Fact-checked by Grok 2 weeks ago

Thioester

A thioester is an organosulfur compound featuring the consisting of a (C=O) bonded to a atom, with the general structure R–C(=O)–S–R', where R and R' are organic substituents. This structure renders thioesters analogous to oxygen esters (R–C(=O)–O–R') but with enhanced reactivity due to the lower and greater of compared to oxygen. Thioesters exhibit a standard free energy of around –33 kJ/mol under physiological conditions, making them thermodynamically favorable for energy transfer in metabolic processes. In , thioesters serve as versatile intermediates in synthetic reactions, including nucleophilic acyl substitutions such as , alcoholysis, aminolysis, and transthioesterification, where they display higher reactivity toward nucleophiles than typical esters but lower than acid chlorides or anhydrides. Their formation often involves activation of carboxylic acids, for instance, through ATP-dependent coupling with thiols like (CoA) to produce thioesters. This reactivity stems from the thioester bond's susceptibility to nucleophilic attack at the carbonyl carbon, facilitated by the atom's ability to stabilize the through and its weaker σ-bond to carbon. Biologically, thioesters are indispensable in central metabolism, acting as activated acyl carriers that enable efficient group transfer without excessive energy input. 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. 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. In protein synthesis pathways, acyl carrier proteins (ACPs) bearing thioester linkages transport growing acyl chains during fatty acid and polyketide assembly. 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. Beyond , thioesters contribute to broader biochemical processes, including the formation of iron-sulfur clusters in proteins and serving as precursors in non-ribosomal . Their high-energy nature allows for the coupling of exergonic to drive endergonic reactions, such as the conversion of thioesters to more stable esters in formation, underscoring their role in and transfer across cellular pathways.

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 groups such as alkyl or aryl substituents. This structure distinguishes thioesters from conventional esters, R–C(=O)–O–R', by replacing the linking oxygen atom with . Analogous to the formation of esters, thioesters arise from the between a and a , releasing water and forming the C–S bond. In systematic IUPAC , thioesters are designated as alkyl (thio)ates, where the alkane portion derives from the and the alkyl group from the , 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 reflects the functional group's role as a derivative of the parent while highlighting the substitution. 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.

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 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 . This is less stabilizing than in oxygen esters because 's lower (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. Physically, thioesters possess higher boiling points than analogous esters due to increased molecular from the polarizable atom; for instance, S-methyl thioacetate boils at 95–96 °C, versus 57 °C for . They exhibit good in organic solvents like and but limited , and many possess a characteristic sulfurous or unpleasant reminiscent of thiols. Chemically, their enhanced electrophilicity leads to greater susceptibility to ; their conjugate acids are strong acids, facilitating rapid under acidic conditions. Spectroscopically, thioesters are distinguished by their () carbonyl stretching frequency at 1680–1700 cm⁻¹, lower than the 1730–1750 cm⁻¹ observed for esters, attributable to the reduced stabilization and heavier sulfur mass effect. In ¹³C (NMR), the carbonyl carbon resonates at 190–200 , deshielded relative to esters (∼170 ) due to diminished from weaker sulfur donation.

Synthesis

Classical Methods

Classical methods for the synthesis of thioesters primarily rely on reactions involving activated derivatives and thiols or their salts, developed in the mid- to late but rooted in foundational 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, of thioacetate salts, and ester-thiol exchange, alongside earlier techniques employing sulfur-transfer reagents. 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. Another established route is the of thioacetate salts, such as thioacetate, with primary alkyl halides, which furnishes S-alkyl thioacetates that can be further manipulated if needed. The 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 , 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. 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 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 in or DMSO at elevated temperatures (60–100°C), yields are moderate (50–70%) and restricted to activated esters or simple systems, as competing often predominates. This method is less favored for routine but provides an when acid chlorides are unsuitable. Early historical methods from the early laid the groundwork, including direct thionation of carboxylic acids with (P4S10 or P2S5) and thiols, which primarily generates dithioesters with thioesters as potential byproducts. These reactions, heated in high-boiling solvents like (, 110°C), suffer from low yields (20–50%) and complex mixtures, limiting their practicality but demonstrating the feasibility of sulfur incorporation without acyl activation. derivatives were also explored in initial syntheses, often via to thiols followed by , though these multistep processes were inefficient by modern standards. These classical techniques offer high efficiency for preparing simple thioesters (RC(O)SR') in good to excellent yields, leveraging the nucleophilicity of and the reactivity of acyl electrophiles. However, they are disadvantaged by the use of toxic or lachrymatory reagents like acid chlorides and , as well as the need for conditions to prevent .

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 of carboxylic acids and thiols under mild temperatures, avoiding toxic intermediates and enabling for pharmaceutical and material applications. One prominent approach utilizes propylphosphonic anhydride (T3P) as a dehydrating agent for the direct coupling of s and thiols to form thioesters, proceeding via activation of the 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 /, 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. Photocatalytic techniques have emerged as odorless alternatives, employing disulfides and carboxylic acids mediated by under visible . In a typical protocol, carboxylic acids react with symmetrical s in the presence of a (e.g., ) and a photocatalyst like fac-Ir(ppy)3 under LED (450 nm), generating acyl radicals via phosphoranyl radical fragmentation and incorporating both sulfur atoms from the into the product. Yields reach up to 90% for diverse aryl and alkyl thioesters, with the process conducted at in solvents, avoiding volatile thiols and enabling late-stage functionalization of complex molecules. Mitsunobu-type reactions provide a stereospecific route for thioester formation, adapting the classic mechanism to couple carboxylic acids and s using azodicarboxylate derivatives and . For instance, 4,4'-azopyridine serves as an electron-deficient azo reagent with PPh3 to promote the , yielding thioesters in 70-95% with inversion of at chiral centers when applicable; the byproduct 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. Post-2010 developments include biocatalytic methods, such as lipase-mediated for chiral thioesters, leveraging selectivity for kinetic or direct . Candida antarctica lipase B (CALB), for example, catalyzes the reaction of thiols with vinyl esters or acids in solvents or ionic liquids, achieving enantiomeric excesses >99% for α-chiral thioesters like fenoprofen derivatives under dynamic kinetic conditions at 30-60°C. These green processes avoid metal catalysts and enable high-purity isolation via simple extraction. Carbonylation routes represent another advance, utilizing transition metals to incorporate into thioesters from thiols and organohalides or alkenes. Palladium-catalyzed thiocarbonylation of aryl iodides with thiols and gas (1-5 atm) in the presence of a like Et3N proceeds at 80-100°C in DMF, furnishing S-aryl thioesters in 80-98% yields with broad scope, including heterocycles. Recent variants employ CO surrogates like CO2 or silanes to enhance and . 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 micelles (e.g., TPGS-750-M) allows reactions in at ambient , yielding up to 85% for peptide thioesters while recycling the catalyst medium over multiple cycles, as reported in 2019 studies. These approaches reduce E-factors by orders of magnitude compared to classical methods, aligning with principles of and waste prevention.

Reactions

Hydrolysis and Stability

Thioesters undergo acid-catalyzed via a involving of the carbonyl oxygen, followed by nucleophilic attack of on the activated carbonyl carbon, leading to a tetrahedral and eventual expulsion of the : RCOSR' + H₃O⁺ → RCOOH + R'SH. This process proceeds approximately 10 to 100 times faster for thioesters than for analogous oxygen esters, primarily due to reduced stabilization of the in thioesters, where the poorer orbital overlap between and the carbonyl π-system results in a more electrophilic carbonyl carbon. Electron-withdrawing substituents on the (R) further accelerate the rate by enhancing carbonyl electrophilicity, while steric hindrance from bulky R' groups can modestly retard it. 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. 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. 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. In neutral at 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 . The activation energy for thioester typically ranges from 15 to 20 kcal/mol, lower than for many esters due to the favorable properties. These stability characteristics necessitate careful handling in practical applications; thioesters are commonly stored under inert atmospheres and anhydrous conditions to minimize moisture-induced hydrolysis. 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.

Nucleophilic Acyl Substitution

Thioesters undergo reactions via an addition-elimination at the carbonyl carbon, where a adds to form a tetrahedral , followed by expulsion of the thiolate . This process is facilitated by the relatively poor overlap between the 3p orbital and the carbonyl π* orbital, which weakens the C-S bond and makes the thiolate (RS⁻) a superior compared to the in esters, rendering thioesters approximately 10³–10⁴ times more reactive toward . represents a specific instance of this substitution with or as the , 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. 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. 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. 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. In , (HOOC-CH₂C(O)-S-CoA) acts as the key two-carbon donor. Formed by of , 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 s from precursors. Acyl carrier proteins (ACPs) are small proteins featuring a 4'-phosphopantetheine that forms thioester linkages with acyl intermediates. In type II synthases and synthases, ACP-bound thioesters shuttle and acyl units between enzymatic domains, enabling modular assembly of s and complex polyketides through repeated condensation and reduction steps. Similarly, in non-ribosomal synthetases, peptidyl carrier proteins (PCPs), which are ACP homologs, carry aminoacyl thioesters for the of natural products like antibiotics and siderophores. Thioester intermediates are also essential in ubiquitin conjugation, a post-translational modification signaling protein degradation. forms a thioester bond with the catalytic of E1 activating enzymes via ATP-dependent adenylation, followed by transthiolation to E2 conjugating enzymes, which then transfer to ligases for attachment to target proteins, marking them for proteasomal degradation. 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 (-30.5 kJ/mol or -7.3 kcal/mol), providing the thermodynamic driving force for coupling to endergonic biosynthetic reactions.

Origin of Life Hypothesis

In the "Thioester World" hypothesis, proposed by biochemist in 1991, thioesters are envisioned as central energy-rich molecules in a prebiotic phase of Earth's history, potentially predating the and serving as precursors to modern high-energy phosphates like ATP by providing comparable for bond formation and . 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 environments, where (CO₂), (H₂S), and iron- or nickel-sulfur minerals interact under reducing conditions, mimicking Fischer-Tropsch-type to form simple thioesters like methyl thioacetate. These reactions leverage geochemical gradients, such as and potentials at alkaline vents, to fix CO₂ into sulfur compounds, providing a plausible abiotic route to activated acyl groups essential for early . Thioesters' enhanced reactivity—stemming from the weak S-C bond and nucleophilic —enables them to participate in non-enzymatic polymerization reactions, such as the formation of proto-peptides from and mercaptoacids, or the activation of , surpassing the lower efficiency of ordinary esters in aqueous prebiotic settings. This property allows thioesters to act as acyl donors in condensation reactions, facilitating the assembly of under mild conditions without dehydrating agents. Experimental support includes simulations of primordial atmospheres and vents yielding thioesters; for instance, geoelectrochemical experiments with Ni-sulfides and CO₂/H₂S mixtures produce derivatives, while modified spark-discharge setups incorporating gases generate sulfur-containing organics akin to thioesters. In the , kinetic studies confirmed thioester stability in acidic, low-temperature aqueous media relevant to , 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 demonstrated that thioesters can catalyze the formation of peptide bonds by linking to in evaporating water droplets, providing evidence that bridges the thioester world hypothesis with the . 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 ; alternatives, such as phosphate anhydrides, are favored for their greater persistence in such settings. Despite these challenges, thioesters like in modern may trace evolutionary roots to these prebiotic intermediates.

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. 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 , which replaces the carbonyl oxygen with : RC(=O)OR' → RC(=S)OR'. Thionoesters exhibit limited stability, often undergoing tautomerization or decomposition, such as conversion to dithiocarboxylic acids under basic conditions via intermediates reacting with . This instability restricts their commercial applications despite early isolations reported in the .

Comparisons to Esters

Thioesters and carboxylate esters differ significantly in reactivity, with thioesters serving as superior acyl donors in reactions. Thioesters undergo substantially faster than esters, often by factors of 10 to 10^5 depending on conditions, due to the greater electrophilicity of their carbonyl carbon. This heightened reactivity stems from the poorer donation by 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. In contrast, esters exhibit greater stability, rendering them suitable for constructing durable polymers such as (), widely used in textiles and packaging. Physically, thioesters display a lower carbonyl in , typically around 1680–1700 cm⁻¹, compared to 1730–1750 cm⁻¹ for esters, reflecting a weaker C=O bond from diminished involvement. Thioesters also tend to have higher points than analogous esters, attributable to increased molecular from the larger atom. In biological contexts, esters predominate in stable structures like triglycerides in , providing through slow , whereas thioesters, such as , function as high-energy activated carriers for rapid transfer in metabolic pathways.
PropertyThioesterCarboxylate Ester
C-X Bond Energy (kJ/mol)C-S: 272C-O: 358
Leaving Group AbilityThiolate (pK_a of RSH ≈ 10, weaker base, better leaving group)Alkoxide (pK_a of ROH ≈ 15–16, stronger base, poorer leaving group)
Example Reaction: Basic HydrolysisRapid (e.g., rate constant k ≈ 0.1–1 M⁻¹ s⁻¹ for simple models, proceeds via tetrahedral intermediate collapse favoring thiol departure)Slower (e.g., rate constant k ≈ 0.1 M⁻¹ s⁻¹, limited by alkoxide departure)

References

  1. [1]
  2. [2]
    Thioester - an overview | ScienceDirect Topics
    Thioester is defined as a compound, such as coenzyme A, that plays a significant role in biosynthesis, characterized by its exchange rate in aqueous solution ...
  3. [3]
    Thioesters provide a plausible prebiotic path to proto-peptides - NIH
    May 11, 2022 · It is known that thiols can condense with carboxylic acids to form thioesters (Fig. 1, top). Further, thioester linkages can be converted to ...
  4. [4]
    S-Methyl thioacetate | C3H6OS | CID 73750 - PubChem - NIH
    S-methyl thioacetate is a thioester. It is functionally related to an ethanethioic S-acid. ... S-Methyl thioacetate has been reported in Allium cepa and Solanum ...
  5. [5]
  6. [6]
    Thioester: Bonding, Synthesis, and Reactions - Chemistry Learner
    Its general formula is R–C(=O)–SR′. [1-3] Thioesters play a central role in both chemistry and biology.Missing: definition | Show results with:definition
  7. [7]
    CHEM 440 - Thioesters
    Dec 10, 2016 · Thioesters are effective acylating agents. (2) Thioesters are also good alkylating agents. Why are thioesters particularly suited for these metabolic roles?Missing: definition | Show results with:definition
  8. [8]
    The geometry of the thioester group and its implications for the ...
    The thioester group's acyl C-S bond is not significantly shorter than other C(sp2)-S bonds, unlike esters, and has little double bond character.
  9. [9]
    Methyl acetate - Wikipedia
    Methyl acetate ; Boiling point, 56.9 °C (134.4 °F; 330.0 K) ; Solubility in water. ~25% (20 °C) ; Vapor pressure, 173 mmHg (20°C) ; Magnetic susceptibility (χ). - ...
  10. [10]
    methyl thioacetate, 1534-08-3 - The Good Scents Company
    S-methyl ethanethioate. CAS Number: 1534-08-3 ... Olfactory investigation of a library of S-methyl thioesters and sensory evaluation of selected components.
  11. [11]
    A carbon-13 nuclear magnetic resonance study of thiol esters
    13 C NMR Chemical shifts and carbon-proton coupling constants of some furocoumarins and furochromones. Phytochemistry 1979, 18 (1) , 139-143. https://doi ...Missing: carbonyl | Show results with:carbonyl
  12. [12]
  13. [13]
  14. [14]
    [PDF] Conversion of Esters to Thioesters under Mild Conditions
    Mar 15, 2021 · thiocarboxylates, condensation of carboxylic acids with thiols,6,7. Mitsunobu reaction of alcohols with thioacetic acids,8 and carbonylation ...
  15. [15]
    [PDF] Phosphorus Pentasulfide: A Mild and Versatile Catalyst/Reagent for ...
    The reaction of carboxylic acids (1) with a variety of thiols or alcohols in the presence of phosphorus pentasulfide (P4S10) as a catalyst and reagent (20−40 ...
  16. [16]
    [PDF] Synthesis of an enantiopure thioester as key substrate for screening ...
    The activation was effected by propylphosphonic anhydride (T3P) with (R)-1 and 4 at three different temperatures. Solvents used for the coupling were adapted to.
  17. [17]
    Bridging the gap between transition metal- and bio-catalysis via ...
    May 15, 2019 · A strategic Ser/Cys exchange in the catalytic triad unlocks an acyltransferase-mediated synthesis of thioesters and tertiary amides. Article ...
  18. [18]
    Kinetics and Mechanisms of Reactions of Thiol, Thiono, and Dithio ...
    Kinetics and Mechanisms of Reactions of Thiol, Thiono, and Dithio Analogues of Carboxylic Esters with Nucleophiles | Chemical Reviews.
  19. [19]
  20. [20]
    Synthesis of (Bio)‐Degradable Poly(β‐thioester)s via Amine ...
    Oct 22, 2012 · Thermal characterization of the polymer shows a melting temperature of 53 °C and good thermal stability of up to 200 °C. The poly(β-thioester) ...
  21. [21]
    Mechanistic investigation of SARS-CoV-2 main protease to ... - Nature
    Dec 5, 2022 · The free energy barrier for this concerted process is ΔG‡ = 14.9 ± 0.1 kcal/mol and with the resulting thioester at ΔG° = 9.7 ± 0.1 kcal/mol ...Cleavage Of The Nsp Fragment · Hydrolysis Of The Thioester... · Md Simulations<|control11|><|separator|>
  22. [22]
    Synthesis of Cyclic Peptides through Direct Aminolysis of ... - NIH
    A promising method for the synthesis of cyclic peptides through the direct aminolysis of peptide thioesters is presented.
  23. [23]
    Synthesis of Proteins by Native Chemical Ligation - Science
    A simple technique has been devised that allows the direct synthesis of native backbone proteins of moderate size.
  24. [24]
    Fukuyama Coupling - Organic Chemistry Portal
    The palladium-catalyzed coupling of organozinc compounds with thioesters to form ketones. This reaction tolerates a variety of functional groups.
  25. [25]
    Acetyl-CoA and the Regulation of Metabolism: Mechanisms ... - NIH
    Feb 20, 2015 · Acetyl-CoA is a metabolite derived from glucose, fatty acid, and amino acid catabolism. During glycolysis, glucose is broken down into two three ...
  26. [26]
    Acetyl Coenzyme A: A Central Metabolite and Second Messenger
    Jun 2, 2015 · Myc-dependent mitochondrial generation of acetyl-CoA contributes to fatty acid biosynthesis and histone acetylation during cell cycle entry.
  27. [27]
    Activity of fatty acid biosynthesis initiating ketosynthase FabH ... - NIH
    The rate of malonyl-CoA decarboxylation by FabH C112Q was ¼ the rate of malonyl-CoA decarboxylative Claisen condensation by the wild-type enzyme at pH 8.
  28. [28]
    Probing the structure and function of acyl carrier proteins to unlock ...
    Acyl carrier proteins (ACPs) play a central role in biosynthesis by shuttling malonyl-based building blocks and polyketide intermediates to catalytic partners.
  29. [29]
    Carrier Protein Structure and Recognition in Polyketide and ...
    Carrier Proteins Involved in Polyketide and Nonribosomal Peptide Biosynthesis. In PK and NRP biosynthesis, the acyl or peptide backbone is built up from acyl ...
  30. [30]
    Structure of a ubiquitin E1-E2 complex - NIH
    Ubiquitin (Ub) conjugation is initiated by an E1 enzyme that catalyzes carboxy-terminal Ub adenylation, thioester bond formation to a catalytic cysteine in the ...Missing: degradation | Show results with:degradation
  31. [31]
    Protein ubiquitination involving an E1–E2–E3 enzyme ... - Nature
    Jan 5, 1995 · This cascade of ubiquitin thioester complexes suggests that E3s have a defined enzymatic activity and do not function simply as docking proteins.
  32. [32]
    Change in Gibbs free energy in hydrolysis of Acetyl CoA - BioNumbers
    Change in Gibbs free energy in hydrolysis of Acetyl CoA ; -35.7 kJ/mol Range: Table - link kJ/mol · Unspecified · Thauer RK, Jungermann K, Decker K. Energy ...
  33. [33]
    Metabolic Energy - The Cell - NCBI Bookshelf - NIH
    In the hydrolysis of ATP to ADP plus phosphate (Pi), ΔG°′= -7.3 kcal/mol. Recall, however, that ΔG°′ refers to “standard conditions,” in which the ...
  34. [34]
    Supercritical carbon dioxide likely served as a prebiotic source of ...
    Oct 3, 2024 · Abiotic synthesis of methanethiol through CO 2 reduction is a presumptive initiation reaction of protometabolism in primordial ocean hydrothermal systems.
  35. [35]
    Thioesters provide a plausible prebiotic path to proto-peptides - Nature
    May 11, 2022 · It is known that thiols can condense with carboxylic acids to form thioesters (Fig. 1, top). Further, thioester linkages can be converted to ...<|separator|>
  36. [36]
    Thioesters Provide a Robust Path to Prebiotic Peptides - ChemRxiv
    Jan 18, 2021 · Thioesters are hypothesized to have played key roles in prebiotic chemistry on early Earth, serving as energy storing molecules, as synthetic ...
  37. [37]
    [PDF] Selected Reactions of Thiocarbonyl Compounds - stoltz2.caltech.edu
    IR spectroscopy: C=S stretching vibration is only of medium intensity, 1100-1300 cm-1 . Enethiols absorb around 2550 cm-1 (SH) and 1640 cm-1 (C=C). 13C NMR ...Missing: thionoesters | Show results with:thionoesters
  38. [38]
    Lawesson's Reagent - Organic Chemistry Portal
    Lawesson's Reagent is a mild and convenient thionating agent for ketones, esters, and amides that allows the preparation of thioketones, thioesters and ...
  39. [39]
    Thionoester - an overview | ScienceDirect Topics
    Direct thionation of esters to the thionoesters with Lawesson's reagent (LR) usually requires high reaction temperatures, long reaction times, and an excess ...
  40. [40]
    [PDF] The Relative Rates of Thiol–Thioester Exchange and Hydrolysis for ...
    Jul 5, 2011 · The ubiquity of thioesters in biology implicates them as prospective compounds of importance to the origin of life. Similar to the case of ...<|control11|><|separator|>
  41. [41]
    Rapid hydrolysis rates of thio- and phosphate esters constrain the ...
    Nov 15, 2024 · But the advantage of thioesters disappears at alkaline pH, as MTA hydrolysis is base catalyzed. At a temperature of 100°C, the hydrolysis rate ...
  42. [42]
    Fmoc-Based Synthesis of Peptide- α Thioesters: Application to the ...
    Here we report an extension of the native chemical ligation method to the total synthesis of a glycosylated protein, the antimicrobial O-linked glycoprotein ...Missing: original | Show results with:original
  43. [43]
    Chapter 10 Polymers containing ester groups in the backbone
    This chapter outlines polymers containing esters groups in the backbone. Poly(esters) are common polymers used as both thermoplastic materials and as fibers.
  44. [44]
    Common Bond Energies (D - Wired Chemist
    Common Bond Energies (D. ) ; C-S, 272, 182 ; C=S · 573, 160 ; C-F, 485, 135 ; C-Cl, 327, 177.Missing: thioester | Show results with:thioester<|separator|>