Fact-checked by Grok 2 weeks ago

Hydroxide

The , denoted as OH⁻, is a diatomic inorganic anion consisting of a single oxygen atom covalently bonded to a , with the negative charge primarily localized on the oxygen. It serves as the conjugate of (H₂O) and is a fundamental species in aqueous chemistry, produced when dissociate in solution according to the Arrhenius definition, where its presence elevates the above 7. With the molecular formula OH⁻ and a of 17.007 g/, the exhibits strong basicity and nucleophilicity due to the high on oxygen, enabling it to accept protons or participate in reactions. Its preferred IUPAC name is (systematic name: oxidanide), and it forms stable ionic compounds known as hydroxides (e.g., , NaOH) that are widely used in industry for adjustment, cleaning, and . In biological systems, trace amounts function as metabolites, while in environmental contexts, it influences in natural s. The 's anomalous diffusion in water, faster than other small s, arises from a proton-hopping mechanism involving hydrogen-bonded networks.

Hydroxide Ion

Properties

The hydroxide (OH⁻) is a monovalent diatomic anion composed of an oxygen atom covalently bonded to a , carrying a negative charge primarily on the oxygen. It possesses an effective of approximately 133 when in a coordination environment of six, consistent with its role in ionic compounds and solvation shells. As the conjugate base of , OH⁻ demonstrates strong basicity in aqueous environments, with a pK_b \approx -1.7, reflecting its high affinity for protons and ability to deprotonate weak acids effectively. In , OH⁻ accepts protons to form H₂O, driving acid-base equilibria toward neutralization in solutions where [OH⁻] exceeds 10⁻⁷ M, resulting in > 7. This basic strength positions OH⁻ as a key in alkaline conditions, though its reactivity is moderated by effects. In aqueous solutions, the hydroxide ion undergoes , forming solvated clusters denoted as [OH(H₂O)_n]⁻, where n ≈ 3–5 constitutes the primary hydration shell through hydrogen bonding from molecules donating protons to the oxygen of OH⁻. These clusters stabilize the and influence its mobility and reactivity, contributing to the elevated in basic media. The equilibrium concentration of OH⁻ is linked to H⁺ via the ion product of , defined as K_w = [\ce{H+}] [\ce{OH-}] = 1.0 \times 10^{-14} at 25°C under standard conditions. This constant arises from the autoionization of (2H₂O ⇌ H₃O⁺ + OH⁻) and shows temperature dependence, increasing to approximately 5.5 × 10^{-14} at 50°C due to the endothermic of the process, thereby shifting neutral below 7 at higher s. The properties of ⁻ are evident in its production during the cathodic reduction of oxygen in alkaline solutions, characterized by the \ce{O2 + 2H2O + 4e- -> 4[OH](/page/Oh)-} with a E^\circ = 0.401 V versus the at 25°C and 14. This potential underscores ⁻'s role in electrochemical processes like and fuel cells, where it serves as a product in oxygen reduction reactions.

Spectroscopic Characteristics

The hydroxide ion (⁻) exhibits distinct vibrational signatures that facilitate its detection across different phases. In the gas phase, the free ⁻ ion displays a sharp O-H stretching vibration in the () spectrum at approximately 3555 cm⁻¹, corresponding to the fundamental mode of the ionic bond. In aqueous or hydrogen-bonded environments, this band broadens and shifts to lower wavenumbers, typically centering around 3000–3600 cm⁻¹ with a continuum extending down to 800 cm⁻¹ due to delocalized proton transfer and effects. complements by revealing symmetric stretching modes; for instance, in concentrated aqueous NaOH solutions, a broad ⁻ stretching band appears near 3500 cm⁻¹, while in solid hydroxides like NaOH, it sharpens to 3633 cm⁻¹. Electronic of OH⁻ reveals (UV) absorption primarily below 200 nm. A characteristic band around 190 nm arises from an n→σ* involving the oxygen to the O-H antibonding orbital, observable in both gas-phase and aqueous measurements of hydroxide solutions. This is intense and shifts slightly in solvated systems due to environmental , but remains a key identifier for the isolated ion. Nuclear magnetic resonance (NMR) spectroscopy provides insights into the electronic environment of OH⁻ via ¹⁷O enrichment. In aqueous solutions, such as NaOH, the ¹⁷O for OH⁻ is approximately 20 ppm relative to , reflecting its high and minimal deshielding. The one-bond scalar ¹J(¹⁷O,¹H) is around 82 Hz, arising from the strong O-H bond and observable in high-resolution spectra despite quadrupolar broadening. In mass spectrometry, particularly electrospray ionization (ESI-MS), the OH⁻ ion is directly detected as the base peak at m/z 17 in negative-ion mode from hydroxide-containing samples. Fragmentation patterns of larger solvated or complexed species often yield this m/z 17 ion via neutral loss of water or other ligands, confirming the presence of intact OH⁻ without extensive dissociation due to ESI's soft ionization. Isotopic substitution highlights mass-dependent effects in OH⁻ spectra. Replacing ¹⁶O with ¹⁸O reduces the O-H stretching frequency by about 5.8%, from ~3555 cm⁻¹ to ~3350 cm⁻¹ in the gas phase, due to the increased reduced mass altering the vibrational potential. Similar downshifts occur in IR and Raman bands for ¹⁸OH⁻ in clusters or solutions, enabling precise quantification of isotopic ratios and aiding structural assignments.

Structural Chemistry

Bonding and Geometry

The hydroxide (OH⁻) is a diatomic anion characterized by a linear and an O–H of 0.964 , as determined from high-resolution spectroscopic measurements. In , the valence electrons occupy a σ orbital formed primarily from the oxygen 2s and 2p orbitals interacting with the 1s orbital, resulting in a of 1 and a closed-shell electronic configuration with eight valence electrons. This contrasts with the neutral OH radical, which has seven valence electrons, placing an in the σ orbital and yielding a slightly longer of 0.971 while maintaining a linear . and (DFT) calculations confirm the electron density distribution, with significant charge accumulation on the oxygen atom, contributing to the ion's high basicity. In coordination chemistry, the hydroxide ion serves as the hydroxo (OH⁻) ligand, typically binding through the oxygen atom in a monodentate fashion or as a bridging μ-OH group between metal centers. Monodentate coordination is exemplified in octahedral hexahydroxo complexes such as [Al(OH)₆]³⁻, where six OH⁻ ligands surround the central aluminum ion, forming regular M–O bonds with M–O–H angles approaching 110° due to the sp³-like hybridization at oxygen. Bridging μ-OH ligands are common in dinuclear and polynuclear complexes, such as [(H₂O)₄Cr(μ-OH)₂Cr(OH₂)₄]⁴⁺, where the OH⁻ group donates to two metal atoms simultaneously, often stabilizing higher oxidation states through interactions. These coordination modes influence the overall , with terminal hydroxo ligands leading to bent M–O–H arrangements and bridging forms adopting linear or asymmetric configurations depending on the metal–metal distance. The hydroxide ion plays a central role in hydrogen bonding networks, acting both as a hydrogen bond donor via its O–H group and as a strong acceptor through the negatively charged oxygen lone pairs. In aqueous solutions, OH⁻ forms extensive three-dimensional networks with molecules, typically engaging in three to four acceptor interactions and one donor interaction per ion, which contributes to its anomalously high mobility despite strong . The energy of individual O–H···O hydrogen bonds involving OH⁻ ranges from 20 to 40 kJ/, reflecting moderate to strong interactions that stabilize ionic structures and influence properties like in alkaline solutions. Quantum chemical calculations, including DFT and coupled-cluster methods, provide insights into the electronic structure of OH⁻. These computations also map the , showing a pronounced with partial negative charge on oxygen (~ -1.2 e) and positive on hydrogen (~ +0.2 e), consistent with the ion's role in proton transfer processes. In acidic conditions, coordinated OH⁻ ligands exhibit amphoteric behavior, readily undergoing to form aqua ligands (M–OH + H⁺ → M–OH₂⁺), which initiates reactions in metal complexes.

Crystal Structures

Hydroxide-containing materials often exhibit layered crystal structures, where metal cations are coordinated by hydroxide ions in octahedral arrangements, forming sheets that are stacked via hydrogen bonding. A prototypical example is , Mg(OH)₂, which adopts a trigonal structure in the P-3m1, consisting of brucite-type layers of edge-sharing MgO₆ octahedra. These layers are held together by weak interlayer hydrogen bonds between the hydroxyl groups of adjacent sheets, resulting in a highly anisotropic structure with cleavage parallel to the layers. Polymorphism is prevalent among hydroxide structures, particularly in aluminum hydroxides, where different stacking arrangements and coordination environments lead to distinct phases. , the α-form of Al(OH)₃, features a monoclinic with double layers of Al(OH)₆ octahedra linked by bonds, forming a dense packing that makes it the most thermodynamically stable polymorph under ambient conditions. In contrast, bayerite (β-Al(OH)₃) has a similar layered motif but with a different interlayer -bonding configuration, leading to a hexagonal , while (γ-AlOOH) represents an oxyhydroxide polymorph with orthorhombic and chains of edge-sharing AlO₆ octahedra connected via bonds. These polymorphs arise from variations in conditions, such as and , influencing their relative and applications in . Ionic hydroxides display diverse types depending on the cation size and charge. , NaOH, forms an orthorhombic in the Cmcm at , with Na⁺ ions surrounded by OH⁻ ions in a distorted octahedral coordination, transitioning to a cubic at elevated temperatures around 575 K where the OH⁻ ions exhibit rotational . In comparison, , Ca(OH)₂, known as , adopts a hexagonal in the P-3m1, isostructural with , featuring layers of CaO₆ octahedra linked by hydrogen bonds between hydroxyl groups pointing toward interlayer spaces. These structural differences reflect the larger of Ca²⁺ compared to Na⁺, promoting layered over cubic packing. Hydrogen bonding networks are crucial for stabilizing hydroxide crystal architectures, often forming infinite chains within layers or three-dimensional frameworks across the structure. In layered hydroxides like and , these networks involve O-H···O bonds with typical donor-acceptor O···O distances ranging from 2.7 to 3.0 , which dictate interlayer cohesion and facilitate properties such as swelling or . Such bonds create a balance between covalent intralayer interactions and weaker intermolecular forces, enabling polymorphism and influencing vibrational spectra. Defects and doping in hydroxide layers significantly alter their electronic and transport properties, particularly in (LDHs) derived from brucite-like structures. Vacancies, such as cation or anion defects, introduce local charge imbalances that enhance ionic by creating pathways for proton or hydroxide migration, as seen in NiFe-LDHs where oxygen vacancies improve rates. Doping with aliovalent ions, like partial substitution of Mg²⁺ with Al³⁺ in LDHs, generates positive layer charges balanced by interlayer anions, further tuning defect densities and boosting applications in electrocatalysis. These modifications, often introduced via or post-treatment, exemplify how structural imperfections can optimize material performance without disrupting the overall .

Inorganic Hydroxides

Alkali and Alkaline Earth Metal Hydroxides

Alkali metal hydroxides, such as those of , sodium, and , are typically synthesized by the direct reaction of the alkali metal with , producing the hydroxide and gas; for example, sodium reacts as 2Na + 2H₂O → 2NaOH + H₂, a highly that generates heat and often ignites the . Alkaline earth metal hydroxides are commonly prepared by the of the corresponding oxides, known as slaking; , for instance, reacts with to form : CaO + H₂O → Ca(OH)₂. Solubility in increases down , with exhibiting lower (12.8 g/100 mL at 20°C) compared to (109 g/100 mL at 20°C) and (112 g/100 mL at 20°C), primarily due to the higher of LiOH arising from the small size of the Li⁺ ion, which makes less favorable despite similar energies. In , increases with increasing cation size, but hydroxides like Ca(OH)₂ remain sparingly soluble overall, with a (K<sub>sp</sub>) of 5.5 × 10<sup>−6</sup> at 25°C, reflecting its limited into Ca²⁺ and OH⁻ ions. These hydroxides display varying thermal stability, with NaOH remaining stable up to its around 323°C before decomposing into Na₂O and H₂O at higher temperatures under certain conditions. In contrast, Mg(OH)₂ decomposes at approximately 350°C to form MgO and : Mg(OH)₂ → MgO + H₂O. Sodium and hydroxides are highly hygroscopic and deliquescent, readily absorbing atmospheric moisture to form hydrates such as NaOH·H₂O, which crystallizes from aqueous solutions in specific temperature ranges. A notable application involves in , a saturated used for detecting ; the reaction Ca(OH)₂ + CO₂ → CaCO₃ + H₂O produces an insoluble white precipitate of , turning the clear solution milky. This property underscores the strong basicity and reactivity of these s-block hydroxides in aqueous environments.

and Hydroxides

Transition and hydroxides exhibit diverse structures, ranging from amorphous gels to crystalline layered materials, and display variable reactivities influenced by the metal's d-electron configuration. These compounds are typically insoluble in , forming gelatinous precipitates that play key roles in qualitative inorganic and . Unlike the highly soluble s-block hydroxides, d-block and post-d-block variants often show amphoteric character, allowing dissolution in either acidic or basic conditions, and can adopt multiple oxidation states leading to mixed hydroxide phases. Precipitation reactions are central to identifying these metals in , particularly through the addition of hydroxide ions to form characteristic insoluble hydroxides. For instance, in qualitative analysis schemes, Fe³⁺ ions as red-brown Fe(OH)₃ upon reaction with OH⁻, as seen in group III cation separations where the hydroxide's color and insolubility distinguish iron from other metals. This reaction, Fe³⁺ + 3OH⁻ → Fe(OH)₃ (red-brown ppt), exemplifies the low solubility products (Ksp ≈ 10⁻³⁸) typical of trivalent hydroxides, enabling selective isolation in analytical procedures. Similar precipitations occur for other ions, such as Cr³⁺ forming green Cr(OH)₃, aiding in systematic metal identification. Amphoteric behavior is prominent in hydroxides of post-transition metals like aluminum and transition metals like zinc, where the precipitates redissolve in excess base to form soluble hydroxy complexes. Aluminum hydroxide, Al(OH)₃, initially forms a white gelatinous precipitate but dissolves in strong base via Al(OH)₃ + OH⁻ → [Al(OH)₄]⁻, demonstrating its ability to act as a Lewis acid by accepting additional hydroxide ligands. This property arises from the borderline acidic character of Al³⁺, with the tetrahydroxoaluminate ion stable in alkaline media (pH > 13). Zinc hydroxide exhibits analogous amphoterism, precipitating as white Zn(OH)₂ before dissolving in excess OH⁻ to yield [Zn(OH)₄]²⁻, a process driven by the coordination chemistry of Zn²⁺ in tetrahedral geometry. These reactions highlight the pH-dependent solubility of such hydroxides, contrasting with purely basic behavior in other metals. Variable oxidation states in transition metals lead to a range of hydroxide phases with distinct colors and stabilities. Manganese provides a classic example: , a pale pink precipitate from Mn²⁺ + 2OH⁻, represents the divalent state and is prone to aerial oxidation, while embodies the trivalent form, occurring as a black in hydrothermal deposits with a structure of edge-sharing Mn(III)O₆ octahedra distorted by Jahn-Teller effects. These phases underscore the versatility of , influencing formation and catalytic applications. Magnetic properties of these hydroxides often stem from unpaired d-electrons in the metal centers, conferring . Fe(OH)₃, with Fe(III) in a high-spin d⁵ , exhibits due to five unpaired electrons, as confirmed in synthetic studies where the material shows susceptibility consistent with isolated Fe³⁺ sites before any antiferromagnetic ordering upon aging or . Similarly, Cr(OH)₃ displays from its Cr(III) d³ electrons (three unpaired), with magnetic moments around 3.8 μB, reflecting weak interactions in the polymeric structure. These properties are exploited in nanomaterial design for . Coprecipitation and aging processes enable the formation of mixed hydroxides, notably (LDHs), which consist of brucite-like sheets of divalent and trivalent metals (e.g., Mg²⁺/Al³⁺ or Zn²⁺/Fe³⁺) with intercalated anions. Synthesized by of metal salts at constant pH (typically 8–10), LDHs undergo aging to crystallize into structures, where positive layer charge is balanced by anions like CO₃²⁻ or Cl⁻. The weak electrostatic binding allows facile anion exchange, enabling applications in remediation by swapping interlayer species for pollutants, with exchange capacities up to 3–4 meq/g depending on layer . This versatility arises from the tunable metal ratios and high surface area (50–200 m²/g) post-aging.

Hydroxides of p-Block Elements

The hydroxides of p-block elements, spanning Groups 13 through 16, exhibit predominantly covalent bonding due to the increasing across the block, leading to molecular or polymeric structures rather than ionic lattices. These compounds are often unstable in isolation, prone to , , or conversion to oxyanions, reflecting the tendency of p-block elements to form stable multiple bonds with oxygen. Unlike the more ionic metal hydroxides, p-block variants frequently display acidic character and , with applications limited by their reactivity and in some cases. In Group 13, , B(OH)<sub>3</sub>, is a prototypical covalent hydroxide, existing as a trigonal planar where acts as a Lewis acid by accepting a to form the tetrahedral [B(OH)<sub>4</sub>]<sup>-</sup> in basic solution. It behaves as a very weak monoprotic acid with an K<sub>a</sub> = 5.8 × 10<sup>−10</sup> at 25°C, owing to the electron-deficient center that weakly polarizes an O–H bond in the hydrated form. Aluminum hydroxide, Al(OH)<sub>3</sub>, marks a transition toward more metallic behavior; it is amphoteric, dissolving in both acids and bases, though detailed and aspects overlap with hydroxides. Group 14 hydroxides highlight the distinction between true inorganic species and organic analogs. For carbon, compounds like methanol (CH<sub>3</sub>OH) incorporate an –OH group but function as alcohols with covalent C–O bonds, lacking the ionic OH<sup>-</sup> character of metal hydroxides and instead undergoing nucleophilic substitution or dehydration. Orthosilicic acid, Si(OH)<sub>4</sub>, is a tetrahedral monomer stable only in dilute aqueous solutions below approximately 100 mg/L SiO<sub>2</sub>; it readily undergoes condensation polymerization, releasing water to form siloxane (Si–O–Si) chains and networks that constitute silicates and silica gels. Tin hydroxides, such as Sn(OH)<sub>2</sub> and Sn(OH)<sub>4</sub>, adopt polymeric or cluster structures for stability; for instance, Sn(II) species form cations like [Sn<sub>3</sub>(OH)<sub>4</sub>]<sup>2+</sup> with bridging hydroxo groups, while Sn(IV) variants exhibit octahedral coordination in extended lattices, contributing to their limited solubility and tendency to precipitate as hydrous oxides. In Group 15, , H<sub>3</sub>PO<sub>3</sub>, features a structure with two ionizable –OH groups attached to phosphorus and a direct P–H bond, rendering it diprotic with pK<sub>a</sub> values of 2.00 and 6.59, distinct from the triprotic H<sub>3</sub>PO<sub>4</sub>. Arsenic trihydroxide, As(OH)<sub>3</sub> (also known as arsenious acid, H<sub>3</sub>AsO<sub>3</sub>), is highly toxic, with acute oral exposure causing severe gastrointestinal distress, , and multiorgan failure at doses as low as 70–180 mg; chronic environmental contamination, particularly in , poses global health risks including carcinogenicity and skin lesions, affecting millions through in food chains. Group 16 elements form oxoacids rather than simple hydroxides; , H<sub>2</sub>SO<sub>4</sub>, is structured as (HO)<sub>2</sub>SO<sub>2</sub> with two –OH groups but is classified as a strong diprotic oxoacid, not a hydroxide, serving as a key industrial source of ions via ionization. For , the selenate dianion SeO<sub>4</sub><sup>2-</sup> from selenic acid H<sub>2</sub>SeO<sub>4</sub> can be conceptualized in hydrated basic media, though discrete Se(OH)<sub>6</sub><sup>2-</sup> species are not commonly isolated, unlike in analogous chemistry.

Reactions and Applications

Industrial and Laboratory Uses

Hydroxide compounds, particularly (NaOH), play a central role in various industrial processes. In soap production, NaOH facilitates , the reaction of fats or oils with alkali to form and , serving as a key raw material for both traditional and synthetic detergents. In the for paper manufacturing, NaOH, combined with , treats wood chips under high pressure to dissolve and separate fibers, enabling pulp production that accounts for the majority of global paper output. Additionally, NaOH is essential in the for alumina extraction, where it dissolves aluminum oxide from ore at elevated temperatures and pressures, forming from which pure alumina is subsequently precipitated. Calcium hydroxide, Ca(OH)₂, is widely employed in for adjustment, softening by precipitating calcium and magnesium ions as carbonates, and neutralizing acidic to prevent environmental harm. In technology, (KOH) acts as the in alkaline batteries, enabling the electrochemical reaction between and ; the overall process can be represented as: \mathrm{Zn + 2MnO_2 + H_2O \rightarrow ZnO + 2MnOOH} This reaction provides high energy density and long shelf life, making KOH-based alkaline batteries a staple in consumer electronics. In laboratory settings, NaOH is a standard reagent for acid-base titrations, where it is used to standardize solutions or determine acid concentrations by reaching equivalence points with indicators like phenolphthalein. It is also integral to aqueous workups in organic extractions, where basic NaOH solutions deprotonate acidic compounds to partition them into the aqueous phase, facilitating purification of reaction mixtures. Magnesium hydroxide, Mg(OH)₂, functions as a non-halogenated in polymers and composites, undergoing endothermic above 300°C to release , which dilutes combustible gases and absorbs from the . This mechanism enhances in applications such as electrical cables and construction materials without generating toxic byproducts.

Role in Acid-Base Chemistry

In acid-base chemistry, the hydroxide ion (OH⁻) plays a central role in defining basicity through its concentration in aqueous solutions. The pOH scale measures this concentration logarithmically as pOH = -log[OH⁻], providing a complementary metric to pH, which quantifies hydronium ion (H₃O⁺) activity. At 25°C, the relationship pH + pOH = 14 arises from the ion product of water (K_w = 1.0 × 10⁻¹⁴), establishing neutrality at pH 7 where [OH⁻] = [H₃O⁺] = 1.0 × 10⁻⁷ M. This framework allows chemists to assess solution basicity; for instance, a solution with [OH⁻] = 0.01 M has pOH = 2 and pH = 12, indicating strong basicity. Hydroxide ions are essential in acid-base titrations, particularly when a strong base like NaOH neutralizes a strong acid such as HCl. The reaction proceeds as NaOH + HCl → NaCl + H₂O, with the equivalence point occurring at 7 due to complete neutralization and formation of water under conditions. In titration curves for strong acid-strong base systems, the rises gradually before the equivalence point, then sharply increases beyond it as excess OH⁻ dominates, enabling precise determination of analyte concentration through stoichiometric ratios. In buffering systems, OH⁻ interacts with weak acids to maintain stable . For example, in an acetate buffer comprising acetic acid (CH₃COOH) and its conjugate base (CH₃COO⁻ from ), added OH⁻ reacts with CH₃COOH to form CH₃COO⁻ + H₂O, shifting the without significant pH change according to the Henderson-Hasselbalch . This resistance to pH variation is crucial for applications requiring constant acidity, such as biological assays, where the buffer's capacity depends on the weak acid's pK_a near the desired pH. Acid-base indicators like rely on OH⁻-induced for visual endpoint detection in titrations. undergoes a color change from colorless (protonated form, HIn) to pink (deprotonated form, In²⁻) over the range 8.2–10.0, as OH⁻ shifts the equilibrium HIn ⇌ H⁺ + In²⁻ toward the colored anion in basic media. This transition aligns well with equivalence points in weak acid-strong base titrations, providing a sharp visual cue for completion. For polyprotic acids, OH⁻ facilitates stepwise neutralization, allowing sequential of multiple acidic protons. (H₃PO₄), a triprotic acid, reacts progressively: H₃PO₄ + OH⁻ → H₂PO₄⁻ + H₂O (pK_{a1} ≈ 2.1), H₂PO₄⁻ + OH⁻ → HPO₄²⁻ + H₂O (pK_{a2} ≈ 7.2), and HPO₄²⁻ + OH⁻ → PO₄³⁻ + H₂O (pK_{a3} ≈ 12.7), with full neutralization requiring three equivalents of base to yield PO₄³⁻ + 3H₂O. curves exhibit distinct inflection points corresponding to each pK_a, enabling selective quantification of acid forms in mixtures like fertilizers or biological fluids.

Organic Chemistry of Hydroxide

Base-Catalyzed Reactions

In base-catalyzed reactions, the hydroxide ion (OH⁻) functions primarily as a Brønsted base by abstracting a proton from an , thereby facilitating transformations such as eliminations, condensations, and isomerizations. These processes are fundamental in , where the strong basicity of OH⁻ in protic solvents like or enables the generation of reactive intermediates like carbanions or enolates. The mechanisms typically proceed via concerted or stepwise pathways, influenced by and substrate structure, and are widely studied for their role in both laboratory and industrial applications. A prominent example is the E2 elimination, a concerted bimolecular process where OH⁻ abstracts a β-proton from an while the departs, forming an . In , this is efficient for primary alkyl bromides, such as the conversion of ethyl bromide to ethylene: \text{CH}_3\text{CH}_2\text{Br} + \text{OH}^- \rightarrow \text{CH}_2=\text{CH}_2 + \text{Br}^- + \text{H}_2\text{O} The depends on composition, with ethanol-water mixtures enhancing the elimination over substitution due to reduced of OH⁻, leading to higher basicity. This requires anti-periplanar geometry for optimal orbital overlap and is favored under kinetic control with strong bases like OH⁻. In , OH⁻ catalyzes the self-addition of carbonyl compounds by deprotonating the α-carbon to form an intermediate, which then attacks another . For , the process begins with enolate formation: \text{CH}_3\text{CHO} + \text{OH}^- \rightleftharpoons ^-\text{CH}_2\text{CHO} + \text{H}_2\text{O} followed by to yield the β-hydroxy , and subsequent under basic conditions to the α,β-unsaturated carbonyl. The rate-limiting step often involves the loss of OH⁻ during dehydration, particularly in aqueous media, highlighting the role of OH⁻ in both initiation and termination. This reaction is versatile for C-C bond formation and exemplifies base-promoted chemistry. The of esters, known as , proceeds via a base-catalyzed where OH⁻ attacks the carbonyl carbon, forming a tetrahedral that expels the , yielding a and : \text{RCOOR'} + \text{OH}^- \rightarrow \text{RCOO}^- + \text{R'OH} This addition-elimination pathway is second-order overall, first-order in both ester and OH⁻ concentrations, with the rate-determining step being the formation of the tetrahedral in aqueous solutions. Theoretical studies confirm multiple pathways, but the BAC2 dominates for simple alkyl esters, driven by the basicity of OH⁻. is industrially significant for production from fats. Hofmann elimination involves the thermal decomposition of quaternary ammonium hydroxides, where OH⁻ abstracts a β-proton in an E2-like manner, leading to an , tertiary , and . The general is: \text{R}_4\text{N}^+ + \text{OH}^- \rightarrow \text{alkene} + \text{R}_3\text{N} + \text{H}_2\text{O} This process favors the least substituted due to steric factors in the and may involve an intermediate for certain substrates, enhancing selectivity for terminal s. It is commonly used for exhaustive followed by elimination to determine structures. Base-induced isomerization of alkenes occurs through reversible proton abstraction by OH⁻, allowing double bond migration toward more stable conjugated or internal positions. For 1-butene, OH⁻ in alcoholic media catalyzes the shift to 2-butene via allylic carbanion intermediates, with solvent polarity influencing the equilibrium. This proton transfer mechanism is equilibrium-controlled and is applied in refining processes to optimize alkene stability.

Nucleophilic Reactions

Hydroxide ion (OH⁻) acts as a in by attacking electron-deficient centers, such as carbon atoms bearing good leaving groups or electrophilic carbonyl carbons, leading to or products. This reactivity is particularly prominent in aqueous or alcoholic media, where OH⁻ is generated from like NaOH or KOH. Unlike its role as a in reactions, here the focus is on direct bond formation via nucleophilic attack. In SN2 reactions, OH⁻ displaces leaving groups from primary or methyl alkyl halides through a concerted backside attack, resulting in inversion of at the carbon center. For example, the reaction of methyl iodide with hydroxide yields and iodide ion: \text{CH}_3\text{I} + \text{OH}^- \rightarrow \text{CH}_3\text{OH} + \text{I}^- This mechanism is favored under basic conditions due to the strong nucleophilicity of OH⁻ and the low steric hindrance at primary carbons. The stereochemical inversion is a hallmark of the SN2 pathway, distinguishing it from SN1 processes that involve intermediates. Epoxide ring opening by OH⁻ proceeds via nucleophilic attack at the less substituted carbon under basic conditions, driven by the strain relief in the three-membered ring and steric accessibility. This contrasts with acid-catalyzed openings, where attack occurs at the more substituted site. A classic example is the of (oxirane) to : \ce{(CH2)2O + OH^- -> HOCH2CH2OH} The reaction is typically carried out in aqueous NaOH and is industrially important for glycol production. Nucleophilic acyl substitution with OH⁻ is a key hydrolysis pathway for activated carboxylic acid derivatives, where the hydroxide adds to the carbonyl carbon, forming a tetrahedral that expels the . Acid chlorides react rapidly with OH⁻ to form salts, which upon acidification yield s: \text{RCOCl} + \text{OH}^- \rightarrow \text{RCOO}^- + \text{Cl}^- This process is significantly faster for acid chlorides than for less reactive , where base hydrolysis requires harsher conditions like heating in concentrated NaOH to cleave the amide bond and produce the and . The reactivity order—acid chlorides > anhydrides > esters > —reflects the quality of the and the stability of the tetrahedral . OH⁻ also adds directly to carbonyl groups in certain aldehydes, particularly , forming gem-diols or their conjugate bases in non-catalyzed or base-promoted . For , the involves OH⁻ addition to yield the hydroxymethoxide ion: \text{HCHO} + \text{OH}^- \rightleftharpoons \text{H}_2\text{C(OH)O}^- This lies far toward the due to the lack of steric hindrance and electron-withdrawing effects in , unlike higher aldehydes where is less favorable. Such additions are relevant in aqueous environments and prebiotic chemistry simulations. To extend OH⁻ nucleophilicity to non-aqueous media, phase-transfer catalysis employs lipophilic quaternary salts (e.g., ) to transport the anion from an aqueous base layer into an organic solvent. This enables efficient SN2 displacements, openings, or acyl substitutions in immiscible systems, enhancing reaction rates by increasing local OH⁻ concentration. The catalysts form pairs with OH⁻, solubilizing it without altering its nucleophilic character.

References

  1. [1]
    Hydroxide | HO- | CID 961 - PubChem - NIH
    Hydroxide is an oxygen hydride. It has a role as a mouse metabolite. It is a conjugate base of a water. ChEBI. Hydroxide is a metabolite found in or produced ...
  2. [2]
    Definitions of Acids and Bases, and the Role of Water
    ... ion. An Arrhenius base is any substance that gives the OH-, or hydroxide, ion when it dissolves in water. Arrhenius acids include compounds such as HCl, HCN ...
  3. [3]
    Hydroxide anion | HO - ChemSpider
    Molecular formula: HO. Average mass: 17.007. Monoisotopic mass: 17.003288. ChemSpider ID: 936. Wikipedia. Download .mol. Cite this record.Missing: properties | Show results with:properties
  4. [4]
    Hydroxide ion | SSHADE
    IUPAC name: Hydroxide ; Secondary names: Hydroxide anion, hydroxyl anion, oxidanide, OH- ; InChI: InChI=1S/H2O/h1H2/p-1 ; InChI key: XLYOFNOQVPJJNP-UHFFFAOYSA-M ...
  5. [5]
    Protons and Hydroxide Ions in Aqueous Systems - PMC
    Understanding the structure and dynamics of water's constituent ions, proton and hydroxide, has been a subject of numerous experimental and theoretical studies ...
  6. [6]
    Radii for OH
    ### Ionic Radii for OH⁻
  7. [7]
    [PDF] Ion product of water substance, 0–1000 °C, 1
    Oct 15, 2009 · The ion product of water (Kw) is the molality of hydrogen ion times the molality of hydroxide ion, defined from 0 to 1000°C and 1 to 10,000 ...
  8. [8]
    Vibrational dynamics of aqueous hydroxide solutions studied using ...
    The infrared spectra of aqueous solutions of NaOH and other strong bases exhibit a broad continuum absorption for frequencies between 800 and 3500 cm-¹, which ...
  9. [9]
    Raman spectra from very concentrated aqueous NaOH and from wet ...
    Mar 16, 2006 · High-temperature, high-pressure Raman spectra were obtained from aqueous NaOH solutions up to 2NaOH∕H2O, with XNaOH=0.667 at 480K.
  10. [10]
    Study on the shift of ultraviolet spectra in aqueous solution with ...
    The absorption peaks of the UV spectra near 191 nm of NaCl, NaOH and PEA in aqueous solution moved in the direction of a red shift while the molar absorption ...
  11. [11]
    17O and 1H NMR chemical shifts of hydroxide and hydronium ion in ...
    The 17 O and 1 H NMR chemical shifts of water [δ( 17 OH 2 ), δ( 1 H 2 O)] in aqueous solutions of NaCl, NaBr, NaI, NaNO 3 , NaClO 4 , NaOH, KCl, KBr, ...
  12. [12]
    Electrospray Ionization Mass Spectrometry: A Technique to Access ...
    The Electrospray Ionization (ESI) is a soft ionization technique extensively used for production of gas phase ions (without fragmentation) of thermally labile ...
  13. [13]
    Effects of O18 isotopic substitution on the rotational spectra and ...
    These fields caused shifts in recorded spectral frequencies by as much as several hundred kilohertz and, while their effects on our previously reported value ...
  14. [14]
    The vibration-rotation spectrum of the hydroxide anion (OH-)
    From this work the bond length of hydroxide was obtained, but the vibrational frequency and associated potential con- stants could not be reliably determined.
  15. [15]
    None
    Nothing is retrieved...<|separator|>
  16. [16]
    Benchmarking Quantum Chemical Methods for the Calculation of ...
    The dipole moments and polarizabilities of a set of 46 molecules were calculated using a broad set of quantum chemical methods and basis sets.
  17. [17]
    [PDF] 65 hydroxyl ion as a ltgand
    A hydroxyl ion acts as a ligand if it participates in the inner coordination sphere of a central particle forming thus a hydroxo complex. It is necessary to.
  18. [18]
    Ab initio molecular dynamics studies of hydroxide coordination of ...
    Jun 7, 2019 · The calculated mean U–Oyl bond distances are longer than those reported experimentally but fall in the middle of the range given by ...<|control11|><|separator|>
  19. [19]
    Impact of hydrogen-bonding interactions on the properties of ...
    May 1, 2024 · The molecular and electronic structures of three Co2(μ-OH)2 complexes containing two, one, and zero intramolecular hydrogen-bonding interactions ...
  20. [20]
    Hydrogen Bond - an overview | ScienceDirect Topics
    The hydrogen bond energy is much lower than the energy of the ordinary covalent bond; it does not exceed 0.42 eV (40 kJ/mol).Missing: hydroxide | Show results with:hydroxide
  21. [21]
    Estimating the Hydrogen Bond Energy - ACS Publications
    As can be seen, the hydrogen bond energies vary by more than 17 kJ/mol. The hydrogen bond distance from the isolated water molecule to the molecule in the ...
  22. [22]
    Different structures give similar vibrational spectra: The case of OH
    Feb 11, 2013 · In summary, IR and Raman measurements of an aqueous hydroxide solution yield a broad spectral band in the OH stretching region. This band has ...
  23. [23]
    High-pressure phase of brucite stable at Earth's mantle transition ...
    Nov 21, 2016 · The crystal structure of brucite consists of Mg2+ cations and OH− anions arranged in layers, in an overall trigonal structure (space group ...
  24. [24]
    Structure of Hydrated Gibbsite and Brucite Edge Surfaces: DFT ...
    Jun 28, 2017 · The structure of brucite, the only existing polymorph of magnesium hydroxide, Mg(OH)2, consists of stacking layers built up by edge-sharing Mg( ...
  25. [25]
    Quantum-chemical study of stable, meta-stable and high-pressure ...
    Jun 24, 2014 · 3 Bayerite – α-Al(OH)3. The structure of bayerite is very similar to the structure of gibbsite regarding the layers. The main difference is ...
  26. [26]
    Reorientational motion of the OH− ion in cubic sodium hydroxide
    Abstract. The rotational motion of the OH− ion was studied in cubic NaOH at 575 K with quasielastic incoherent neutron scattering.<|separator|>
  27. [27]
    mp-23879: Ca(HO)2 (Trigonal, P-3m1, 164) - Materials Project
    Ca(OH)₂ crystallizes in the monoclinic C2 space group. The structure is two-dimensional and consists of one Ca(OH)₂ sheet oriented in the (0, 0, 1) direction.
  28. [28]
    Importance of interlayer H bonding structure to the stability ... - Nature
    Oct 16, 2017 · Layered (oxy) hydroxide minerals often possess out-of-plane hydrogen atoms that form hydrogen bonding networks which stabilize the layered ...
  29. [29]
    Cation Vacancy Strategy Activated Layered Double Hydroxides
    ... vacancies (NC(A)L). The cation vacancy defects exposed more active sites and enhanced conductivity, resulting in a remarkable specific capacity of 1434.0 C ...
  30. [30]
    Review Defects engineering of layered double hydroxide-based ...
    Introducing metal vacancies on LDH can modify the oxidation state of metal species in the electrocatalytic process, providing more high-valent metal active ...
  31. [31]
    Lithium Hydroxide | LiOH | CID 3939 - PubChem - NIH
    3 Solubility. Solubility in water, g/100ml at 20 °C: 12.8 (good). ILO-WHO International Chemical Safety Cards (ICSCs). 3.2.4 Density. 2.54 g/cm³. ILO-WHO ...
  32. [32]
    Sodium Hydroxide | NaOH | CID 14798 - PubChem - NIH
    Sodium hydroxide is a white crystalline odorless solid that absorbs moisture from the air. It is a manufactured substance.Missing: cubic | Show results with:cubic
  33. [33]
    Potassium Hydroxide | KOH | CID 14797 - PubChem - NIH
    Solubility in water (g KOH/100 g H2O): 97 at 0 °C; 103 at 10 °C; 112 at 20 °C; 126 ... Solubility in water, g/100ml at 25Â °C: 110 (very good). ILO-WHO ...
  34. [34]
    Ksp Table
    Calcium hydrogen phosphate CaHPO4 1×10–7. Calcium hydroxide Ca(OH)2 5.5×10–6. Calcium oxalate CaC2O4 2.7×10–9. Calcium phosphate Ca3(PO4)2 2.0×10–29. Calcium ...
  35. [35]
    Thermal Dissociation of Sodium Hydroxide upon Evacuation
    Kinetics of dehydration of sodium hydroxide melt upon evacuation was studied in relation to the temperature. The kinetic parameters of thermal decompositio.Missing: stability | Show results with:stability
  36. [36]
    Investigation on the Initial Stage of the Dehydration Process in Mg ...
    Apr 28, 2024 · The dehydration process of Mg(OH)2 to MgO typically occurs in the 300–400 °C range, but it is necessary to lower the dehydration temperature to ...Introduction · Methods · Results and Discussion · Conclusion
  37. [37]
    Sodium hydroxide--water (1/1/1) | H3NaO2 - PubChem
    Sodium hydroxide--water (1/1/1) | H3NaO2 | CID 23719242 - structure, chemical names, physical and chemical properties, classification, patents, literature, ...
  38. [38]
    [PDF] POTASSIUM HYDROXIDE CAS N°: 1310-58-3
    KOH is a white and deliquescent solid with a low vapour pressure. It is a strong alkaline substance that dissociates completely in water to potassium and.
  39. [39]
    [PDF] Limewater and Carbon Dioxide - Flinn Scientific Canada
    Limewater is a saturated solution of calcium hydroxide. When carbon dioxide is bubbled through limewater, a precipitate forms.
  40. [40]
    One-pot synthesis of paramagnetic iron(III) hydroxide nanoplates ...
    Aug 15, 2014 · In this study, Fe(OH) 3 nanoplates and Fe 3 O 4 nanoparticles were successfully prepared using a one-pot hydrothermal process.
  41. [41]
    [PDF] Experiment 2-3 Qualitative Analysis of Metal Ions in Solution
    The iron (III) and Ni (II) are separated as hydroxides by precipitation with an excess of sodium hydroxide in the presence of hydrogen peroxide. Ni2+ + 2OH-. Ni ...
  42. [42]
    Qualitative Analysis - Wired Chemist
    The color of the solid is characteristic of the Fe3+ ion. Since the compound forms an insoluble precipitate with hydroxide, it eliminates the alkali metals, ...
  43. [43]
    Amphoteric Nature of Aluminum Hydroxide | Department of Chemistry
    Aluminum hydroxide is amphoteric, acting as both an acid and a base, dissolving in both NaOH and HCl.
  44. [44]
    Effect of pH and impurities on the surface charge of zinc oxide in ...
    Zinc oxide and zinc hydroxide are amphoteric, dissolving in acids to form salts containing the hydrated zinc (II) cation which may be more or less complexed ...<|separator|>
  45. [45]
    Manganese oxide minerals: Crystal structures and economic and ...
    The manganite (γ-MnOOH) crystal structure is similar to that of pyrolusite but all of the Mn is trivalent and one-half of the O atoms are replaced by hydroxyl ...
  46. [46]
    A novel synthetic route to "iron trihydroxide, Fe(OH)3"
    A novel synthetic route to "iron trihydroxide, Fe(OH)3": characterization and magnetic properties | Inorganic Chemistry.
  47. [47]
    Formation of oxo-centered trinuclear chromium carboxylate ...
    Paramagnetic 2H NMR is a viable technique to study the formation and degradation of trinuclear basic Cr(III) carboxylate assemblies in a variety of solvents.
  48. [48]
    Synthesis, characterization, and applications of layered double ...
    Layers of LDHs are expanded due to the high exchangeability of the anionic species between the layers of the LDHs, and no linking between the hydroxide layers, ...
  49. [49]
    A review on effect of synthesis conditions on the formation of layered ...
    The most common method for the synthesis of layered double hydroxides (LDHs) is co-precipitation. This review is focused on the summarizing of results ...
  50. [50]
    Kraft Pulp Mills: New Source Performance Standards (NSPS)
    Wood chips are treated with a mixture of sodium hydroxide and sodium sulfide and cooked under high pressure to disintegrate the chips into smaller fibers. The ...
  51. [51]
    TENORM: Bauxite and Alumina Production Wastes | US EPA
    May 23, 2025 · The Bayer Process; Waste generation, disposal ... Bauxite ore is dissolved in sodium hydroxide, or lye, at a high temperature and pressure.
  52. [52]
    [PDF] Alkaline Manganese Dioxide - Energizer
    An alkaline battery produces electricity when the manganese dioxide cathode is reduced and the zinc anode becomes oxidized. The equation for a simple alkaline ...
  53. [53]
    [PDF] Overview of Flame Retardants Including Magnesium Hydroxide
    1) The endothermic decomposition commencing at about 330oC for magnesium hydroxide (versus about 230oC for ATH) withdraws heat from the substrate, slowing the ...
  54. [54]
    Flame retardant effects of magnesium hydroxide - ScienceDirect.com
    It can be made synthetically with high purity and in a range of useful morphologies, responds well to surface modifiers and decomposes endothermically with ...
  55. [55]
    15.2 pH and pOH – Chemistry Fundamentals - UCF Pressbooks
    pH is the logarithmic measure of hydronium ion concentration, and pOH is the logarithmic measure of hydroxide ion concentration. pH + pOH = 14 at 25°C.
  56. [56]
    pH, pOH, pK a , and pK b
    pH is calculated as -log[H3O+], pOH as -log[OH-], pKa as -log(Ka), and pKb as -log(Kb). 'p' means the negative of the log of.
  57. [57]
    pH and pOH - Chembook
    pH is a logarithm scale for hydrogen ion concentration, and pOH is related to hydroxide concentration. pH and pOH always sum to 14.
  58. [58]
    15.7 Acid-Base Titrations – Chemistry Fundamentals
    For the titration of a strong acid with a strong base, the equivalence point occurs at a pH of 7.00 and the points on the titration curve can be calculated ...
  59. [59]
    [PDF] Lecture 23: Acid-Base Titrations Part I - MIT OpenCourseWare
    Nov 3, 2014 · ... strong base, or a strong base with a strong acid, the pH changes slowly initially, changes rapidly through pH 7 at the equivalence point and ...
  60. [60]
    The titration of a STRONG base with a STRONG acid
    Titration involves titrating a strong base with a strong acid, where the pH changes from strongly basic to neutral at the equivalence point, then to strongly ...
  61. [61]
    Buffer Solutions
    Example: A buffer solution was made by dissolving 10.0 grams of sodium acetate in 200.0 mL of 1.00 M acetic acid. Assuming the change in volume when the sodium ...
  62. [62]
    15.6 Buffers – Chemistry Fundamentals - UCF Pressbooks
    1). A solution of acetic acid and sodium acetate (CH3COOH + CH3COONa) is an example of a buffer that consists of a weak acid and its salt.
  63. [63]
    8.8: Buffers – CHM130 Fundamental Chemistry
    For example, a buffer can be composed of dissolved HC2H3O2 (a weak acid) and NaC2H3O2 (the salt derived from that weak acid). Another example of a buffer is a ...
  64. [64]
    4.5.1: Acid-Base Indicators - Chemistry LibreTexts
    Dec 7, 2024 · As the pH increase between 8.2 to 10.0 the color becomes red because of the equilibrium shifts to form mostly In- ions. magicpitcher.GIF. The ...Missing: mechanism deprotonation
  65. [65]
    Acid and Base pH Indicator Reference Guide - Alfa Chemistry
    Their color change arises from structural modifications due to protonation or deprotonation. The general dissociation reaction of an indicator (HIn) can be ...<|control11|><|separator|>
  66. [66]
    Diprotic and Triprotic Acids and Bases
    Several important acids can be classified as polyprotic acids, which can lose more than one H+ ion when they act as Brnsted acids. Diprotic acids, such as ...Missing: neutralization NaOH
  67. [67]
    [PDF] Equilibrium Unit Thomas Wenzel Department of Chemistry Bates ...
    For example, consider the series of equations for phosphoric acid. ... pH versus ml titrant for the titration of phosphoric acid (0.1 M, 30 mL) with NaOH.
  68. [68]
    [PDF] Potentiometric titration of a HCl–H3PO4 mixture
    As pH changes, protons are removed from (or added to) the phosphoric acid in a stepwise manner. The complete removal of each proton will be signaled by a ...
  69. [69]
    The influence of solvent composition on the rates of E2 elimination ...
    The overall rates of 2 elimination of hydrogen bromide from β-phenylethyl bromide and ethylene dibromide by alkoxide and hydroxide ions in methanol– and ethanol
  70. [70]
    Theoretical studies of elimination reactions. 2. The importance of ...
    Evidence of a Borderline Region between E1cb and E2 Elimination Reaction Mechanisms ... Carbon−Oxygen Hydrogen Bonding in Dehydrohalogenation Reactions: PM3 ...
  71. [71]
    None
    Nothing is retrieved...<|separator|>
  72. [72]
    Kinetics of the Saponification of the Ethyl Esters of Normal Aliphatic ...
    Determination of the kinetics of biodiesel saponification in alcoholic hydroxide solutions. ... Esterification and ester hydrolysis. 1969, 505-588. https ...
  73. [73]
    Mechanism of the Hofmann Elimination Reaction: An Ylide ...
    Synthesis of trialkylamines with extreme steric hindrance and their decay by a Hofmann-like elimination reaction.
  74. [74]
    SOLVENT EFFECTS IN THE BASE-CATALYZED ISOMERIZATION ...
    Theoretical study of the double bond migration mechanism with participation of hydroxide ion. ... DOUBLE BOND MIGRATION OF ALLYL ETHERS OVER SOLID BASE CATALYSTS.
  75. [75]
    Chapter 7: Nucleophilic attack at the carbonyl carbon: – OCLUE
    Since hydroxide is more reactive than water, the carbonyl does not need to be activated by protonation.
  76. [76]
    [PDF] CHAPTER 9. SUBSTITUTION REACTIONS - Organic Chemistry
    An SN2 mechanism: Here the hydroxide ion in this reaction is acting as a nucleophile (an electron-rich, nucleus-loving species), the carbon ...
  77. [77]
    [PDF] Chapter 9 Alcohols, Ethers, and Epoxides - UCI Open
    PROTONATE THE EPOXIDE FIRST! Notice the regioselectivity in this reaction! • With strong nucleophiles, attack occurs at the less highly substituted carbon. Q ...<|separator|>
  78. [78]
    [PDF] Nucleophilic ring opening of 1,2-epoxides in aqueous medium
    The attack of nucleophile on substituted oxirane ring occurs mainly on the least substituted carbon except when the substituent is an aryl group. X. O. OBz. X.
  79. [79]
    [PDF] Carboxylic Acid Derivatives: Nucleophilic Acyl Substitution 20.1
    Aminolysis: Reaction of acid chlorides with ammonia, 1° or 2° amines to afford amides. 4. Hydrolysis: Acid chlorides react with water to afford carboxylic acids.
  80. [80]
    [PDF] 100 Chapter 21. Carboxylic Acid Derivatives and Nucleophilic Acyl ...
    ... nucleophilic acyl substitution ... acid chlorides with ammonia,. 1° or 2° amines gives amides. ... Hydrolysis- Amides are hydrolyzed to the carboxylic acids and.
  81. [81]
    [PDF] Chapter 17: Aldehydes and Ketones
    Hydrate Formation is Base-Catalyzed. Reaction faster than just with water (pH 7). Hydroxide is much more nucleophilic than water ... Addition to Carbonyls. H3C.Missing: non- | Show results with:non-<|separator|>
  82. [82]
    [PDF] Phase-Transfer Catalysis (PTC) - Macmillan Group
    Apr 10, 2008 · □ Several reactions have been done using chiral quaternary ammonium salts as asymmetric phase transfer catalysts. □ These reactions ...
  83. [83]
    [PDF] Application of phase-transfer catalysis (PTC) to reactions of C-H ...
    ... hydroxide in the presence of a catalyst, a quaternary ammonium salt 1,2 or in aprotic dipolar solvent 3 (phase- transfer catalysis, PTC 4-7). Both CA and ...