Fact-checked by Grok 2 weeks ago

Organic compound

An organic compound is a member of a vast class of chemical substances characterized by the presence of carbon atoms covalently bonded to atoms, often in combination with other elements such as oxygen, , , , , or , forming stable chains, rings, and complex structures essential to and . These compounds, numbering over 20 million known variants, exclude simple carbon-containing substances like , carbonates, carbides, and cyanides, which are classified as inorganic. According to the International Union of Pure and Applied Chemistry (IUPAC), any structure containing at least one carbon atom qualifies as for purposes, highlighting carbon's unique tetravalency and ability that enables diverse molecular architectures. The historical development of organic chemistry began in the late amid the doctrine, which asserted that organic compounds required a mystical "vital force" from living organisms and could not be synthesized from inorganic materials. This view was decisively refuted in 1828 by German chemist , who heated —an inorganic salt—to produce , the first organic compound synthesized without biological intervention, thus bridging organic and and ushering in the era of synthetic organic chemistry. Subsequent advancements, such as August Kekulé's 1858 proposal of carbon's quadrivalent bonding and 1865 benzene ring structure, along with Jacobus van 't Hoff's 1874 tetrahedral carbon model, laid the foundations for understanding molecular and reactivity. Organic compounds are classified primarily by functional groups—specific arrangements of atoms responsible for characteristic chemical behaviors—into categories such as hydrocarbons (e.g., alkanes like , CH₄), alcohols (e.g., , C₂H₅OH), carboxylic acids (e.g., acetic acid, CH₃COOH), and amines, among others. This diversity arises from carbon's capacity for (forming long chains) and isomerism (multiple structures with the same formula), resulting in compounds ranging from simple gases to complex biomolecules like proteins and DNA. Their significance permeates daily life and science: organic compounds form the basis of all known biological processes, serving as fuels (e.g., ), pharmaceuticals (e.g., aspirin), polymers (e.g., plastics like ), and agrochemicals (e.g., pesticides). In , they enable targeting diseases; in , they enhance crop yields; and in , they drive innovations in sustainable alternatives to petroleum-based products. Ongoing research emphasizes green synthesis to address environmental impacts, underscoring organic chemistry's role in advancing , , and .

Fundamentals

Definition and Scope

Organic compounds are chemical compounds containing carbon atoms, typically covalently bonded to , oxygen, , , or other elements, but excluding simple carbon-containing substances such as carbon oxides (e.g., CO and CO₂), carbonates, cyanides, carbides, and elemental allotropes like and . This exclusion reflects the focus of on molecules with carbon-carbon (C-C) and carbon-hydrogen (C-H) bonds, which allow for diverse, stable structures. A defining feature of carbon in these compounds is , its capacity to form long chains, rings, or branched networks through strong covalent bonds with itself, enabling the vast array of organic molecules observed in nature and synthesis. The scope of organic chemistry primarily covers hydrocarbons—compounds composed exclusively of carbon and —and their derivatives, in which some hydrogen atoms are substituted by heteroatoms such as oxygen, , , , or . These heteroatoms introduce functional groups that impart specific reactivity and properties to the molecules. Organometallic compounds, which feature direct carbon-to-metal bonds, occupy a position between and inorganic realms, often studied in both fields due to their hybrid nature. This contemporary definition evolved from the discrediting of —the outdated belief that organic compounds required a vital force from living organisms—with Friedrich Wöhler's synthesis of from inorganic serving as a pivotal demonstration that such molecules could arise through standard chemical processes. Certain ambiguities persist in classification; for example, CO₂ is deemed inorganic as a simple gaseous lacking C-H bonds, while , despite its extended C-C network, is an elemental allotrope without the heteroatom diversity typical of .

Key Characteristics

Organic compounds are predominantly characterized by covalent , where carbon atoms form strong, directional bonds with other , primarily through sharing of electrons. This tetravalency of carbon arises from its four electrons, enabling it to form up to four covalent bonds. The of these bonds is influenced by hybridization: in saturated compounds, carbon adopts sp³ hybridization, resulting in a tetrahedral arrangement with bond angles of approximately 109.5°, as seen in (CH₄). In compounds with double bonds, such as alkenes, carbon exhibits sp² hybridization, leading to a trigonal planar with 120° bond angles, while triple bonds in alkynes involve sp hybridization, producing a linear structure with 180° angles. A defining trait is carbon's exceptional ability to undergo catenation, the self-linking of carbon atoms via covalent bonds to form extended chains or rings. This property stems from the high strength of carbon-carbon bonds (approximately 348 kJ/mol), allowing stable structures ranging from straight chains to complex branched or cyclic forms. Catenation underpins the structural versatility of organic molecules, enabling the construction of polymers and biomolecules like proteins and DNA. Isomerism further amplifies this versatility, as organic compounds with identical molecular formulas can exhibit different arrangements of atoms, leading to distinct physical and chemical properties. Structural isomers differ in connectivity, such as chain isomers; for example, butane (CH₃CH₂CH₂CH₃) features a straight chain, while isobutane ((CH₃)₂CHCH₃) has a branched structure. Stereoisomers involve the same connectivity but vary in spatial orientation, including enantiomers (mirror images) and diastereomers, while conformational isomers arise from rotation around single bonds, like the staggered and eclipsed forms of ethane. These phenomena contribute to the functional diversity observed in nature and synthesis. The molecular diversity of organic compounds is immense, with over 20 million known structures cataloged, as of 2025, far exceeding the number of inorganic compounds. This vast array results from variable lengths, , and the combinatorial possibilities of and isomerism, allowing for an estimated chemical space of up to 10⁶⁰ possible small molecules. Compared to inorganic compounds, organic molecules typically exhibit lower and points due to weaker intermolecular forces like van der Waals interactions rather than strong ionic or metallic bonds; for instance, boils at -161°C, while melts at 801°C. Solubility in organic compounds is governed by , following the principle that "like dissolves like," where polar solutes dissolve in polar solvents through dipole-dipole interactions or , and nonpolar solutes dissolve in nonpolar solvents via dispersion forces. For example, (polar due to its hydroxyl group) is miscible with , but (nonpolar) is not, instead dissolving in nonpolar organic solvents like . This rule influences the behavior of organic compounds in biological and industrial contexts.

Historical Context

Vitalism and Early Concepts

The concept of in chemistry emerged from longstanding philosophical traditions that distinguished living matter from inanimate substances. Rooted in Aristotelian philosophy, which viewed materials as instruments of the soul within living bodies, vitalism posited that life possessed an inherent, non-material principle animating it. Aristotle's ideas influenced early modern thought by emphasizing a formative "soul" or vital essence that transformed primordial substances into living forms, setting the stage for later chemical interpretations. In the , Lavoisier's advancements in provided empirical tools for studying substances but were primarily applied to inorganic compounds, reinforcing the perceived boundary between lifeless minerals and the products of life. Early chemists observed a clear distinction between compounds derived from living organisms and those from mineral sources, attributing the former's complexity to biological origins. Substances like sugars extracted from plants and proteins isolated from animal tissues were seen as uniquely tied to vital processes, incapable of replication outside living systems. For instance, Carl Wilhelm Scheele's isolation of tartaric and citric acids from fruits in the 1770s, and Michel Eugène Chevreul's separation of fatty acids from animal fats in the early 1800s, highlighted these materials' intricate structures, which seemed beyond artificial imitation. This empirical divide fueled the belief that organic compounds embodied a special essence absent in inorganic matter. In 1815, Swedish chemist Jöns Jacob Berzelius formalized vitalism in chemistry through his "vital force" (Lebenskraft) theory, asserting that organic substances required an external life force to form and could not be synthesized from inorganic precursors. Berzelius, who had coined the term "organic chemistry" in 1806 to describe compounds from biological sources, maintained this stance despite growing evidence of elemental similarities between organic and inorganic materials, viewing the vital force as a regulative principle beyond physical laws. He staunchly opposed attempts at synthesis, dismissing proposals by contemporaries like Jean-Baptiste Dumas as incompatible with the theory's core tenet that life alone could arrange atoms into organic forms. Berzelius's influence extended to critiquing early synthetic efforts, such as Joseph Louis Gay-Lussac's investigations into fulminates and cyanates in the 1820s, which Berzelius argued failed to bridge the vital-inorganic gap despite producing complex products. These attempts, often starting from borderline materials, underscored the era's challenges and Berzelius's insistence on the impossibility of lab-based creation without vital intervention. This doctrine persisted until Friedrich Wöhler's 1828 of from inorganic salts challenged its foundations.

Emergence of Modern Organic Chemistry

The pivotal moment in the emergence of modern organic chemistry occurred in 1828 when Friedrich Wöhler synthesized urea from inorganic ammonium cyanate, demonstrating that organic compounds could be produced without a vital force. By heating ammonium cyanate (\ce{NH4OCN}), Wöhler obtained urea (\ce{(NH2)2CO}), a substance previously thought to require biological processes for its formation.60740-X/fulltext) This experiment, detailed in his publication in Annalen der Physik und Chemie, challenged the doctrine of vitalism and laid the groundwork for viewing organic chemistry as a branch of general chemistry amenable to laboratory synthesis. In the following decades, key advancements solidified this shift. advanced the field in the 1840s through his work on agricultural and plant chemistry, emphasizing the role of mineral nutrients in organic processes and applying analytical methods to study plant metabolism.62112-5/fulltext) proposed the cyclic structure of in 1865, envisioning a six-carbon ring with alternating double bonds to explain its stability and reactivity, as outlined in his seminal paper in the Bulletin de la Société Chimique de . Emil Fischer's investigations into sugars during the 1890s further expanded structural understanding; he achieved the synthesis of glucose, , and from , elucidating their configurations and through systematic degradation and synthesis. Institutional developments paralleled these discoveries, fostering a systematic approach. Charles Gerhardt published one of the first comprehensive textbooks, Précis de Chimie Organique, in the , which organized compounds by and equivalents, influencing pedagogical standards. The rise of structural theory culminated in Jacobus Henricus van 't Hoff's 1874 proposal of the tetrahedral carbon atom in La Chimie dans l'Espace, providing a spatial model that explained isomerism and optical activity in molecules. The 20th century saw integrate with biochemistry and . Hermann Staudinger's work in the 1920s established by demonstrating that polymers like rubber and were long-chain macromolecules rather than aggregates, as argued in his 1920 paper in Berichte der Deutschen Chemischen Gesellschaft. By the 1950s, insights into biochemical organics advanced rapidly, with the elucidation of DNA's double-helix structure by and in 1953 revealing the molecular basis of genetic information storage. These milestones transformed into a foundational discipline for understanding life's molecular machinery and synthetic innovations.

Nomenclature

Systematic Naming Conventions

The systematic naming of organic compounds follows the principles established by the International Union of Pure and Applied Chemistry (IUPAC), primarily through substitutive , which generates unambiguous names by identifying a parent hydride chain and modifying it with prefixes and suffixes for substituents and s. The core process begins with selecting the longest continuous carbon chain as the parent hydride, ensuring it includes the principal ; for example, in unbranched alkanes, this yields names like for the five-carbon chain CH₃CH₂CH₂CH₂CH₃. Substituents are then numbered along the chain to assign the lowest possible locants, starting from the end that gives priority to the principal characteristic group or, if none, to multiple bonds or substituents, with sets of locants compared term by term for minimality. Functional group priority dictates the choice of suffix for the parent chain, following a hierarchical order where higher-ranking groups receive the suffix while lower ones are treated as prefixes; carboxylic acids, for instance, take precedence over alcohols, resulting in names like 2-hydroxypropanoic acid for the compound with both a carboxylic acid and a hydroxyl group on a three-carbon chain (commonly known as lactic acid). Specific suffixes include -oic acid for carboxylic acids (with the carbonyl carbon as position 1), -ol for alcohols (with the lowest locant for the hydroxy group), and -one for ketones (with the lowest locant for the carbonyl). For hydrocarbons, alkanes end in -ane (e.g., pentane), while alkenes and alkynes incorporate -ene or -yne with locants indicating the position of the double or triple bond, such as propene for CH₃CH=CH₂ (implying the double bond at position 1) or but-1-yne for the terminal alkyne HC≡CCH₂CH₃. When multiple functional groups or unsaturations are present, composite suffixes like -en-oic acid are used, and prefixes such as hydroxy- or oxo- denote subordinate groups, always arranged in alphabetical order. The 2013 IUPAC recommendations, detailed in the , refined these rules by introducing Preferred IUPAC Names (PINs) for greater consistency in scientific and regulatory contexts, expanding multiplicative for assemblies, and providing dedicated systems for specialized classes. For polymers, structure-based identifies the constitutional repeating unit (CRU) and names it using organic substitutive rules, such as poly(1-chloroethane-1,2-diyl) for , while source-based relies on names like poly(). Isotopically labeled compounds are named by adding symbols in square brackets for mixtures (e.g., [²H]methane for deuterated ) or parentheses for fully substituted ones (e.g., (²H₄)methane), with locants and specifications for specific, selective, or uniform labeling to indicate the extent and position of isotopic modification. These updates ensure systematic names remain adaptable to emerging areas like isotopic tracing and macromolecular chemistry without relying on retained common names for precision.

Historical and Common Names

Many historical names for organic compounds derive from their natural sources or sensory properties, reflecting the early isolation of these substances from biological materials. For instance, acetic acid originates from the Latin word acetum, meaning , as it was first obtained from fermented sources like wine. Similarly, the name glucose stems from the Greek glykys, meaning sweet, due to its taste when isolated from or fruits. These etymological roots highlight how 19th-century chemists often named compounds based on their origins or characteristics rather than structural features. The International Union of Pure and Applied Chemistry (IUPAC) permits certain traditional names, known as retained names, for simplicity in general , even as systematic naming becomes standard. Examples include acetone (retained instead of propan-2-one) and acetic acid (instead of ethanoic acid), which are widely used due to their familiarity and brevity. Aspirin, systematically named 2-acetoxybenzoic acid or acetylsalicylic acid, also retains its common usage from early pharmaceutical contexts, illustrating how retained names bridge historical practice with modern conventions. These retained names are explicitly listed in IUPAC recommendations to avoid confusion while allowing practical application. Trivial names remain prevalent in pharmaceuticals and natural products, where systematic names can be cumbersome for complex structures. Ibuprofen, a common drug with the systematic name 2-(4-(2-methylpropyl)phenyl)propanoic acid, exemplifies this, as does , a retained from its discovery in gallstones. However, such names can introduce , particularly for intricate molecules like natural products, where multiple isomers or sources might share similar trivial designations, complicating without additional . This persists despite efforts to standardize, as trivial names prioritize memorability over precision in applied fields. The evolution of organic nomenclature began with ad hoc 19th-century naming practices, often inconsistent and source-based, leading to the need for IUPAC standardization. , in his 1865 work on aromatic compounds, referred to as "," derived from isolated from resins, reflecting the era's descriptive approach before structural theory advanced. This shift from informal names like to the modern "" underscores the progression toward systematic rules, initiated by early international committees in the late 1800s to resolve growing ambiguities in an expanding field.

Classification

Structural Classification

Organic compounds are primarily classified structurally based on the arrangement of their carbon atoms and the presence of functional groups, providing a systematic framework for understanding their diversity and reactivity. This classification emphasizes the carbon skeleton as the foundational structure, upon which heteroatoms and functional groups are attached, influencing the compound's chemical behavior. The International Union of Pure and Applied Chemistry (IUPAC) endorses this approach in its nomenclature guidelines, prioritizing the carbon framework and substituent priorities for naming and categorization. The carbon skeleton forms the backbone of organic molecules and is categorized into acyclic, cyclic, and polycyclic types. Acyclic compounds, also known as open-chain or aliphatic compounds, feature carbon atoms arranged in straight or branched chains without rings, such as (C₂H₆) or . Cyclic compounds contain one or more rings; monocyclic examples include (C₆H₁₂), a saturated six-carbon ring, while unsaturated cyclic structures like (C₆H₆) exhibit aromatic stability. Polycyclic compounds involve multiple fused or bridged rings, exemplified by (C₁₀H₈), which consists of two fused rings and is classified as a . This skeletal classification highlights how ring formation increases molecular rigidity and alters electronic properties compared to linear chains. Functional groups provide a hierarchical overlay on the carbon skeleton, determining the compound's class and reactivity; hydrocarbons form the base, followed by those incorporating oxygen, , and other heteroatoms. Hydrocarbons are divided into alkanes (saturated, general C_nH_{2n+2}, e.g., CH₄), alkenes (with one , C_nH_{2n}, e.g., ethene C₂H₄), alkynes (, C_nH_{2n-2}, e.g., ethyne C₂H₂), and aromatics. Oxygen-containing compounds include alcohols (R-OH, e.g., CH₃CH₂OH) and ethers (R-O-R', e.g., CH₃CH₂OCH₂CH₃), which introduce polarity and hydrogen bonding capabilities. Nitrogen-containing classes encompass amines (R-NH₂, e.g., CH₃NH₂), classified as primary, secondary, or tertiary based on substitution. This prioritizes the highest-ranking for , as per IUPAC rules, ensuring consistent classification across diverse structures. Heterocyclic compounds represent a specialized structural class where rings incorporate one or more heteroatoms (e.g., , oxygen, ) replacing carbon, combining cyclic skeleton features with heteroatom reactivity. Common examples include (C₅H₅N), a six-membered ring with , which mimics benzene's but adds basic properties due to the on . Heterocycles are subcategorized by and heteroatom type, such as five-membered (oxygen) or six-membered (two nitrogens). Their prevalence in pharmaceuticals underscores their biological importance; over 85% of all biologically active compounds are heterocycles or comprise a heterocycle, enabling targeted interactions with enzymes and receptors through tuned electronics and sterics. Seminal studies highlight heterocycles' role in antiviral and anticancer agents, as in the rings of analogs. Stereochemistry integrates into structural classification by accounting for spatial arrangements around chiral centers, particularly in compounds with tetrahedral carbons bearing four different substituents. A chiral center imparts optical activity, leading to enantiomers distinguishable by the rules, which assign (rectus, clockwise) or S (sinister, counterclockwise) configurations based on priorities. In , the alpha carbon is typically chiral (except ), with natural L-amino acids corresponding to the S configuration for most (e.g., L-alanine is (S)-2-aminopropanoic acid), influencing and drug . This stereochemical layer refines classification, as diastereomers and meso compounds arise from multiple chiral centers, impacting biological recognition without altering the core skeleton or functional groups.

Origin-Based Classification

Organic compounds can be classified based on their , which provides a practical framework for understanding their production sources and implications in fields like , , and . This classification divides them into , synthetic, and biotechnologically derived categories, with notable overlaps due to advances in chemical and . organic compounds are those produced by living organisms, including , animals, and microorganisms, often as secondary metabolites serving ecological roles such as defense or attraction. For instance, alkaloids like are extracted from plant sources such as beans ( species), where they occur as derivatives contributing to the plant's mechanisms. , another prominent class, form the primary constituents of essential oils in like fruits or , with examples including from peels, which imparts characteristic aromas and potential properties. These compounds highlight the of natural sources, where many feature heterocyclic structures adapted over evolutionary time. Synthetic organic compounds are entirely lab-produced through chemical reactions, enabling the creation of materials and drugs not readily available from nature. , a , exemplifies this through of monomers like and , yielding strong fibers used in textiles and applications. In pharmaceuticals, synthetic derivatives of natural scaffolds, such as from penicillin, are manufactured via targeted modifications to enhance stability and efficacy against bacterial infections. Biotechnologically derived organic compounds bridge natural and synthetic realms, utilizing engineered biological systems like enzymes or genetically modified organisms for production. Recombinant human insulin, first achieved in 1978 by inserting synthetic human insulin genes into bacteria, revolutionized treatment by enabling scalable, animal-free manufacturing. By the 2020s, CRISPR-Cas9 gene editing has facilitated semi-synthetic organics, such as enhanced specialized metabolites in edited , improving yield and nutritional profiles without traditional . Overlaps and ambiguities arise in origin classification, particularly with bioidentical synthetics that chemically mirror natural counterparts. Synthetic (L-ascorbic acid) is identical in structure and to its plant-derived form, allowing production to meet global demands while raising questions about "natural" labeling. Environmental organics, like hydrocarbons, originate from ancient biological remains—primarily and transformed under geological pressure—blurring lines between natural and abiotic processes, as these fossil-derived alkanes serve as feedstocks for further . This classification aids in regulatory contexts, such as patenting biotech innovations or assessing environmental impacts, distinct from purely structural categorizations.

Properties

Physical Properties

Organic compounds exhibit a wide range of physical states at , primarily determined by their molecular weight and intermolecular forces. Low-molecular-weight organic compounds, such as and , are typically gases or liquids, while higher-molecular-weight ones like waxes and polymers tend to be solids. For instance, most hydrocarbons with fewer than five carbon atoms are gases, those with five to seventeen carbons are liquids, and longer chains form solids. Solubility in varies significantly with molecular . Hydrocarbons, being nonpolar, are generally hydrophobic and insoluble in due to weak interactions with polar molecules. In contrast, polar compounds like alcohols exhibit greater ; , for example, is completely miscible with owing to hydrogen bonding between its hydroxyl group and molecules. Melting and boiling points of organic compounds increase with molecular weight and chain length because of stronger van der Waals forces. For alkanes, this trend is evident: methane has a boiling point of -161°C, while decane boils at 174°C. Branching in the molecular structure reduces these points by decreasing surface area and intermolecular interactions; n-pentane boils at 36.1°C, compared to 27.7°C for isopentane. Most organic compounds melt below 200°C, though some decompose prior to melting. Optical properties arise from molecular structure, particularly conjugation and . Extended conjugation in systems like beta-carotene, with eleven double bonds, lowers the energy gap between molecular orbitals, absorbing visible in the blue-green region and appearing orange-red. Chiral organic compounds, lacking a plane of , rotate plane-polarized , a phenomenon known as optical activity; for example, enantiomers of chiral molecules exhibit equal but opposite rotations. Organic compounds are generally less dense than water, with densities ranging from 0.7 to 0.99 g/cm³ for many liquids like hydrocarbons and oils, causing them to float on . Viscosity, a measure of resistance to flow, varies widely; nonpolar hydrocarbons like have low viscosity similar to , while more viscous oils and syrups result from higher molecular weights or stronger intermolecular forces.

Chemical Reactivity

Organic compounds exhibit a wide range of chemical reactivity primarily driven by the formation and breaking of covalent bonds, particularly those involving carbon-carbon and carbon-heteroatom linkages. This reactivity is governed by the functional groups present, which dictate specific patterns such as , , and elimination reactions. These processes often proceed through mechanisms that involve , , or intermediates, influenced by electronic and steric factors. Addition reactions are characteristic of unsaturated compounds like s, where electrophiles add across the carbon-carbon . For instance, the addition of (HBr) to an follows , in which the attaches to the carbon with more hydrogens, leading to the more stable . reactions, common in alkyl halides, occur via SN1 or SN2 pathways; SN2 involves a concerted backside attack by a on primary or secondary halides, resulting in inversion of , while SN1 proceeds through a for halides, allowing . Elimination reactions, such as E1 and E2, convert saturated compounds to s by removing a and a beta-; E2 is a bimolecular process favored by strong bases and anti-periplanar geometry, whereas E1 is unimolecular and occurs under solvolytic conditions with formation. The acidity and basicity of organic compounds are determined by the stability of conjugate acids or bases, often quantified by values. Carboxylic acids, with values around 5, are significantly more acidic than alcohols ( 15-18) due to resonance stabilization of the carboxylate anion, where the negative charge is delocalized over two oxygen atoms. This effect enhances the willingness of the proton to dissociate, contrasting with alcohols where the alkoxide ion lacks such delocalization. Oxidation and reduction reactions frequently target carbonyl groups; for example, aldehydes can be oxidized to carboxylic acids using (KMnO4), a process that involves the addition of oxygen or removal of hydrogen equivalents. Stability factors play a crucial role in dictating reactivity trends across functional groups. , the delocalization of electrons from adjacent C-H bonds into an empty p-orbital, stabilizes carbocations, with tertiary carbocations being more stable than secondary or primary due to more available alkyl groups. Inductive effects, where electron-withdrawing or donating groups influence charge distribution through bonds, further modulate ; for instance, electron-donating alkyl groups stabilize positive charges inductively. A general for functional groups in nucleophilic substitutions ranks acid chlorides and anhydrides as highly reactive, followed by esters and amides, reflecting the ability and electrophilicity of the carbonyl carbon. Physical properties, such as in polar solvents, can influence reaction rates by facilitating approach in substitution processes.

Synthesis and Reactions

Synthetic Methods

Synthetic methods in encompass a range of strategies for assembling complex molecules from simpler precursors, emphasizing the formation of carbon-carbon bonds, introduction of functional groups, and control over and selectivity. These approaches have evolved from classical techniques to modern catalytic and continuous processes, enabling the efficient construction of diverse compounds essential for pharmaceuticals, materials, and synthesis. Key methods prioritize versatility, mild conditions, and high yields to facilitate multi-step syntheses while minimizing side reactions. Carbon-carbon bond formation is fundamental to building molecular skeletons, with organometallic reagents playing a central role. The , developed by in 1900, involves the of an organomagnesium (RMgX) to a carbonyl compound, such as an or , yielding a secondary or tertiary after acidic workup; for example, RMgX reacts with R'₂C=O to produce R'₂C(OH)R. This method is widely used due to its broad substrate scope and operational simplicity, though it requires anhydrous conditions to avoid of the . Complementing this, the , introduced by Georg Wittig in 1954, converts carbonyls to alkenes via reaction with a phosphonium ylide, such as Ph₃P=CH₂ + R₂C=O → R₂C=CH₂ + Ph₃P=O, providing stereoselective access to olefins critical for chain extension in . Both reactions exemplify how organometallics enable precise connectivity in organic frameworks. Functionalization strategies introduce heteroatoms or modify existing groups to enhance reactivity or solubility. Free radical halogenation of alkanes, typically under UV light or heat, substitutes a hydrogen with a halogen via a chain mechanism: initiation by Cl₂ homolysis, propagation through H abstraction (e.g., CH₄ + Cl• → CH₃• + HCl) and Cl addition (CH₃• + Cl₂ → CH₃Cl + Cl•), and termination; chlorination of methane (CH₄ + Cl₂ → CH₃Cl) illustrates this for preparing alkyl halides as synthetic intermediates, though selectivity is low for complex alkanes. In contrast, hydroboration-oxidation of alkenes achieves anti-Markovnikov hydration with syn stereochemistry: BH₃ adds across the double bond (B to less substituted carbon), followed by H₂O₂/OH⁻ oxidation to yield the alcohol, providing regioselective access to primary alcohols from terminal alkenes without carbocation rearrangements. These methods expand the functional diversity of hydrocarbons. Protecting groups are indispensable in multi-step syntheses to mask reactive functionalities temporarily, preventing unwanted side reactions. For instance, carbonyl groups can be protected as acetals under acidic conditions with alcohols, which are stable to bases and nucleophiles but cleaved under mild acidic , allowing selective manipulation of other sites in polyfunctional molecules. This strategy, detailed in comprehensive references, ensures in complex assemblies by enabling sequential deprotection without affecting the core structure. Modern synthetic methods address efficiency, scalability, and chirality. Asymmetric synthesis, such as the developed in the 1980s, uses titanium tartrate catalysts with t-BuOOH to convert allylic alcohols to epoxy alcohols with high enantioselectivity (>90% ee), revolutionizing the preparation of chiral building blocks for drugs and natural products; this work earned K. Barry Sharpless the 2001 for chirally catalyzed oxidations. chemistry, a continuous processing technique, enhances scalability by pumping reagents through microreactors, improving heat/mass transfer and safety for exothermic reactions, as seen in multistep syntheses yielding pharmaceuticals with reduced waste and higher throughput compared to batch methods. These innovations underscore the shift toward sustainable, precise .

Functional Group Transformations

Functional group transformations involve the selective conversion of one functional group into another, enabling the construction of complex organic molecules from simpler precursors in synthetic routes. These reactions exploit the inherent reactivity of functional groups, often under mild conditions, to achieve high yields and . Common transformations include interconversions between alcohols, carbonyl compounds, amines, and esters, which are foundational in for building carbon skeletons and introducing heteroatoms. A key transformation converts alcohols to alkyl halides, which serves as a versatile for further substitutions. Thionyl chloride (SOCl₂) is widely used for this purpose, reacting with primary or secondary alcohols to form alkyl s while releasing SO₂ and HCl as byproducts. The proceeds via formation of a chlorosulfite , followed by , minimizing carbocation rearrangements in secondary alcohols when conducted with . For example, reacts with SOCl₂ to yield quantitatively under conditions. This method is preferred over alternatives like HCl or PCl₃ due to its cleaner byproduct profile and applicability to sensitive substrates. Carbonyl compounds undergo reductions to alcohols using mild agents like (NaBH₄), which selectively reduces aldehydes and ketones without affecting esters or carboxylic acids. NaBH₄ delivers hydride to the carbonyl carbon, forming a tetrahedral intermediate that protonates upon to yield primary (from aldehydes, RCHO → RCH₂OH) or secondary alcohols (from ketones, RCOR' → RCH(OH)R'). This operates in protic solvents like at , achieving near-quantitative yields for aliphatic and aromatic carbonyls, as demonstrated in early studies on its scope. Conversely, primary alcohols can be oxidized to aldehydes using (PCC), a chromium(VI)-based oxidant that halts at the aldehyde stage by avoiding over-oxidation to carboxylic acids. PCC, prepared from , HCl, and CrO₃, effects this transformation in , with converting to in 90-95% yield, preserving acid-sensitive groups. Amine synthesis often employs or specialized alkylations. Catalytic hydrogenation reduces nitro groups (RNO₂) to primary (RNH₂) using H₂ and (Pd/C), proceeding through and intermediates under mild pressure (1-3 atm) and temperature (25-50°C). This method is highly selective for aromatic nitroarenes, yielding anilines in >95% efficiency, and tolerates halides or esters. Alternatively, the provides a route to primary alkyl from alkyl halides, involving of to form its , which undergoes SN₂ (e.g., with ) to give an N-alkylphthalimide. or hydrazinolysis then cleaves the phthalimide, liberating the primary while avoiding over-alkylation common in direct . This classical approach, yielding 70-90% for unhindered primaries, remains valuable for pharmaceutical intermediates. Ester hydrolysis cleaves the acyl-oxygen bond to generate carboxylic s and alcohols, catalyzed by or . -catalyzed hydrolysis, using HCl or H₂SO₄ in aqueous media, protonates the carbonyl oxygen to facilitate nucleophilic attack by , following an AAc2 mechanism reversible under equilibrium conditions; for , rates increase with concentration, achieving 80-90% conversion in refluxing conditions. -catalyzed hydrolysis, or , employs NaOH or KOH to form the tetrahedral intermediate irreversibly, as the product is deprotonated and insoluble in for salts. This process is pivotal in production, where triglycerides (fatty acid esters) react with alkali to yield and fatty salts, with industrial yields exceeding 95% for tallow-based soaps under heating. values quantify ester content in fats, guiding formulation for cleansing agents.

Analysis and Determination

Spectroscopic Techniques

Nuclear magnetic resonance (NMR) spectroscopy is a cornerstone technique for determining the of organic compounds by exploiting the magnetic properties of atomic nuclei, particularly hydrogen-1 (^1H) and (^13C). In ^1H NMR, the provides information on the electronic environment of protons, with typical values for protons appearing around 0.9 due to their shielding in non-polar environments. Coupling constants, measured in hertz (Hz), reveal connectivity between neighboring protons through scalar couplings, such as the vicinal coupling (^3J) in ethyl groups showing a characteristic quartet and triplet pattern. For ^13C NMR, the broader range of s—spanning over 200 —allows distinction of carbon types; for instance, carbonyl carbons resonate around 200 , reflecting their deshielded, electron-deficient nature. These spectra, often acquired in deuterated solvents like CDCl_3, enable the assignment of carbon skeletons and are essential for verifying molecular connectivity in complex organics. Infrared (IR) spectroscopy identifies s in organic s by measuring the of corresponding to molecular vibrations, particularly bond stretches. The O-H stretch in alcohols and carboxylic acids appears as a broad band between 3200 and 3600 cm^{-1}, arising from hydrogen bonding that broadens the peak. The C=O stretch, a sharp and intense band near 1700 cm^{-1}, is characteristic of carbonyls in ketones, aldehydes, and esters, with slight variations depending on conjugation or . IR is particularly valuable for purity checks, as impurities introduce additional peaks in the functional group region (4000–1500 cm^{-1}), while the (1500–600 cm^{-1}) provides a unique pattern for compound identification. Mass spectrometry (MS) determines the molecular weight and fragmentation patterns of organic compounds by ionizing molecules and analyzing the (m/z) of resulting . The molecular (M^+) corresponds to the intact ionized , offering direct evidence of the molecular formula, often confirmed by high-resolution MS for exact mass. Fragmentation patterns reveal structural details; for example, the McLafferty rearrangement in carbonyl-containing compounds like aldehydes produces a prominent at m/z 44 (from CH_2=CH-OH^+•), indicating the presence of a gamma-hydrogen. Common ionization methods like electron impact () generate characteristic fragments, aiding in the elucidation of functional groups and skeletal arrangements without destroying the sample bulk. Ultraviolet-visible (UV-Vis) detects conjugated systems and chromophores in organic compounds by measuring in the 200–800 range, corresponding to π→π^* and n→π^* electronic . In , the λ_max for the primary band is approximately 255 , attributed to the ^1B_{2u} ← ^1A_{1g} in its conjugated π-system. Extended conjugation, as in polyenes or aromatic dyes, shifts λ_max to longer wavelengths (bathochromic shift) and increases molar absorptivity (ε), quantifying the extent of delocalization; for instance, styrene shows λ_max at 244 compared to . This technique is crucial for studying and unsaturation in natural products and synthetic dyes.

Structural Elucidation Methods

X-ray crystallography is a pivotal technique for determining the three-dimensional atomic structure of organic compounds in the crystalline state, yielding precise bond lengths, angles, and stereochemical details that are crucial for understanding molecular architecture. This method relies on the diffraction of X-rays by electrons in a single crystal, producing a pattern that is analyzed to reconstruct the electron density map and atomic positions. It is particularly valuable for complex molecules where solution-based methods fall short, though it requires the growth of suitable crystals, which can be challenging for some organic substances. A landmark application was the 1951 determination of the absolute configuration of sodium rubidium (+)-tartrate, the first such determination using anomalous X-ray scattering to distinguish enantiomers. Another iconic example is the 1953 model of the DNA double helix, derived from X-ray fiber diffraction data that revealed helical parameters and base pairing in this biomolecular organic polymer. Chromatographic methods play a complementary role in structural elucidation by enabling the separation and identification of organic compounds within complex mixtures, often coupled with detection systems for enhanced characterization. Gas chromatography-mass spectrometry (GC-MS) is ideal for volatile and semi-volatile organics, where separation occurs based on boiling point and polarity, followed by mass spectrometric fragmentation that provides molecular weight and structural fragments for identification. For instance, electron ionization in GC-MS generates characteristic mass spectra that allow comparison to libraries for confirming structures of unknowns like hydrocarbons or pharmaceuticals. High-performance liquid chromatography (HPLC), suited for non-volatile or polar compounds, separates analytes on the basis of interactions with a stationary phase and can be interfaced with mass spectrometry (HPLC-MS) or UV detectors to assess purity and infer structural features through retention times and spectral data. These techniques ensure sample integrity prior to more definitive analyses, with GC-MS often resolving structures in environmental or forensic samples. Computational modeling, particularly density functional theory (DFT), offers a powerful approach to predict and refine organic compound structures when experimental constraints limit direct measurement, focusing on quantum mechanical calculations of electronic structure and geometry optimization. DFT approximates the many-electron problem using electron density functionals to compute ground-state properties, identifying energy minima that correspond to stable conformations. For organic molecules, it excels in predicting bond lengths, angles, and torsional preferences, as seen in studies of reaction intermediates or conformers where crystal data is unavailable. These models are routinely applied to large datasets of organic compounds, providing theoretical spectra or geometries that guide experimental design and validate hypotheses about reactivity or stability. When integrated with sparse experimental inputs, such as from diffraction, DFT enhances accuracy for flexible or amorphous organics. Elemental analysis through combustion provides foundational data for structural elucidation by quantifying the percentages of key elements—primarily carbon, hydrogen, nitrogen, and oxygen—in organic compounds, enabling the derivation of from mass ratios. In this classical method, a sample is combusted in excess oxygen, converting carbon to CO₂ and hydrogen to H₂O, which are trapped and weighed to calculate their contributions. is often determined via the or Kjeldahl digestion, while oxygen is found by difference after accounting for other elements. For example, a compound yielding 3.99 g CO₂ and 1.63 g H₂O from 2.00 g sample indicates an empirical formula of C₂H₄O, establishing the elemental skeleton before advanced techniques refine the full structure. Modern variants use automated analyzers for precise determinations, essential for confirming purity and composition in synthetic organics. These methods often serve as inputs for computational or chromatographic confirmations, ensuring comprehensive structural validation.

Resources and Applications

Chemical Databases

Chemical databases serve as essential repositories for organic compound data, enabling researchers to access, retrieve, and analyze structural, property, and reactivity information to support discovery and validation in . These resources aggregate vast datasets from experimental and computational sources, often integrating tools for searching by name, formula, or structure, which facilitates interdisciplinary applications in , , and . Key databases emphasize curated, high-quality entries to ensure reliability, with ongoing updates to incorporate new findings and advanced search capabilities. PubChem, maintained by the (NIH), is a freely accessible database launched in 2004 that contains 119 million unique chemical compounds (as of 2025), the majority of which are . It provides comprehensive data on molecular structures, physical and chemical properties, and bioactivities, including 295 million biological test results derived from and . Users can search by structure, name, or identifier, and the database links to related patents and , aiding in the exploration of compound applications. Recent enhancements as of 2025 include improved integration with structural elucidation tools for data deposition. ChemSpider, owned by the Royal Society of Chemistry and launched in 2007, is a database with over 130 million chemical structures, primarily , aggregated from more than 270 public and proprietary sources. It supports free text and structure-based searches, incorporating spectral data such as NMR and spectra, alongside links to patents, publications, and property predictions to enhance data validation through community contributions. The platform's crowdsourcing model allows users to deposit and curate entries, ensuring dynamic updates and broad coverage of organic molecules. Reaxys, developed by and introduced in 2009, builds on the historic Beilstein database of , which originated in 1881 and was digitized starting in 1988, now encompassing experimentally validated data on over 298 million substances and over 68 million reactions (as of October 2025). It focuses on reaction retrieval and synthetic route prediction, allowing users to query multi-step transformations with and condition details, drawn from peer-reviewed and patents since the . This resource is particularly valuable for planning, with workflow tools that map retrosynthetic pathways. CAS SciFinder, provided by the (CAS), offers access to the world's largest collection of chemical information, including over 200 million organic substances and 150 million reactions indexed from journals and patents. Launched as an evolution of earlier CAS tools, it enables structure, substructure, and reaction searches with detailed experimental conditions, , and property data to support organic compound identification and synthesis design. The database's , updated regularly, assist in forecasting reaction outcomes based on historical precedents. For biomolecules, the (PDB), established in 1971 at and now managed by the Research Collaboratory for Structural Bioinformatics (RCSB), archives over 227,000 three-dimensional structures of proteins, nucleic acids, and organic cofactors (as of 2024), with annual depositions of approximately 9,500. It includes atomic coordinates derived from , NMR, and cryo-EM, linked to functional annotations for organic components like ligands. As of 2025, PDB integrates AI-driven search tools for enhanced querying of biomolecular interactions involving organic compounds.

Biological and Industrial Importance

Organic compounds form the foundation of life, serving as the primary building blocks of biomolecules essential for biological processes. Proteins, which are polymers composed of amino acids—organic molecules containing amine and carboxylic acid functional groups—perform critical roles in enzymatic catalysis, structural support, and cellular signaling. Lipids, another class of organic compounds including fats, phospholipids, and steroids, are vital for forming cell membranes, storing energy, and acting as signaling molecules. Nucleic acids, such as DNA and RNA, consist of nucleotide monomers and store genetic information, enabling heredity and protein synthesis. In , compounds are central to and development, targeting specific biological pathways to treat diseases. Statins, a of molecules derived from natural products like , inhibit to lower levels and reduce cardiovascular risk, with their discovery rooted in fungal metabolites identified in the . Antibiotics such as penicillin, an compound produced by the fungus Penicillium notatum, revolutionized infection treatment by disrupting bacterial cell wall synthesis; it was discovered by in 1928 and became widely available after 1945. Industrially, organic compounds underpin manufacturing on a massive scale, with polymers like —a simple alkene-derived chain—produced in excess of 100 million tons annually for , , and consumer goods. Fuels such as , a of alkanes ranging from C5 to C12 hydrocarbons, power internal combustion engines and derive primarily from refining. Agrochemicals, including the —an organophosphorus compound—control weeds in agriculture, with widespread use since its registration as a in 1974. Organic compounds also pose environmental challenges and drive efforts. Polychlorinated biphenyls (PCBs), synthetic organochlorine compounds used historically in electrical equipment, are persistent organic pollutants that bioaccumulate in food chains, causing in and humans, including endocrine disruption and carcinogenicity. To mitigate such impacts, principles, formalized in the 1990s by and , promote sustainable synthesis by emphasizing waste prevention, , and safer , reducing hazardous byproducts in .

References

  1. [1]
    Organic Chemistry - American Chemical Society
    Organic chemistry is the study of the structure, properties, composition, reactions, and preparation of carbon-containing compounds. Most organic compounds ...Missing: authoritative source
  2. [2]
    Organic Compound - an overview | ScienceDirect Topics
    An organic compound is defined as a chemical compound primarily formed by carbon (C), hydrogen (H), oxygen (O), and nitrogen (N), and it may also contain ...
  3. [3]
    Blue Book chapter P-1 - IUPAC nomenclature
    For nomenclature purposes, a structure containing at least one carbon atom is considered to be an organic compound and can be named using the principles of ...
  4. [4]
    [PDF] The Phlogiston Theory - Yale Chemistry
    O (organic) An example of isomers into urea Begins the downfall of Vitalism The Beginning of the End for Vitalism Page 11 I cannot, so to say, hold my chemical ...
  5. [5]
    The Chemistry of Carbon - Purdue University
    Organic chemistry is often defined as the chemistry of carbon. But this definition would include calcium carbonate (CaCO3) and graphite, which more closely ...
  6. [6]
    I—The Chemistry of Carbon, 1 Biotic and Abiotic Carbon Compounds
    Historically, chemists have referred to all compounds of carbon except the oxides, carbonates (e.g., limestone and marble), metallic carbides, and elemental ...Missing: excluding | Show results with:excluding
  7. [7]
    Organic Compound - an overview | ScienceDirect Topics
    An organic compound is defined as a chemical compound that primarily contains carbon-to-carbon or carbon-to-hydrogen bonds, such as methane or benzene, ...
  8. [8]
    Chapter 13 - Organic Chemistry - CHE 110 - Textbook
    Aug 18, 2025 · It makes fairly strong bonds with itself (a characteristic called catenation), allowing for the formation of large molecules; it also forms ...
  9. [9]
    What Do You Learn in Chemistry? | North Central College
    Dec 29, 2021 · If organic chemistry relates to organic compounds, inorganic chemistry is the opposite. ... organometallic compounds contain carbon). Work ...
  10. [10]
    Self-reinforcing Mechanisms Driving the Evolution of the Chemical ...
    Oct 1, 2023 · For example, organic chemistry has been often regarded as a field that began with Wöhler's synthesis of urea in 1828. This view has been ...
  11. [11]
  12. [12]
    Types of Hybridization in Carbon - BYJU'S
    The different types of hybridization (sp3, sp2, sp) in carbon determines the structure and reactivity of carbon compounds that are formed. · Hybridized orbitals ...
  13. [13]
    What is Catenation and Tetravalency? - GeeksforGeeks
    Nov 29, 2022 · Catenation is the property of the carbon element that allows its atoms to link together to form long carbon chains.
  14. [14]
    Types of Isomers: Constitutional, Stereoisomers, Enantiomers, and ...
    Sep 10, 2018 · For example, there is only one possible isomer for CH4 (methane), C2H6 (ethane) and propane (C3H8), and only two are possible for C4H10 (2- ...
  15. [15]
    [PDF] 12.1 The Nature of Organic molecules
    More than 18 millions organic compounds, that are known today, are classified into just a few general families on the basis of chemistry that follows a simple ...
  16. [16]
  17. [17]
    Solubility of Organic Compounds - Chemistry Steps
    And this observation is formulated in the golden rule of solubility: “like dissolves like”. This means polar compounds dissolve in polar solvents, and nonpolar ...
  18. [18]
    Origins of Organic Chemistry and Organic Synthesis - Wentrup - 2022
    Mar 23, 2022 · Although a growing number of compounds was being isolated and identified as discrete organic molecules, Berzelius was at first highly sceptical.
  19. [19]
    A Critique Of Vitalism And Its Implications For Integrative Medicine
    Aristotle's vitalistic theory proposed that living organisms consisted of a primordial substance (soul) and form, which transformed it into a specific thing.
  20. [20]
    Antoine Laurent Lavoisier The Chemical Revolution - Landmark
    Antoine-Laurent Lavoisier forever changed the practice and concepts of chemistry by forging a new series of laboratory analyses that would bring order to the ...
  21. [21]
    1.1.1: History - Chemistry LibreTexts
    Aug 7, 2022 · Jöns Jacob Berzelius, a physician by trade, first coined the term “organic chemistry” in 1806 for the study of compounds derived from biological sources.Missing: 1815 | Show results with:1815
  22. [22]
    Kekulé's Benzene Theory and the Appraisal of Scientific Theories
    In three papers published in 1865 and 1866, August Kekulé, professor of chemistry at the University of Ghent, proposed a theory of the structure of benzene ...
  23. [23]
    Emil Fischer – Biographical - NobelPrize.org
    His greatest success was his synthesis of glucose, fructose and mannose in 1890, starting from glycerol. This monumental work on the sugars, carried out between ...
  24. [24]
    (PDF) Charles Fréderic Gerhardt - ResearchGate
    Aug 9, 2025 · Abstract Charles Fréderic Gerhardt (1816-1856) was one of the most important chemists of the nineteenth century and whose researches and ...
  25. [25]
    The Young J. H. van 't Hoff: The Background to the Publication of his ...
    The Young J. H. van 't Hoff: The Background to the Publication of his 1874 Pamphlet on the Tetrahedral Carbon Atom, Together with a New English Translation.
  26. [26]
    Hermann Staudinger – Facts - NobelPrize.org
    In the early 1920s Herman Staudinger claimed they could be very large; tens or even hundreds of thousands of atoms in size. He showed how small molecules ...
  27. [27]
    The Discovery of the Double Helix, 1951-1953 | Francis Crick
    The discovery in 1953 of the double helix, the twisted-ladder structure of deoxyribonucleic acid (DNA), by James Watson and Francis Crick marked a milestone ...
  28. [28]
    [PDF] Brief Guide to the Nomenclature of Organic Chemistry - IUPAC
    IUPAC nomenclature uses substitutive nomenclature, combining a parent compound with a principal characteristic group suffix and prefixes for other groups. The ...
  29. [29]
    Blue Book | International Union of Pure and Applied Chemistry
    The rules given in the Nomenclature of Organic Chemistry, commonly known as the “Blue Book”, emphasize the generation of unambiguous names.
  30. [30]
    Brief Guide to Polymer Nomenclature
    Monomers can be named using IUPAC recommendations, or well-established traditional names. Should ambiguity arise, class names can be added.[6] For example ...
  31. [31]
  32. [32]
    Organic Nomenclature - MSU chemistry
    Organic nomenclature uses a systematic system to name compounds, indicating carbon bonding and functional groups. IUPAC rules use a root, suffix, and ...
  33. [33]
    Glucose - Etymology, Origin & Meaning
    Glucose, coined by Eugène Péligot in 1840 from French and Greek gleukos meaning "sweet wine," refers to a group of sugars or sugar-syrup derived from ...
  34. [34]
    Origin and Evolution of Organic Nomenclature - ACS Publications
    Jul 22, 2009 · Chemical nomenclature as we know it today had its beginning when Guyton de Morveau, Lavoisier, Berthollet, and Fourcroy published their Methode de Nomenclature ...Missing: historical | Show results with:historical
  35. [35]
    Benzene - American Chemical Society
    Jul 3, 2023 · In 1865, German chemist Friedrich August Kekulé published a paper in which he described benzene as consisting of a ring of six carbon atoms, ...
  36. [36]
    Naphthalene | C10H8 | CID 931 - PubChem - NIH
    Naphthalene appears as a white crystalline volatile solid with a strong coal-tar odor. The solid is denser than water and insoluble in water.
  37. [37]
    Heterocyclic Compounds - MSU chemistry
    The reactivity of partially unsaturated compounds depends on the relationship of the double bond and the heteroatom (e.g. 3,4-dihydro-2H-pyran is an enol ether) ...
  38. [38]
    Prescribed drugs containing nitrogen heterocycles: an overview - PMC
    ... pharmaceutical agents. Heterocyclic compounds, as the most important organic compounds, are frequently present in molecules of interest in medicinal chemistry.
  39. [39]
    Coffee | Linus Pauling Institute | Oregon State University
    Caffeine (1,3,7-trimethylxanthine) is a purine alkaloid that occurs naturally in coffee beans (Figure 2). ... Of note, no association was reported with non-coffee ...
  40. [40]
    Terpenes | Waterhouse Lab - UC Davis
    Terpenes are a large class of organic compounds produced by plants, and they are the main components of essential oils.Missing: examples | Show results with:examples
  41. [41]
  42. [42]
    The Pharmaceutical Collection - Ampicillin - Molecular Expressions
    Jul 2, 2018 · Ampicillin is a synthetic derivative of penicillin, a highly effective medication to treat many bacterial infections. Side effects include ...
  43. [43]
    Making, Cloning, and the Expression of Human Insulin Genes ... - NIH
    In the mid- to late 1970s, recombinant deoxyribonucleic acid methods for cloning and expressing genes in E. coli were under intense development.
  44. [44]
    Possible EU futures for CRISPR-edited plants - PubMed Central - NIH
    This article addresses the scenarios that may be encountered by the first application for pre-market approval of a CRISPR-edited plant in the EU.Missing: organic compounds
  45. [45]
    Supplemental Forms | Linus Pauling Institute | Oregon State University
    Natural and synthetic L-ascorbic acid are chemically identical, and there are no known differences in their biological activity. The possibility that the ...Natural vs. Synthetic · Mineral Ascorbates · Vitamin C with Bioflavonoids · Ester-C®Missing: bioidentical | Show results with:bioidentical
  46. [46]
    2.4: Origins of Oil and Natural Gas | PNG 301 - Dutton Institute
    Research indicates that petroleum and natural gas originated from zooplankton (single-celled plants found drifting freely in fresh or brackish water) and algae.Missing: compounds | Show results with:compounds
  47. [47]
    [PDF] Introduction to Organic Molecules - Utah Tech University
    If the boiling point of a compound is higher than 25o C, the compound will be a liquid (or a solid) at room temperature. If the boiling point is lower than 25o ...
  48. [48]
    Boiling & Melting Points - MSU chemistry
    Indeed, many of the physical characteristics of compounds that are used to identify them (e.g. boiling points, melting points and solubilities) are due to ...
  49. [49]
    Alcohols and Ethers
    Solubilities of Alcohols​​ As a result, hydrocarbons don't dissolve in water. They are often said to be immiscible (literally, "not mixable") in water. Alcohols, ...
  50. [50]
    [PDF] Synthesis and Structure of Alcohols
    Solubility of Alcohols. The hydroxyl groups in alcohols can form hydrogen bonds with water, and many low molecular weight alcohols are miscible with water.
  51. [51]
    CHE 120 - Introduction to Organic Chemistry - Textbook: Chapter 1
    Aug 18, 2025 · Although all three have the same molecular formula, they have different properties, including boiling points: pentane, 36.1°C; isopentane, 27.7° ...Alcohols, Phenols, Thiols, Ethers · Chapter 3 - Aldehydes, Ketones
  52. [52]
    Physical Properties - MSU chemistry
    Melting Points. Most organic compounds have melting points below 200 ºC. Some decompose before melting, a few sublime, but a majority undergo repeated ...
  53. [53]
    [PDF] A Brief Discussion of Color
    The more highly conjugated ß-carotene and lycopene, each having eleven conjugated double bonds absorb in the lower energy blue and green portions of the ...
  54. [54]
    Optical Activity – Organic Chemistry - Maricopa Open Digital Press
    Rotation of plane polarized light by a chiral compound is known as optical activity. Achiral compounds are optically inactive. Racemic mixtures are optically ...
  55. [55]
    Chapter: 2 Chemical and Physical Properties of Crude Oils
    Given that the density of fresh water is 1.00 g/cm3 at environmental temperatures and the densities of crude oils commonly range from 0.7 to 0.99 g/cm3 (see ...<|control11|><|separator|>
  56. [56]
    [PDF] Physical Properties of Organic Solvents - Utah Tech University
    which solvents dissolve in water, and which won't. Most organic solvents have a lower density than water, while a few are more dense than water. The density ...
  57. [57]
    Chemical Reactivity - MSU chemistry
    Functional groups are atoms or small groups of atoms (usually two to four) that exhibit a characteristic reactivity when treated with certain reagents. To view ...Missing: series | Show results with:series
  58. [58]
    Addition Reactions of Alkenes - MSU chemistry
    The oxymercuration reaction gives the product predicted by Markovnikov's rule; hydroboration on the other hand gives the "anti-Markovnikov" product.
  59. [59]
    [PDF] Chapter 7 Alkyl Halides and Nucleophilic Substitution
    For 2° alkyl halides the rates of SN1 vs. SN2 are much more competitive! • 2° alkyl halides react with good nucleophiles by SN2 reactions. • If a 2° alkyl ...
  60. [60]
    Carboxylic Acid Reactivity - MSU chemistry
    However, inductive effects also play a role. For example, alcohols have pKa's of 16 or greater but their acidity is increased by electron withdrawing ...
  61. [61]
    [PDF] 17. Oxidation and Reduction Reactions - Organic Chemistry
    We obtain the highest yields of carbonyl compound products using KMnO4 dissolved in benzene containing a Crown ether. It permits KMnO4 to dissolve in ...
  62. [62]
    [PDF] 1 Chapter 6. Alkenes: Structure and Stability Degrees of ...
    Hyperconjugation stabilizes carbocations. Thus, more substituted carbocations are more stable. vacant p-orbital. The C-H s-bond on the neighboring carbon lines ...Missing: compounds | Show results with:compounds
  63. [63]
    [PDF] The Inductive Effect in Organic Chemistry - UNL Digital Commons
    In 18, CF3 uses a short orbital to form the C-C bond, thus placing charge density away from C- and closer to CF3. Unfortunately, hyperconjugation in 11 and ...
  64. [64]
    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: source | Show results with:source
  65. [65]
    The modern interpretation of the Wittig reaction mechanism
    May 14, 2013 · In this review, we gather together the huge body of evidence that has been amassed to show that the Li salt-free Wittig reactions of non- ...
  66. [66]
    15.1: Free Radical Halogenation of Alkanes - Chemistry LibreTexts
    Jun 5, 2019 · The radical chain mechanism is characterized by three steps: initiation, propagation and termination. Initiation requires an input of energy but ...Introduction · Radical Chain Mechanism · Problems with the Chlorination...
  67. [67]
    Hydroboration-Oxidation of Alkenes - Chemistry LibreTexts
    Jun 7, 2023 · Hydroboration-oxidation is a two step pathway used to produce alcohols. The reaction proceeds in an Anti-Markovnikov manner.The Borane Complex · The Mechanism · Oxidation of the Trialkylborane...
  68. [68]
    The Nobel Prize in Chemistry 2001 - Popular information
    Many scientists have identified Sharpless' epoxidation as the most important discovery in the field of synthesis during the past few decades.
  69. [69]
    A field guide to flow chemistry for synthetic organic chemists
    Mar 15, 2023 · Flow chemistry uses a continuous stream of reagents in a reactor, offering enhanced mass transfer, safety, and efficiency compared to batch ...
  70. [70]
    Thionyl Chloride - an overview | ScienceDirect Topics
    First, a nucleophilic oxygen atom of the alcohol displaces a chloride ion from thionyl chloride to form a protonated alkyl chlorosulfite intermediate.
  71. [71]
    Investigation of Methods for Preparing Pure Secondary Alkyl Halides
    Conversion to a pure alkyl chloride in relatively low yield is possible by use of thionyl chloride and pyridine. Pure sec-alkyl bromides may be prepared in ...
  72. [72]
    Reduction of Aldehydes, Ketones and Acid Chlorides by Sodium ...
    A multi-outcome experiment: microwave-promoted reduction of unknown aldehydes and ketones. 2021, 3-12.
  73. [73]
    Sodium Borohydride - an overview | ScienceDirect Topics
    Ethanolic solutions of sodium borohydride reduce aldehydes and ketones in the presence of epoxides, esters, lactones, acids, nitriles, or nitro groups.
  74. [74]
    Stoichiometry of the oxidation of primary alcohols with pyridinium ...
    Oxidation of Aliphatic Aldehydes by Tetraethylammonium Chlorochromate: A Kinetic and Mechanistic Study. ... Oxidation Studies with Pyridiniumchlorochromate (PCC) ...
  75. [75]
    the reduction of aromatic nitro compounds to amines with hydrogen ...
    The reduction of aromatic nitro compounds to amines with hydrogen and platinum-oxide platinum black as a catalyst.
  76. [76]
    Gabriel Synthesis - an overview | ScienceDirect Topics
    The Gabriel synthesis is a classical but still useful procedure for the preparation of primary amines. The method consists of alkylation of phthalimide anion ...
  77. [77]
    The Mechanisms of Acid Catalyzed Ester Hydrolysis, Esterification ...
    Acid-catalyzed carboxylic acid esterification and ester hydrolysis ... Activation Energy of Acid-Base Catalysis. 1940, 249-271. https://doi.org ...
  78. [78]
    Catalyzed Hydrolysis - an overview | ScienceDirect Topics
    Base-catalyzed hydrolysis of ester occurs by SN2 pathway and is irreversible, because the end product of base-catalyzed hydrolysis of ester produces alcohol and ...
  79. [79]
    Saponification - an overview | ScienceDirect Topics
    Saponification is a hydrolysis reaction where free hydroxide breaks the ester bonds between the fatty acids and glycerol of a triglyceride, resulting in free ...
  80. [80]
    NMR Chemical Shifts of Trace Impurities: Industrially Preferred ...
    Feb 19, 2016 · Tables of 1H and 13C NMR chem. shifts have been compiled for common org. compds. often used as reagents or found as products or contaminants in ...
  81. [81]
    Students Constructing for Themselves the Concept of Chemical Shift ...
    Dec 20, 2023 · Using 13C{1H} NMR spectroscopy, students can learn how chirality, sterics, and electronegativity influence the no. and chem. shift of carbons.
  82. [82]
    Infrared (IR) spectroscopy | Resource - RSC Education
    Infrared stretching absorptions (A level). Simulation ... Infrared spectroscopy can be used to identify certain functional groups in an organic compound.
  83. [83]
    Examining Student Learning of Infrared Spectroscopy in Organic ...
    May 30, 2025 · Functional groups have characteristic vibrational frequencies that fall into well-defined regions in IR spectroscopy. Since there are specific ...
  84. [84]
    Mass Spectrometric Analysis. Molecular Rearrangements
    The role of surface complexes in ketene formation from fatty acids via pyrolysis over silica: from platform molecules to waste biomass.
  85. [85]
    Mass spectrometry of organic compounds. IX. McLafferty ...
    Gross. Fragmentation of Organic Ions and Interpretation of EI Mass Spectra. ... A new fragmentation pathway of cis‐1‐decalone involving a double McLafferty ...
  86. [86]
    Intramolecular hydrogen transfer in mass spectra. II. McLafferty ...
    Studies of Fragmentation and Structure of Several Organic Ions by Using MIKE Spectrometry and Thermochemistry. Journal of the Mass Spectrometry Society of ...
  87. [87]
    [PDF] Ultraviolet absorption spectra of seven substituted benzenes
    The ultraviolet absorption spectra of seven substituted benzenes have been measured in the region of 220 to 300 millimicrons.
  88. [88]
    14.8: Interpreting Ultraviolet Spectra- The Effect of Conjugation
    Sep 25, 2024 · One of the most important factors affecting the wavelength of UV absorption by a molecule is the extent of conjugation. Molecular orbital ...
  89. [89]
    The structure and UV spectroscopy of benzene–water (Bz–W 6 ...
    Sep 1, 2014 · UV spectrum of benzene shows three well-known absorption bands at 4.9, 6.20, and 6.94 eV related to three electronic excitations from the ...
  90. [90]
    X-Ray Crystallography of Chemical Compounds - PMC - NIH
    X-ray crystallography is a comprehensive technique for determining molecular structure, providing 3D parameters for drug design and determining absolute ...Missing: seminal | Show results with:seminal
  91. [91]
    [PDF] Demystifying X-ray Crystallography - stoltz2.caltech.edu
    Jan 5, 2018 · Early X-ray Crystal Structures of Organic Compounds. 1951. J. M. Bijvoet (strychnine) first time crystallography decided between proposed ...Missing: seminal | Show results with:seminal
  92. [92]
    A Structure for Deoxyribose Nucleic Acid - Nature
    The determination in 1953 of the structure of deoxyribonucleic acid (DNA), with its two entwined helices and paired organic bases, was a tour de force in ...
  93. [93]
    [PDF] Structure Determination Of Organic Compounds
    X-ray Crystallography. When a pure crystal of an organic compound can be obtained, X-ray crystallography offers the most definitive structure determination.Missing: seminal papers<|control11|><|separator|>
  94. [94]
    (PDF) Recent Applications of GC with high-resolution MS
    These features designate gas chromatography with high-resolution mass spectrometry as the first choice for identification and structure elucidation of unknown ...<|control11|><|separator|>
  95. [95]
    Advanced Chromatography Analytical Methods for the Isolation and ...
    Advanced HPLC can be coupled with mass spectrometry (MS), a combination known as HPLC-MS, which provides structural elucidation and quantification of compounds.
  96. [96]
    Techniques for Structure Elucidation of Unknowns
    Even gas chromatography (GC) could be used, as GC–MS (with EI ionization) allows comparing the spectrum you obtain from a suspect sample to those of known ...
  97. [97]
    Best‐Practice DFT Protocols for Basic Molecular Computational ...
    This work provides best‐practice guidance on the numerous methodological and technical aspects of DFT calculations in three parts.
  98. [98]
    Applying Density Functional Theory to Common Organic Mechanisms
    Nov 30, 2022 · A computational experiment investigating common organic chemistry mechanisms has been developed and implemented in a junior/senior-level physical chemistry ...
  99. [99]
    Machine learning assisted prediction of organic salt structure ... - arXiv
    Feb 14, 2024 · In this work, we demonstrate an approach to accelerate CSP by constructing machine learning (ML) models which can predict properties of DFT- ...
  100. [100]
    3.8: Determine Empirical Formula from Combustion Analysis
    Feb 11, 2023 · One of the most common ways to determine the elemental composition of an unknown hydrocarbon is an analytical procedure called combustion analysis.
  101. [101]
    Combustion Analysis - ChemTeam
    Determine the Empirical Formula Using Combustion Analysis​​ Bonus Example #1: A 2.00 g sample of a compound was found to produce 3.99 g of CO2 and 1.63 g of H2O ...
  102. [102]
    4.1 Biological Molecules – Human Biology
    There are four major classes of biological macromolecules (carbohydrates, lipids, proteins, and nucleic acids), and each is an important component of the cell.
  103. [103]
    CH103 - Chapter 8: The Major Macromolecules - Chemistry
    ... organic molecules are called biological macromolecules. There are four major classes of biological macromolecules (carbohydrates, lipids, proteins, and nucleic ...
  104. [104]
    Biological molecules
    b) Proteins are used primarily for enzymatic activities, signaling, and dynamic structural components. c) Nucleic acids are used for genetic information ...
  105. [105]
    A historical perspective on the discovery of statins - PMC - NIH
    Scientists have developed several lines of evidence that establish the causal connection between blood cholesterol, atherosclerosis, and coronary heart disease.
  106. [106]
    Penicillin: Opening the Era of Antibiotics - USDA ARS
    Feb 8, 2024 · Penicillin was discovered in 1928 by bacteriologist Alexander Fleming, working at St. Mary's Hospital in London. It was not until 1939 that ...
  107. [107]
    Synthesis and Deconstruction of Polyethylene-type Materials
    Feb 26, 2024 · Polyethylene is the largest produced synthetic polymer, with an annual production of more than 100 million tons. Polyethylene alone accounts for ...
  108. [108]
    Chapter 7 - Alkanes and Halogenated Hydrocarbons - Chemistry
    Alkanes are organic compounds that consist entirely of single-bonded carbon ... The octane is one of the molecules found in petrol (gasoline). Looking ...
  109. [109]
    Glyphosate | US EPA
    Glyphosate is a widely used herbicide that controls broadleaf weeds and grasses. It has been registered as a pesticide in the US since 1974.
  110. [110]
    Learn about Polychlorinated Biphenyls | US EPA
    Taken together, the data strongly suggest that PCBs are probable human carcinogens. PCBs are one of the most widely studied environmental contaminants. Many ...Inadvertent PCBs · Commercial Uses for PCBs · PCB Mixtures and Trade Names
  111. [111]
    Basics of Green Chemistry | US EPA
    Feb 4, 2025 · The 12 Principles of Green Chemistry · 1. Prevent waste · 2. Maximize atom economy · 3. Design less hazardous chemical syntheses · 4. Design safer ...