Thioester
A thioester is an organosulfur compound featuring the functional group consisting of a carbonyl group (C=O) bonded to a sulfur atom, with the general structure R–C(=O)–S–R', where R and R' are organic substituents.[1] This structure renders thioesters analogous to oxygen esters (R–C(=O)–O–R') but with enhanced reactivity due to the lower electronegativity and greater polarizability of sulfur compared to oxygen.[1] Thioesters exhibit a standard free energy of hydrolysis around –33 kJ/mol under physiological conditions, making them thermodynamically favorable for energy transfer in metabolic processes.[2] In organic chemistry, thioesters serve as versatile intermediates in synthetic reactions, including nucleophilic acyl substitutions such as hydrolysis, alcoholysis, aminolysis, and transthioesterification, where they display higher reactivity toward nucleophiles than typical esters but lower than acid chlorides or anhydrides.[1] Their formation often involves activation of carboxylic acids, for instance, through ATP-dependent coupling with thiols like coenzyme A (CoA) to produce acyl-CoA thioesters.[1] This reactivity stems from the thioester bond's susceptibility to nucleophilic attack at the carbonyl carbon, facilitated by the sulfur atom's ability to stabilize the transition state through resonance and its weaker σ-bond to carbon.[2] Biologically, thioesters are indispensable in central metabolism, acting as activated acyl carriers that enable efficient group transfer without excessive energy input.[2] The most prominent example is acetyl-CoA, a thioester derivative of coenzyme A that links glycolysis to the citric acid cycle by transferring the acetyl group from pyruvate, generated via the pyruvate dehydrogenase complex.[1] Thioesters also drive fatty acid β-oxidation, where acyl-CoA molecules are sequentially shortened to produce acetyl-CoA units for energy production, and they facilitate lipid biosynthesis by activating fatty acids for esterification to glycerol backbones.[1] In protein synthesis pathways, acyl carrier proteins (ACPs) bearing thioester linkages transport growing acyl chains during fatty acid and polyketide assembly.[2] Additionally, thioester bonds appear in the complement system of innate immunity, where internal thioesters in proteins like C3 and C4 enable covalent tagging of pathogens.[2] Beyond metabolism, thioesters contribute to broader biochemical processes, including the formation of iron-sulfur clusters in proteins and serving as precursors in non-ribosomal peptide synthesis.[2] Their high-energy nature allows for the coupling of exergonic hydrolysis to drive endergonic reactions, such as the conversion of thioesters to more stable esters in triglyceride formation, underscoring their role in energy conservation and transfer across cellular pathways.[1]Fundamentals
Definition and Nomenclature
Thioesters are a class of organosulfur compounds characterized by the general formula R–C(=O)–S–R', where R and R' represent organic groups such as alkyl or aryl substituents.[2] This structure distinguishes thioesters from conventional carboxylate esters, R–C(=O)–O–R', by replacing the linking oxygen atom with sulfur. Analogous to the formation of esters, thioesters arise from the condensation reaction between a carboxylic acid and a thiol, releasing water and forming the C–S bond.[3] In systematic IUPAC nomenclature, thioesters are designated as alkyl alkane(thio)ates, where the alkane portion derives from the carboxylic acid and the alkyl group from the thiol, with the sulfur linkage specified by the "thio" prefix and an "S-" locator. For instance, the simple thioester CH₃C(O)SCH₃ is named S-methyl ethanethioate. This naming convention reflects the functional group's role as a derivative of the parent carboxylic acid while highlighting the sulfur substitution.[4] Common names persist in specialized contexts, particularly biochemistry, where thioesters like acetyl-coenzyme A (acetyl-CoA) denote biologically relevant examples with the acyl group bound to the thiol moiety of coenzyme A. Thioesters entered organic synthesis literature in the early 20th century, with initial reports focusing on their preparation and reactivity as acyl transfer agents.[5]Structure and Properties
Thioesters feature a carbonyl group (C=O) covalently linked to a sulfur atom, forming the general structure R-C(=O)-S-R', where R and R' are alkyl or aryl groups. The carbonyl exhibits partial double-bond character due to resonance delocalization, while the C-S bond is longer than typical C-O bonds in esters owing to the larger atomic radius of sulfur. Bond angles around the carbonyl carbon are nearly planar, reflecting sp² hybridization. Resonance in thioesters involves contributions from canonical structures, including the primary form R-C(=O)-S-R' and a minor form R-C(-O⁻)=S⁺-R', where the negative charge resides on oxygen and the positive charge on sulfur. This resonance is less stabilizing than in oxygen esters because sulfur's lower electronegativity (2.58 vs. oxygen's 3.44) reduces effective electron donation to the carbonyl, resulting in greater electrophilicity at the carbonyl carbon. Consequently, thioesters display heightened reactivity toward nucleophiles compared to esters.[6][7] Physically, thioesters possess higher boiling points than analogous esters due to increased molecular polarity from the polarizable sulfur atom; for instance, S-methyl thioacetate boils at 95–96 °C, versus 57 °C for methyl acetate.[4][8] They exhibit good solubility in organic solvents like dichloromethane and ethanol but limited water solubility, and many possess a characteristic sulfurous or unpleasant odor reminiscent of thiols.[9] Chemically, their enhanced electrophilicity leads to greater susceptibility to nucleophilic acyl substitution; their conjugate acids are strong acids, facilitating rapid hydrolysis under acidic conditions. Spectroscopically, thioesters are distinguished by their infrared (IR) carbonyl stretching frequency at 1680–1700 cm⁻¹, lower than the 1730–1750 cm⁻¹ observed for esters, attributable to the reduced resonance stabilization and heavier sulfur mass effect.[10] In ¹³C nuclear magnetic resonance (NMR), the carbonyl carbon resonates at 190–200 ppm, deshielded relative to esters (∼170 ppm) due to diminished electron density from weaker sulfur resonance donation.[11]Synthesis
Classical Methods
Classical methods for the synthesis of thioesters primarily rely on nucleophilic acyl substitution reactions involving activated carboxylic acid derivatives and thiols or their salts, developed in the mid- to late 20th century but rooted in foundational organic chemistry principles. These approaches prioritize straightforward laboratory procedures using readily available reagents, though they often require handling of moisture-sensitive or odorous compounds. Key routes include the reaction of acid chlorides with thiolates, alkylation of thioacetate salts, and ester-thiol exchange, alongside earlier techniques employing sulfur-transfer reagents.[12] The most common classical method involves the reaction of acid chlorides with thiolates, proceeding via nucleophilic attack on the carbonyl carbon to displace chloride. The general equation is: \text{RCOCl} + \text{R'S}^- \rightarrow \text{RC(O)SR'} + \text{Cl}^- This reaction is typically conducted in an aprotic solvent such as diethyl ether or dichloromethane at room temperature, often in the presence of a base like triethylamine to neutralize HCl and generate the thiolate in situ from the thiol. Yields commonly exceed 90% for simple aliphatic and aromatic substrates, making it highly efficient for preparing unsubstituted thioesters.[13] Another established route is the alkylation of thioacetate salts, such as potassium thioacetate, with primary alkyl halides, which furnishes S-alkyl thioacetates that can be further manipulated if needed. The reaction follows an SN2 mechanism: \text{CH}_3\text{COS}^- \text{K}^+ + \text{R-X} \rightarrow \text{CH}_3\text{C(O)SR} + \text{X}^- Performed in polar aprotic solvents like DMF at room temperature, this method achieves high yields (often >80%) with unactivated primary alkyl bromides or iodides, while secondary halides are avoided to prevent elimination side reactions. Its mild conditions and broad tolerance of functional groups contribute to its utility in multistep syntheses.[14] Ester-thiol exchange offers a direct transacylation approach but has limited utility due to the poor leaving group ability of alkoxides compared to thiolates, requiring forcing basic conditions to drive equilibrium toward the thioester. The general transformation is: \text{RC(O)OR''} + \text{R'SH} \rightarrow \text{RC(O)SR'} + \text{R''OH} Typically conducted with strong bases like sodium ethoxide in ethanol or DMSO at elevated temperatures (60–100°C), yields are moderate (50–70%) and restricted to activated esters or simple systems, as competing hydrolysis often predominates. This method is less favored for routine synthesis but provides an alternative when acid chlorides are unsuitable.[12] Early historical methods from the early 20th century laid the groundwork, including direct thionation of carboxylic acids with phosphorus pentasulfide (P4S10 or P2S5) and thiols, which primarily generates dithioesters with thioesters as potential byproducts. These reactions, heated in high-boiling solvents like toluene (reflux, 110°C), suffer from low yields (20–50%) and complex mixtures, limiting their practicality but demonstrating the feasibility of sulfur incorporation without acyl activation. Thiourea derivatives were also explored in initial syntheses, often via hydrolysis to thiols followed by acylation, though these multistep processes were inefficient by modern standards.[15] These classical techniques offer high efficiency for preparing simple thioesters (RC(O)SR') in good to excellent yields, leveraging the nucleophilicity of sulfur and the reactivity of acyl electrophiles. However, they are disadvantaged by the use of toxic or lachrymatory reagents like acid chlorides and phosphorus sulfides, as well as the need for anhydrous conditions to prevent hydrolysis.Modern Methods
Modern methods for thioester synthesis emphasize environmentally benign conditions, high efficiency, and reduced waste compared to classical approaches like the acid chloride route, which often requires harsh reagents and generates significant byproducts. These contemporary strategies typically involve direct coupling of carboxylic acids and thiols under mild temperatures, avoiding toxic intermediates and enabling scalability for pharmaceutical and material applications. One prominent approach utilizes propylphosphonic anhydride (T3P) as a dehydrating agent for the direct coupling of carboxylic acids and thiols to form thioesters, proceeding via activation of the carboxylic acid to a mixed anhydride intermediate. The reaction, represented as \ce{RC(O)OH + R'SH -> RC(O)SR' + H2O}, achieves yields up to 85% under low-temperature conditions (down to -78°C) in solvents like ethyl acetate/acetonitrile, with high enantioselectivity and minimal epimerization for chiral substrates. This method's low toxicity and ease of handling make it suitable for large-scale synthesis, as demonstrated in the preparation of enantiopure thioesters for enzymatic screening.[16] Photocatalytic techniques have emerged as odorless alternatives, employing disulfides and carboxylic acids mediated by phosphine under visible light irradiation. In a typical protocol, carboxylic acids react with symmetrical disulfides in the presence of a phosphine (e.g., tributylphosphine) and a photocatalyst like fac-Ir(ppy)3 under blue LED light (450 nm), generating acyl radicals via phosphoranyl radical fragmentation and incorporating both sulfur atoms from the disulfide into the product. Yields reach up to 90% for diverse aryl and alkyl thioesters, with the process conducted at room temperature in organic solvents, avoiding volatile thiols and enabling late-stage functionalization of complex molecules.[17] Mitsunobu-type reactions provide a stereospecific route for thioester formation, adapting the classic dehydration mechanism to couple carboxylic acids and thiols using azodicarboxylate derivatives and triphenylphosphine. For instance, 4,4'-azopyridine serves as an electron-deficient azo reagent with PPh3 to promote the reaction, yielding thioesters in 70-95% with inversion of configuration at chiral thiol centers when applicable; the byproduct pyridine facilitates easy purification. This variant operates under mild conditions (0-25°C in THF) and tolerates sensitive functional groups, offering advantages over traditional Mitsunobu for sulfur nucleophiles.[18] Post-2010 developments include biocatalytic methods, such as lipase-mediated transesterification for chiral thioesters, leveraging enzyme selectivity for kinetic resolution or direct synthesis. Candida antarctica lipase B (CALB), for example, catalyzes the reaction of thiols with vinyl esters or acids in organic solvents or ionic liquids, achieving enantiomeric excesses >99% for α-chiral thioesters like fenoprofen derivatives under dynamic kinetic resolution conditions at 30-60°C. These green processes avoid metal catalysts and enable high-purity isolation via simple extraction.[19] Carbonylation routes represent another advance, utilizing transition metals to incorporate CO into thioesters from thiols and organohalides or alkenes. Palladium-catalyzed thiocarbonylation of aryl iodides with thiols and CO gas (1-5 atm) in the presence of a base like Et3N proceeds at 80-100°C in DMF, furnishing S-aryl thioesters in 80-98% yields with broad substrate scope, including heterocycles. Recent variants employ CO surrogates like CO2 or silanes to enhance safety and sustainability.[12] Sustainability is a core focus in these methods, with innovations like halide-free activations and aqueous micellar systems minimizing organic solvent use. For example, enzyme-catalyzed thioester formation in designer surfactant micelles (e.g., TPGS-750-M) allows reactions in water at ambient temperature, yielding up to 85% for peptide thioesters while recycling the catalyst medium over multiple cycles, as reported in 2019 green chemistry studies. These approaches reduce E-factors by orders of magnitude compared to classical methods, aligning with principles of atom economy and waste prevention.[20]Reactions
Hydrolysis and Stability
Thioesters undergo acid-catalyzed hydrolysis via a mechanism involving protonation of the carbonyl oxygen, followed by nucleophilic attack of water on the activated carbonyl carbon, leading to a tetrahedral intermediate and eventual expulsion of the thiol leaving group: RCOSR' + H₃O⁺ → RCOOH + R'SH.[21] This process proceeds approximately 10 to 100 times faster for thioesters than for analogous oxygen esters, primarily due to reduced resonance stabilization of the ground state in thioesters, where the poorer orbital overlap between sulfur and the carbonyl π-system results in a more electrophilic carbonyl carbon.[7] Electron-withdrawing substituents on the acyl group (R) further accelerate the rate by enhancing carbonyl electrophilicity, while steric hindrance from bulky R' groups can modestly retard it.[21] In basic conditions, thioesters undergo saponification through a bimolecular base-catalyzed mechanism (BAc₂), where hydroxide attacks the carbonyl to form a tetrahedral intermediate, expelling the thiolate ion: RCOSR' + OH⁻ → RCOO⁻ + R'SH.[21] The thiolate serves as an excellent leaving group owing to its lower basicity (pKₐ of conjugate acid ≈10 for thiols) compared to alkoxide (pKₐ ≈15-16 for alcohols), rendering thioesters significantly more reactive than esters under alkaline conditions—often by factors exceeding 10³. Kinetic studies indicate second-order rate constants for base-mediated hydrolysis on the order of 10⁻¹ to 10⁰ M⁻¹ s⁻¹ at 23°C, depending on the substituents.[22] Thioesters exhibit good thermal stability, remaining intact up to approximately 200°C in the absence of nucleophiles, but they are sensitive to hydrolytic degradation in aqueous environments.[23] In neutral water at pH 7 and 23°C, simple alkyl thioesters like S-methyl thioacetate have half-lives of 100-150 days, substantially shorter than those of corresponding esters (which can exceed years), reflecting their heightened susceptibility to uncatalyzed or mildly catalyzed hydrolysis.[22] The activation energy for thioester hydrolysis typically ranges from 15 to 20 kcal/mol, lower than for many esters due to the favorable leaving group properties.[24] These stability characteristics necessitate careful handling in practical applications; thioesters are commonly stored under inert atmospheres and anhydrous conditions to minimize moisture-induced hydrolysis.[21] In peptide synthesis, such as native chemical ligation, controlled hydrolysis is leveraged for desulfurization steps, but rapid reaction setups and pH buffering are essential to prevent unintended degradation.[21]Nucleophilic Acyl Substitution
Thioesters undergo nucleophilic acyl substitution reactions via an addition-elimination mechanism at the carbonyl carbon, where a nucleophile adds to form a tetrahedral intermediate, followed by expulsion of the thiolate leaving group.[21] This process is facilitated by the relatively poor overlap between the sulfur 3p orbital and the carbonyl π* orbital, which weakens the C-S bond and makes the thiolate (RS⁻) a superior leaving group compared to the alkoxide in esters, rendering thioesters approximately 10³–10⁴ times more reactive toward nucleophiles.[21] Hydrolysis represents a specific instance of this substitution with water or hydroxide as the nucleophile, but thioesters' enhanced reactivity enables efficient synthetic transformations under milder conditions. Aminolysis of thioesters with amines proceeds readily to form amides, as illustrated by the reaction RC(=O)SR' + R''NH₂ → RC(=O)NHR'' + R'SH. This process is a key method for amide bond formation in peptide synthesis, often achieving yields of 70–95% under mild, neutral conditions without racemization.[25] Native chemical ligation exploits thioester reactivity for chemoselective peptide bond formation, coupling a C-terminal thioester peptide with an N-terminal cysteine residue on another fragment. The mechanism begins with transthioesterification, where the cysteine thiol attacks the thioester carbonyl to form a thioester intermediate: \text{RC(=O)SR' + HS-CH}_2\text{-Cys-peptide} \rightarrow \text{RC(=O)S-CH}_2\text{-Cys-peptide + R'SH} This intermediate then undergoes an intramolecular S-to-N acyl shift via a five-membered ring transition state, yielding the native amide-linked product: \text{RC(=O)S-CH}_2\text{-Cys-peptide} \rightarrow \text{RC(=O)NH-CH(CH}_2\text{SH)-peptide} This method has enabled the total synthesis of proteins up to moderate sizes with high fidelity.[26] The Fukuyama coupling provides a convergent route to ketones by reacting thioesters with organozinc reagents under palladium catalysis, as in RC(=O)SR' + R''₂Zn → RC(=O)R'' + R'SZnR''. This reaction tolerates diverse functional groups and proceeds via oxidative addition of the thioester to Pd(0), transmetalation with the organozinc, and reductive elimination, offering a mild alternative to traditional ketone syntheses.[27] Thioesters can also be reduced selectively depending on the reagent: LiAlH₄ delivers two equivalents of hydride to afford primary alcohols, RC(=O)SR' → RCH₂OH + R'SH, analogous to ester reduction. Selective reduction to aldehydes requires specialized reagents such as diisobutylaluminum hydride (DIBAL-H) or silane-mediated methods, often under controlled low-temperature conditions.Biological Significance
Role in Metabolism
Thioesters play a pivotal role in metabolism as high-energy activated acyl carriers, facilitating the transfer of acyl groups in various biosynthetic and catabolic pathways. Their reactivity stems from the thioester bond, which is more labile than oxygen esters due to the poorer orbital overlap between sulfur and carbonyl carbon, enabling efficient nucleophilic acyl substitutions under physiological conditions. A prime example is acetyl-coenzyme A (acetyl-CoA), with the structure CH₃C(O)-S-CoA, where the acetyl group is linked via a thioester to the sulfhydryl group of coenzyme A. Acetyl-CoA serves as a central intermediate in the tricarboxylic acid (TCA) cycle, where it condenses with oxaloacetate to form citrate, initiating the cycle that generates reducing equivalents for ATP production. It is also crucial in fatty acid β-oxidation, where sequential cleavage of acyl-CoA thioesters produces acetyl-CoA units from fatty acids, fueling energy production. Additionally, acetyl-CoA donates acetyl groups for acetylation reactions, including histone modification and the synthesis of acetylcholine and other metabolites.[28][29] In fatty acid synthesis, malonyl-CoA (HOOC-CH₂C(O)-S-CoA) acts as the key two-carbon donor. Formed by carboxylation of acetyl-CoA, malonyl-CoA undergoes decarboxylative condensation with growing acyl chains on acyl carrier proteins, extending the chain by two carbons while releasing CO₂ to drive the reaction forward. This process, repeated iteratively, builds long-chain fatty acids from acetyl-CoA precursors.[30] Acyl carrier proteins (ACPs) are small proteins featuring a 4'-phosphopantetheine prosthetic group that forms thioester linkages with acyl intermediates. In type II fatty acid synthases and polyketide synthases, ACP-bound thioesters shuttle malonyl and acyl units between enzymatic domains, enabling modular assembly of fatty acids and complex polyketides through repeated condensation and reduction steps. Similarly, in non-ribosomal peptide synthetases, peptidyl carrier proteins (PCPs), which are ACP homologs, carry aminoacyl thioesters for the synthesis of peptide natural products like antibiotics and siderophores.[31][32] Thioester intermediates are also essential in ubiquitin conjugation, a post-translational modification signaling protein degradation. Ubiquitin forms a thioester bond with the catalytic cysteine of E1 activating enzymes via ATP-dependent adenylation, followed by transthiolation to E2 conjugating enzymes, which then transfer ubiquitin to E3 ligases for attachment to target proteins, marking them for proteasomal degradation.[33][34] The energetics of thioester hydrolysis underscore their metabolic utility; for acetyl-CoA, the standard free energy change (ΔG°') for hydrolysis to acetate and CoA is approximately -35.7 kJ/mol (-8.5 kcal/mol), slightly more exergonic than ATP hydrolysis (-30.5 kJ/mol or -7.3 kcal/mol), providing the thermodynamic driving force for coupling to endergonic biosynthetic reactions.[35][36]Origin of Life Hypothesis
In the "Thioester World" hypothesis, proposed by biochemist Christian de Duve in 1991, thioesters are envisioned as central energy-rich molecules in a prebiotic phase of Earth's history, potentially predating the RNA world and serving as precursors to modern high-energy phosphates like ATP by providing comparable free energy for bond formation and catalysis. This scenario suggests that thioesters, with their activated carbonyl groups, could have driven primitive metabolic cycles and polymerizations without enzymatic assistance, bridging geochemical processes to the emergence of replicative biochemistry. Prebiotic synthesis of thioesters is proposed to occur in hydrothermal vent environments, where carbon dioxide (CO₂), hydrogen sulfide (H₂S), and iron- or nickel-sulfur minerals interact under reducing conditions, mimicking Fischer-Tropsch-type catalysis to form simple thioesters like methyl thioacetate. These reactions leverage geochemical gradients, such as pH and redox potentials at alkaline vents, to fix CO₂ into organic sulfur compounds, providing a plausible abiotic route to activated acyl groups essential for early metabolism.[37] Thioesters' enhanced reactivity—stemming from the weak S-C bond and nucleophilic sulfur—enables them to participate in non-enzymatic polymerization reactions, such as the formation of proto-peptides from amino acids and mercaptoacids, or the activation of nucleotides, surpassing the lower efficiency of ordinary esters in aqueous prebiotic settings.[38] This property allows thioesters to act as acyl donors in condensation reactions, facilitating the assembly of biopolymers under mild conditions without dehydrating agents.[39] Experimental support includes simulations of primordial atmospheres and vents yielding thioesters; for instance, geoelectrochemical experiments with Ni-sulfides and CO₂/H₂S mixtures produce thioacetic acid derivatives, while modified spark-discharge setups incorporating sulfur gases generate sulfur-containing organics akin to thioesters. In the 2010s, kinetic studies confirmed thioester stability in acidic, low-temperature aqueous media relevant to early Earth, with half-lives extending to hours under vent-like conditions, though rapid thiol-thioester exchange supports dynamic polymerization cycles. More recently, in August 2025, researchers at University College London demonstrated that thioesters can catalyze the formation of peptide bonds by linking amino acids to RNA in evaporating water droplets, providing evidence that bridges the thioester world hypothesis with the RNA world.[40] Criticisms of the hypothesis center on thioesters' vulnerability to hydrolysis in neutral-to-alkaline aqueous environments, where half-lives drop to seconds at elevated temperatures, potentially undermining their accumulation in a global ocean; alternatives, such as phosphate anhydrides, are favored for their greater persistence in such settings. Despite these challenges, thioesters like acetyl-CoA in modern metabolism may trace evolutionary roots to these prebiotic intermediates.Related Compounds
Thionoesters
Thionoesters are organic compounds featuring a thiocarbonyl functional group with the general structure R–C(=S)–OR', where the sulfur replaces the oxygen in the carbonyl of regular esters, making them isomers of thioesters (R–C(=O)–SR'). This arrangement positions the sulfur in the double bond, as exemplified by methyl thionobenzoate (C₆H₅C(=S)OCH₃). The thiocarbonyl moiety in thionoesters enhances reactivity toward nucleophilic attack at the C=S bond due to the lower bond energy and increased electrophilicity compared to C=O bonds. In infrared spectroscopy, the characteristic C=S stretching vibration appears in the 1100–1300 cm⁻¹ range.[41] Synthesis of thionoesters typically involves the esterification of thioacyl chlorides with alcohols: RC(=S)Cl + R'OH → RC(=S)OR' + HCl. An alternative route employs thionation of conventional esters using Lawesson's reagent, which replaces the carbonyl oxygen with sulfur: RC(=O)OR' → RC(=S)OR'.[42] Thionoesters exhibit limited stability, often undergoing tautomerization or decomposition, such as conversion to dithiocarboxylic acids under basic conditions via carbanion intermediates reacting with carbon disulfide. This instability restricts their commercial applications despite early isolations reported in the 1950s.[43]Comparisons to Esters
Thioesters and carboxylate esters differ significantly in reactivity, with thioesters serving as superior acyl donors in nucleophilic acyl substitution reactions. Thioesters undergo hydrolysis substantially faster than esters, often by factors of 10 to 10^5 depending on conditions, due to the greater electrophilicity of their carbonyl carbon.[44][45] This heightened reactivity stems from the poorer resonance donation by sulfur compared to oxygen; the larger 3p orbitals of sulfur provide less effective overlap with the carbonyl carbon's 2p orbitals, reducing ground-state stabilization and lowering the activation barrier for nucleophilic attack. Consequently, thioesters are more prone to transacylation and exchange reactions, making them valuable in synthetic applications like native chemical ligation for peptide assembly, where their reactivity enables efficient coupling without harsh conditions.[46] In contrast, esters exhibit greater stability, rendering them suitable for constructing durable polymers such as polyethylene terephthalate (PET), widely used in textiles and packaging.[47] Physically, thioesters display a lower carbonyl stretching frequency in infrared spectroscopy, typically around 1680–1700 cm⁻¹, compared to 1730–1750 cm⁻¹ for esters, reflecting a weaker C=O bond from diminished resonance involvement. Thioesters also tend to have higher boiling points than analogous esters, attributable to increased molecular polarizability from the larger sulfur atom. In biological contexts, esters predominate in stable structures like triglycerides in lipids, providing energy storage through slow hydrolysis, whereas thioesters, such as acetyl-CoA, function as high-energy activated carriers for rapid acyl group transfer in metabolic pathways.[48]| Property | Thioester | Carboxylate Ester |
|---|---|---|
| C-X Bond Energy (kJ/mol) | C-S: 272[49] | C-O: 358[49] |
| Leaving Group Ability | Thiolate (pK_a of RSH ≈ 10, weaker base, better leaving group)[6] | Alkoxide (pK_a of ROH ≈ 15–16, stronger base, poorer leaving group)[6] |
| Example Reaction: Basic Hydrolysis | Rapid (e.g., rate constant k ≈ 0.1–1 M⁻¹ s⁻¹ for simple models, proceeds via tetrahedral intermediate collapse favoring thiol departure)[44] | Slower (e.g., rate constant k ≈ 0.1 M⁻¹ s⁻¹, limited by alkoxide departure)[50] |