Fact-checked by Grok 2 weeks ago

Side chain

In , a side chain is a shorter chain or group of atoms attached to a principal chain or to a ring in a molecule. Side chains, often synonymous with substituents, are fundamental to molecular under IUPAC rules, where they are identified and numbered relative to the longest continuous carbon chain to ensure systematic naming. Their presence influences key molecular behaviors, such as reactivity and steric hindrance during chemical reactions. In biochemistry, side chains—known as R groups in —represent the variable portion beyond the common α-amino and carboxyl groups, conferring unique chemical properties like , charge, and hydrophobicity to each of the 20 standard . These properties dictate , stability, enzymatic activity, and intermolecular interactions, with hydrophobic side chains often driving core formation in globular proteins while charged or polar ones facilitate and binding. For instance, the nonpolar side chain of (a ) contrasts with the acidic side chain of (a carboxyl group), enabling diverse roles from to . In , side chains grafted onto a main backbone modify material properties, including crystallinity, , , and strength, which are critical for applications in plastics, coatings, and electronics. Flexible or bulky side chains can enhance and processability in conjugated polymers used for , while their affects aggregation and charge transport.

Fundamentals

Definition

In , a side chain is a chemical group attached to the parent structure or main of a , often represented by the generic R to denote its variable nature. This distinguishes it from the core backbone, which forms the primary structural framework. Side chains are substituents that branch off from the principal , contributing to the overall architecture without being integral to the longest continuous connectivity path. The parent chain, or backbone, is defined as the longest continuous carbon chain in the molecule, serving as the reference for nomenclature and structural analysis. In contrast, side chains are any atomic groups appended to this backbone, such as alkyl, aryl, or functional moieties. For instance, in a linear alkane like pentane (CH₃-CH₂-CH₂-CH₂-CH₃), replacing a hydrogen on the central carbon with a methyl group (-CH₃) creates 3-methylpentane, where the -CH₃ acts as a side chain modifying the core scaffold. Similarly, a propyl group (-CH₂CH₂CH₃) attached to a benzene ring yields propylbenzene, illustrating how side chains can introduce aliphatic character to the aromatic parent structure. These appendages are non-essential to the fundamental chain length but alter the molecule's identity and behavior. Side chains exhibit diverse compositions, including saturated alkyl groups like ethyl (-CH₂CH₃), unsaturated alkenyl groups, aryl systems such as phenyl, or those incorporating heteroatoms and functional groups like hydroxyl (-OH) or amino (-NH₂). They profoundly influence key molecular : alkyl side chains enhance nonpolar solubility by increasing hydrophobic surface area, while polar functional side chains improve water through hydrogen bonding. Sterically bulky side chains, such as tert-butyl (-C(CH₃)₃), induce steric hindrance that can reduce reactivity in reactions requiring close approach, like nucleophilic substitutions, by shielding reactive sites on the backbone. Aryl side chains contribute to conjugation effects, stabilizing delocalization and affecting electronic . These modifications occur without disrupting the primary covalent linkages of the backbone, allowing precise tuning of molecular characteristics in .

Conventions and Nomenclature

In , side chains are commonly represented using notation systems that simplify structural formulas. The R-group serves as a placeholder for a generic alkyl or aryl attached to a or backbone, allowing chemists to focus on reactive sites without specifying the entire chain. This convention is widely adopted in depictions of molecules like alcohols (R-OH) or alkyl halides (R-X). IUPAC recommendations for naming s emphasize systematic approaches but retain certain prefixes for branched alkyl groups, such as iso- for structures like isopropyl (a chain with a methyl branch at the second carbon) and neo- for neopentyl (a tert-butyl-like extension), though these are considered retained names rather than preferred for complex . Diagrammatic representations follow established conventions in skeletal formulas, where bonds are depicted as lines connecting implied carbon atoms at vertices or ends, and side chains appear as branches diverging from the main chain. Carbon and hydrogen atoms are typically omitted for clarity, with only heteroatoms or functional groups explicitly labeled. The main chain is prioritized by the longest continuous carbon chain rule, selecting the longest sequence of carbons—even if it includes fewer branches—to serve as the parent structure. In cases of equal length, the chain with the greater number of substituents or multiple bonds is chosen. Common abbreviations streamline discussions of side chains in reaction mechanisms and synthetic contexts. In , R denotes a general alkyl or , while X represents a (e.g., Cl, Br), and Nu stands for a in reactions. These symbols facilitate concise notation without altering the underlying structural rules. Distinctions between side chains and the main chain rely on selection criteria that ensure systematic naming. Side chains are identified as any groups shorter than or branching from the parent chain, particularly those with lower-priority s. The parent chain is selected first by maximum length, then by inclusion of the highest-priority (e.g., carboxylic acids over alcohols), ensuring the principal characteristic group defines the . If multiple options exist, priority favors the chain with the most skeletal atoms or the lowest for attachments.

Historical Development

Early Concepts in Organic Chemistry

The concept of side chains emerged in the mid-19th century as part of the foundational shift toward structural , where chemists began representing molecules as assemblies of atoms connected by specific bonds rather than mere aggregates of elements. played a pivotal role in this development, proposing in that carbon atoms exhibit tetravalency and can link together to form extended chains or skeletons, to which other atoms or groups—effectively side chains—could be attached, enabling the representation of complex molecular architectures such as those in derivatives and aliphatic compounds. This structural approach built on earlier ideas from predecessors like Charles Gerhardt, who in the 1840s introduced symbolic notations for organic radicals (using "" to denote variable alkyl groups), laying groundwork for distinguishing main frameworks from appended substituents. A key milestone in recognizing side chains came through investigations into isomerism during the 1850s and 1860s, as chemists observed that molecules with identical elemental compositions could differ in properties due to variations in chain branching. For instance, in studies of alcohols, normal butanol (n-butanol, with a straight four-carbon chain) was distinguished from (a branched with a three-carbon main chain and a methyl side chain), first isolated via by in 1852, highlighting how side chain rearrangements produced distinct compounds like these C4H9OH . Aleksandr Butlerov advanced this understanding in 1861 by formalizing the theory of , asserting that a molecule's properties arise from the specific spatial arrangement of atoms, including branched side chains, which he used to predict and explain such isomeric forms in alcohols and carboxylic acids. These early ideas profoundly influenced valence theory, reinforcing carbon's tetravalent nature as the basis for molecular diversity, where side chains allowed for myriad substitutions and branchings beyond linear sequences. Kekulé's chain model, elaborated in his textbook, demonstrated how attaching side groups to a carbon could account for the vast array of isomers observed experimentally, shifting chemistry from empirical formulas to predictive structural representations. Butlerov's contributions further emphasized that chemical reactivity and isomerism stem from these structural variations, providing a theoretical framework that explained why seemingly similar compounds, such as straight- versus branched-chain acids, exhibited different behaviors. Early applications of side chain concepts appeared in the analysis of products, particularly fats, where long alkyl chains were identified as key structural elements. Michel Eugène Chevreul's pioneering work from 1811 to 1823 decomposed soaps and fats into and fatty acids, classifying the latter—such as stearic (C18 straight ) and oleic (unsaturated )—based on their side chains attached to the carboxyl group, which influenced early taxonomic schemes for and underscored chains' role in molecular classification. This recognition of chain variations in substances foreshadowed broader uses of side chain theory in understanding diversity.

Advancements in Biochemistry

In the early , pioneered the integration of side chain concepts into chemistry, establishing the foundational understanding of as chains of with variable R-groups. By 1901, Fischer synthesized the first , glycylglycine, and coined the term "" in 1902, recognizing that the diverse side chains of influenced and properties. His work extended to synthesizing longer polypeptides, such as an octadecapeptide using and , where protecting groups like the ethoxycarbonyl were employed to manage reactive side chains during coupling reactions. This period also highlighted early challenges in , as the variability and reactivity of side chains complicated efforts to determine the order of in natural proteins, a problem that persisted into the 1930s until later methodological advances. Mid-20th-century advancements shifted focus toward the structural roles of side chains in proteins, notably through Pauling's 1951 proposals for secondary structures. Pauling, along with Corey and Herman Branson, introduced the α-helix model, a right-handed with 3.7 residues per turn stabilized by bonds, where side chain (R-group) properties were treated as stereochemically equivalent but critical for overall helix stability and solvent interactions. They also described the β-sheet, a pleated structure with antiparallel or parallel chains, in which side chain interactions influenced the rise per residue (3.3 Å) and sheet conformation, extending beyond mere substituents to dictate folding patterns. Concurrently, the discovery of dynamic side chain modifications, such as , revealed their regulatory functions in enzymes; in 1956, Edwin G. Krebs and demonstrated that of serine residues in converted it between active and inactive forms, marking a pivotal recognition of side chains as switches in metabolic control. By the 1970s, advancements provided atomic-level insights into side chain interactions, transforming their perception from passive organic appendages to essential functional elements in . Improved techniques, including early sources, enabled higher-resolution structures of proteins like rubredoxin (1970), illuminating how side chains in active sites formed hydrogen bonds, ionic interactions, and hydrophobic clusters critical for and substrate . This era built on prior organic foundations, evolving the view of side chains—initially seen as inert R-groups in syntheses—into dynamic contributors to protein specificity, folding, and signaling, as evidenced by phosphorylation's role in activation and structural studies revealing their involvement in .

Applications in Organic Chemistry

Structural Role

Side chains play a crucial role in determining the of compounds by introducing steric hindrance, which arises from the spatial repulsion between non-bonded atoms or groups. Bulky side chains, such as the tert-butyl group (-C(CH₃)₃), can restrict bond rotations and favor specific conformations, thereby influencing the overall shape and reactivity of the molecule. For instance, in substituted alkanes, the presence of a tert-butyl group often leads to gauche or anti conformations to minimize steric strain, altering the molecule's three-dimensional architecture. Side chains also modulate key physical properties of organic molecules, including , points, and . Polar side chains like hydroxyl (-OH) increase molecular and enhance hydrogen bonding, raising points and improving in polar solvents such as , whereas nonpolar alkyl groups like methyl (-CH₃) reduce , lower points relative to chain length, and favor in nonpolar solvents. In branched hydrocarbons, increasing the degree of branching with alkyl side chains decreases surface area and van der Waals interactions, resulting in lower points compared to linear isomers, while simultaneously affecting melting points by packing efficiency. In various compound classes, side chains exemplify these structural effects; for example, in alkanes like , the branched methyl side chain alters compared to n-butane; in alkenes such as isobutene, it influences double-bond ; and in aromatics like (isopropylbenzene), it modifies ring substitution patterns. A notable case is the isopropyl side chain (-CH(CH₃)₂), which stabilizes adjacent through from its alpha hydrogens, enhancing stability in secondary carbocation intermediates during reactions. Furthermore, side chains contribute to chirality by creating asymmetric centers when a carbon atom bears four different substituents, including the side chain. For instance, attaching a methyl side chain to a carbon already bound to , hydroxyl, and carboxyl groups in a like results in a chiral center, leading to enantiomers with distinct optical activities. This asymmetry arises solely from the spatial arrangement imposed by the unique side chain, without altering the core connectivity.

Synthetic Importance

Side chains play a pivotal role in by enabling the controlled introduction of functional groups that influence reactivity and product selectivity. One common functionalization technique involves using Grignard reagents, which are organomagnesium halides prepared from alkyl halides and magnesium, allowing the addition of alkyl side chains to carbonyl compounds such as aldehydes, ketones, or esters to form alcohols or carboxylic acids after . These are particularly valuable for extending carbon chains in complex molecules due to their high nucleophilicity and compatibility with a wide range of electrophiles. For aryl side chains, the Suzuki-Miyaura provides an efficient method, coupling aryl or vinyl boronic acids with aryl or vinyl halides in the presence of a catalyst and base to form biaryl or styryl linkages under mild aqueous conditions. This reaction's tolerance for functional groups makes it ideal for late-stage attachment of aryl side chains in polyfunctionalized targets. Activating side chains, such as alkyl groups like methyl (-CH₃), exert directing effects in (EAS) reactions by donating through and inductive effects, thereby accelerating the reaction rate and favoring - and - positions relative to the . For instance, undergoes or predominantly at the and positions due to the methyl group's ortho-para directing influence, which stabilizes the intermediate via involvement of the side chain. This is crucial for synthesizing substituted aromatics with precise side chain orientations, minimizing the need for separation of regioisomers. In , side chain modifications often employ protecting groups to mask reactive functionalities during multi-step sequences, ensuring and preventing unwanted side reactions. Common protecting groups for or side chains include silyl ethers (e.g., ) or carbamates (e.g., Cbz), which are orthogonally removable under mild conditions like fluoride treatment or . Chain elongation strategies, such as iterative or , further utilize side chains as handles for extending molecular frameworks, particularly in pharmaceutical intermediates where precise carbon chain length dictates . These approaches allow modular , reducing synthetic steps and improving overall yields in complex . A representative example is the of ibuprofen, a , where the isobutyl side chain on the aromatic ring is introduced early via Friedel-Crafts acylation of isobutylbenzene, followed by and steps to build the moiety. This side chain is critical for ibuprofen's biological activity, as modifications to the isobutyl group can enhance potency by altering steric interactions with the . However, synthetic challenges arise in reactions involving beta-functionalized alkyl side chains, where E2 elimination can compete with substitution, leading to loss of the side chain and formation of alkenes under basic conditions due to anti-periplanar alignment of the and beta-hydrogen. Controlling such eliminations often requires bulky bases or low temperatures to favor desired pathways.

Applications in Biochemistry

Amino Acid Side Chains

In biochemistry, the side chains of the 20 standard are the variable substituent groups attached to the α-carbon atom of each , distinguishing them from one another and imparting unique chemical properties. These side chains, also known as R-groups, range from a simple in to more complex structures incorporating heteroatoms like oxygen, , and . The classification of side chains is primarily based on their and charge at physiological (around 7.4), which influences their interactions in aqueous environments. Non-polar side chains are hydrophobic and typically consist of moieties, while polar and charged side chains enable , ionic interactions, and in . Non-polar side chains include those of (R = H), (R = -CH₃), (R = -CH(CH₃)₂), (R = -CH₂CH(CH₃)₂), (R = -CH(CH₃)CH₂CH₃), (R = a ring fused to the α-amino group), (R = -CH₂C₆H₅, benzyl with an aromatic ring), (R = -CH₂CH₂SCH₃), and (R = -CH₂-indole). These groups are generally small to medium in size, with (e.g., Kyte-Doolittle values) ranging from -1.6 for to +4.5 for , promoting burial in protein interiors to avoid water. Proline's cyclic structure restricts backbone flexibility, while aromatic rings in and confer π-π stacking capabilities. Polar uncharged side chains, found in serine (R = -CH₂OH), threonine (R = -CH(OH)CH₃), cysteine (R = -CH₂SH), tyrosine (R = -CH₂C₆H₄OH, phenolic), asparagine (R = -CH₂CONH₂), and glutamine (R = -CH₂CH₂CONH₂), feature hydroxyl, thiol, or amide groups that can form bonds. These side chains have moderate sizes and hydrophobicity values around -3.5 to -0.7, with the hydroxyl and groups exhibiting nucleophilic reactivity. Cysteine's is notable for its low (≈8.3), allowing reversible bond formation, while tyrosine's hydroxyl (pKa ≈10.1) can be phosphorylated. Acidic side chains are present in (R = -CH₂COOH) and (R = -CH₂CH₂COOH), both containing groups with pKa values of ≈3.9 and ≈4.3, respectively. These medium-sized chains (hydrophobicity ≈ -3.5) are negatively charged at physiological pH, contributing to electrostatic repulsion or attraction. Basic side chains include those of (R = -(CH₂)₄NH₂, pKa ≈10.5), (R = -(CH₂)₃NHC(NH)NH₂, guanidino group with pKa ≈12.5), and (R = -CH₂-imidazole, pKa ≈6.0). These longer chains have hydrophobicity values ranging from -4.5 () to -3.2 () and are positively charged, with 's imidazole ring uniquely able to toggle near neutral pH, enabling pH-sensitive . The chemical reactivity of side chains is largely governed by ionizable groups, which exhibit pH-dependent states according to the Henderson-Hasselbalch . For instance, acidic side chains deprotonate below their to form anions (COO⁻), while basic side chains protonate above their to form cations (e.g., NH₃⁺ for ). This behavior affects , , and reactivity; non-ionizable non-polar chains remain neutral across ranges, whereas polar and charged ones participate in acid-base equilibria that can alter protein conformation or activity. Sulfur-containing and side chains also show sensitivity, with cysteine's prone to oxidation.
Amino AcidAbbreviationSide Chain (R)ClassificationKey Properties (Hydrophobicity, pKa if applicable)
Gly (G)-HNon-polarSmallest; hydrophobicity -0.4; no ionizable group
(A)-CH₃Non-polarSmall; hydrophobicity 1.8; non-ionizable
Val (V)-CH(CH₃)₂Non-polarBranched; hydrophobicity 4.2; non-ionizable
Leu (L)-CH₂CH(CH₃)₂Non-polarHydrophobic; hydrophobicity 3.8; non-ionizable
Ile (I)-CH(CH₃)CH₂CH₃Non-polarBranched; hydrophobicity 4.5; non-ionizable
Pro (P)Cyclic ()Non-polarRestricts flexibility; hydrophobicity -1.6; non-ionizable
Phe (F)-CH₂C₆H₅Non-polarAromatic; hydrophobicity 2.8; non-ionizable
Trp (W)-CH₂-indoleNon-polarBulky aromatic; hydrophobicity -0.9; non-ionizable
Met (M)-CH₂CH₂SCH₃Non-polarSulfur-containing; hydrophobicity 1.9; non-ionizable
SerineSer (S)-CH₂OHPolar unchargedHydroxyl; hydrophobicity -0.8; pKa (OH) ~13
Thr (T)-CH(OH)CH₃Polar unchargedHydroxyl; hydrophobicity -0.7; pKa (OH) ~13
Cys (C)-CH₂SHPolar uncharged; hydrophobicity 2.5; pKa (SH) 8.3
Tyr (Y)-CH₂C₆H₄OHPolar uncharged; hydrophobicity -1.3; pKa (OH) 10.1
Asn (N)-CH₂CONH₂Polar uncharged; hydrophobicity -3.5; non-ionizable
Gln (Q)-CH₂CH₂CONH₂Polar uncharged; hydrophobicity -3.5; non-ionizable
Asp (D)-CH₂COOHAcidicCarboxyl; hydrophobicity -3.5; pKa 3.9
Glu (E)-CH₂CH₂COOHAcidicCarboxyl; hydrophobicity -3.5; pKa 4.3
Lys (K)-(CH₂)₄NH₂BasicAmino; hydrophobicity -3.9; pKa (NH₃⁺) 10.5
Arg (R)-(CH₂)₃NHC(NH)NH₂BasicGuanidino; hydrophobicity -4.5; pKa 12.5
His (H)-CH₂-Basic; hydrophobicity -3.2; pKa 6.0
This table summarizes the structures and properties, with hydrophobicity values derived from the Kyte-Doolittle scale (positive indicating hydrophobicity) and for ionizable groups at 25°C.

Protein Function and Interactions

side chains play a pivotal in protein function by mediating non-covalent interactions that dictate folding, stability, and interactions with other molecules. Through hydrophobic packing, polar bonding, and electrostatic forces, these side chains enable proteins to adopt specific three-dimensional structures essential for . For instance, nonpolar side chains like that of cluster in the protein interior, excluding and stabilizing the core via the . Hydrophobic interactions, exemplified by side chains in protein cores, drive the burial of nonpolar residues away from aqueous environments, promoting compact folding and enhancing thermodynamic stability. Complementary packing of these side chains minimizes voids and optimizes van der Waals contacts, contributing up to several kcal/mol to folding . On protein surfaces, polar side chains such as form hydrogen bonds with or other residues, facilitating and enabling specific recognition events. These bonds, often involving the group of , can stabilize surface loops and contribute to overall protein rigidity even in flexible regions. Electrostatic interactions, including salt bridges between positively charged and negatively charged glutamate side chains, further reinforce by providing directional stabilization, with each bridge adding approximately 3-5 kcal/mol to stability in folded states. In enzymatic , side chains in active sites directly participate in binding and reaction chemistry; the imidazole ring of , for example, facilitates proton transfer by acting as a general acid-base catalyst, as seen in where His64 shuttles protons during CO2 hydration, accelerating the reaction by over 10^6-fold. Post-translational modifications on side chains, such as of serine and hydroxyl groups, introduce moieties that modulate protein localization, stability, and interactions, often competing with to regulate signaling pathways. These modifications can alter side chain polarity, enhancing resistance to or enabling cell-surface recognition. Mutations in side chains can disrupt function, as illustrated by sickle cell anemia, where substitution of for at position 6 of the β-globin chain in creates a hydrophobic patch that promotes deoxyhemoglobin polymerization under low oxygen conditions, leading to sickling and vaso-occlusive crises. Enzyme specificity also relies on side chain complementarity, where precise steric and electrostatic matches between residues and substrates ensure selective ; in serine proteases, for instance, the oxyanion hole formed by backbone and side chain hydrogens stabilizes transition states, discriminating substrates by up to 10^4-fold in rate enhancements. Side chains contribute to tertiary structure stability through van der Waals forces, which arise from close packing of atoms in the core and provide cumulative stabilization estimated at 50-100 kcal/mol for a typical protein. Covalent disulfide bonds between cysteine side chains further lock tertiary and quaternary structures, particularly in extracellular proteins, preventing unfolding under oxidative or mechanical stress and increasing melting temperatures by 10-20°C. These bonds form selectively during folding, guided by proximity in the native state, and are crucial for maintaining active conformations in enzymes like ribonucleases.

Applications in Other Fields

Polymer Chemistry

In polymer chemistry, side chains, also known as groups, are substituents attached to the repeating units along the main backbone, influencing the overall macromolecular structure and behavior. These groups can range from simple atoms or small moieties to complex molecular segments, and they are formed during processes such as or reactions. For instance, in , a serves as the side chain attached to every other carbon in the backbone, contributing to the polymer's rigidity and . The length, branching, and nature of side chains significantly affect key material properties, including crystallinity, flexibility, and . Longer or more branched side chains disrupt chain packing, reducing crystallinity by increasing free volume and hindering intermolecular interactions; for example, in , isotactic configurations with regularly aligned methyl side chains promote high crystallinity and a around 160–170°C, whereas atactic variants with randomly oriented methyl groups result in amorphous structures with lower density and flexibility./29%3A_Polymers/29.05%3A_Correlation_of_Polymer_Properties_with_Structure) Similarly, increasing side chain length lowers Tg by enhancing chain mobility and entropic freedom, as observed in poly(3-alkylthiophenes) where longer alkyl side chains shift Tg to lower temperatures, improving processability but potentially reducing mechanical strength. Branching also enhances flexibility in otherwise rigid polymers by introducing steric hindrance that prevents close packing. Side chains in polymers can be classified into types such as linear branches or more complex architectures like those in graft copolymers, where entire chains serve as side groups attached to a primary backbone. Linear polymers like (LDPE) feature short alkyl branches (e.g., ethyl or butyl groups) that arise during free-radical , leading to a branched with about 2–50 branches per 1000 carbon atoms and reduced density compared to linear . In graft copolymers, the side chains are distinct polymeric segments, enabling tailored properties such as improved compatibility in blends. Representative examples illustrate the functional impact of side chains. In (PVC), atoms act as small polar side chains, comprising 56.8% of the polymer's weight, which imparts inherent flame retardancy by releasing during and inhibiting further burning once the heat source is removed. architectures, with their highly branched side chains radiating from a , enhance in organic solvents like or , as seen in dendrimer-ligated aromatic polymers, facilitating applications in where poor of the backbone would otherwise limit utility.

Pharmaceutical Design

In pharmaceutical design, side chains are strategically incorporated into drug molecules to enhance desirable properties such as , , and target specificity, while minimizing undesirable traits like rapid clearance or off-target effects. These modifications allow medicinal chemists to fine-tune the pharmacokinetic profile of , ensuring they can effectively reach therapeutic sites in the body. For instance, lipophilic alkyl chains are often added to improve permeability and oral , as seen in statins like simvastatin, where such side chains contribute to higher by facilitating passive across lipid bilayers. Conversely, polar functional groups, such as hydroxyl or amino moieties on side chains, are employed to boost aqueous , addressing challenges in hydrophilic drugs that otherwise exhibit poor and limited systemic exposure. Structure-activity relationship (SAR) studies play a pivotal role in optimizing side chains, systematically varying their structure to correlate chemical modifications with , potency, and resistance profiles. In , alterations to the variable R-group side chain directly influence and susceptibility to bacterial resistance mechanisms, such as hydrolysis or altered penicillin-binding protein affinity. For example, in penicillin derivatives, the benzyl side chain in (penicillin G) confers a narrow primarily effective against Gram-positive cocci like streptococci and staphylococci, by optimizing binding to bacterial synthesis enzymes. Similarly, in HIV inhibitors such as , bulky cyclic or aromatic side chains are designed to occupy the enzyme's S1 and subsites, mimicking the natural and preventing viral maturation while evading mutations that confer resistance. Despite these benefits, side chain design must contend with metabolic liabilities and potential toxicity. Cytochrome P450 enzymes, particularly and , frequently oxidize side chains—such as alkyl or aromatic groups—leading to reactive metabolites that can cause idiosyncratic toxicities like or reactions. Unintended reactivity from electrophilic side chain fragments, often generated via P450-mediated oxidation, poses additional risks, necessitating proactive screening in to mitigate covalent binding to proteins and subsequent adverse events.

References

  1. [1]
    How to name organic compounds using the IUPAC rules
    b) the chain whose substituents have the lowest- numbers. c) the chain having the greatest number of carbon atoms in the smaller side chain. d)the chain having ...
  2. [2]
    Substituents in Organic Chemistry | Identification & Examples
    Substituents may also be called side chains or pendent groups. These terms have the meaning of being a group attached to the parent chain. Side chain is a term ...
  3. [3]
    Illustrated Glossary of Organic Chemistry - Amino acid side chain ...
    Amino acid side chain: The portion of alpha-amino acid structure other than the amino group (or the ammonium group in an ionic form), the alpha carbon, ...Missing: definition | Show results with:definition
  4. [4]
    The Structure of Proteins
    The unique side chains confer unique chemical properties on amino acids, and dictate how each amino acid interacts with the others in a protein. Amino acids ...
  5. [5]
    [PDF] 1-1 Amino Acids
    The chemical characters of the amino-acid side chains have important consequences for the way they participate in the folding and functions of proteins. The ...
  6. [6]
    The Amino Acid Collection - Molecular Expressions
    Nov 13, 2015 · The amino acid backbone determines the primary sequence of a protein, but the nature of the side chains determines the protein's properties.
  7. [7]
    Impact of Side-Chain Polarity and Symmetry on the Structure and ...
    Conformationally flexible side-chains on conjugated polymers promote solution processability while significantly impacting solution aggregation, solid-state ...
  8. [8]
  9. [9]
    Branching, and Its Affect On Melting and Boiling Points
    Jul 9, 2010 · Starting with the simplest branched compound, as you increase branching, you will increase the melting point, but decrease the boiling point.Branching, And Its Affect On... · 2. Higher Surface Area =... · 4. Some More Experimental...<|control11|><|separator|>
  10. [10]
    Impact of the alkyl side-chain length on solubility, interchain packing ...
    Aug 20, 2024 · Increasing the length of alkyl side chains is a typical way to improve the solubility of π-conjugated polymers designed for use in solution-processed devices.Missing: reactivity | Show results with:reactivity
  11. [11]
    3.1: Generic (Abbreviated) Structures (aka R Groups)
    May 30, 2020 · The 'R' group is a convenient way to abbreviate the structures of large biological molecules, especially when we are interested in something that is occurring ...
  12. [12]
    [PDF] Brief Guide to the Nomenclature of Organic Chemistry - IUPAC
    Identify the substituents and arrange the corresponding prefixes in alphabetical order. f. Insert multiplicative prefixes, without changing the already.Missing: iso neo
  13. [13]
    Nomenclature of Alkanes - Chemistry LibreTexts
    Jan 22, 2023 · These common names make use of prefixes, such as iso-, sec-, tert-, and neo-. The prefix iso-, which stands for isomer, is commonly given to 2- ...
  14. [14]
    skeletal formula (08208) - IUPAC Gold Book
    skeletal formula ... Two-dimensional representation of a molecular entity in which bonds are indicated as lines between vertices representing octet carbon atoms ...
  15. [15]
    2.2: Nomenclature of Alkanes - Chemistry LibreTexts
    Jun 30, 2024 · If the chains are unequal, the longest chain is the parent chain, even if it does not contain a double or triple bond, e.g., the first example ...
  16. [16]
    Functional Groups In Organic Chemistry
    Oct 6, 2010 · Note that “R” is a placeholder for a generic carbon substituent. ... Alkyl halides have functional group R-F, R-Cl, R-Br, R-I where R is an alkyl ...
  17. [17]
    Organic Nomenclature - MSU chemistry
    A general and useful generic notation that complements the use of R- for an alkyl group is Ar- for an aryl group (any aromatic ring).
  18. [18]
    2.4: IUPAC Naming of Organic Compounds with Functional Groups
    Jun 30, 2024 · This chain determines the parent name of the compound. Change the ... The numbers on the chain should start from the left side to ensure that ...
  19. [19]
    Table of Functional Group Priorities for Nomenclature
    Feb 14, 2011 · Highest priority groups are carboxylic acids, sulfonic acids, esters, acid halides, and amides. Next are nitrile, aldehyde, ketone, alcohol, ...
  20. [20]
    August Kekule von Stradonitz | German Chemist & Organic ...
    Sep 23, 2025 · Kekule argued that tetravalent carbon atoms could link together to form what he called a “carbon chain” or a “carbon skeleton,” to which other ...Missing: side | Show results with:side
  21. [21]
    200th Birthday: Charles Frédéric Gerhardt - ChemistryViews
    Aug 21, 2016 · Gerhardt's work brought some much-needed order to organic chemistry: He came up with the concept of homologous series (related compounds that ...Missing: R | Show results with:R
  22. [22]
    Aleksandr Butlerov | Organic Chemistry, Stereochemistry, Synthesis
    Sep 23, 2025 · In 1861 Butlerov stated his concept of chemical structure: that the chemical nature of a molecule is determined not only by the number and type ...Missing: side chains
  23. [23]
    Development of Structural Theory - ACS Publications
    He coined the term "chemical structure," recognized the existence of struc- tural isomers and the number of theoretically possible isomers for a given ...
  24. [24]
    August Kekulé and Archibald Scott Couper - Science History Institute
    Couper invented a symbolic language to represent carbon linkage. Both made significant contributions to the field of structural chemistry.
  25. [25]
    Kekulé, Butlerov, and the Historiography of the Theory of Chemical ...
    Jan 5, 2009 · Kekulé, Butlerov, and the Historiography of the Theory of Chemical Structure. Published online by Cambridge University Press: 05 January 2009.
  26. [26]
    Contribution of Chevreul to lipid chemistry | OCL
    Apr 20, 2023 · In this key book, Chevreul proposed for the first time a classification of all lipids known at his time, a classification that could still be ...
  27. [27]
    [PDF] One Hundred Years of Peptide Chemistry*
    Emil Fischer is considered to be the founding father of the field of peptide chemistry and originator of the term peptide. In the beginning of the 20th ...
  28. [28]
    The discovery of the α-helix and β-sheet, the principal structural ...
    PNAS papers by Linus Pauling, Robert Corey, and Herman Branson in the spring of 1951 proposed the α-helix and the β-sheet, now known to form the backbones ...
  29. [29]
    From phosphoproteins to phosphoproteomes: a historical account
    Jan 12, 2017 · The first phosphoprotein (casein) was discovered in 1883, yet the enzyme responsible for its phosphorylation was identified only 130 years ...
  30. [30]
    Review A Glimpse of Structural Biology through X-Ray Crystallography
    Nov 20, 2014 · Structural elucidation of DNA and protein is arguably the most important scientific discovery in the 20th century. Proposal of the double ...
  31. [31]
    Steric Hindrance - an overview | ScienceDirect Topics
    Steric hindrance refers to the spatial arrangement of molecular groups that affects the performance of devices by modifying the morphology and ...Appendix Ii Similarities And... · 1.15. 1.2. 3 Ring-Closing... · 1.08. 1.3. 3. (iii) Acyclic...
  32. [32]
    Importance of Side-Chains on Molecular Characteristics of ...
    It is emphasized that the presence of side-chains brings about the diversity of molecular characteristics in interacting molecules. This may point out the ...Missing: reactivity solubility
  33. [33]
    Intermolecular Force and Physical Properties of Organic Compounds
    Dec 15, 2021 · Intermolecular forces are the attractive force between molecules and that hold the molecules together; it is an electrical force in nature.
  34. [34]
    3.2 Solubility – Introductory Organic Chemistry
    If the solvent is polar, like water, then a larger dipole moment, indicating greater molecular polarity, will tend to increase the solubility of a substance in ...
  35. [35]
    22.2: Alkanes, Cycloalkanes, Alkenes, Alkynes, and Aromatics
    Jun 5, 2019 · A two-carbon chain is called ethane; a three-carbon chain, propane; and a four-carbon chain, butane. Longer chains are named as follows: pentane ...
  36. [36]
    3 Factors That Stabilize Carbocations - Master Organic Chemistry
    Jan 29, 2025 · Three main factors increase the stability of carbocations: Increasing the number of adjacent carbon atoms (methyl < primary < secondary ...
  37. [37]
    24.7: Chirality in Organic Chemistry
    Jul 7, 2023 · Molecules that are nonsuperimposable mirror images of each other are said to be chiral (pronounced “ky-ral,” from the Greek cheir, meaning “hand”).
  38. [38]
    4.1. Chirality | Organic Chemistry 1: An open textbook
    A non-superimposable set of two molecules that are mirror images of one another. The existence of these molecules is determined by concept known as chirality.Missing: creating | Show results with:creating
  39. [39]
    The Grignard Reagents | Organometallics - ACS Publications
    The Grignard reagents probably have been the most widely used organometallic reagents. Most of them are easily prepared in ethereal solution.Missing: introduction | Show results with:introduction
  40. [40]
    Organoborane coupling reactions (Suzuki coupling) - PMC - NIH
    In this article, the cross-coupling reactions of organoboron compounds are described briefly in the order of 1-alkenyl, aryl, 1-alkyl-, and 1-alkynylboron ...
  41. [41]
    Acid Zeolite Catalyzed Methylation of Benzene and Toluene
    ... EAS ... The values of ΔEbar follow the same order as for homogeneous reactions: the methyl substituent of toluene is an activating ortho-/para-directing group.Missing: chains | Show results with:chains
  42. [42]
    Practical N-to-C peptide synthesis with minimal protecting groups
    Oct 26, 2023 · Side chain protection was not necessary for amino acids containing functional groups of moderate nucleophilicity (Trp: 2bg, Ser: 2bi, Thr ...
  43. [43]
    Recent Advances in the Synthesis of Ibuprofen and Naproxen - NIH
    Herein, we review the recent progress in the synthesis of representative nonsteroidal anti-inflammatory drugs (NSAIDs), ibuprofen and naproxen.
  44. [44]
    Modification of ibuprofen to improve the medicinal effect; structural ...
    Mar 5, 2024 · The alteration of the propionic acid, aryl and/or the isobutyl moieties of ibuprofen has been observed to strengthen the anti-inflammatory, ...
  45. [45]
    SN2 versus E2 Competition of F– and PH2– Revisited
    Oct 20, 2020 · The anti-E2 elimination prevails over the SN2 substitution. This is ascribed to the lower energy and entropy barrier for the anti-E2 elimination ...
  46. [46]
    Energetics of complementary side-chain packing in a protein ...
    May 30, 1989 · The energetics of complementary packing of nonpolar side chains in the hydrophobic core of a protein were analyzed by protein engineering experiments.
  47. [47]
    Hydrogen bonding interactions between glutamine and asparagine ...
    We conclude that even in highly mobile, surface-exposed regions, hydrogen bonds can significantly stabilise proteins, provided that their geometric requirements ...
  48. [48]
    The structure and IR signatures of the arginine-glutamate salt bridge ...
    Jun 7, 2015 · Salt bridges and ionic interactions play an important role in protein stability, protein-protein interactions, and protein folding.
  49. [49]
    Role of histidine 64 in the catalytic mechanism of human carbonic ...
    To test the hypothesis that histidine 64 in the active site of human carbonic anhydrase II functions as a proton-transfer group in the catalysis of CO2 ...
  50. [50]
    Dynamic interplay between O-glycosylation and O-phosphorylation ...
    The glycosylation of serine and threonine residues with beta-O-linked N-acetylglucosamine (O-GlcNAc) is an abundant posttranslational modification of ...
  51. [51]
    Molecular mechanism of red cell "sickling" - PubMed
    Precision scale models of sickle-cell hemoglobin molecules indicate that the genetic substitution of valine for glutamic acid at the 6th position in the two ...
  52. [52]
    Electrostatic complementarity within the substrate-binding pocket of ...
    We calculated the changes in the transition-state energies for various enzyme-substrate pairs to reflect electrostatic and hydrogen-bond interactions.
  53. [53]
    Side-chain contributions to membrane protein structure and stability
    Side-chain contributions to membrane protein structure and stability ... van der Waals packing is the dominant force that drives membrane protein folding.
  54. [54]
    Mechanism of protein folding. III. Disulfide bonding - PubMed
    Among the possible combination of cysteine residues some definite pairs are realized in the tertiary structure. ... cysteine residues to make a disulfide bond.
  55. [55]
    Polystyrene - Polymer Science Learning Center
    Polystyrene is a vinyl polymer. Structurally, it is a long hydrocarbon chain, with a phenyl group attached to every other carbon atom. Polystyrene is produced ...
  56. [56]
    Glass transition temperature from the chemical structure of ... - Nature
    Feb 14, 2020 · As the side chain length of poly(3-alkylthiophene) (P3AT) increases, the glass transition temperature of the backbone shifts to a lower ...
  57. [57]
    Effects of flexibility and branching of side chains on the mechanical ...
    This paper describes effects of the flexibility, length, and branching of side chains on the mechanical properties of low-bandgap semiconducting polymers.
  58. [58]
    Poly(ethene) (Polyethylene) - The Essential Chemical Industry
    LDPE is generally amorphous and transparent with about 50% crystallinity. The branches prevent the molecules fitting closely together and so it has low density.
  59. [59]
    Polymeric Chain - an overview | ScienceDirect Topics
    Block copolymers have two or more homopolymer subunits linked by covalent bonds, while graft copolymers contain side chains attached to the main polymer chains.
  60. [60]
    [PDF] Fire Properties of Polyvinyl Chloride - The Vinyl Institute
    However, PVC will typically not burn once the source of heat or flame is removed. This results from PVC having 56.8% chlorine in its base polymer weight and it ...
  61. [61]
    Synthesis, properties, and material hybridization of bare aromatic ...
    Sep 16, 2022 · Notably, dendrimer-ligated bare aromatic polymers are highly soluble in common organic solvents such as chloroform, tetrahydrofuran (THF) and ...
  62. [62]
    Lipophilic Conjugates of Drugs: A Tool to Improve Drug ...
    Aug 31, 2021 · Highly hydrophilic drugs can suffer from poor permeability, and this may lead to pharmacokinetic issues including low oral bioavailability, poor ...
  63. [63]
    Hydrophilic or Lipophilic Statins? - PMC - NIH
    Drugs can be classified as hydrophilic or lipophilic depending on their ability to dissolve in water or in lipid-containing media.
  64. [64]
    [PDF] Functional Group Characteristics and Roles - ASHP
    Each functional group has an electronic effect, a solubility effect, and a steric effect that needs to be considered when evaluating the overall pharmacodynamic.
  65. [65]
    Structure-activity relationships of different beta-lactam antibiotics ...
    We compared the affinities of different beta-lactam antibiotics to a modified soluble form of a resistant Enterococcus faecium PBP5 (Delta1-36 rPBP5).
  66. [66]
    Benzylpenicillin: Uses, Interactions, Mechanism of Action - DrugBank
    Benzylpenicillin (Penicillin G) is narrow spectrum antibiotic used to treat infections caused by susceptible bacteria. It is a natural penicillin antibiotic ...Missing: chain | Show results with:chain
  67. [67]
    Recent Progress in the Development of HIV-1 Protease Inhibitors for ...
    HIV-1 protease inhibitors continue to play an important role in the treatment of HIV/AIDS, transforming this deadly ailment into a more manageable chronic ...
  68. [68]
    A history of the roles of cytochrome P450 enzymes in the toxicity of ...
    Aug 18, 2020 · P450s affect drug toxicity mainly by either reducing exposure to the parent molecule or, in some cases, by converting the drug into a toxic entity.
  69. [69]
    Survey of Human Oxidoreductases and Cytochrome P450 Enzymes ...
    Dec 8, 2014 · Regarding drug metabolism, three-fourths of the human P450 reactions can be accounted for by a set of five P450s: 1A2, 2C9, 2C19, 2D6, and 3A4, ...