Printed from https://www.webqc.org

Properties of C9H8O2

Properties of C9H8O2 (Cinnamic acid):

Compound NameCinnamic acid
Chemical FormulaC9H8O2
Molar Mass148.15862 g/mol

Chemical structure
C9H8O2 (Cinnamic acid) - Chemical structure
Lewis structure
3D molecular structure
Physical properties
AppearanceWhite monoclinic crystals
OdorHoney-like
Solubility0.5 g/100mL
Density1.2475 g/cm³
Helium 0.0001786
Iridium 22.562
Melting133.00 °C
Helium -270.973
Hafnium carbide 3958
Boiling300.00 °C
Helium -268.928
Tungsten carbide 6000

Alternative Names

''trans''-Cinnamic acid
Phenylacrylic acid
Cinnamylic acid
3-Phenylacrylic acid
(''E'')-Cinnamic acid
Benzenepropenoic acid
Isocinnamic acid

Elemental composition of C9H8O2
ElementSymbolAtomic weightAtomsMass percent
CarbonC12.0107972.9598
HydrogenH1.0079485.4425
OxygenO15.9994221.5977
Mass Percent CompositionAtomic Percent Composition
C: 72.96%H: 5.44%O: 21.60%
C Carbon (72.96%)
H Hydrogen (5.44%)
O Oxygen (21.60%)
C: 47.37%H: 42.11%O: 10.53%
C Carbon (47.37%)
H Hydrogen (42.11%)
O Oxygen (10.53%)
Mass Percent Composition
C: 72.96%H: 5.44%O: 21.60%
C Carbon (72.96%)
H Hydrogen (5.44%)
O Oxygen (21.60%)
Atomic Percent Composition
C: 47.37%H: 42.11%O: 10.53%
C Carbon (47.37%)
H Hydrogen (42.11%)
O Oxygen (10.53%)
Identifiers
CAS Number140-10-3
SMILESO=C(O)\C=C\c1ccccc1
Hill formulaC9H8O2

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

Related
Molecular weight calculator
Oxidation state calculator

Cinnamic acid (C9H8O2): Chemical Compound

Scientific Review Article | Chemistry Reference Series

Abstract

Cinnamic acid, systematically named (2E)-3-phenylprop-2-enoic acid, is an unsaturated carboxylic acid with the molecular formula C9H8O2 and a molar mass of 148.16 g/mol. This white crystalline solid exhibits a characteristic honey-like odor and melts at 133 °C. The compound exists predominantly in the trans configuration due to thermodynamic stability, with the cis isomer (allocinnamic acid) being less common. Cinnamic acid demonstrates limited water solubility (500 mg/L) but dissolves readily in many organic solvents. With a pKa of 4.44, it functions as a weak organic acid. The compound serves as a fundamental building block in organic synthesis and a precursor to numerous commercially important derivatives including fragrances, flavoring agents, and pharmaceutical intermediates.

Introduction

Cinnamic acid represents a significant organic compound belonging to the phenylpropanoid class, characterized by a phenyl group attached to a three-carbon acrylic acid chain. This structural motif confers distinctive chemical properties that have established the compound as an important intermediate in both synthetic organic chemistry and industrial applications. The compound was first isolated in 1872 from cinnamon oil, from which it derives its common name, though it occurs naturally in various plant resins including storax and shea butter. The systematic IUPAC nomenclature designates the most stable isomer as (2E)-3-phenylprop-2-enoic acid, reflecting the trans configuration about the carbon-carbon double bond. Cinnamic acid occupies a central position in organic chemistry due to its conjugated system, which enables diverse reactivity patterns and serves as a prototype for studying electronic effects in unsaturated carboxylic acids.

Molecular Structure and Bonding

Molecular Geometry and Electronic Structure

The molecular structure of trans-cinnamic acid features a planar arrangement resulting from conjugation between the phenyl ring and the acrylic acid moiety. X-ray crystallographic analysis reveals that the trans isomer crystallizes in the monoclinic space group P21/c with unit cell parameters a = 7.441 Å, b = 9.709 Å, c = 14.178 Å, and β = 97.25°. The carboxylic acid group lies nearly coplanar with the vinyl group, forming a dihedral angle of approximately 6.5° between the planes of the carboxyl and phenyl groups. This planarity facilitates extensive π-electron delocalization throughout the molecule.

The carbon atoms of the acrylic chain exhibit sp2 hybridization, with bond angles of approximately 120° around each atom. The C=C bond length measures 1.34 Å, characteristic of a carbon-carbon double bond, while the C-C bond connecting the vinyl group to the phenyl ring measures 1.46 Å. The carboxylic acid group displays typical dimensions with C-O bond lengths of 1.23 Å (C=O) and 1.32 Å (C-OH). The molecule possesses Cs point group symmetry in its optimal trans configuration, with the mirror plane bisecting the carboxylic acid group and passing through the vinyl-phenyl connecting bond.

Electronic structure analysis reveals significant resonance stabilization between the phenyl ring and the unsaturated carboxylic system. The highest occupied molecular orbital (HOMO) demonstrates electron density distributed across the entire conjugated system, while the lowest unoccupied molecular orbital (LUMO) shows greater localization on the acrylic acid portion. This electronic distribution accounts for the compound's ultraviolet absorption characteristics and its reactivity toward electrophilic addition.

Chemical Bonding and Intermolecular Forces

Cinnamic acid exhibits conventional covalent bonding patterns with bond dissociation energies typical of aromatic and unsaturated aliphatic systems. The carbon-carbon double bond dissociation energy measures approximately 264 kJ/mol, while the carbon-phenyl bond dissociation energy is approximately 410 kJ/mol. The carboxylic acid group displays characteristic bonding with O-H bond dissociation energy of 494 kJ/mol.

Intermolecular interactions in crystalline cinnamic acid are dominated by hydrogen bonding between carboxylic acid groups, forming characteristic dimeric structures common to carboxylic acids. The O-H···O hydrogen bond distance measures 2.64 Å, with an O-H···O angle of 176°. These dimers further interact through weaker van der Waals forces and π-π stacking interactions between phenyl rings of adjacent molecules, with an interplanar spacing of approximately 3.6 Å. The compound exhibits a calculated dipole moment of 2.45 Debye, oriented along the long molecular axis from the phenyl ring toward the carboxylic acid group.

The trans isomer demonstrates greater thermodynamic stability than the cis form by approximately 12 kJ/mol, primarily due to reduced steric hindrance between the phenyl ring and carboxylic acid group. This energy difference accounts for the predominance of the trans isomer under standard conditions.

Physical Properties

Phase Behavior and Thermodynamic Properties

Cinnamic acid presents as white monoclinic crystals at room temperature with a density of 1.2475 g/cm3. The compound undergoes a sharp phase transition at 133 °C, melting to form a clear, colorless liquid. The enthalpy of fusion measures 21.5 kJ/mol. The boiling point occurs at 300 °C at atmospheric pressure, with an enthalpy of vaporization of 62.8 kJ/mol. The heat capacity of solid cinnamic acid follows the equation Cp = 0.124 + 0.00257T J/g·K between 25 °C and 130 °C.

The sublimation pressure of cinnamic acid follows the relationship log10(P/mmHg) = 12.23 - 4980/T, where T is temperature in Kelvin. The compound exhibits limited solubility in water (0.5 g/L at 25 °C) but demonstrates high solubility in polar organic solvents including ethanol (1.2 g/mL), acetone (1.5 g/mL), and ethyl acetate (1.4 g/mL). The refractive index of crystalline cinnamic acid measures 1.555 at 589 nm and 20 °C.

Spectroscopic Characteristics

Infrared spectroscopy of cinnamic acid reveals characteristic absorption bands corresponding to its functional groups. The O-H stretching vibration appears as a broad band between 2500-3300 cm-1, while the carbonyl stretching vibration appears at 1685 cm-1. The C=C stretching vibration of the vinyl group appears at 1635 cm-1, and the aromatic C-H stretching vibrations occur between 3000-3100 cm-1. The out-of-plane C-H bending vibrations of the disubstituted alkene appear at 985 cm-1 (trans) and 690 cm-1 (cis), providing diagnostic bands for configuration determination.

Proton nuclear magnetic resonance spectroscopy in deuterated chloroform shows characteristic chemical shifts: the vinyl protons appear as doublets at δ 6.43 ppm (J = 15.9 Hz) and δ 7.69 ppm (J = 15.9 Hz) for the trans isomer, demonstrating typical coupling constants for trans-disubstituted alkenes. The aromatic protons appear as a multiplet between δ 7.35-7.50 ppm, integrating to five protons. The carboxylic acid proton appears broadly at approximately δ 11.5 ppm. Carbon-13 NMR reveals signals at δ 172.5 ppm (carbonyl carbon), δ 144.8 ppm and δ 116.3 ppm (vinyl carbons), and δ 134.2, 129.8, 128.9, and 127.6 ppm (aromatic carbons).

Ultraviolet-visible spectroscopy shows strong absorption at 273 nm (ε = 21,000 M-1cm-1) corresponding to the π→π* transition of the conjugated system. Mass spectrometric analysis exhibits a molecular ion peak at m/z 148, with major fragment ions at m/z 103 (loss of COOH), m/z 77 (phenyl cation), and m/z 51 (C4H3+).

Chemical Properties and Reactivity

Reaction Mechanisms and Kinetics

Cinnamic acid demonstrates characteristic reactivity patterns of both aromatic compounds and α,β-unsaturated carboxylic acids. The conjugated system undergoes electrophilic addition reactions with halogens, with bromination occurring preferentially at the β-position of the double bond. The reaction with bromine follows second-order kinetics with a rate constant of 2.3 × 10-3 M-1s-1 at 25 °C in acetic acid.

Hydrogenation of the double bond proceeds catalytically with palladium on carbon, yielding 3-phenylpropanoic acid with complete stereoselectivity. The reaction follows Langmuir-Hinshelwood kinetics with an activation energy of 45 kJ/mol. The carboxylic acid group undergoes standard derivatization reactions including esterification with rate constants comparable to benzoic acid derivatives. Esterification with methanol catalyzed by sulfuric acid follows second-order kinetics with a rate constant of 8.7 × 10-5 M-1s-1 at 60 °C.

Thermal decarboxylation occurs at elevated temperatures (above 200 °C) with an activation energy of 135 kJ/mol, producing styrene as the primary product. Photodimerization occurs in the solid state upon ultraviolet irradiation, forming truxillic and truxinic acids through [2+2] cycloaddition reactions with quantum yields up to 0.8 in crystalline form.

Acid-Base and Redox Properties

Cinnamic acid functions as a weak organic acid with a pKa of 4.44 in water at 25 °C. The acid dissociation constant follows the typical temperature dependence with ΔH° = -5.2 kJ/mol and ΔS° = -91 J/mol·K. The compound forms stable salts with alkali metals and ammonium ions, with sodium cinnamate exhibiting solubility of 1.2 g/mL in water at 25 °C.

Electrochemical analysis reveals a reversible one-electron oxidation wave at +1.23 V versus standard hydrogen electrode, corresponding to formation of the radical cation. Reduction occurs at -1.45 V versus standard hydrogen electrode, involving two-electron reduction of the double bond. The compound demonstrates stability in acidic media but undergoes gradual decomposition in strongly alkaline conditions above pH 12, with a half-life of 48 hours at pH 13 and 25 °C.

Synthesis and Preparation Methods

Laboratory Synthesis Routes

The Perkin reaction represents the classical laboratory synthesis of cinnamic acid, involving the condensation of benzaldehyde with acetic anhydride in the presence of sodium acetate catalyst. The reaction proceeds through formation of a mixed anhydride intermediate, followed by aldol condensation and dehydration. Typical reaction conditions employ equimolar quantities of benzaldehyde and acetic anhydride with 5-10% sodium acetate by weight, heated at 180 °C for 4-8 hours. This method yields cinnamic acid with 65-75% conversion and requires purification by recrystallization from hot water.

The Knoevenagel condensation provides an alternative synthetic route employing benzaldehyde and malonic acid in the presence of pyridine as both solvent and base. This method proceeds at milder temperatures (80-100 °C) and offers higher yields (85-90%) of trans-cinnamic acid. The mechanism involves formation of a malonic acid anion, nucleophilic addition to benzaldehyde, dehydration, and subsequent decarboxylation of the intermediate β-carboxycinnamic acid.

Modern laboratory synthesis often utilizes the Wittig reaction between benzaldehyde and (carbethoxymethylene)triphenylphosphorane, yielding ethyl cinnamate followed by saponification with aqueous sodium hydroxide. This method provides excellent stereoselectivity for the trans isomer (>95%) and overall yields of 80-85% after purification.

Industrial Production Methods

Industrial production of cinnamic acid primarily employs the Perkin reaction on a large scale, utilizing continuous flow reactors with residence times of 2-3 hours at 170-190 °C. Process optimization has increased yields to 78-82% through careful control of water content and implementation of fractional distillation for acetic acid recovery. Annual global production exceeds 10,000 metric tons, with major manufacturing facilities in China, Germany, and the United States.

An alternative industrial process involves oxidation of cinnamaldehyde using silver oxide or manganese dioxide catalysts in aqueous medium. This method offers the advantage of utilizing cinnamaldehyde derived from cinnamon oil, providing a natural product designation for the resulting cinnamic acid. The oxidation proceeds with 85-90% conversion and selectivity, with catalyst recovery and regeneration systems minimizing environmental impact.

Economic analysis indicates production costs of approximately $12-15 per kilogram for synthetic cinnamic acid, with raw material costs constituting 60-65% of total production expenses. Environmental considerations include acetic acid recovery systems with >95% efficiency and treatment of aqueous waste streams containing sodium acetate.

Analytical Methods and Characterization

Identification and Quantification

Chromatographic methods provide the primary means for cinnamic acid identification and quantification. Reverse-phase high performance liquid chromatography with UV detection at 273 nm offers a limit of detection of 0.1 mg/L and linear response between 1-1000 mg/L. Typical chromatographic conditions employ a C18 column with mobile phase consisting of acetonitrile:water:acetic acid (45:54:1 v/v) at flow rate of 1.0 mL/min. Gas chromatography with mass spectrometric detection provides complementary analysis with limits of detection of 0.01 mg/L after derivatization with diazomethane.

Titrimetric analysis using standard sodium hydroxide solution with phenolphthalein indicator offers a simple quantitative method for pure samples, with relative error less than 1% for samples containing >98% cinnamic acid. Spectrophotometric quantification at 273 nm provides rapid analysis with molar absorptivity of 21,000 M-1cm-1 in ethanol.

Purity Assessment and Quality Control

Pharmaceutical-grade cinnamic acid must conform to specifications outlined in various pharmacopeias, typically requiring minimum purity of 99.0% by HPLC analysis. Common impurities include benzoic acid (limit <0.2%), phenylpropionic acid (limit <0.3%), and traces of heavy metals (limit <10 ppm). Melting point determination provides a rapid purity assessment, with pure trans-cinnamic acid melting sharply within a 1 °C range.

Stability testing indicates that cinnamic acid remains stable for at least three years when stored in sealed containers protected from light at room temperature. Accelerated aging studies at 40 °C and 75% relative humidity show no significant decomposition over six months. Photodegradation becomes significant only upon direct ultraviolet exposure, with a half-life of 120 hours under simulated sunlight.

Applications and Uses

Industrial and Commercial Applications

Cinnamic acid serves primarily as a chemical intermediate in the production of various derivatives with commercial significance. Esterification produces methyl, ethyl, and benzyl cinnamates, which find extensive application in the fragrance and flavor industry due to their balsamic and floral notes. Global production of cinnamate esters exceeds 5,000 metric tons annually, with market value exceeding $50 million.

The compound functions as a precursor in the synthesis of the artificial sweetener aspartame through enzymatic amination to phenylalanine. This application consumes approximately 2,000 metric tons of cinnamic acid annually. Additional industrial applications include use as a monomer in specialty polymers, particularly those requiring ultraviolet-absorbing properties, and as an intermediate in the synthesis of pharmaceuticals including cinchocaine and other local anesthetics.

Research Applications and Emerging Uses

Cinnamic acid derivatives continue to attract research interest for various advanced applications. Vinyl cinnamate monomers undergo photopolymerization to form cross-linked polymers with applications in holographic data storage and optical devices. The compound serves as a building block in molecular electronics due to its conjugated system and ability to form self-assembled monolayers on metal surfaces.

Recent research explores cinnamic acid as a precursor to novel liquid crystalline materials with smectic phases between 120-180 °C. Derivatives with extended conjugation show promise as organic semiconductor materials with charge carrier mobility up to 0.02 cm2/V·s. Electrooptical applications include use as nonlinear optical materials with second harmonic generation efficiency approximately 0.5 times that of urea.

Historical Development and Discovery

Cinnamic acid was first isolated in 1872 by the Italian chemist Luigi Chiozza from cinnamon oil, though its correct empirical formula was not established until 1877. The compound's structure was elucidated through the work of several nineteenth-century chemists, including William Henry Perkin Sr., who developed the synthetic reaction that bears his name in 1877. The stereochemistry of cinnamic acid became a subject of intensive investigation following the discovery of geometric isomerism by van't Hoff and Le Bel in 1874.

The photodimerization of cinnamic acid, discovered in the early twentieth century, played a crucial role in understanding solid-state photochemical reactions and topochemical principles. Gerhard Schmidt's systematic investigation of cinnamic acid crystal structures in the 1960s established fundamental relationships between molecular packing and photoreactivity that continue to influence materials design. Industrial production began in the 1920s to meet demand from the fragrance industry, with significant process improvements occurring in the 1950s with the development of continuous reaction systems.

Conclusion

Cinnamic acid represents a structurally simple yet chemically rich organic compound that continues to serve as a valuable intermediate in both industrial and research contexts. Its conjugated system, combining aromatic and unsaturated carboxylic acid functionalities, provides diverse reactivity patterns that have been extensively studied and utilized. The compound's physical properties, particularly its crystalline behavior and spectroscopic characteristics, make it an excellent model system for investigating solid-state chemistry and molecular interactions.

Future research directions likely include development of more sustainable production methods, particularly enzymatic synthesis routes utilizing phenylalanine ammonia lyase. Advanced applications in materials science, including photoresponsive polymers and organic electronic devices, continue to expand the utility of this classic organic compound. The fundamental understanding gained from studying cinnamic acid's chemical behavior provides insights that extend to more complex conjugated systems throughout organic chemistry.

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?