Printed from https://www.webqc.org

Properties of C4H6O6

Properties of C4H6O6 (Tartaric acid):

Compound NameTartaric acid
Chemical FormulaC4H6O6
Molar Mass150.08684 g/mol

Chemical structure
C4H6O6 (Tartaric acid) - Chemical structure
Lewis structure
3D molecular structure
Physical properties
AppearanceWhite powder
Solubility1330.0 g/100mL
Density1.7370 g/cm³
Helium 0.0001786
Iridium 22.562
Melting6.00 °C
Helium -270.973
Hafnium carbide 3958

Alternative Names

2,3-Dihydroxysuccinic acid
Threaric acid
Racemic acid
Uvic acid
Paratartaric acid
Winestone
(2''R'', 3''R'')-Threaric acid ( )
(2''S'', 3''S'')-Threaric acid (-)
''meso''-Erythraric acid

Elemental composition of C4H6O6
ElementSymbolAtomic weightAtomsMass percent
CarbonC12.0107432.0100
HydrogenH1.0079464.0294
OxygenO15.9994663.9606
Mass Percent CompositionAtomic Percent Composition
C: 32.01%H: 4.03%O: 63.96%
C Carbon (32.01%)
H Hydrogen (4.03%)
O Oxygen (63.96%)
C: 25.00%H: 37.50%O: 37.50%
C Carbon (25.00%)
H Hydrogen (37.50%)
O Oxygen (37.50%)
Mass Percent Composition
C: 32.01%H: 4.03%O: 63.96%
C Carbon (32.01%)
H Hydrogen (4.03%)
O Oxygen (63.96%)
Atomic Percent Composition
C: 25.00%H: 37.50%O: 37.50%
C Carbon (25.00%)
H Hydrogen (37.50%)
O Oxygen (37.50%)
Identifiers
CAS Number147-71-7
SMILESO=C(O)C(O)C(O)C(=O)O
Hill formulaC4H6O6

Related compounds
FormulaCompound name
CHOColanic acid
CH2OFormaldehyde
H2CO3Carbonic acid
C3H8OPropanol
CH2COKetene
C4H8OTetrahydrofuran
CH3OHMethanol
CH2O2Formic acid
C3H6OPropionaldehyde
C7H8OAnisole

Sample reactions for C4H6O6
EquationReaction type
C4H6O6 + NaOH = C4NaHO4 + H2Odouble replacement
C4H6O6 + O2 = CO2 + H2Ocombustion

Related
Molecular weight calculator
Oxidation state calculator

Tartaric Acid (C₄H₆O₆): Chemical Compound

Scientific Review Article | Chemistry Reference Series

Abstract

Tartaric acid (C₄H₆O₆), systematically named 2,3-dihydroxybutanedioic acid, represents a naturally occurring dicarboxylic acid with significant industrial and scientific importance. This crystalline organic compound exists in three stereoisomeric forms: dextrorotatory L-(+)-tartaric acid, levorotatory D-(-)-tartaric acid, and the achiral meso-tartaric acid. The naturally predominant L-(+)-enantiomer exhibits a melting point of 169-172 °C and density of 1.737 g/cm³. Tartaric acid demonstrates diprotic acid behavior with pKa values of 2.89 and 4.40 at 25 °C. Its aqueous solubility varies considerably among stereoisomers, with L-(+)-tartaric acid dissolving at 1.33 kg/L while the racemic mixture shows reduced solubility of 0.21 kg/L. The compound serves as a fundamental chiral building block in asymmetric synthesis and finds extensive application in food science, pharmaceutical preparation, and industrial processes.

Introduction

Tartaric acid constitutes an alpha-hydroxy dicarboxylic acid belonging to the aldaric acid class, formally recognized as a dihydroxyl derivative of succinic acid. This organic compound occurs naturally in numerous fruits, with particularly high concentrations in grapes (Vitis vinifera) and tamarinds (Tamarindus indica). The compound's historical significance in chemistry stems from its pivotal role in the discovery of molecular chirality by Jean-Baptiste Biot in 1832 and subsequent crystallographic investigations by Louis Pasteur in 1847. Industrial production primarily focuses on the L-(+)-enantiomer, which is isolated from wine-making byproducts through classical inorganic salt transformations. Tartaric acid's unique combination of hydroxyl and carboxylic acid functionalities, coupled with its chiral nature, establishes it as a versatile compound with applications spanning from food additives to sophisticated chiral auxiliaries in asymmetric synthesis.

Molecular Structure and Bonding

Molecular Geometry and Electronic Structure

Tartaric acid molecules adopt specific conformations dictated by their stereochemistry. The L-(+)- and D-(-)-enantiomers possess C₂ molecular symmetry with the two chiral centers maintaining (2R,3R) and (2S,3S) configurations respectively. These enantiomers exhibit a staggered conformation with torsion angles of approximately 180° between the hydroxyl groups, minimizing steric interactions. Bond lengths within the carbon skeleton measure 1.54 Å for C-C bonds, 1.42 Å for C-O bonds in hydroxyl groups, and 1.21 Å for C=O bonds in carboxylic groups. The central carbon atoms demonstrate sp³ hybridization with bond angles of approximately 109.5°, while the carboxylic carbon atoms show sp² hybridization with bond angles near 120°. Meso-tartaric acid, possessing (2R,3S) configuration, exhibits a different molecular geometry with a plane of symmetry bisecting the molecule, resulting in achiral characteristics despite containing chiral centers.

Chemical Bonding and Intermolecular Forces

The electronic structure of tartaric acid features polarized covalent bonds with oxygen atoms exhibiting significant electronegativity. The hydroxyl oxygen atoms carry partial negative charges (δ⁻ = -0.65) while hydrogen atoms show partial positive charges (δ⁺ = +0.42). This charge distribution facilitates extensive intermolecular hydrogen bonding, with O-H···O distances measuring 2.70 Å in crystalline forms. The molecular dipole moment measures 4.2 D for the enantiomeric forms, reflecting the polar nature of the molecule. Carboxylic acid groups engage in typical dimerization through double hydrogen bonding with interaction energies of approximately 30 kJ/mol. The presence of multiple hydrogen bond donors and acceptors enables the formation of complex hydrogen-bonded networks in solid state structures, contributing to the relatively high melting points and crystalline nature of the compounds.

Physical Properties

Phase Behavior and Thermodynamic Properties

Tartaric acid stereoisomers demonstrate distinct phase behavior and thermodynamic characteristics. L-(+)-Tartaric acid crystallizes as white orthorhombic crystals with density 1.737 g/cm³ and melting point range 169-172 °C. The racemic DL-form exhibits higher density (1.79 g/cm³) and elevated melting point (206 °C) due to formation of a racemic compound rather than a conglomerate. Meso-tartaric acid shows the highest density at 1.886 g/cm³ with melting at 165-166 °C. Enthalpy of fusion measurements yield 35.2 kJ/mol for L-(+)-tartaric acid, 38.1 kJ/mol for the racemate, and 32.8 kJ/mol for the meso form. Specific heat capacity at 25 °C measures 1.21 J/g·K for all forms. The compounds sublime at temperatures above 150 °C under reduced pressure (0.1 mmHg) with sublimation enthalpy of 98.5 kJ/mol. Solubility in water displays significant stereochemical dependence: L-(+)-form dissolves at 1330 g/L, meso-form at 1250 g/L, while racemic acid shows markedly reduced solubility at 210 g/L at 20 °C.

Spectroscopic Characteristics

Infrared spectroscopy reveals characteristic vibrational modes: O-H stretching at 3200-3500 cm⁻¹, C=O stretching at 1720 cm⁻¹, C-O stretching at 1250 cm⁻¹, and O-H bending at 1420 cm⁻¹. Proton NMR spectroscopy in D₂O shows chemical shifts at δ 4.38 ppm for methine protons and δ 3.92 ppm for hydroxyl protons, with coupling constants JHH = 6.2 Hz between vicinal protons. Carbon-13 NMR displays signals at δ 178.2 ppm (carbonyl carbons) and δ 72.4 ppm (chiral carbons). UV-Vis spectroscopy indicates no significant absorption above 210 nm due to absence of chromophores. Mass spectrometric analysis shows molecular ion peak at m/z 150 with major fragmentation peaks at m/z 87 [M-COOH-CO₂]⁺, m/z 73 [M-COOH-CH(OH)COOH]⁺, and m/z 44 [CO₂]⁺.

Chemical Properties and Reactivity

Reaction Mechanisms and Kinetics

Tartaric acid participates in characteristic reactions of both carboxylic acids and secondary alcohols. Esterification reactions proceed with rate constants of k = 2.3 × 10⁻⁴ L/mol·s at 25 °C in acidic media, forming mono- and di-esters. Oxidation with periodic acid cleaves the vicinal diol system with second-order rate constant k₂ = 1.8 L/mol·s at 20 °C. Dehydration reactions under strong acidic conditions produce unsaturated fumaric acid derivatives through elimination of water. Thermal decomposition above 200 °C generates pyruvic acid and carbon dioxide through decarboxylation mechanisms. Complexation with metal ions, particularly copper(II), forms stable chelates with stability constant log β₂ = 6.45 for the Cu(II)-tartrate complex. The hydroxyl groups undergo selective protection through silylation and acetylation reactions with preference for formation of cyclic acetonic derivatives under acidic conditions.

Acid-Base and Redox Properties

Tartaric acid functions as a diprotic acid with dissociation constants pKa1 = 2.89 and pKa2 = 4.40 for L-(+)-tartaric acid at 25 °C. The meso-isomer exhibits slightly different acidity: pKa1 = 3.22 and pKa2 = 4.85. Buffer solutions prepared from tartaric acid and its salts maintain effective pH ranges of 2.8-4.4. Redox properties include standard reduction potential E° = -0.87 V for the tartaric acid/dihydroxymaleic acid couple. Electrochemical oxidation at platinum electrodes occurs at +1.2 V versus standard hydrogen electrode. The compound demonstrates stability in aqueous solutions between pH 2-8, with decomposition observed under strongly alkaline conditions (pH > 10) through retro-aldol reactions. Tartaric acid forms soluble complexes with many metal ions, preventing precipitation of hydroxides in basic media, a property utilized in Fehling's solution for sugar detection.

Synthesis and Preparation Methods

Laboratory Synthesis Routes

Racemic tartaric acid synthesis proceeds through epoxidation of maleic acid using hydrogen peroxide with potassium tungstate catalyst (0.5 mol%) in aqueous medium at 60 °C for 6 hours, yielding 85-90% conversion. Subsequent hydrolysis of the epoxide intermediate with water at pH 7.0 and 80 °C provides racemic tartaric acid with overall yield of 78%. Meso-tartaric acid preparation involves stereoisomerization of L-(+)-tartaric acid by heating in water at 165 °C for 48 hours, producing an equilibrium mixture containing 60% meso and 40% racemic acid. Alternative synthesis from dibromosuccinic acid employs silver hydroxide in aqueous ethanol at reflux temperature for 4 hours, yielding meso-tartaric acid with 75% isolation yield after crystallization. Enantioselective synthesis of L-(+)-tartaric acid utilizes microbial oxidation of cis-epoxysuccinic acid with specific enzymes from Pseudomonas putida, achieving enantiomeric excess greater than 99%.

Industrial Production Methods

Industrial production of L-(+)-tartaric acid primarily utilizes wine-making byproducts, specifically argol (crude potassium bitartrate) deposited during wine fermentation. Processing involves treatment of argol with calcium hydroxide slurry at 80 °C to form calcium tartrate according to the stoichiometry: 2KHC₄H₄O₆ + Ca(OH)₂ → CaC₄H₄O₆ + K₂C₄H₄O₆ + 2H₂O. The calcium tartrate precipitate is separated by filtration and treated with sulfuric acid (70% w/w) at 50 °C to liberate tartaric acid: CaC₄H₄O₆ + H₂SO₄ → H₂C₄H₄O₆ + CaSO₄. Crude tartaric acid undergoes purification through activated carbon treatment, crystallization from water, and vacuum drying, yielding pharmaceutical grade product with 99.5% purity. Global production exceeds 50,000 metric tons annually, with major manufacturing facilities located in wine-producing regions including France, Italy, and Spain. Production costs average $3.50-4.00 per kilogram, with energy consumption approximately 15 kWh per kilogram of final product.

Analytical Methods and Characterization

Identification and Quantification

Tartaric acid identification employs multiple analytical techniques. Polarimetric analysis distinguishes enantiomers through specific rotation measurements: [α]D20 = +12.0° (c = 20, H₂O) for L-(+)-form and [α]D20 = -12.0° for D-(-)-form. Chiral HPLC using amylose-based stationary phases provides separation of enantiomers with resolution factor Rs > 2.5. Quantitative analysis utilizes acid-base titration with sodium hydroxide (0.1 M) using phenolphthalein indicator, achieving accuracy within ±0.5%. Spectrophotometric methods based on complex formation with vanadate ions enable detection limits of 0.1 μg/mL. Gas chromatography of trimethylsilyl derivatives provides simultaneous quantification of tartaric acid and related organic acids with precision of ±2%. Capillary electrophoresis with UV detection at 185 nm offers rapid analysis with separation efficiency exceeding 200,000 theoretical plates.

Purity Assessment and Quality Control

Pharmaceutical grade tartaric acid must conform to stringent purity specifications outlined in pharmacopeial standards. Limits include: assay ≥ 99.5%, sulfate ash ≤ 0.05%, heavy metals ≤ 10 ppm, arsenic ≤ 3 ppm, and oxalate content undetectable. Moisture content determination by Karl Fischer titration must not exceed 0.5%. Optical purity requirements specify enantiomeric excess ≥ 99.0% for L-(+)-tartaric acid. Residual solvent analysis by headspace gas chromatography must show ethanol content < 5000 ppm and isopropanol < 500 ppm. Microbiological testing requires total viable aerobic count < 100 CFU/g and absence of Escherichia coli and Salmonella species. Stability studies indicate shelf life of 36 months when stored in sealed containers at room temperature with protection from moisture. Accelerated stability testing at 40 °C and 75% relative humidity for 6 months demonstrates no significant degradation or moisture uptake.

Applications and Uses

Industrial and Commercial Applications

Tartaric acid serves numerous industrial functions based on its chemical properties. The food industry utilizes it as acidulant (E334) in beverages, confectionery, and baked goods, with global consumption exceeding 30,000 tons annually. Pharmaceutical applications include use as chiral resolving agent for racemic amine salts, with demand estimated at 5,000 tons yearly. The compound functions as complexing agent in electroplating solutions for copper and nickel deposition, preventing anode passivation. Textile industry applications include use as mordant in dyeing processes, particularly for wool and nylon fabrics. Construction industry employs tartaric acid derivatives as set retarders in gypsum plaster and cement formulations, controlling hydration kinetics. Metal cleaning formulations utilize tartaric acid for removal of rust and scale from iron and steel surfaces through chelation of metal ions. Emerging applications include use as template molecule in molecular imprinting technology and as component in chiral stationary phases for chromatographic separations.

Research Applications and Emerging Uses

Tartaric acid derivatives find extensive application in asymmetric synthesis as chiral auxiliaries and ligands. Tartrate-modified titanium catalysts effect enantioselective epoxidation of allylic alcohols with enantiomeric excess up to 95%. Diisopropyl tartrate serves as ligand in the Sharpless asymmetric dihydroxylation reaction, providing access to enantiomerically pure diols. Research applications include use as building block for synthesis of chiral macrocyclic compounds and molecular tweezers. Materials science investigations explore tartaric acid as organic component in hybrid organic-inorganic materials and metal-organic frameworks. Electrochemical studies utilize tartaric acid as complexing agent in copper electrodeposition for microelectronics applications. Emerging technologies investigate tartaric acid-based polymers as biodegradable materials with chiral recognition capabilities. Patent literature discloses applications in optical resolution processes, nonlinear optical materials, and chiral ionic liquids.

Historical Development and Discovery

The history of tartaric acid spans centuries of chemical investigation. Early winemakers observed crystalline deposits of potassium bitartrate in wine vessels as early as the 8th century. Systematic chemical investigation began with Carl Wilhelm Scheele's isolation of tartaric acid from argol in 1769. The compound's optical activity was discovered by Jean-Baptiste Biot in 1832, marking the first observation of molecular chirality. Louis Pasteur's meticulous crystallographic studies in 1847 demonstrated that sodium ammonium tartrate crystallizes as two distinct crystal forms exhibiting mirror-image morphology, leading to the first manual separation of enantiomers. This work established the fundamental principles of molecular chirality and stereochemistry. The 20th century witnessed development of industrial production methods and expansion of applications in food and pharmaceutical industries. Recent decades have seen sophisticated applications in asymmetric synthesis and materials science, solidifying tartaric acid's position as a fundamental chiral building block in chemical research and industry.

Conclusion

Tartaric acid represents a chemically versatile compound with unique stereochemical properties and diverse applications. Its structural characteristics, featuring two chiral centers and multiple functional groups, enable participation in numerous chemical transformations and molecular recognition processes. The distinct physical and chemical properties of its stereoisomers provide valuable insights into structure-property relationships in chiral molecules. Industrial production methods efficiently utilize renewable resources from wine production, demonstrating sustainable chemical manufacturing. Ongoing research continues to develop new applications in asymmetric catalysis, materials science, and pharmaceutical technology. The compound's historical significance in establishing fundamental principles of stereochemistry underscores its continuing importance in chemical education and research. Future investigations will likely focus on developing new tartaric acid-derived chiral catalysts and advanced materials with tailored properties for specific technological applications.

Chemical Compound Properties Database

This database contains physical properties and alternative names for thousands of chemical compounds. In chemical formula you may use:
  • Any chemical element. Capitalize the first letter in chemical symbol and use lower case for the remaining letters: Ca, Fe, Mg, Mn, S, O, H, C, N, Na, K, Cl, Al.
  • Functional groups: D, T, Ph, Me, Et, Bu, AcAc, For, Tos, Bz, TMS, tBu, Bzl, Bn, Dmg
  • parenthesis () or brackets [].
  • Common compound names.
Examples: H2O, CO2, CH4, NH3, NaCl, CaCO3, H2SO4, C6H12O6, water, carbon dioxide, methane, ammonia, sodium chloride, calcium carbonate, sulfuric acid, glucose.

The database includes melting points, boiling points, densities, and alternative names collected from various chemical sources.

What are compound properties?

Chemical compound properties include physical characteristics such as melting point, boiling point, and density, which are important for chemical identification and applications. Alternative names help identify the same compound when referenced by different naming conventions.

How to use this tool?

Enter a chemical formula (like H2O) or compound name (like water) to look up available properties and alternative names. The tool will search through the database and display any available physical properties and known alternative names for the compound.
Please let us know how we can improve this web app.
Menu Balance Molar mass Gas laws Units Chemistry tools Periodic table Chemical forum Symmetry Constants Contribute Contact us
How to cite?