Fact-checked by Grok 2 weeks ago

Addition reaction

An addition reaction is a type of in which atoms or groups from two or more reactant molecules are added across a carbon-carbon double or triple bond, typically in alkenes or alkynes, resulting in a single product with a saturated carbon . These reactions are exothermic due to the relatively weak (approximately 63 kcal/mol) in the multiple bond, which is broken during the process. Addition reactions are fundamental in , enabling the conversion of unsaturated compounds into functionalized derivatives used in pharmaceuticals, polymers, and materials. The most common addition reactions are classified by their mechanisms, including , , and concerted addition. In electrophilic addition, an such as a proton from HX (where X is a ) attacks the pi electrons of the , forming a intermediate that is then captured by a ; this often follows , where the adds to the less substituted carbon to yield the more stable . Notable examples include (e.g., HBr addition to propene forming ) and (e.g., Br₂ addition to ethene producing via a bromonium intermediate, resulting in anti ). Concerted additions, such as hydroboration-oxidation, proceed without discrete intermediates in a single step, often exhibiting syn stereochemistry and anti-Markovnikov ; this method, developed by (, 1979), adds (BH₃) across the double bond followed by oxidation to yield alcohols. Free radical additions, like the peroxide-initiated addition of HBr to alkenes, also favor anti-Markovnikov products through a chain mechanism involving radical intermediates. These reactions' and are crucial for controlling product outcomes in .

Fundamentals

Definition

An addition reaction in is a process in which two or more reactant molecules combine to form a single product, typically by adding across a multiple —such as a carbon-carbon double (C=C) or triple (C≡C) —in an unsaturated compound, without the elimination of any small byproduct like or . This reaction type is characteristic of unsaturated systems, where the multiple serves as the reactive site, leading to an increase in and molecular complexity. The general scheme for an can be illustrated generically as follows, where the multiple between atoms C and D is broken and new single bonds are formed with the adding A-B: \ce{A-B + C=D -> A-C-D-B} This schematic represents the of the pi component of the multiple bond, resulting in a product with no net loss of atoms beyond the reactants. Addition reactions differ fundamentally from substitution reactions, which involve the replacement of one or group in a by another without altering the overall level, and from elimination reactions, which remove adjacent atoms or groups to generate a multiple bond and often produce a small , thereby increasing unsaturation. In addition reactions, the process instead builds the by incorporating the reactants directly across the unsaturated site. Prerequisite to understanding these reactions is the concept of pi bonds, which arise from the sideways overlap of p orbitals in multiple bonds and provide the that makes such sites reactive toward addition, based on .

Key Characteristics

Addition reactions are thermodynamically favorable in many cases because they involve the conversion of a weaker into stronger σ bonds, resulting in an with a negative change (ΔH < 0). For typical additions to alkenes, such as hydrogenation, the reaction releases energy due to the higher bond energy of the two new σ bonds (approximately 80-100 kcal/mol each) compared to the single π bond (about 63 kcal/mol). This can be illustrated in a simple energy diagram where the reactants (alkene and addend) occupy a higher energy state than the products (saturated compound), with the downward arrow representing the exothermic ΔH:
          Reactants
             |
             | ΔH < 0
             v
         Products
The degree of exothermicity varies with substitution; for example, more substituted yield slightly less exothermic additions due to the stability of the alkene itself. Regioselectivity in addition reactions to unsymmetrical substrates is often governed by , an empirical principle stating that in electrophilic additions, the hydrogen atom adds to the carbon of the double bond bearing the greater number of hydrogens, while the electrophile attaches to the more substituted carbon. This arises from the formation of the more stable carbocation intermediate during the mechanism. In contrast, anti-Markovnikov regioselectivity can occur via free-radical mechanisms, such as the peroxide-initiated addition of HBr to , where the bromine adds to the less substituted carbon. The reactivity of unsaturated substrates toward addition follows the order alkenes > alkynes, primarily due to differences in π-electron density and orbital hybridization. Alkenes, with sp² hybridization, exhibit higher in their π bonds, making them more susceptible to electrophilic attack than alkynes (sp hybridized, with tighter-held electrons). While many addition reactions are irreversible under standard conditions, some—particularly nucleophilic additions to carbonyl compounds—can be reversible, establishing an between the carbonyl and the addition product. This reversibility depends on the of the nucleophile and , allowing the tetrahedral intermediate to expel the nucleophile if the favors the starting materials.

Classification by Mechanism

Electrophilic Addition

Electrophilic addition reactions involve the attack of an electron-deficient , known as an , on the electron-rich of an or , leading to the formation of new sigma bonds. This process is a cornerstone of reactivity in , typically proceeding through a two-step mechanism. In the first step, the pi electrons of the carbon-carbon donate to the (E⁺), breaking the pi bond and generating a on one of the carbons. This can be represented conceptually with curved notation, where arrows indicate the of the pi toward the and the bonding of the to the less substituted carbon, resulting in a positively charged carbon . The second step involves the addition of a (often an anion or solvent molecule) to the electron-deficient , forming the final product and restoring neutrality. The intermediate is planar and sp² hybridized, with the empty p orbital allowing for stabilization if adjacent groups permit. Stability increases with substitution: carbocations (3°) are more stable than secondary (2°), which are more stable than primary (1°), due to and inductive effects from s donating electron density. This stability hierarchy drives rearrangements, such as 1,2-hydride or methyl shifts, where a or alkyl group from an adjacent carbon migrates to the carbocation center, forming a more stable isomer. According to , the of the rate-determining first step resembles the carbocation intermediate in structure and energy, particularly since this step is endergonic; thus, pathways leading to more stable carbocations have lower activation energies and predominate. Common electrophiles in these reactions include protons (H⁺) generated from strong acids, which initiate addition by protonating the . Halogens (X₂) act via electrophilic halonium ions, three-membered cyclic intermediates that bridge the two carbons and prevent direct formation in some cases, though the mechanism retains an electrophilic character. In metal-catalyzed additions, such as oxymercuration, electrophiles form π-complexes with the , coordinating the metal to the pi electrons before nucleophilic attack. These variations highlight the versatility of while maintaining the core principle of pi bond activation by electron-deficient species. Reaction outcomes can be under kinetic or thermodynamic control, depending on conditions. At low temperatures, the kinetic product—formed via the faster pathway through the more stable —is favored, even if it is the less stable overall. Higher temperatures allow equilibration, promoting the thermodynamic product, which is the most stable. This control is particularly relevant in cases where multiple pathways or rearrangements are possible, aligning with in regioselective additions.

Nucleophilic Addition

Nucleophilic addition reactions to carbon-carbon multiple bonds occur when the is activated by an (EWG), such as a carbonyl, making the β-carbon electrophilic. This allows a to attack the β-carbon, forming a new C-Nu bond and generating a intermediate stabilized by the EWG (often as an ). The mechanism proceeds in two steps: first, the donates its to the β-carbon, breaking the and placing the negative charge on the α-carbon; second, of the yields the neutral addition product. This can be represented as: \ce{Nu^- + R-CH=CH-EWG -> R-CH(Nu)-CH^- -EWG ->[H^+] R-CH(Nu)-CH_2-EWG} Common nucleophiles include enolates (carbon nucleophiles) in the Michael reaction, as well as amines, thiols, and alkoxides. For unactivated alkenes, such additions are rare due to insufficient electrophilicity, but they are more feasible with alkynes or conjugated systems. Specific examples, such as conjugate additions to α,β-unsaturated carbonyls, are detailed in later sections. The addition is often under kinetic control and irreversible under typical conditions, particularly with carbon nucleophiles, due to the stability of the intermediate. In basic media, the reaction favors the addition product, while acidic conditions may promote elimination instead. Variants include aza-Michael additions with amines, leading to β-amino carbonyls, and thiol-Michael additions, which are radical-mediated in some cases but follow the nucleophilic pathway here.

Free-Radical Addition

Free-radical addition reactions involve the addition of atoms or groups across a carbon-carbon multiple bond via a mechanism mediated by neutral intermediates, distinguishing them from polar ionic pathways. These reactions typically proceed through three stages: , , and termination, and are commonly observed in the addition of hydrogen halides to alkenes under radical conditions. Initiation occurs when a radical source, such as a , undergoes homolysis to generate initiating , often under or . For example, dialkyl peroxides decompose as follows: \text{ROOR} \rightarrow 2 \text{RO}^\bullet These radicals can then abstract a bromine atom from HBr to form a bromine radical: \text{RO}^\bullet + \text{HBr} \rightarrow \text{ROH} + \text{Br}^\bullet This step generates the reactive species that start the chain, with peroxides like benzoyl peroxide commonly used as initiators. In the propagation phase, the bromine radical adds to the at the less substituted carbon, forming a new carbon-centered that is stabilized by adjacent alkyl groups. For propene (CH_3CH=CH_2), this yields: \text{Br}^\bullet + \text{CH}_2=\text{CHCH}_3 \rightarrow \text{BrCH}_2\text{CH}^\bullet\text{CH}_3 The resulting secondary then abstracts a from HBr, regenerating the bromine and forming the product: \text{BrCH}_2\text{CH}^\bullet\text{CH}_3 + \text{HBr} \rightarrow \text{BrCH}_2\text{CH}_2\text{CH}_3 + \text{Br}^\bullet This cycle continues, leading to efficient chain growth and anti-Markovnikov , where the attaches to the less substituted carbon due to the preference for forming the more intermediate. Termination halts the chain when two radicals combine, such as through coupling to form a or where one abstracts a from another. Examples include: $2 \text{BrCH}_2\text{CH}^\bullet\text{CH}_3 \rightarrow (\text{BrCH}_2\text{CHCH}_3)_2 or \text{BrCH}_2\text{CH}^\bullet\text{CH}_3 + \text{BrCH}_2\text{CHCH}_3^\bullet \rightarrow \text{BrCH}_2\text{CH}_2\text{CH}_3 + \text{BrCH}_2\text{CH=CH}_2 These steps are less frequent than but limit the overall length. in free-radical additions arises from the stability of the , favoring addition that places the radical on the more substituted carbon (tertiary > secondary > primary), though steric factors can influence outcomes; this contrasts with ionic mechanisms by lacking charge-based polarization. These reactions are generally less regioselective than their electrophilic counterparts in non-symmetrical cases beyond HX additions. Typical conditions involve initiators like peroxides or UV light at moderate temperatures, often under an inert atmosphere (e.g., ) to prevent oxygen inhibition, as O_2 rapidly reacts with radicals to form unreactive peroxides. Free-radical additions form the basis for of alkenes, such as from , where by peroxides leads to repeated across C=C bonds, building long chains until termination.

Pericyclic Addition

Pericyclic addition reactions represent a class of concerted cycloadditions where multiple π-bonds form new σ-bonds simultaneously through a cyclic , adhering strictly to the principles of orbital symmetry conservation. These processes are inherently stereospecific, proceeding without discrete intermediates or charge separation, in contrast to ionic or mechanisms. The theoretical foundation for their feasibility and is provided by the Woodward-Hoffmann rules, which analyze the symmetry properties of the frontier molecular orbitals (FMOs) involved in the reaction under thermal or photochemical excitation. A prototypical example is the thermal [4+2] , exemplified by the Diels-Alder reaction between a conjugated and a dienophile. According to the Woodward-Hoffmann rules, this suprafacial process is thermally allowed because the symmetry of the diene's highest occupied (HOMO, specifically the ψ₂ orbital of s-cis-butadiene) matches that of the dienophile's lowest unoccupied (LUMO, the π* orbital of ), enabling constructive overlap along the . The HOMO-LUMO interaction illustrates this bonding alignment:
Diene HOMO (ψ₂): lobes with alternating phases across the four p-orbitals
          +     -     +     -
Dienophile LUMO (π*): symmetric antibonding lobes
          - -         + +
Resulting overlap: Two new σ-bonds form suprafacially, conserving orbital [symmetry](/page/Symmetry).
This synchronous involves a pericyclic transition state where both new bonds develop concurrently, typically with partial character but no stable intermediates, as confirmed by computational and experimental kinetic studies. Other common pericyclic additions include photochemical [2+2] cycloadditions and [3+2] 1,3-dipolar cycloadditions. The [2+2] variant, forbidden thermally due to mismatched orbital symmetries in the , becomes allowed under photochemical conditions by exciting one reactant to a (via ), which alters the FMO symmetries to permit bonding overlap and often proceeds through a 1,4-diradical intermediate before cyclizing to a cyclobutane product. Representative examples involve enones and , yielding strained four-membered rings with high diastereoselectivity. In contrast, [3+2] 1,3-dipolar cycloadditions feature a 1,3-dipole (such as an or ) reacting with a dipolarophile (e.g., an or ) in a concerted manner to form five-membered heterocycles; the reaction is thermally suprafacial and symmetry-allowed, with the dipole's zwitterionic resonance forms facilitating the pericyclic . Stereospecificity is a hallmark of pericyclic additions, arising from the rigid orbital overlap requirements. In the Diels-Alder reaction, the Alder endo rule dictates preferential formation of the endo , where electron-withdrawing groups on the dienophile orient toward the in the ; this selectivity stems from favorable secondary orbital interactions between the dienophile's LUMO and the 's HOMO coefficients, lowering the by approximately 2-5 kcal/mol compared to the exo pathway. The rule holds for most thermal Diels-Alder reactions involving cyclic s and unsaturated dienophiles, ensuring retention of from both reactants.

Specific Examples and Applications

Hydrohalogenation and Hydrogenation

involves the addition of , such as (HBr) or (HCl), to alkenes, resulting in the formation of alkyl . This proceeds via an mechanism, where the alkene's π-bond acts as a , attacking the electrophilic of HX to form a intermediate, followed by anion capture. For unsymmetrical alkenes, the addition follows , directing the to the carbon with more hydrogens and the to the carbon with fewer, as seen in the of propene with HBr yielding : \text{CH}_3\text{CH}=\text{CH}_2 + \text{HBr} \rightarrow \text{CH}_3\text{CHBr}\text{CH}_3 In the presence of peroxides, HBr addition to alkenes switches to a free-radical , leading to anti-Markovnikov where attaches to the less substituted carbon. This peroxide effect is unique to HBr due to the favorable bond dissociation energies in the radical propagation steps and does not occur with HCl or . Hydrogenation is the addition of molecular hydrogen (H₂) to alkenes or alkynes, catalyzed by transition metals, converting unsaturated compounds to saturated alkanes. The reaction proceeds with syn , adding both hydrogens from the same face of the , and is typically carried out under mild conditions with catalysts like platinum (Pt) or palladium (Pd). A general example is the of an internal to an : \text{RCH}=\text{CHR} + \text{H}_2 \xrightarrow{\text{catalyst}} \text{RCH}_2\text{CH}_2\text{R} Catalysts for hydrogenation are divided into heterogeneous types, such as Pd or Pt supported on carbon or alumina, which facilitate complete reduction of alkenes to alkanes or alkynes to alkanes under atmospheric pressure, and homogeneous types like Wilkinson's catalyst (RhCl(PPh₃)₃), which enable selective partial reduction of alkynes to cis-alkenes. Heterogeneous catalysts are preferred for industrial scalability due to ease of separation, while homogeneous ones offer higher selectivity in complex syntheses. Industrially, is pivotal for synthesis in processes, such as converting olefin streams from cracking to stable for fuels and feedstocks, and for stereoselective reductions in pharmaceutical production. For instance, Pd/C-catalyzed saturates vegetable oils to produce solid fats, demonstrating its broad utility in converting unsaturated hydrocarbons to on a massive scale.

Halogenation and Dihydroxylation

of alkenes involves the of molecular , such as (Br₂) or (Cl₂), across the carbon-carbon double bond to form vicinal dihalides. The reaction proceeds via a three-membered cyclic intermediate, first proposed by Roberts and Kimball in 1937 to explain the observed . In this , the polarized halogen molecule approaches the , with the electrophilic halogen bridging the two carbon atoms, forming the ; a then attacks from the opposite face, leading to trans addition. A representative example is the addition of Br₂ to , yielding : \ce{CH2=CH2 + Br2 -> BrCH2-CH2Br} This process is typically carried out in an inert solvent like dichloromethane to favor dihalide formation. In aqueous media, the reaction diverges to produce halohydrins instead of dihalides, as water acts as a nucleophile to open the halonium ion intermediate, with the hydroxyl group attaching to the more substituted carbon following Markovnikov-like regioselectivity. For instance, propene reacts with Br₂ in water to give 1-bromo-2-propanol predominantly. This solvent effect highlights the role of the nucleophile in determining product outcome, with non-nucleophilic solvents promoting halide attack and protic solvents enabling oxygen incorporation. Dihydroxylation achieves the syn addition of two hydroxyl groups across an to form vicinal s, or glycols, using reagents like (OsO₄) or (KMnO₄). The OsO₄-mediated process, mechanistically elucidated by Criegee in 1936, involves a [3+2] cycloaddition to form a cyclic osmate intermediate, which is then hydrolyzed to the diol. For a general alkene RCH=CHR', the reaction sequence is: \ce{RCH=CHR' + OsO4 ->[cyclic osmate ester] ->[hydrolysis] RCH(OH)-CH(OH)R'} This method ensures stereospecific syn addition, preserving the alkene's geometry in the product. KMnO₄ in cold, dilute, basic conditions similarly effects syn dihydroxylation, though it is less selective for sensitive substrates and can lead to over-oxidation with heating. The Upjohn process enhances practicality by employing catalytic OsO₄ with a stoichiometric co-oxidant like N-methylmorpholine N-oxide (NMO), regenerating the osmium catalyst in situ and minimizing toxic osmium waste. Developed in 1976, this variant is widely adopted for large-scale syntheses due to its efficiency and mild conditions. Stereochemistry is a defining feature of these additions: halogenation yields anti products due to back-side attack on the halonium ion, resulting in meso compounds from trans alkenes or racemic mixtures from cis alkenes. In contrast, dihydroxylation delivers syn addition, producing enantiomeric diols from trans alkenes and meso diols from cis alkenes, directly reflecting the alkene's configuration. These outcomes are critical for controlling chirality in synthesis, with the alkene geometry dictating diastereoselectivity. Applications of these reactions abound in . Vicinal dihalides serve as precursors for alkynes via double dehydrohalogenation and for alkenes through stereospecific elimination, enabling carbon chain manipulations. Halohydrins from aqueous are key intermediates for formation under basic conditions, useful in constructing complex oxygen heterocycles. Dihydroxylation products, glycols, are vital for producing polyesters, antifreeze agents like , and pharmaceutical intermediates, with the Upjohn process facilitating scalable production of chiral diols.

Conjugate Additions

Conjugate additions, also referred to as 1,4-additions, occur when a nucleophile attacks the β-carbon of an α,β-unsaturated carbonyl compound, such as an enone or enal, rather than the carbonyl carbon directly. This regioselectivity arises from the conjugation between the alkene and the carbonyl group, which delocalizes the electron density and stabilizes the resulting enolate intermediate at the α-position. The general mechanism begins with nucleophilic addition to the β-carbon, forming a β-substituted enolate; protonation then occurs at the α-carbon to yield the saturated carbonyl product. A representative equation is: \ce{CH2=CH-C(=O)R + Nu^- ->[1,4-addition] Nu-CH2-CH=C(OH)R ->[tautomerization/protonation] Nu-CH2-CH2-C(=O)R} This process is thermodynamically favored due to the formation of the more stable enolate and is widely used in organic synthesis for constructing carbon-carbon bonds in 1,5-dicarbonyl systems. The Michael addition represents a classic subtype of conjugate addition, involving enolate nucleophiles derived from carbon-based donors, such as active methylene compounds, adding to α,β-unsaturated carbonyl acceptors under basic conditions. These donors, including malonates or β-ketoesters, are deprotonated by a base to generate stabilized enolates that preferentially undergo 1,4-addition. A seminal example is the base-catalyzed reaction of diethyl malonate with acrolein, yielding diethyl 2-(3-oxopropyl)propanedioate after addition and protonation, which serves as a precursor for further synthetic elaborations. This reaction, first described by Arthur Michael in 1887, is reversible and often driven to completion by using excess donor or removing byproducts, highlighting its utility in building complex carbon frameworks. Organocopper , particularly Gilman reagents of the form (R₂CuLi), enable conjugate additions of alkyl or aryl groups to α,β-unsaturated carbonyls, providing a method for anti-Markovnikov-like with soft nucleophiles. These form a copper(III) intermediate upon coordination to the β-carbon, followed by to deliver the R group and regenerate the , which is then protonated. The general is: \ce{(R2CuLi) + CH2=CH-C(=O)R -> R-CH2-CH2-C(=O)R + RCu + Li^+} Developed in the , this approach contrasts with harder organometallics like Grignards, which favor 1,2-addition, and has been pivotal in due to its compatibility with diverse functional groups. plays a key role in enhancing efficiency and selectivity; basic conditions (e.g., alkoxides or amines) promote additions by generating the donor, while salts facilitate organocopper reactions by activating the . Asymmetric variants employ chiral ligands, such as phosphines or N-heterocyclic carbenes, with catalysts to achieve high enantioselectivity (often >90% ee) in the creation of stereocenters at the β-position, as reviewed in foundational studies on stereocontrolled conjugate additions. These methods underscore the versatility of conjugate additions in stereoselective synthesis.

Stereochemistry and Scope

Stereochemical Outcomes

Addition reactions to alkenes exhibit distinct stereochemical outcomes depending on the , primarily manifesting as or anti addition. In addition, the two substituents add to the same face of the , resulting in cis relative for the product. This occurs in catalytic , where both atoms approach from the same side of the adsorbed on the catalyst surface. Similarly, pericyclic additions, being concerted, proceed suprafacially and yield . In contrast, anti addition delivers substituents to opposite faces, producing trans relative , as seen in electrophilic via a intermediate, where the attacks from the backside. Diastereoselectivity arises when the substrate or reaction conditions favor one over another. For cyclic alkenes, syn to in yields the cis-1,2-diol, while anti in produces the trans-1,2-dihalide. In acyclic systems with a chiral center adjacent to the reacting or carbonyl, Cram's rule predicts the preferred based on steric control, where the approaches from the less hindered side opposite the largest in the conformer with the carbonyl coordinated to a metal or hydrogen-bonded. This model, established in 1952, typically gives moderate to good diastereoselectivity in nucleophilic additions to α-chiral aldehydes. For example, to 2-phenylpropanal follows Cram's chelate rule variant under coordinating conditions, enhancing selectivity. Enantioselectivity is achieved using chiral catalysts or auxiliaries, enabling the formation of enantioenriched products from achiral substrates. The Sharpless asymmetric dihydroxylation exemplifies this, employing osmium tetroxide with chiral cinchona alkaloid ligands to perform syn dihydroxylation with high enantiomeric excess (ee). For trans-alkenes like trans-stilbene, it delivers the (R,R)-diol with up to 99% ee using the DHQD ligand system. This 1988 ligand-accelerated catalysis method has broad scope, often exceeding 90% ee for various alkenes, and is pivotal in asymmetric synthesis. In cases without chiral control, such as uncatalyzed additions, products are typically racemic, highlighting the role of chiral auxiliaries in resolving stereoisomers.

Substrate Scope and Limitations

Addition reactions encompass a broad range of substrates, primarily depending on the mechanistic pathway involved. For electrophilic additions, electron-rich alkenes and alkynes serve as ideal substrates due to their nucleophilic π-bonds, which readily interact with electrophiles like H+ or halonium ions. In contrast, nucleophilic additions typically target electron-poor carbonyl compounds (e.g., aldehydes, ketones) and imines, where the polarized C=O or C=N bonds facilitate attack by nucleophiles such as Grignard reagents or organolithiums. Pericyclic additions, including , favor conjugated dienes and dienophiles with electron-withdrawing groups to enhance efficiency. Free-radical additions are versatile, accommodating unactivated alkenes and alkynes via mechanisms such as the peroxide-initiated anti-Markovnikov addition of HBr. Despite this versatility, steric and electronic factors impose significant limitations on substrate compatibility. Highly substituted alkenes, such as trisubstituted or tetrasubstituted ones, often exhibit reduced reactivity in electrophilic additions due to steric hindrance around the , leading to slower reaction rates or incomplete conversions. Electronically deactivated substrates, like fluoroalkenes, resist electrophilic attack because the electronegative withdraws electron density from the π-system, stabilizing it against or . In carbocation-mediated pathways, such as acid-catalyzed additions, β-elimination can compete with addition, particularly in substrates prone to forming stable carbocations, resulting in rearrangement or elimination products instead of the desired adducts. Side reactions further constrain the scope of addition reactions. Radical additions to dienes or polyenes frequently lead to , as chain propagation can generate multiple addition sites, diminishing yields of monomeric products. For alkynes, over-addition is a common issue in both electrophilic and nucleophilic processes, where the initial adduct undergoes a second addition to form geminal dihalides or saturated compounds, requiring careful control of or protective groups. Advancements as of 2024 have expanded the substrate scope beyond traditional limitations. Strained cyclic alkenes, such as norbornene, undergo facile additions due to angle strain that lowers the activation energy for π-bond breaking, enabling applications in polymer synthesis. Allenes and cumulenes have emerged as viable substrates for stereoselective additions, particularly in metal-catalyzed variants. In bioorthogonal chemistry, copper-free click reactions represent a modern extension, allowing strain-promoted azide-alkyne cycloadditions to strained cyclooctynes in aqueous environments without interfering with biological substrates. For instance, cobalt-catalyzed stereoselective hydrofunctionalization of alkenes and alkynes enables the synthesis of enantioenriched functionalized products with high efficiency. These developments highlight ongoing efforts to mitigate inherent limitations through catalyst design and substrate engineering.