetravalency and its ability to form chains, which formalized the idea of substituents as atomic groups attached to carbon skeletons in hydrocarbons. Couper's work, published in and English, illustrated substitutions through valence-based diagrams, while Kekulé's contributions highlighted how such groups influenced molecular architecture. This shift from abstract radicals to explicit structural representations marked a foundational in the substituent concept. The modern definition of a substituent as "a group produced by removal of a hydrogen atom from a parent hydride" was codified in IUPAC's 1993 Nomenclature of Organic Chemistry recommendations, which systematized substitutive nomenclature for organic compounds. These were refined in the 2013 IUPAC Blue Book, incorporating updates for complex substituents and preferred names to ensure consistency in an expanding field.

Naming Conventions

IUPAC Nomenclature

In IUPAC , substituents are named systematically as prefixes derived from parent hydrides by removing one or more atoms and adding appropriate es to indicate the valency and bonding type. The primary for a monovalent substituent formed by removal of a single atom is "-yl", as specified in the IUPAC Recommendations 2013 (, P-29.1). For example, the substituent derived from by removing one is named ethyl (CH₃-CH₂-). For divalent substituents where two atoms are removed from the same atom, implying a double bond-like attachment, the "-ylidene" is used (, P-33.3, Table 3.4); ethylidene (=CH-CH₃) illustrates this for a group from . Similarly, the "-ylidyne" denotes a trivalent substituent from removal of three atoms from one atom, such as in methylidyne (≡CH) (, P-33.3, Table 3.4). These es ensure unambiguous description of the attachment mode in substitutive . Rules for assigning locants to substituents prioritize the lowest possible numbers for the point of attachment and any structural features, with locants placed immediately before the part of the name to which they refer (, P-14.3.4). For multiple identical substituents, multiplicative prefixes such as "di-", "tri-", or "tetra-" are employed, and identical locant sets are cited in ascending order (, P-14.5). Complex substituents, which themselves contain branches or additional features, are named by treating the substituent as a hydride and enclosing the full name in parentheses; for instance, the branched group from known systematically as (1-methylethyl) is used when it substitutes a parent chain (, P-29.3.2.1). This approach allows hierarchical naming, where the complex substituent's locants are numbered starting from the attachment point. The integration of substituents into parent chains follows the seniority order of functional groups outlined in the (P-41), which determines the principal chain and the expression of the senior group as a , relegating others to status. Seniority descends from cations and acids to alcohols, amines, and hydrocarbons, ensuring that substituents do not override the principal characteristic group (, P-41, Table 4.4). For unsaturated substituents, the is retained in the name, with indicated atoms if necessary; the ethenyl group (commonly , CH₂=CH-) exemplifies this, derived from ethene (, P-31.1.4.1). Cyclic substituents are named from the corresponding cyclic parent hydride, such as (C₆H₅-) from , which is a retained name for use as a preferred IUPAC (Blue Book, P-29.3.2.2, P-58.2). These guidelines, detailed in the IUPAC 2013, promote consistency across organic compounds.

Common and Traditional Names

In , common and traditional names for substituents provide a concise alternative to systematic IUPAC , facilitating communication in research, , and . These names often originate from historical usage or structural simplicity and are retained by IUPAC for substituents that are frequently encountered. For instance, the group derived from at the central carbon is commonly called isopropyl rather than 1-methylethyl, and the branched group from is known as tert-butyl instead of 1,1-dimethylethyl. Historical common names reflect early discoveries or natural occurrences, such as tolyl for the methyl-substituted (from ), benzyl for the phenylmethyl group (from ), and allyl for the prop-2-en-1-yl group (from allyl compounds in ). These names persist due to their widespread adoption in literature and patents, despite the preference for systematic names in formal contexts. IUPAC guidelines permit the use of retained names for unsubstituted substituents in general , particularly in educational texts and preliminary communications, but recommend systematic names for indexing and official documentation to ensure unambiguity. Retained names are acceptable only for the parent structures without further substitution, and their application is limited to avoid confusion with complex molecules. Frequently used traditional names are categorized below for alkyl, aryl, and halo substituents, highlighting those with broad practical utility:

Alkyl Substituents

  • Methyl (CH₃–)
  • Ethyl (CH₃CH₂–)
  • Isopropyl ((CH₃)₂CH–)
  • tert-Butyl ((CH₃)₃C–)
  • Neopentyl ((CH₃)₃CCH₂–)
  • Phenyl (C₆H₅–)
  • Tolyl (CH₃C₆H₄–, with -, -, or - isomers)
  • Benzyl (C₆H₅CH₂–)
  • Naphthyl (C₁₀H₇–, with 1- or 2- positions)

Halo Substituents

  • Fluoro (F–)
  • Chloro (Cl–)
  • Bromo (Br–)
  • Iodo (I–)
These examples illustrate the preference for brevity in everyday chemical discourse, where retained names enhance readability without sacrificing essential meaning.

Representation

Symbolic Notation

In , substituents are often represented using abstract symbols and abbreviations in equations, formulas, and discussions to denote generic or specific groups without detailing their full structure. This symbolic notation simplifies communication and emphasizes key functional aspects, particularly for alkyl, aryl, or heteroatom-based substituents. A substituent, as a fragment replacing a in a parent molecule, is typically implied through these conventions to focus on reactivity patterns rather than exhaustive structural description. The most fundamental symbol is R-, introduced by French chemist Charles Gerhardt in his 1844 work Précis de chimie organique to represent generic radicals or substituents, such as alkyl (e.g., -CH₃) or aryl groups, in generalized formulas like R-H for alkanes or R-OH for alcohols. Gerhardt chose R likely as an abbreviation for "," a term then used for reactive molecular fragments, allowing chemists to denote unspecified chains without enumeration; this notation became standardized by the mid-19th century through adoption by contemporaries like August Laurent and . Specific substituents derived from common groups are abbreviated as Ph- for phenyl (-C₆H₅) and Me- for methyl (-CH₃), conventions that facilitate concise representation in structural formulas and reaction schemes. For electronegative substituents, particularly (F, Cl, Br, I), the X- is conventionally used to indicate any such atom, as in RX for alkyl halides, highlighting their similar reactivity profiles without specifying the element. This arbitrary yet widespread designation arose from the need to generalize halogen-based reactivity in substitution reactions. In mechanistic contexts involving substituents, Nu- denotes a (electron-pair donor) and E an (electron-pair acceptor), terms coined by Christopher Ingold in to describe roles in electronic theories of , replacing earlier descriptors like "anionoid" and "cationoid." These symbols appear in reaction arrows, such as Nu: attacking an electrophilic carbon bearing a substituent, to illustrate formation or . Skeletal formulas, also known as line-angle or bond-line notation, employ conventions where carbon atoms and their attached hydrogens are implied at line intersections and endpoints, with explicit symbols for non-hydrogen substituents to avoid clutter. For instance, a line represents the carbon chain, and attachments like - or - are shown directly, assuming standard valences; this minimalist approach, rooted in 19th-century structural diagrams, prioritizes substituent positions for clarity in complex molecules.

Structural Depiction

In , substituents are visually represented through various diagrammatic methods to convey their , , and electronic features within a . These depictions range from simplified line drawings to detailed electron-dot structures, facilitating the understanding of spatial and bonding arrangements without relying solely on abstract symbols like R-. Condensed structural formulas provide a compact way to depict substituents by writing the atomic composition in a linear format, often using parentheses for branching, as opposed to expanded structural formulas that explicitly show every bond. For instance, the ethyl substituent is represented as CH₃CH₂- in condensed form, implying the carbon-carbon and attached hydrogens, whereas the expanded form draws out all individual bonds: H₃C-CH₂-. This condensed approach saves space while preserving the sequence of atoms, making it suitable for quick sketches and textual descriptions. For substituents containing heteroatoms, such as or oxygen, Lewis structures are employed to illustrate covalent bonds as lines and lone pairs of electrons as dots, ensuring the distribution is clear. In the hydroxyl substituent (-), for example, the oxygen atom is shown with a to the parent structure and three lone pairs (six electrons total), highlighting its potential for hydrogen bonding or nucleophilicity. This representation is essential for heteroatom-containing groups like -NH₂ or -Cl, where lone pairs influence reactivity and are explicitly depicted to avoid ambiguity. When substituents introduce , wedge-dash notation is used in two-dimensional diagrams to indicate the three-dimensional orientation around tetrahedral centers. Solid wedges represent bonds projecting out of the plane toward the viewer, while dashed lines denote bonds receding into the plane, as seen in chiral alkyl substituents like the (R)-1-methylethyl group. This convention allows for the depiction of stereoisomers without full . In computational and software-based representations, substituents are often encoded using SMILES notation, a text-based system for generating graphical structures. The methyl substituent, for example, is denoted simply as "C," which software interprets as -CH₃ when attached to a parent chain, enabling automated visualization and database storage. This linear notation supports branches and through symbols like "@" for chiral centers.

Examples

Substituents Derived from Methane

Substituents derived from provide foundational examples in , demonstrating how the simplest , CH₄, yields groups by systematic removal or replacement of atoms. These groups vary in valency and bonding, ranging from monovalent radicals like the to polyvalent linkers, and extend to halogenated variants formed via reactions. Such derivatives are crucial for building more complex structures, with reflecting the number of attachment points and bond types. The monovalent methyl group, -CH₃, arises from by excising one , serving as a ubiquitous alkyl substituent in countless compounds. Halogenated variants include chloromethyl (-CH₂Cl), dichloromethyl (-CHCl₂), and trichloromethyl (-CCl₃), produced through successive radical chlorination of , where hydrogens are replaced by chlorines. Disubstituted derivatives feature two attachment sites. The divalent , -CH₂-, equivalent to minus two hydrogens, functions as a bridging unit in chains or rings, with the carbon typically sp³ hybridized. A variant with a is methylidene, =CH₂, used for exocyclic unsaturation. Trisubstituted groups involve three bonds from the central carbon. The trivalent , >CH-, forms by removing three hydrogens from , with the carbon bound to three non-hydrogen atoms and often sp³ hybridized in branched structures. Unsaturated analogs include methanylidene, =CH-, featuring a and a , and methylidyne, ≡CH, with a , both commonly encountered in reactive intermediates or coordination compounds. The tetravalent tetrayl group, C, formed by removing all four hydrogens from , consists of a carbon atom bound to four non-hydrogen atoms via single bonds. It is rare as a simple substituent in conventional organic molecules but appears in structures with carbon centers. The following table summarizes key methane-derived substituents, categorized by remaining hydrogens and bond multiplicity (valency), highlighting representative names and structures:
Remaining HMonovalent (1 attachment)Divalent (2 attachments)Trivalent (3 attachments)Tetravalent (4 attachments)
3-CH₃ (methyl)N/AN/AN/A
2N/A-CH₂- (methanediyl/methylene)
=CH₂ (methylidene)
N/AN/A
1N/AN/A>CH- (methanetriyl/methine)
=CH- (methanylidene)
≡CH (methylidyne)
N/A
0N/AN/AN/AC (methanetetrayl)
Halogenated variants follow similar patterns for monovalent groups, with prefixes like chloro- indicating substitutions (e.g., -CH₂Cl, -CHCl₂, -CCl₃).

Other Common Substituents

Beyond the simple derived from , alkyl substituents commonly include longer or branched chains that modify the properties of parent hydrocarbons or other molecules. The (CH₃CH₂–) is a fundamental alkyl substituent, often seen in compounds like (C₆H₅CH₂CH₃), where it attaches to a ring. Propyl groups appear in straight-chain (n-propyl, CH₃CH₂CH₂–) and branched (isopropyl, (CH₃)₂CH–) forms; for instance, isopropyl is a key substituent in molecules such as isopropylbenzene (). Butyl substituents exhibit greater isomerism, including n-butyl (CH₃CH₂CH₂CH₂–), isobutyl ((CH₃)₂CHCH₂–), sec-butyl (CH₃CH₂CH(CH₃)–), and tert-butyl ((CH₃)₃C–), which are frequently encountered in branched alkanes and synthetic intermediates like . Aryl substituents derive from aromatic systems and introduce ring structures to parent chains. The (C₆H₅–) is the most prevalent, as in (C₆H₅CH₃), where it serves as the core with an alkyl attachment, or in (C₆H₅C₆H₅). Naphthyl (C₁₀H₇–), from , appears in more complex aromatics like naphthyl , contributing to polycyclic structures in dyes and pharmaceuticals. Heteroatom-based substituents incorporate non-carbon elements and are vital in functional group chemistry. The hydroxy group (–OH) is common in phenols like phenol (C₆H₅OH), enhancing polarity. The amino group (–NH₂) features in anilines such as aniline (C₆H₅NH₂), a precursor for dyes. The nitro group (–NO₂) appears in nitrobenzene (C₆H₅NO₂), used in explosive and pharmaceutical synthesis. Carbonyl derivatives like the acetyl group (CH₃CO–) are exemplified in acetophenone (C₆H₅COCH₃), where it imparts ketone functionality.

Properties and Effects

Influence on Molecular Properties

Substituents exert significant influence on the physical and electronic properties of molecules primarily through , , and . The involves the transmission of along sigma bonds due to differences in , where electron-withdrawing groups (-I) like pull from the molecular framework, while electron-donating groups (+I) such as alkyl groups push it toward adjacent atoms. For instance, in haloacids, exhibits the strongest -I effect among , lowering the of fluoroacetic acid to 2.59 compared to acetic acid's 4.76, thereby increasing acidity by stabilizing the conjugate base. Conversely, methyl groups display a +I effect, as seen in the progressive stabilization of carbocations in tertiary > secondary > primary alkyl systems, which enhances at the positive center. The mesomeric effect, also known as the resonance effect, operates through pi-bond conjugation and lone pair delocalization, allowing substituents to donate or withdraw electrons over longer distances. Electron-donating groups with +M character, such as the hydroxyl (-OH) in phenol, increase electron density on the ring via resonance from the oxygen lone pair, raising the pKa of phenol to 10.0 relative to benzene derivatives without such donation. Electron-withdrawing groups like nitro (-NO₂) exhibit -M effects by delocalizing electrons into the substituent, as in 4-nitrophenol (pKa 7.23), where resonance stabilizes the phenolate anion more effectively than in meta-nitrophenol (pKa 8.36), thus enhancing acidity. These effects collectively alter properties like dipole moments and boiling points; for example, nitrobenzene has a higher dipole moment (3.9 D) than benzene (0 D) due to the -M pull of the nitro group. Steric effects arise from the spatial bulk of substituents, causing non-bonded repulsions that influence molecular conformation and overall shape without direct electronic involvement. Larger groups like impose greater steric hindrance than smaller ones like methyl, leading to preferred conformations that minimize strain, such as in the axial avoidance in derivatives where the for tert-butyl is 4.9 kcal/ versus 1.7 kcal/ for methyl. This bulkiness can distort angles and elongate bonds, increasing steric energy and affecting properties like ; for instance, has a lower (169°C) than expected due to reduced molecular packing efficiency compared to less hindered isomers. Quantitative assessment of these substituent influences, particularly on acidity and basicity, is provided by Hammett (σ) constants, which correlate effects in - and -substituted benzoic acids. The σ_m values primarily capture inductive effects, while σ_p values incorporate both inductive and contributions, as defined in the : log(K/K₀) = ρσ, where ρ measures reaction sensitivity. For example, the nitro group has σ_m = 0.71 and σ_p = 0.81, reflecting its strong electron withdrawal that lowers the of p-nitrobenzoic acid to 3.44 from benzoic acid's 4.20, whereas methoxy has σ_m = 0.12 and σ_p = -0.27, increasing to 4.49 by donation. These constants, compiled from extensive measurements, enable prediction of substituent impacts on equilibria and are foundational for understanding property variations in aromatic systems./26:_More_on_Aromatic_Compounds/26.06:_Correlations_of_Structure_with_Reactivity_of_Aromatic_Compounds)

Role in Reactivity

Substituents play a crucial role in directing the of () reactions on derivatives. Electron-donating groups, such as the amino group (-NH₂), activate the ring and direct incoming electrophiles preferentially to the and positions relative to the substituent, as these positions allow for greater stabilization of the positively charged Wheland intermediate through . In contrast, electron-withdrawing groups like the nitro group (-NO₂) deactivate the ring and direct electrophiles to the meta position, where the intermediate experiences less destabilization from the substituent's inductive withdrawal of electrons. In electrophilic additions to s, substituents influence product distribution according to , which states that in the addition of hydrogen halides (HX) to an unsymmetrical , the adds to the carbon with more hydrogen substituents, placing the on the more substituted carbon to form the more stable intermediate. For example, in the reaction of propene (CH₃-CH=CH₂) with HBr, the major product is , as the forms on the secondary carbon. Substituents also govern in elimination reactions, as described by Zaitsev's rule, which predicts that the major product of E1 or E2 eliminations from alkyl halides is the with the most substituted , due to its greater thermodynamic stability from and inductive effects. In the E2 elimination of with a , the predominant product is (E)- and (Z)-2-butene rather than , as the trisubstituted is more stable. In nucleophilic substitution reactions, bulky substituents impose steric hindrance that significantly affects the reaction pathway. Primary alkyl halides with minimal steric bulk undergo SN2 reactions readily via backside attack, but as substituents increase in size—such as in secondary or tertiary systems like isopropyl or tert-butyl halides—the SN2 rate decreases due to crowding at the transition state, often favoring the SN1 mechanism where a carbocation intermediate forms without direct nucleophilic approach./11:_Reactions_of_Alkyl_Halides-_Nucleophilic_Substitutions_and_Eliminations/11.03:_Characteristics_of_the_SN2_Reaction) For instance, tert-butyl bromide reacts via SN1 in polar protic solvents, as the bulky methyl groups hinder SN2 but stabilize the tertiary carbocation.

Statistical Analysis

Distribution in Organic Compounds

An analysis of over 3 million molecules from chemical databases identified 849,574 unique substituents, suggesting an estimated total of approximately 3.1 million known substituents when extrapolated. This dataset highlights the vast but skewed distribution of substituents in documented , with a small appearing frequently across compounds. The most prevalent substituents by occurrence include methyl (the most common), followed by phenyl, , methoxy, and hydroxyl groups. These top substituents account for a significant portion of attachments in bioactive and synthetic molecules, reflecting their utility in building diverse structures. In a complementary analysis of 314,525 compounds from literature, methyl, phenyl, and hydroxyl ranked among the top four R-groups by frequency, with and methoxy also prominent within the top 20. Substituents typically contain up to 12 non-hydrogen atoms, with the majority being smaller fragments that facilitate modular synthesis. Elemental composition is dominated by carbon, hydrogen, oxygen, and nitrogen, alongside halogens; for instance, in a dataset of 143,180 substituents from approximately 700,000 bioactive ChEMBL molecules, carbon comprised 68.7%, nitrogen 21.6%, oxygen 6.1%, and sulfur 3.2% of atoms, underscoring the prevalence of C-, N-, O-, and halogen-based groups. The observed distribution is influenced by factors such as synthetic accessibility, where simpler, commercially available building blocks like alkyl and aryl halides predominate due to ease of incorporation via standard reactions. This bias toward accessible substituents shapes the chemical space explored in databases, limiting representation of more complex or rare groups.

Applications in Data Analysis

Substituents play a central role in quantitative structure-activity relationship (QSAR) models, where they are used to predict molecular properties and biological activities by quantifying their electronic, steric, hydrophobic, and hydrogen-bonding effects. In cheminformatics analyses of large datasets, such as those derived from millions of compounds, substituents are characterized through calculated descriptors that form the basis for QSAR regressions, enabling the identification of patterns that correlate structural variations with outcomes like or receptor . For instance, the Free-Wilson approach treats substituents additively, assigning independent contributions to activity based on their occurrence in datasets, which has been foundational for early QSAR applications in . Substituent libraries are integral to and , where they serve as modular building blocks attached to core scaffolds to generate diverse chemical spaces for . In , these libraries are enumerated by varying substituents at defined positions (e.g., R-groups), allowing rapid exploration of structure-activity relationships without physical , as demonstrated in scaffold-based designs yielding libraries of thousands of potential inhibitors. Such approaches facilitate pharmacological screening by prioritizing candidates with favorable predicted properties, reducing experimental costs in early-stage hit identification. Computational tools like RDKit and enable the analysis of substituent patterns across large molecular datasets, supporting tasks from fragmentation to QSAR descriptor generation. RDKit's fragmentation modules, such as RECAP and , decompose molecules into substituent-like fragments marked by dummy atoms, allowing cataloging and fingerprinting for similarity searches and in cheminformatics workflows. Similarly, 's cheminformatics suite provides R-group analysis, matched molecular pair profiling, and over 400 molecular descriptors to evaluate substituent effects on drug-likeness and diversity in virtual libraries. In , substituent data informs adherence to guidelines like , which assesses oral by limiting molecular weight, , donors, and acceptors—properties often modulated by substituent complexity to avoid excessive or size. This rule guides the selection of substituent combinations in library design, ensuring candidates maintain drug-like profiles while optimizing potency, as seen in analyses where simpler substituents enhance permeability without compromising activity.

References

  1. [1]
    Illustrated Glossary of Organic Chemistry - Substituent
    Substituent: An atom or group other than hydrogen on a molecule. (The atom or group has substituted for the missing hydrogen.) Cyclohexane has no substituents.
  2. [2]
    Organic Nomenclature - MSU chemistry
    If two different substituents are present on the ring, they are listed in alphabetical order, and the first cited substituent is assigned to carbon #1. The ...
  3. [3]
    How to name organic compounds using the IUPAC rules
    Other groups which are attached to the parent chain are called substituents. Alkanes - saturated hydrocarbons. The names of the straight chain saturated ...
  4. [4]
    Substitution Reactions of Benzene and Other Aromatic Compounds
    The first is the inductive effect of the substituent. Most elements other than metals and carbon have a significantly greater electronegativity than hydrogen.
  5. [5]
    Common Definitions and Terms in Organic Chemistry
    Sep 29, 2025 · [The substituent either induces charge towards or away from itself with the formation of a dipole.]
  6. [6]
    Unit 4: Free Energy Relationships
    When the reaction site is actually in the aromatic ring, the substituent effects, as reflected by the rho value, are even larger than in the standard reaction ...
  7. [7]
    What Are Substituents in Organic Chemistry?
    Substituents (methyl, ethyl) in organic chemistry are atoms or groups of atoms that replace a hydrogen atom on the parent chain of a molecule.
  8. [8]
    Substitution Reaction | Research Starters - EBSCO
    Substitution reactions are easily recognized by the replacement of one substituent—an atom or functional group—in a reactant compound with a different group at ...
  9. [9]
    27.2: Introduction to Substitution Reactions - Chemistry LibreTexts
    Jul 12, 2023 · We are going to be discussing some general aspects of nucleophilic substitution reactions, and in doing so it will simplify things greatly.Associative nucleophilic... · Dissociative nucleophilic... · Protonation states and...
  10. [10]
    Chapter 7 - Alkanes and Halogenated Hydrocarbons - Chemistry
    For the methane – chlorine reaction this is the formation of the chloromethane (CH3Cl). Termination Reaction. In summary, radical reactions occur in three ...
  11. [11]
    [PDF] Lecture Notes Chem 51A S. King Chapter 4 Alkanes I. Alkanes
    Carbons in alkanes or other organic compounds can be classified as primary (1°), secondary (2°), tertiary (3°), and quaternary: • a primary carbon is bound to ...
  12. [12]
    Illustrated Glossary of Organic Chemistry - Inductive effect
    Inductive effect: The effect on electron density in one portion of a molecule due to electron-withdrawing or electron-donating groups elsewhere in the molecule.
  13. [13]
    Three-Step Method to Teach Inductive Effect - ACS Publications
    13 jun 2023 · The inductive effect is one of the very important concepts of electronic effects in organic chemistry. In traditional teaching methods, ...
  14. [14]
    Ring Conformations - MSU chemistry
    Due to steric hindrance in the axial location, substituent groups prefer to be equatorial and that chair conformer predominates in the equilibrium.<|control11|><|separator|>
  15. [15]
    [PDF] Organic Chemistry in the Nineteenth Century : theory of radicals to ...
    Mar 30, 2020 · The radical theory was forced to be modified when new experimental results showed halogens replacing hydrogen, within organic compounds, ...
  16. [16]
    Before Radicals Were Free – the Radical Particulier of de Morveau
    Apr 20, 2020 · (a) Joseph Louis Gay-Lussac (1778 – 1850) showed that CN was a compound radical and opened the doors to the Radical Theory of organic chemistry.
  17. [17]
    None
    ### Summary of Historical Development of Notation for Substituents in Organic Chemistry
  18. [18]
    History of Chemical Notations from Alchemy to Psycho‐Chemistry
    Aug 12, 2022 · In 1874, Dutch chemist van′t Hoff and independently French chemist Le Bel concurrently proposed that the carbon atom bonded to four substituent ...
  19. [19]
    Foundations of the structural theory - ACS Publications
    The Kekule-Couper Centennial Lecture celebrates the discovery of the tetravalence of carbon and the self-linking of carbon atoms, reached independently by ...
  20. [20]
    Couper - New Chemical Theory - Le Moyne
    As to the first: this theory indeed brings every chemical combinate under a certain comparative point of view with every other. Herein apparently is its merit.
  21. [21]
    REVISED NOMENCLATURE FOR RADICALS, IONS ... - iupac
    The names for the substituent groups HS-,. HSe-, and HTe- as prefixes are "sulfanyl-", "selanyl-I!, and "tellanyl-" in these recommendations, rather than " ...<|separator|>
  22. [22]
    None
    Nothing is retrieved...<|separator|>
  23. [23]
    [PDF] Brief Guide to the Nomenclature of Organic Chemistry - IUPAC
    Locants indicate the position of substituents or other structural features. They are generally placed before the part of the name that indicates the ...
  24. [24]
    Nucleophilicity Prediction via Multivariate Linear Regression Analysis
    Feb 3, 2021 · The term nucleophile (and electrophile) was first introduced by Ingold in 1933. (1,2) Since then, ...
  25. [25]
    Condensed Formulas: Deciphering What the Brackets Mean
    Jan 8, 2025 · Brackets in condensed formulas reduce work, remove ambiguity, show double bonds (like C=O), and show branching (like CH(CH3)CH2CH3).
  26. [26]
    Lewis Structures in Organic Chemistry
    Here is a simple view on this important topic. In Lewis structures, we show each covalent bond as two dots, which represent a pair of electrons.
  27. [27]
    Ch 3 - Understanding Diagrams - University of Calgary
    Bonds are represented by lines and lone pairs as pairs of dots on the corresponding atom. In a complete Lewis structure, the covalent bonds are also ...
  28. [28]
    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.
  29. [29]
    Wedge And Dash Convention For Tetrahedral Carbon
    Mar 12, 2013 · The way it's done in chemistry is with heavy “wedges” to depict bonds in the foreground (pointing out of the page) and light “dashes” to depict bonds in the ...
  30. [30]
    3. SMILES - A Simplified Chemical Language
    SMILES, Name, SMILES, Name. CC, ethane, [OH3+], hydronium ion. O=C=O, carbon dioxide, [2H]O[2H], deuterium oxide. C#N, hydrogen cyanide, [235U], uranium-235.
  31. [31]
    12.5 IUPAC Nomenclature - Organic Chemistry - Lumen Learning
    For example, the CH 3 group derived from methane (CH 4) results from subtracting one hydrogen atom and is called a methyl group.
  32. [32]
    Illustrated Glossary of Organic Chemistry - Methylene group
    Methylene group: A portion of a molecular structure equivalent to methane minus two hydrogen atoms: -CH2-. The methylene carbon can be sp or sp.
  33. [33]
    Bring the power of IUPAC naming to your desktop! - ACD/Labs
    The name methylene is used instead of methanediyl; the name methylidene is used for the group when doubly bonded to another part of the structure.
  34. [34]
    Illustrated Glossary of Organic Chemistry - Methylene group
    Methine group: A portion of molecular structure equivalent to methane minus three hydrogen atoms: CH. The methine carbon can be sp, sp, or sp.
  35. [35]
    Trichloromethyl Group - an overview | ScienceDirect Topics
    The trichloromethyl group is defined as a versatile substituent that can undergo partial or complete reduction, particularly in the context of β-sultam ...
  36. [36]
    Methylidyne | chemical compound | Britannica
    is trivalent since it is combined with three substituents instead of four, and its unshared electron is represented by a dot. Free radicals of the ...Missing: substituent | Show results with:substituent
  37. [37]
    [PDF] Short Summary of IUPAC Nomenclature of Organic Compounds
    Now that the functional groups and substituents from Groups A, B, and C have been described, a modified set of steps for naming organic compounds can be applied ...
  38. [38]
    15.1: Naming Aromatic Compounds - Chemistry LibreTexts
    Mar 17, 2024 · draw the Kekulé, condensed or shorthand structure of an aromatic compound in which the phenyl group is regarded as a substituent, given its ...
  39. [39]
    What is the difference between aryl and alkyl phenyl? - askIITians
    Mar 8, 2025 · Common examples of aryl groups include phenyl (-C6H5), naphthyl (-C10H7), and biphenyl. Alkyl phenyl: Alkyl phenyl refers to a molecular ...
  40. [40]
    Common Functional Groups in Organic Chemistry - ThoughtCo
    Oct 1, 2019 · Examples of functional groups include the hydroxyl group, ketone group, amine group, and ether group. Hydroxyl Functional Group.Hydroxyl Functional Group · Aldehyde Functional Group · Phenyl Functional Group
  41. [41]
    Table of Functional Group Priorities for Nomenclature
    Feb 14, 2011 · If carbon-carbon multiple bonds are present in the molecule, they are considered as substituents with a priority (or “seniority”, according to ...
  42. [42]
    Inductive and Resonance (Mesomeric) Effects - Chemistry Steps
    Both inductive and mesomeric (resonance) effects tell us whether one part of the molecule withdraws or donates electron density to another part of the molecule.
  43. [43]
    steric effect (S05997) - IUPAC Gold Book
    Steric effects arise from contributions ascribed to strain as the sum of (1) non-bonded repulsions, (2) bond angle strain and (3) bond stretches or compressions ...
  44. [44]
    A survey of Hammett substituent constants and resonance and field ...
    A survey of Hammett substituent constants and resonance and field parameters | Chemical Reviews.
  45. [45]
  46. [46]
    Steric Hindrance in SN2 and SN1 Reactions - Chemistry Steps
    However, the bulky alkyl groups that make the substrate sterically hindered also stabilize the intermediate carbocation, which in turn accelerates the reaction.
  47. [47]
    Cheminformatics Analysis of Organic Substituents: Identification of ...
    Steric substituent properties were characterized simply by topological size (number of non-hydrogen atoms) and maximum topological length. Substituent ...
  48. [48]
    R-group replacement database for medicinal chemistry - PMC
    Jun 30, 2021 · We have generated a searchable database of frequent R-groups and their preferred replacements. To our knowledge, the database represents the first R-group ...
  49. [49]
    Page Not Found
    **Summary:**
  50. [50]
    Virtual Combinatorial Chemistry and Pharmacological Screening
    Jan 30, 2022 · The design of virtual combinatorial libraries (VCLs) is a critical part in the early phases of the drug discovery process as these libraries are ...
  51. [51]
    Virtual combinatorial libraries: Dynamic generation of molecular and ...
    The virtual combinatorial library approach may represent a powerful methodology for the discovery of substrates, inhibitors, receptors, catalysts, and carriers ...
  52. [52]
    Getting Started with the RDKit in Python
    Molecular Fragments¶. The RDKit contains a collection of tools for fragmenting molecules and working with those fragments. Fragments are defined to be made up ...Reading, Drawing, And... · Working With Molecules · Drawing Molecules
  53. [53]
    Molecular Operating Environment (MOE) | MOEsaic | PSILO
    Integrated computer-aided molecular design platform for small molecule and biological therapeutics. Common platform for Chemists, Biologists and ...Download Request · CCG CCG SVL Exchange · (CCG) | Support · Research
  54. [54]
    Experimental and computational approaches to estimate solubility ...
    In the discovery setting 'the rule of 5' predicts that poor absorption or permeation is more likely when there are more than 5 Hbond donors, 10 Hbond ...