Fact-checked by Grok 2 weeks ago

Anilide

Anilides are a of compounds that are derivatives of , characterized by the general structure R–C(=O)–NH–C₆H₅, where R represents a or an such as alkyl or aryl. These compounds are formed by replacing the hydroxyl group of an oxoacid with an anilino group (–NH–C₆H₅) or an N-substituted anilino group. In , anilides are distinguished from salts derived from , such as sodium anilide (NaNHPh), which result from of the amino group and represent a secondary usage of the term. Anilides are commonly prepared through reactions, where acts as a attacking the of an , carboxylic anhydride, or . For instance, (CH₃C(O)NHC₆H₅), the simplest and most representative anilide, is synthesized by heating with or , often in the presence of a to neutralize the released acid. This step is frequently employed as a strategy in to moderate the reactivity of the aniline amino group, preventing unwanted side reactions during . Anilides play significant roles in industrial and pharmaceutical applications due to their stability and versatility. , for example, serves as a key intermediate in the manufacture of dyes, rubber antioxidants, and precursors to antibiotics like . Historically, it was used as an and drug from the late until the 1940s, when it was largely replaced by safer alternatives like acetaminophen due to risks of . In modern contexts, substituted anilides are explored for agrochemical uses, such as herbicides, and in synthetic methodologies including palladium-catalyzed arylation and enzyme inhibition studies. (Note: Propionamide is not accurate; better: wait, no. Actually, for herbicides, cite )

Definition and Structure

General Formula

Anilides are organic compounds characterized by the general formula R-C(=O)-NH-C₆H₅, where R represents a , an , or an aryl group derived from a or its derivative. This structure consists of a (C=O) bonded to a nitrogen atom, which is further attached to a phenyl ring (C₆H₅), forming the characteristic N-phenyl linkage. The amide functionality in anilides features significant resonance stabilization, arising from the delocalization of the nitrogen lone pair into the π* orbital of the carbonyl group. This resonance imparts partial double-bond character to the C-N bond, resulting in a planar configuration around the amide group and restricted rotation. Additionally, the attached phenyl ring enables further conjugation, allowing electron delocalization involving the nitrogen, carbonyl, and aromatic system, which enhances overall stability. This delocalization affects the availability of the lone pair for . Anilides are slightly more than their aliphatic counterparts, as evidenced by the pKₐ of the conjugate acid for (a representative anilide) at approximately 0.6, compared to -0.5 for simple aliphatic amides like . The standard of an anilide illustrates the planar moiety with the carbonyl carbon connected to R, double-bonded to oxygen, single-bonded to NH, and the linked to the phenyl ring; the is often depicted with curved arrows showing the lone pair movement from N to the C-O bond, alongside orthogonal conjugation into the phenyl π-system for emphasis on aromatic involvement.

Nomenclature and Classification

Anilides are systematically named according to IUPAC recommendations as N-phenyl derivatives of carboxylic , typically using the prefix "N-phenyl" followed by the name of the parent , such as N-phenylalkanamide for those derived from alkanoic acids or N-phenylbenzamide for aromatic variants. For instance, the compound derived from acetic acid is designated N-phenylacet. Retained traditional names employing the suffix "-anilide" in place of "-amide" are also permitted for general use, particularly for simple derivatives. Historical common names, such as for N-phenylacetamide, continue to appear frequently in , industrial contexts, and pharmaceutical references due to their established usage. Anilides are classified based on the substitution pattern at the atom of the group. Primary anilides refer to those where the bears the and the phenyl substituent with one hydrogen remaining (R-C(O)-NH-C₆H₅), functioning as secondary s in standard terminology. Secondary anilides arise when the is further substituted with an (R-C(O)-NR'-C₆H₅, where R' is alkyl), rendering them tertiary s, while tertiary anilides would involve two additional non-phenyl substituents, though these are less common. This classification distinguishes anilides from related compounds like sulfonanilides (R-SO₂-NH-C₆H₅), which are derivatives rather than carboxamides derived from oxoacids. Anilides are further categorized by the nature of the R group in the general formula R-C(O)-NH-C₆H₅. Alkyl anilides feature an alkyl substituent for R, as exemplified by (R = CH₃), which is prevalent in applications like analgesics and intermediates due to its straightforward synthesis and reactivity. Aryl anilides, where R is an , include benzanilide (R = C₆H₅), commonly encountered in and as synthetic intermediates, though less ubiquitous than their alkyl counterparts in commercial use.

History

Discovery and Early Synthesis

The emergence of anilides coincided with the rapid expansion of in the mid-19th century, particularly through the exploration of derivatives in the burgeoning German chemical industry. , isolated from as early as 1841 and pivotal to the synthesis of the first artificial dye, , by in 1856, served as the foundational starting material for these compounds. This period saw intense research into aniline's reactions, driven by industrial demands for dyes and intermediates, laying the groundwork for derivatives like anilides. The first documented synthesis of an occurred in 1852, when French chemist prepared (N-phenylacetamide) by reacting with , a stemming from his studies on anhydrides. Gerhardt's approach involved treating the anhydride with the base, yielding the stable as part of broader experiments on acylations. This marked the initial entry into the class of anilides, though the compound remained obscure for decades, primarily noted in academic contexts without immediate practical application. The initial synthetic route—heating with —quickly became standard and notably produced as an unintended byproduct during aniline dye manufacturing processes in the late . Early isolations and characterizations of such compounds were reported in prominent journals, including Justus Liebigs Annalen der Chemie, where researchers documented purification techniques and basic properties amid the era's focus on aromatic amines. A pivotal advancement came in 1886, when physicians Arnold Cahn and Paul Hepp serendipitously isolated pure from a dye factory residue mistaken for , revealing its effects and spurring its recognition beyond mere chemical curiosity.

Development in Organic Chemistry

Following the establishment of foundational syntheses in the late , anilide chemistry integrated into broader theories during the and as emerged from advancements. This period saw anilides recognized for their resonance-stabilized structures, where the lone pair delocalizes into the carbonyl, imparting partial double-bond character to the C-N linkage and influencing . A pivotal contribution came from in the late 1930s, who applied resonance theory to explain the planar amide configuration and consequent reduced reactivity of the nitrogen atom compared to typical amines. Pauling's analysis, detailed in his seminal work, demonstrated how this delocalization stabilizes the bond, hindering nucleophilic attack and rotation, which accounted for the lower basicity and slower rates observed in anilides. These insights, verified through and spectroscopic studies, unified anilide behavior within general chemistry and influenced subsequent research. Post-World War II, anilide chemistry scaled industrially alongside the pharmaceutical boom, driven by demand for analgesics and antipyretics amid expanded chemical manufacturing capabilities. Anilide derivatives played a central role, exemplified by the development of (acetaminophen) in the 1950s as a safer successor to early anilides like , which had been limited by toxicity concerns. This shift, spurred by metabolic studies in the late identifying as the active metabolite of , enabled mass production for global use in fever reduction and pain relief, marking anilides' transition to high-volume therapeutic synthesis. In recent years up to 2025, computational modeling has advanced understanding of anilide conformations, employing (DFT) and multivariate to predict bond formation efficiencies and rotational barriers. These models reveal how substituents affect nitrogen pyramidalization and bond twist in anilides, aiding design of conformationally flexible variants for drug-like molecules. Concurrently, green synthesis routes have emphasized , utilizing Brønsted acidic ionic liquids in aqueous media for metal-free anilide assembly with high yields and catalyst recyclability. Organocatalysis has further propelled anilide formation, with innovations like the photoactive Cat-Se catalyst enabling direct amidation of anilines under mild blue LED irradiation, achieving up to 93% yields in minutes without racemization. Modern researchers, including those advancing boron-mediated and chiral organocatalysts, have focused on selective N-arylation routes that minimize waste and enhance stereocontrol. These developments, building on Pauling's foundational resonance concepts, underscore anilide chemistry's ongoing evolution toward efficient, eco-friendly applications.

Synthesis

Acylation of Aniline

The acylation of represents the classical and most widely employed route for synthesizing anilides, involving the reaction of (C₆H₅NH₂) with an acylating agent to form N-phenylamides of the general formula RCONHC₆H₅. In the standard laboratory procedure, is treated with an (RCOCl) in the presence of a base such as , which serves both as a scavenger and , to afford the anilide product and HCl as a byproduct. The reaction is typically conducted at or mild heating in an inert atmosphere to minimize side reactions, with equimolar ratios ensuring high selectivity for monoacylation due to the reduced nucleophilicity of the resulting relative to the starting . The mechanism proceeds via : the on the of attacks the electrophilic carbonyl carbon of the , forming a tetrahedral ; subsequent elimination of and by the base yields the anilide. This addition-elimination pathway is facilitated by the high reactivity of acyl chlorides toward nucleophiles like aniline. The overall transformation can be represented as: \text{C}_6\text{H}_5\text{NH}_2 + \text{RCOCl} \xrightarrow{\text{pyridine}} \text{RCONHC}_6\text{H}_5 + \text{HCl} Yields for this method are generally excellent, often exceeding 90%, owing to the efficiency of the acyl chloride as an activating group. Alternative acylating agents include acid anhydrides, such as (RCO)₂O, which react with aniline under milder conditions, often in aqueous or ethanolic media with a buffering base like sodium acetate to neutralize the carboxylic acid byproduct. For example, acetic anhydride acetylates aniline to acetanilide in an exothermic process, followed by precipitation and recrystallization for purification. The mechanism mirrors that of acyl chlorides, involving nucleophilic attack on one carbonyl of the anhydride, tetrahedral intermediate formation, and elimination of RCOOH. Yields with anhydrides typically range from 80% to 95%, comparable to acyl chlorides but with the advantage of generating less corrosive byproducts. Esters can also serve as acylating agents for anilide synthesis, though this requires elevated temperatures (140–160°C) to overcome the lower reactivity due to the poorer leaving group (). The reaction, known as aminolysis, proceeds via a similar nucleophilic addition-elimination mechanism and is often performed in high-boiling solvents like , yielding 80–95% of the anilide after extended heating. Optimization of these acylations frequently involves solvent selection to enhance and ; for instance, () is commonly used with acyl chlorides and due to its low polarity and ability to dissolve both reactants while facilitating of HCl. Catalysts are rarely required for monoacylation, as the intrinsic in nucleophilicity between and the product inherently favors the monosubstituted outcome, though excess acylating agent can be avoided by precise .

Other Synthetic Routes

Catalytic methods have emerged as green alternatives, exemplified by palladium-catalyzed aminocarbonylation reactions that incorporate CO surrogates to form anilides from s and s. In one such process, an (ArX) couples with and a CO surrogate (e.g., or ) under Pd catalysis, generating the anilide via , CO insertion, and , as represented by the general equation: \ce{ArX + C6H5NH2 + CO -> ArCONHC6H5 + HX} These reactions typically proceed at 80-120°C in solvent-free or low-solvent conditions, yielding 60-85% of the product and emphasizing atom economy by avoiding acylating agents like chlorides. The use of CO surrogates eliminates the need for toxic gaseous CO, aligning with green chemistry principles, and is especially useful for scale-up in pharmaceutical synthesis where complex aryl groups are tolerated.

Properties

Physical Characteristics

Most anilides appear as white crystalline solids at , reflecting their molecular structure that favors solid-state packing through hydrogen bonding and π-π interactions. Their melting points generally fall within the range of 100–200 °C, influenced by the size and nature of the ; for instance, melts at 114 °C. Anilides display low solubility in water, typically less than 0.5 g/100 mL at 25 °C due to the hydrophobic aromatic ring dominating over the polar functionality, though this can vary slightly with substituents. In contrast, they dissolve well in polar organic solvents such as , acetone, and . Partition coefficients () for representative anilides, like acetanilide at 1.16, indicate moderate hydrophobicity, generally spanning 1–3 for common derivatives. Infrared spectroscopy reveals a characteristic carbonyl (C=O) stretching absorption for anilides at 1671–1686 cm⁻¹, shifted to lower wavenumbers compared to aliphatic amides due to resonance delocalization of the nitrogen lone pair into the aromatic ring, which weakens the C=O bond. Proton NMR spectra show a broad signal for the NH proton around 8–9 , arising from hydrogen bonding and , alongside aromatic protons in the 7–8 . Anilides exhibit good thermal stability, often decomposing only above 250 °C, with many subliming or near 300 °C before full ; , for example, has a of 304 °C and onset above this . Their vapor pressures are low, typically less than 0.1 mmHg at 100 °C, as seen with at approximately 0.3 mmHg under similar conditions, limiting volatility at ambient temperatures.

Chemical Reactivity

Anilides exhibit reduced nucleophilicity at the nitrogen atom compared to amines, primarily due to resonance delocalization of the nitrogen lone pair into the carbonyl group, which stabilizes the amide bond and diminishes the availability of the lone pair for interactions such as protonation./07%253A_Acid-base_Reactions/7.06%253A_Acid-base_properties_of_nitrogen-containing_functional_groups) This resonance effect renders the nitrogen in anilides much less basic, with the pKa of the conjugate acid of the protonated amide typically around 0, in contrast to the pKa values of approximately 10-11 for protonated amines. The -NHCOR on the aromatic of anilides acts as an ortho-para in reactions, owing to the electron-donating effect from the , which outweighs the electron-withdrawing of the . This group is moderately activating, weaker than the free -NH₂ due to the partial delocalization of the lone pair into the carbonyl, leading to a preference for substitution at ortho and para positions while avoiding excessive reactivity. Anilides demonstrate greater resistance to than esters under both acidic and basic conditions, attributed to the poor leaving group ability of compared to or phenoxide ions in esters. This stability arises from the resonance-stabilized bond, which reduces the electrophilicity of the carbonyl carbon and hinders nucleophilic attack, requiring harsher conditions for cleavage./17:_Carboxylic_Acids_and_their_Derivatives/17.04:_Hydrolysis_of_Esters_and_Amides) In terms of redox behavior, anilides are generally stable toward oxidation, lacking the facile electron donation from nitrogen that makes anilines prone to oxidative dimerization or polymerization. However, they show sensitivity to strong reducing agents, such as silanes in the presence of catalysts, which can promote deacylative cleavage of the C-N bond to yield amines and aldehydes.

Reactions and Derivatives

Hydrolysis and Cleavage

Anilides undergo hydrolysis under acidic conditions to afford the corresponding carboxylic acid and aniline. Treatment with concentrated hydrochloric acid or sulfuric acid at reflux temperature facilitates this transformation, typically requiring several hours for completion. The mechanism proceeds via protonation of the carbonyl oxygen, which enhances the electrophilicity of the carbon, followed by nucleophilic attack by water to form a tetrahedral intermediate; subsequent proton transfers and expulsion of aniline restore the carbonyl, yielding the products. Basic hydrolysis of anilides is generally slower and demands harsher conditions compared to acidic hydrolysis, often involving concentrated at elevated temperatures around 150°C. The reaction follows an addition-elimination pathway, where adds to the carbonyl to form a tetrahedral , leading to expulsion of and formation of the : \mathrm{RCONHC_6H_5 + OH^- \rightarrow RCOO^- + C_6H_5NH_2} This process is less efficient for anilides due to their inherent under media, unless electron-withdrawing groups are present on the acyl moiety to accelerate cleavage. In protecting group chemistry, serve to temporarily mask the amino group of during , allowing selective on the ring; deprotection is achieved via targeted under acidic conditions to regenerate and the . hydrolysis is slower than that of corresponding alkyl amides, primarily due to enhanced stabilization involving the aromatic ring, which reduces the electrophilicity of the .

Electrophilic Substitution on the Aromatic Ring

The -NHCOR group in anilides acts as an and activator in reactions due to donation from the , which increases on the and positions of the aromatic ring, despite some counteracting inductive withdrawal by the carbonyl. This directing effect moderates the ring's reactivity compared to free anilines, preventing over-substitution while favoring substitution at the position over due to steric hindrance. A representative example is the bromination of acetanilide, where treatment with in glacial acetic acid generates Br₂ as the , leading predominantly to the para-bromo derivative. The reaction proceeds under mild conditions at , yielding the as the major product with high . of anilides, such as , employs a of concentrated nitric and sulfuric acids to produce the nitronium (NO₂⁺), resulting in para-nitroanilides as the major products with approximately 79% para and 19% ortho selectivity. Overall yields for the nitration process are typically around 70-80%, depending on purification. The mechanism involves electrophilic attack on the electron-rich aromatic ring, forming a resonance-stabilized complex () where the positive charge is delocalized, with significant stabilization at the and positions by the -NHCOR group through involving the nitrogen . then restores . A simplified for bromination is: \mathrm{RC(O)NHC_6H_5 + Br_2 \rightarrow p\text{-Br-C_6H_4NHC(O)R + HBr}} These substitutions enable regioselective functionalization of anilides, which is valuable in synthesizing intermediates for dyes and pharmaceuticals, as the preserved amide group allows further transformations while controlling product distribution.

Applications

Pharmaceutical and Medicinal Uses

Anilides have played a significant role in pharmaceutical development, particularly as analgesics and antipyretics in the late 19th and early 20th centuries. Acetanilide, introduced in 1886 by French chemists Arnold Cahn and Paul Hepp, was the first synthetic compound recognized for its fever-reducing properties and soon thereafter for its pain-relieving effects, marketed under names like Antifebrin. It served as a widely used alternative to natural remedies like salicylates until the 1940s, when its metabolism to acetaminophen was identified as the active metabolite responsible for therapeutic benefits. This discovery paved the way for the development of acetaminophen as a safer derivative, highlighting acetanilide's role as a foundational compound in analgesic therapy. In applications, anilide derivatives have been employed to combat bacterial and parasitic infections. , a salicylanilide, is an FDA-approved agent used to treat tapeworm infections by disrupting parasite mitochondrial function and energy production, demonstrating broad-spectrum potential including activity against like . Other anilide-based compounds, such as , function as bacteriostatic and fungistatic agents in topical formulations, inhibiting microbial growth through disruption of integrity. These derivatives underscore the utility of the anilide moiety in enhancing and targeting infectious pathogens in clinical settings. Contemporary medicinal uses of anilides extend to and inflammatory disorders through (HDAC) inhibitors. (suberoylanilide hydroxamic acid, SAHA), the first FDA-approved HDAC inhibitor in 2006, treats by promoting histone acetylation, leading to tumor cell differentiation, growth arrest, and ; clinical trials have shown objective response rates of 24-30% in advanced cases. Analogs of , retaining the anilide core, are under investigation for broader cancer applications and exhibit anti-inflammatory efficacy, as is being investigated in a phase II trial (NCT03167437) for potential reduction of inflammation in patients through modulation of proinflammatory cytokines. These developments emphasize anilides' versatility in epigenetic modulation for therapeutic intervention. Early anilides like were associated with significant toxicity, notably inducing by oxidizing hemoglobin's iron, which impairs oxygen transport and can lead to and in susceptible individuals. This adverse effect, reported in clinical use from the 1890s onward, prompted regulatory scrutiny and the shift to safer alternatives by the 1950s, including acetaminophen, which avoids methemoglobin formation while retaining efficacy. Modern anilide-based drugs incorporate structural modifications to mitigate such risks, ensuring improved safety profiles in long-term therapies.

Industrial and Agricultural Applications

Anilides play a significant role as intermediates in the synthesis of azo dyes and pigments, particularly functioning as components that react with diazonium salts to produce vibrant colorants. For instance, acetoacetanilides are commonly employed in this process, enabling the formation of stable azo compounds used extensively in textile dyeing for fabrics like and , where they provide excellent color fastness to light and washing. These pigments also find applications in inks and coatings, leveraging the electron-rich nature of anilides to facilitate efficient reactions that yield high-purity, photoreceptor-grade materials suitable for colorants. In , anilide-based fungicides such as carboxin (5,6-dihydro-2-methyl-1,4-oxathiin-3-carboxanilide) have been utilized since primarily as treatments to combat basidiomycete pathogens, including smuts and bunts in crops like , , and . Carboxin's efficacy stems from its targeted inhibition of , a key in the fungal mitochondrial respiratory chain, disrupting energy production and leading to pathogen control without broad-spectrum effects on beneficial microbes. This selective mode of action has made it a staple in , often formulated with for enhanced protection against seedling diseases. Anilides are incorporated as stabilizers in plastics to mitigate oxidative degradation during processing and use, with compounds like oxalanilides serving as effective UV absorbers that scavenge free radicals and prevent polymer chain scission in materials such as polyamides and polyolefins. These additives extend the service life of plastic products, including automotive parts and outdoor gear, by absorbing harmful UV radiation and inhibiting photo-oxidation. Global production of anilide-based polymer stabilizers reaches thousands of tons annually, supporting the expansive that consumes over 36 million tons of additives yearly. The environmental impact of anilide pesticides, exemplified by carboxin, involves moderate biodegradability, with complete mineralization to CO₂ and biomass achievable by specialized bacteria like Delftia sp. HFL-1 under aerobic conditions at 30–42°C, though rates vary by soil type and microbial community. In natural settings, carboxin exhibits half-lives of 7–14 days in soil due to hydrolysis and microbial metabolism, reducing persistence. The U.S. EPA regulates anilide fungicides under FIFRA, classifying carboxin as having low acute toxicity to birds and bees but moderate to high acute toxicity to aquatic organisms (LC₅₀ 1–7 mg/L for fish), with guidelines mandating buffer zones and application limits to protect waterways and non-target species.

Notable Anilides

Acetanilide

, with the chemical formula CH3CONHC6H5 (or C8H9NO), represents the prototypical anilide as the N-acetyl derivative of . It was among the first major anilides synthesized in the mid-19th century, initially prepared through the of , establishing it as a foundational compound in for protecting amine groups during . Historically, gained prominence when introduced as Antifebrin in 1886 by German physicians Arnold Cahn and Paul Hepp, who discovered its antipyretic and properties while experimenting with derivatives for treating fever in a patient with typhoid. Marketed widely for pain relief and fever reduction, its therapeutic use was eventually curtailed due to reports of excessive toxicity, including leading to and , prompting its gradual replacement in clinical practice by the 1940s, when safer alternatives like became available. Nonetheless, played a pivotal role in the discovery of (); in the 1940s, researchers identified p-hydroxy—a of —as the primary active responsible for its beneficial effects, leading to the development and widespread adoption of the safer drug. Acetanilide exhibits a of 114°C and limited in , approximately 0.5 g/100 mL at 25°C, rendering it more soluble in solvents like , acetone, and . These properties contribute to its current niche applications as an intermediate in , particularly for rubber accelerators, where it facilitates cross-linking in production. Additionally, it serves as a precursor in the manufacture of dyes, pharmaceuticals, and other fine chemicals, including penicillin analogs, while finding limited use as an analytical reagent in such as electrophilic substitutions. Carboxin, systematically named 5,6-dihydro-2-methyl-1,4-oxathiin-3-carboxanilide, is an anilide-class systemic primarily used as a to protect crops from soilborne and seedborne fungal pathogens. First reported in and introduced commercially in 1969 by Uniroyal (now part of Agriscience), it marked one of the early examples of targeted protectants for crops, offering protection against diseases that affect and early growth. The fungicide's involves binding to the ubiquinone site on , a key in complex II of the mitochondrial , thereby disrupting fungal respiration and energy production. This inhibition is particularly effective against basidiomycete fungi, including smuts (such as Ustilago nuda in ) and bunts in cereals, as well as certain rusts, providing systemic uptake from treated seeds to protect emerging seedlings. Related compounds include oxycarboxin, the 4,4-dioxide () analog of carboxin, which exhibits greater stability in tissues and a broader spectrum of activity, extending efficacy to foliar rusts on cereals, , and ornamentals while maintaining control over smuts. Globally, carboxin and its analogs have been applied as seed protectants in , often in mixtures with other fungicides like for enhanced performance. However, due to moderate environmental persistence (soil DT50 of 14–42 days) and potential ecological risks, registration has lapsed in the , with carboxin withdrawn from approval on May 31, 2024, leading to phase-out in that region during the 2020s, with ongoing regulatory reviews elsewhere and the development of alternative inhibitors (SDHIs).

References

  1. [1]
    anilides (A00356) - IUPAC Gold Book
    Salts formed by replacement of a nitrogen-bound hydron of aniline by a metal or other cation, e.g. NaNHPh sodium anilide. ... International Union of Pure and ...
  2. [2]
    Anilide - an overview | ScienceDirect Topics
    Anilide is a class of compounds with an aniline group (phenyl attached to amino), and can inhibit certain enzymes.
  3. [3]
    Aluminum-catalyzed efficient synthesis of anilides by the acylation of ...
    Nov 19, 2012 · Amides and anilides are also important in pharmaceutical, agrochemical industries, and are used as protecting groups in organic synthesis 5 6.
  4. [4]
    Acetanilide | C8H9NO | CID 904 - PubChem - NIH
    It has a role as an analgesic. It is a member of acetamides and an anilide. It is functionally related to an acetic acid.
  5. [5]
    Blue Book P-66-69 - IUPAC nomenclature
    N-Phenyl derivatives of primary amides are called 'anilides' and may be named using the term 'anilide ... definition given in P-67.2. As the reported ...
  6. [6]
    [PDF] Classes of Amides that Undergo Selective N–C Amide Bond Activa
    Resonance energy of the parent N,N-Ph/Me anilide is 13.5 kcal/mol, which indicates a lower degree of amide bond destabilization compared with other classes of ...<|control11|><|separator|>
  7. [7]
    [PDF] PDF - IUPAC nomenclature
    N-Phenyl derivatives of primary amides are called 'anilides' and may be named using the term 'anilide' in place of. 'amide' in systematic or retained names of ...<|control11|><|separator|>
  8. [8]
    R-5.7.8 Amides, imides, and hydrazides - ACD/Labs
    N-Phenyl derivatives of primary amides are called "anilides" and may be named using the suffix "-anilide" in place of the suffix "-amide".
  9. [9]
    Sir William Henry Perkin | Organic synthesis, Dye-making, Aniline
    In 1858 he and B.F. Duppa synthesized glycine in the first laboratory preparation of an amino acid. They synthesized tartaric acid in 1860. After Graebe and ...
  10. [10]
    The Project Gutenberg eBook of History of Chemistry, Volume 2 (of 2)
    Oct 23, 2024 · Acetanilide C6H5NH.COCH3, an aniline derivative, was discovered by Gerhardt in 1853. Phenacetin is a derivative of para-aminophenol: An ...
  11. [11]
    Acetanilide - an overview | ScienceDirect Topics
    They discovered two alternative drugs, acetanilide and phenacetin. Figure ... It was reported that French chemist Charles Gerhardt synthesized paracetamol in 1852 ...
  12. [12]
    Pain relief: from coal tar to paracetamol | Feature - RSC Education
    Jun 30, 2005 · Cahn and Hepp published their discovery in 1886, cryptically attributing it to 'a fortunate accident',4 and the following year acetanilide was ...<|control11|><|separator|>
  13. [13]
    Drug Therapeutics & Regulation in the U.S. - FDA
    Jan 31, 2023 · Discovered in 1886 by French scientists Arnold Cahn and Paul Hepp, acetanilide became the first drug product with both analgesic and antipyretic ...
  14. [14]
    [PDF] Linus Pauling - National Academy of Sciences
    LINUS CARL PAULING was born in Portland, Oregon, on. February 28, 1901, and died at his ranch at Big Sur,. California, on August 19, 1994.
  15. [15]
    All Quotes (60) - Linus Pauling and the Structure of Proteins: A ...
    ... resonance structure - observations that fit precisely with Pauling's studies of the amide bond in urea during the early 1930s." Lily E. Kay. The Molecular ...Missing: reduced reactivity primary
  16. [16]
    [PDF] Linus Pauling - Nobel Lecture
    Knowledge of the structure of amides and also of the amino acids, provided by the theory of resonance and verified by extensive careful experimental studies ...Missing: reduced reactivity 1930s primary source
  17. [17]
    Pauling's Conceptions of Hybridization and Resonance in Modern ...
    Jul 6, 2021 · Linus Pauling's qualitative conceptions of directional hybridization and resonance delocalization are manifested in all known variants of modern computational ...
  18. [18]
    Paracetamol (Acetaminophen) - Pharmaceutical Drugs - NCBI - NIH
    Paracetamol is available in pure form as numerous trade-name preparations for oral use. It is also found combined in over 200 preparations with other drugs.Missing: WWII | Show results with:WWII
  19. [19]
    Paracetamol: past, present, and future - PubMed
    As a consequence, paracetamol became the mainstay analgesic and antipyretic for children with a subsequent reduction in the incidence of Reye's syndrome.Missing: acetanilide post WWII
  20. [20]
    Predicting relative efficiency of amide bond formation using ... - PNAS
    Apr 11, 2022 · This study outlines a data science–based workflow for effective statistical modeling with sparse experimental data.Missing: anilide | Show results with:anilide
  21. [21]
    A Small Change in Structure, a Big Change in Flexibility - MDPI
    Abstract. Studies of the rotational barrier energy of the amide bond using quantum computing and nuclear magnetic resonance (NMR) are focused mainly on its use ...Missing: anilide | Show results with:anilide
  22. [22]
    Green Synthesis of Primary Aniline-Based Indolylmethanes via One ...
    Aug 2, 2025 · In this study, a practical, sustainable, and metal-free methodology for synthesizing these derivatives was presented. This approach utilized a ...
  23. [23]
    Design and Development of an Organocatalyst for Light Accelerated Amide and Peptide Synthesis
    ### Summary of Organocatalyst for Amide Synthesis, Especially Anilides
  24. [24]
    Recent Advances in Direct Amidation Via Organocatalysis - Hussein
    Dec 8, 2024 · This review comprehensively discusses catalytic amidation reactions published up to January 2024, focusing on both organocatalysis and boron-based catalysis.
  25. [25]
    Acylation of Aniline: Videos & Practice Problems - Pearson
    The acetylation of aniline typically involves the use of an acyl chloride (such as acetyl chloride) and a base. Pyridine is a commonly used base in this ...
  26. [26]
    Reactions of Aniline - Chemistry Steps
    The amide is still an activator and an ortho, para, but not as strong as to cause over-substitution. The steric hindrance by the acetyl group directs the ...
  27. [27]
    Preparation of Acetanilide: Step-by-Step Lab Guide - Vedantu
    Rating 4.2 (373,000) Acetanilide is prepared by the acetylation of aniline with acetic anhydride in the presence of glacial acetic acid or concentrated sulfuric acid as a catalyst.Preparation Of Acetanilide... · Step-By-Step Procedure · Other Preparative Routes
  28. [28]
    1: Acetylation of Aniline (Experiment) - Chemistry LibreTexts
    Aug 16, 2021 · Acetanilide is prepared from aniline using an acetylation reaction. Acetylation is often used to place an acetyl protecting group on primary or secondary ...
  29. [29]
    US5130443A - Process for the preparation of anilides
    The anilide couplers can be prepared in an embodiment by first converting the carboxylic acid to an aromatic ester and then reacting the aromatic ester with an ...<|control11|><|separator|>
  30. [30]
    Synthesis of Anilides by Aminolysis of Unactivated Esters using ...
    May 9, 2023 · For non-cyclic alkylated anilines, the reaction temperature was increased to 160 °C to afford 3 an in 80% yield.
  31. [31]
    What are the roles of pyridine and DCM in the acylation of an alcohol?
    Feb 7, 2016 · Dichloromethane (DCM) is a common organic solvent and this is its role in both reactions (and in most cases where it is used).
  32. [32]
    Synthesis of N-substituted aryl amidines by strong base activation of ...
    Abstract. We describe an efficient method for the direct preparation of N-substituted aryl amidines from nitriles and primary amines.
  33. [33]
    A New Approach for the Synthesis of N-Arylamides Starting ... - MDPI
    The requisite amidine substrates were prepared from amines and nitriles by applying the Pinner reaction approach [21]. Considering the easy access of amidines ...Missing: variant | Show results with:variant
  34. [34]
    Aminocarbonylation using CO surrogates - RSC Publishing
    Dec 4, 2024 · This review looks at the various CO substitutes used in aminocarbonylation reactions. These include metal carbonyls, acids, formates, chloroform, and others ...
  35. [35]
    CO Surrogates: A Green Alternative in Palladium-Catalyzed CO Gas ...
    Carbonylation reactions with carbon monoxide (CO) provide efficient and attractive routes for the synthesis of bulk and fine chemicals.
  36. [36]
    [PDF] Tables of Derivatives
    ANILIDE p-TOLUIDIDE. 118. 82. 114. 147. 185. 171. 167 o-Anisic (methoxybenzoic) 100 ... Aniline. 183. 114. 160. 154. Benzylamine. 184. 60. 105-. 156. N- ...
  37. [37]
    Acetanilide | 103-84-4 - ChemicalBook
    Jan 27, 2025 · Melting point, 113-115 °C(lit.) Boiling point, 304 °C(lit.) Density ... Acetanilide is prepared from aniline by acetylating it with acetic ...
  38. [38]
    SOLUBILITIES OF SOME NORMAL ALIPHATIC AMIDES, ANILIDES ...
    Extracting chemical information from initial partial charges and electronic properties of reactants to predict order of chemical reactivity: Heterogeneous ...
  39. [39]
    [PDF] ACETANILIDE CAS N°: 103-84-4
    Physical-chemical properties of acetanilide are as follows: melting point 113.7 ºC, boiling point 304 ºC at 760 mmHg, water solubility 4 g/L at 20 ºC, Log Pow ...
  40. [40]
    [PDF] Infrared spectroscopic studies of amides and anilides
    There are two possible resonance structures in amides as indicated below. ... Infrared Spectroscopic Studies of Amides and Anilide~. TABLE V lntegrated ...
  41. [41]
    [PDF] I. Acetanilide, N-Methylacetanilide and Related Compounds
    in Table 4, the N-H proton resonance moves upfield in this series from 8.58. Table 4. NMR data for various anilides measured on 10 mole % chloroform-d solutions ...<|separator|>
  42. [42]
    A Spectroscopic Overview of Intramolecular Hydrogen Bonds of NH ...
    The chemical shifts of the hydrogen bonded NH protons in these compounds are in the 8–9 ppm range [11]. Figure 3. Figure 3 · Open in a new tab. Bis(6-amino-1,3- ...
  43. [43]
    Acetanilide - Phoenix College
    Decomposition. (). Decomposition Temperature: >304°C (22) Decomposition Products: carbon monoxide, carbon dioxide and nitrogen oxides. (22). Chemical ...
  44. [44]
    Acetanilide CAS#: 103-84-4 - ChemicalBook
    Acetanilide ; bulk density, 370kg/m3 ; vapor density, 4.65 (vs air) ; vapor pressure, 1 mm Hg ( 114 °C) ; refractive index, 1.5700 (estimate) ; Flash point, 173 °C.
  45. [45]
    Basicity of Amines - Chemistry Steps
    Amides are generally weaker bases than amines. The pKa of the conjugate acid of a typical amide is about zero, and there are two factors contributing to this ...
  46. [46]
    16.4 Substituent Effects in Electrophilic Substitutions - OpenStax
    Sep 20, 2023 · All activating groups are ortho- and para-directing, and all deactivating groups other than halogen are meta-directing. Halogens are unique in ...Missing: anilide | Show results with:anilide
  47. [47]
    [PDF] Hydrolysis of Anilides - Acta Chemica Scandinavica
    Phenyl- and thiolesters have good leaving groups whereas amides and anilides have poor ones. This fact alone is a sufficient reason for the appearance of large ...
  48. [48]
    Highly Chemoselective Reduction of Amides (Primary, Secondary ...
    Jan 24, 2014 · The reaction proceeds with C–N bond cleavage in the carbinolamine intermediate, shows excellent functional group tolerance, and delivers the ...
  49. [49]
    THE HYDROLYSIS OF ACETANILIDE. - ACS Publications
    The data and calculations are given in Tables 2 and 3. When the reaction takes place in acetic acid solution the speeds of the reaction in the two directions ...
  50. [50]
    Acid and base-catalyzed hydrolysis of amides (video) - Khan Academy
    Aug 15, 2014 · Let's look at the mechanism for acid-catalyzed hydrolysis of amides. The first step is protonation of the carbonyl oxygen.Missing: anilide | Show results with:anilide
  51. [51]
    None
    ### Summary: Anilides or Aryl Amides as Protecting Groups for Carboxylic Acids
  52. [52]
    [PDF] A mild alkaline hydrolysis of N- and N,N-substituted amides ... - Arkivoc
    We realized, as expected, that the hydrolysis rate of aryl amides is slower than that for aliphatic amides. Moreover, the hydrolysis rate is accelerated with ...
  53. [53]
  54. [54]
    [PDF] A Study on Electrophilic Aromatic Substitution of Acetanilide - ijarsct
    The electrophilic aromatic substitution reactions of acetanilide—specifically nitration, bromination, and sulfonation— demonstrate the significant influence ...
  55. [55]
    History of antipyretic analgesic therapy - PubMed
    Phenacetin was developed in 1886. Acetaminophen has been in use since the 1890s.
  56. [56]
    Acetaminophen | Research Starters - EBSCO
    In the late 1940s, researchers at Yale University determined that acetaminophen was the major metabolite of acetanilide, a known analgesic and antipyretic, and ...
  57. [57]
    Niclosamide: Uses, Interactions, Mechanism of Action - DrugBank
    Niclosamide is an anthelmintic indicated in the treatment of beef, pork, fish, and dwarf tapeworm infections in adults and children.
  58. [58]
    Anilide - an overview | ScienceDirect Topics
    Anilides, such as triclocarban and tribromsalan, are chemical compounds used as preservatives and in antiseptics. They are bacteristatic and fungistatic.Missing: sulfonanilides distinction
  59. [59]
    Vorinostat: Uses, Interactions, Mechanism of Action | DrugBank Online
    Jun 13, 2005 · Vorinostat is a histone deacetylase (HDAC) inhibitor used for the treatment of cutaneous manifestations in patients with progressive, persistent ...
  60. [60]
  61. [61]
    Acetanilide - an overview | ScienceDirect Topics
    Acetaminophen was first described in the scientific literature in 1878. However, in 1899, Karl Morner of Germany discovered that acetanilide is metabolized ...
  62. [62]
    Azo dye - Wikipedia
    Azo dyes are also prepared by the condensation of nitrated aromatic compounds with anilines followed by reduction of the resulting azoxy intermediate: ArNO2 ...
  63. [63]
    Classifications, properties, recent synthesis and applications of azo ...
    The azo dyes are generally characterized by a chemical groups capable of forming covalent bonds with the textile substrates. The energy required for the rupture ...Missing: acetoacetanilide | Show results with:acetoacetanilide
  64. [64]
    Azo pigments and their intermediates. A facile synthesis of ...
    The synthesized anilides are of high purity and can be used to synthesize photoreceptor-grade (IR-sensitive) azo pigments without further purification.
  65. [65]
    Carboxin - Wikipedia
    Carboxin is a narrow-spectrum fungicide used as a seed treatment in agriculture to protect crops from fungal diseases. It was first marketed by Uniroyal in ...
  66. [66]
    Carboxin | C12H13NO2S | CID 21307 - PubChem - NIH
    Systemic fungicide. /Used for/ seed treatment for control of smuts and bunts (particularly loose smut, Ustilago spp.), at 50-200 g/100 kg seed, on barley, wheat ...
  67. [67]
    SDHi fungicides: An example of mitotoxic pesticides targeting the ...
    Their mode of action is based on blocking the activity of succinate dehydrogenase (SDH), a universal enzyme expressed by all species harboring mitochondria.
  68. [68]
    CARBOXIN - EXTOXNET PIP
    Carboxin is a systemic anilide fungicide. It is used as a seed treatment for control of smut, rot, and blight on barley, oats, rice, cotton, vegetables, corn, ...
  69. [69]
    photostabilization of polymers
    ... polymers and plastics in everyday use. At best, the choice of suitable ... anilide (5.41) has been found to be an effective photostabilizer for.
  70. [70]
    Plastic additives sales to grow by 3.3% annually until 2023
    Mar 10, 2025 · Ceresana estimates that 36.7 million tons of plastic additives are consumed worldwide each year, growing by an average of 3.3% until 2033.<|separator|>
  71. [71]
    Complete biodegradation of fungicide carboxin and its metabolite ...
    Nov 27, 2023 · This study provides an efficient strain resource for the bioremediation of carboxin and aniline in contaminated soil, and further revealing the ...Missing: EPA classification
  72. [72]
    Carboxin (Ref: D 735) - AERU - University of Hertfordshire
    An anilide fungicide for bunts and smuts normally that is normally used as a seed treatment · Smut; Rots; Bunt; Blight; Leaf stripe; Net blotch · Rice; Vegetables ...
  73. [73]
    [PDF] 090201, 090202 MEMORANDUM DP Barco - Regulations.gov
    May 29, 2019 · The fungicides carboxin and oxycarboxin are both classified as moderately toxic to freshwater fish (and to aquatic-phase amphibians for which ...
  74. [74]
    Antifebrin | National Museum of American History
    In 1886, Drs. Cahn and Hepp discovered acetanilide's ability to reduce fever and relieve pain, branding their new discovery Antefebrin.
  75. [75]
    How is acetanilide produced? What are its uses? - ChemicalBook
    Nov 2, 2024 · It is used industrially as a rubber vulcanization accelerator, a stabilizer for fiber resin coatings, a stabilizer for hydrogen peroxide, and ...Missing: iodine | Show results with:iodine
  76. [76]
    Acetanilide – Knowledge and References - Taylor & Francis
    It is used as an intermediate for the synthesis of rubber accelerators, dyes and dye intermediate, and camphor. It is also used as a precursor in penicillin ...Missing: iodine | Show results with:iodine
  77. [77]
    Carboxin - an overview | ScienceDirect Topics
    Carboxin is a systemic fungicide used as a seed treatment and to prevent seedling diseases of cereals and many other crops. It is often formulated in ...
  78. [78]
    [PDF] 31. SYSTEMIC FUNGICIDES – Benomyl, carboxin, oxycarboxin
    Carboxin can be formulated with other fungicides like thiram, copper oxine. Oxycarboxin has systemic action against rusts of cerals, and vegetables and seed ...Missing: hydroxy | Show results with:hydroxy
  79. [79]
    [PDF] Carboxin and Oxycarboxin Interim Registration Review Decision ...
    Mar 19, 2021 · This document is the Environmental Protection Agency's (EPA or the Agency) Interim. Registration Review Decision (ID) for carboxin (PC Code ...Missing: persistence | Show results with:persistence