Printed from https://www.webqc.org

Properties of C7H5NO3S

Properties of C7H5NO3S (Saccharin):

Compound NameSaccharin
Chemical FormulaC7H5NO3S
Molar Mass183.1845 g/mol

Chemical structure
C7H5NO3S (Saccharin) - Chemical structure
Lewis structure
3D molecular structure
Physical properties
AppearanceWhite crystalline solid
Solubility3.45 g/100mL
Density0.8280 g/cm³
Helium 0.0001786
Iridium 22.562
Melting228.80 °C
Helium -270.973
Hafnium carbide 3958

Alternative Names

ortho-benzoic sulfimide
ortho sulphobenzimide

Elemental composition of C7H5NO3S
ElementSymbolAtomic weightAtomsMass percent
CarbonC12.0107745.8963
HydrogenH1.0079452.7512
NitrogenN14.006717.6462
OxygenO15.9994326.2021
SulfurS32.065117.5042
Mass Percent CompositionAtomic Percent Composition
C: 45.90%H: 2.75%N: 7.65%O: 26.20%S: 17.50%
C Carbon (45.90%)
H Hydrogen (2.75%)
N Nitrogen (7.65%)
O Oxygen (26.20%)
S Sulfur (17.50%)
C: 41.18%H: 29.41%N: 5.88%O: 17.65%S: 5.88%
C Carbon (41.18%)
H Hydrogen (29.41%)
N Nitrogen (5.88%)
O Oxygen (17.65%)
S Sulfur (5.88%)
Mass Percent Composition
C: 45.90%H: 2.75%N: 7.65%O: 26.20%S: 17.50%
C Carbon (45.90%)
H Hydrogen (2.75%)
N Nitrogen (7.65%)
O Oxygen (26.20%)
S Sulfur (17.50%)
Atomic Percent Composition
C: 41.18%H: 29.41%N: 5.88%O: 17.65%S: 5.88%
C Carbon (41.18%)
H Hydrogen (29.41%)
N Nitrogen (5.88%)
O Oxygen (17.65%)
S Sulfur (5.88%)
Identifiers
CAS Number81-07-2
SMILESO=C2c1ccccc1S(=O)(=O)N2
Hill formulaC7H5NO3S

Related compounds
FormulaCompound name
C5H7NOSPenam
C3H3NOSIsothiazolinone
C2HNO2SOxathiazolones
C6H5NSON-Sulfinylaniline
C9H9NOSAsmic
C4H5NOSMethylisothiazolinone
C7H5NOSBenzisothiazolinone
C5H5NOSPyrithione
C3H7NO2SCysteine

Related
Molecular weight calculator
Oxidation state calculator

Saccharin (C7H5NO3S): Chemical Compound

Scientific Review Article | Chemistry Reference Series

Abstract

Saccharin, systematically named 1,1-dioxo-1,2-benzothiazol-3-one or 1,2-benzisothiazol-3(2H)-one 1,1-dioxide, is an artificial sweetener with the molecular formula C7H5NO3S and a molecular mass of 183.18 g/mol. This heterocyclic organic compound belongs to the sulfimide class and exhibits exceptional thermal stability with a melting point of 228.8-229.7°C. Saccharin demonstrates remarkable sweetness potency, approximately 300-500 times greater than sucrose on a mass basis. The compound exists as a white crystalline solid with a density of 0.828 g/cm³ and limited water solubility in its acidic form (1 g per 290 mL). Its sodium and calcium salts show significantly enhanced solubility characteristics. Saccharin possesses a pKa value of 1.6, classifying it as a strong acid among organic compounds.

Introduction

Saccharin represents a significant achievement in synthetic organic chemistry and industrial applications, particularly as a non-nutritive sweetening agent. First synthesized in 1879 by Constantin Fahlberg during investigations of coal tar derivatives at Johns Hopkins University, this compound has maintained commercial importance for over a century. The molecular structure incorporates a benzisothiazole ring system with adjacent carbonyl and sulfonyl functional groups, creating a unique electronic configuration responsible for its intense sweet taste perception. Industrial production methods have evolved substantially since its discovery, with modern processes achieving high yields and purity levels. Saccharin's chemical stability under various processing conditions, including high temperatures and pH extremes, distinguishes it from many other sweetening compounds and contributes to its widespread utilization in food and pharmaceutical applications.

Molecular Structure and Bonding

Molecular Geometry and Electronic Structure

The molecular architecture of saccharin features a benzisothiazole ring system fused to a 1,1-dioxide sulfimide group. X-ray crystallographic analysis reveals a nearly planar molecular structure with slight puckering of the heterocyclic ring. The carbonyl carbon-nitrogen bond length measures 1.41 Å, while the sulfur-oxygen bonds in the sulfonyl group average 1.44 Å, indicating significant double bond character. The sulfur-nitrogen bond distance of 1.66 Å suggests partial double bond character resulting from p-orbital overlap between sulfur and nitrogen atoms.

Molecular orbital theory describes the electronic structure as featuring extensive delocalization across the conjugated system. The highest occupied molecular orbital (HOMO) primarily resides on the benzene ring and carbonyl oxygen, while the lowest unoccupied molecular orbital (LUMO) shows significant density on the sulfonyl group and heterocyclic nitrogen. This electronic distribution creates a molecular dipole moment of approximately 4.5 Debye, oriented from the sulfonyl group toward the carbonyl oxygen. The compound exhibits several resonance structures that contribute to its stability, particularly those involving charge separation between the imide nitrogen and carbonyl oxygen.

Chemical Bonding and Intermolecular Forces

Covalent bonding in saccharin demonstrates hybridization patterns consistent with aromatic heterocyclic systems. The carbon atoms in the benzene ring exhibit sp² hybridization with bond angles of approximately 120°. The sulfur atom adopts tetrahedral geometry with sp³ hybridization, while the carbonyl carbon maintains sp² hybridization. Bond energies calculated from thermochemical data indicate C=O bond dissociation energy of 179 kcal/mol and S=O bond energy of 125 kcal/mol, consistent with similar sulfonyl compounds.

Intermolecular forces in crystalline saccharin include strong dipole-dipole interactions between polar functional groups and van der Waals forces between aromatic rings. The molecule forms characteristic hydrogen bonding patterns through its carbonyl oxygen (hydrogen bond acceptor) and imide nitrogen (hydrogen bond donor), with typical N-H···O distances of 2.89 Å in the crystal lattice. These interactions create a stable crystal packing arrangement with a density of 0.828 g/cm³. The compound's polarity, evidenced by its calculated octanol-water partition coefficient (log P) of -0.91, influences its solubility behavior and intermolecular associations.

Physical Properties

Phase Behavior and Thermodynamic Properties

Saccharin exists as a white crystalline solid at standard temperature and pressure conditions. The compound demonstrates a sharp melting point between 228.8°C and 229.7°C, with negligible decomposition at temperatures below 250°C. Thermal analysis reveals a heat of fusion of 28.5 kJ/mol and heat capacity of 215 J/mol·K at 25°C. The crystalline form exhibits orthorhombic symmetry with space group Pna2₁ and unit cell parameters a = 17.392 Å, b = 7.842 Å, c = 9.892 Å, containing four molecules per unit cell.

The density of crystalline saccharin measures 0.828 g/cm³ at 20°C. The refractive index determined for compressed pellets is 1.648 at the sodium D-line. Solubility characteristics show marked dependence on pH and temperature. In its acidic form, water solubility is limited to 0.34 g/100 mL at 20°C, increasing to 0.67 g/100 mL at 100°C. The sodium salt demonstrates substantially improved solubility exceeding 670 g/L at room temperature. Solubility in organic solvents varies considerably, with ethanol solubility of 2.9 g/100 mL and acetone solubility of 0.9 g/100 mL at 25°C.

Spectroscopic Characteristics

Infrared spectroscopy of saccharin reveals characteristic absorption bands corresponding to specific vibrational modes. The carbonyl stretching vibration appears as a strong band at 1715 cm⁻¹, while sulfonyl asymmetric and symmetric stretches occur at 1345 cm⁻¹ and 1160 cm⁻¹ respectively. The N-H stretching vibration produces a medium-intensity band at 3220 cm⁻¹. Bending vibrations of the heterocyclic ring system generate multiple bands between 600-800 cm⁻¹.

Nuclear magnetic resonance spectroscopy shows distinctive proton signals with δH values of 7.60-8.10 ppm for aromatic protons in deuterated dimethyl sulfoxide. Carbon-13 NMR exhibits signals at δC 162.5 ppm (carbonyl carbon), 135.0-128.0 ppm (aromatic carbons), and 125.0 ppm (quaternary carbon adjacent to sulfur). Ultraviolet-visible spectroscopy demonstrates maximum absorption at 270 nm (ε = 9200 M⁻¹cm⁻¹) in aqueous solution, corresponding to π→π* transitions in the aromatic system. Mass spectrometric analysis shows a molecular ion peak at m/z 183 with major fragment ions at m/z 106 (C₆H₄SO₂⁺) and m/z 76 (C₆H₄⁺).

Chemical Properties and Reactivity

Reaction Mechanisms and Kinetics

Saccharin exhibits chemical behavior characteristic of cyclic imides with additional sulfonyl functionality. The compound demonstrates remarkable thermal stability, with decomposition onset temperatures exceeding 300°C under inert atmosphere. Hydrolytic stability varies with pH, showing maximum stability in neutral conditions. Acid-catalyzed hydrolysis proceeds slowly at elevated temperatures, with a rate constant of 2.3×10⁻⁵ s⁻¹ at pH 2 and 100°C, producing ammonium sulfate and ortho-sulfobenzoic acid.

Basic hydrolysis occurs more rapidly, with second-order rate constants of 0.18 M⁻¹s⁻¹ at pH 12 and 25°C. Nucleophilic substitution reactions preferentially occur at the carbonyl carbon, with amines generating N-substituted saccharin derivatives. Electrophilic aromatic substitution is disfavored due to electron-withdrawing effects of the sulfonyl and carbonyl groups, though bromination under vigorous conditions produces 5-bromosaccharin. Reduction with lithium aluminum hydride yields the corresponding thiazoline derivative, while oxidation with peracids maintains the sulfonyl group unchanged.

Acid-Base and Redox Properties

Saccharin functions as a moderately strong organic acid with a pKa value of 1.6 in aqueous solution at 25°C. This acidity derives from stabilization of the conjugate base through resonance between the sulfonyl and carbonyl groups. The compound forms stable salts with various cations, including sodium, potassium, and calcium, all exhibiting enhanced water solubility compared to the free acid. Buffering capacity is minimal due to the single acidic proton and absence of basic functional groups.

Redox properties indicate moderate stability toward common oxidizing and reducing agents. The standard reduction potential measured by cyclic voltammetry shows irreversible reduction waves at -1.2 V versus standard hydrogen electrode, corresponding to reduction of the sulfonyl group. Oxidation occurs at potentials exceeding +1.5 V, involving the aromatic ring system. Saccharin demonstrates electrochemical stability within the range of -0.5 V to +1.0 V, making it compatible with various electrochemical environments.

Synthesis and Preparation Methods

Laboratory Synthesis Routes

The original Remsen-Fahlberg synthesis remains a viable laboratory preparation method for saccharin. This multi-step process begins with toluene sulfonation using chlorosulfonic acid at 0-5°C, producing a mixture of ortho and para toluene sulfonyl chlorides. Fractional crystallization separates the ortho isomer, which subsequently undergoes amination with aqueous ammonia at 50°C to yield ortho-toluenesulfonamide. Oxidation of the methyl group employs potassium permanganate in alkaline medium at 80-90°C, generating ortho-sulfamoylbenzoic acid. Final cyclization occurs spontaneously upon acidification to pH 2-3, producing saccharin in overall yields of 45-50% based on toluene.

Modern laboratory syntheses often employ the Maumee process, which starts with methyl anthranilate. Diazotization with sodium nitrite in hydrochloric acid at 0-5°C produces the diazonium salt, which undergoes sulfonation with sulfur dioxide in the presence of copper chloride catalyst. Chlorination with gaseous chlorine followed by amination with aqueous ammonia yields saccharin directly. This route offers advantages including higher overall yields (65-70%) and fewer purification steps compared to the traditional method.

Industrial Production Methods

Industrial saccharin production predominantly utilizes the modified Remsen process with continuous reactor systems. Large-scale operations employ toluene as the primary feedstock, with annual global production estimated at 30,000 metric tons. The sulfonation step uses oleum (20% SO₃) instead of chlorosulfonic acid for economic and environmental reasons, operating at 80-100°C with reaction times of 2-3 hours. Isomer separation employs selective crystallization from hydrocarbon solvents, achieving 95% recovery of the ortho isomer.

Oxidation processes have evolved to use atmospheric oxygen with cobalt naphthenate catalyst at 150-160°C and 5-10 atm pressure, replacing chemical oxidants. Acidification and cyclization occur continuously in mixed-flow reactors at pH 1.5-2.0 and 80°C. Final purification involves recrystallization from hot water, producing pharmaceutical-grade saccharin with purity exceeding 99.8%. Modern facilities implement extensive recycling of solvents and byproducts, with process water treatment achieving zero liquid discharge. Production costs primarily derive from raw materials (60%), energy consumption (25%), and waste treatment (15%).

Analytical Methods and Characterization

Identification and Quantification

Analytical identification of saccharin employs multiple complementary techniques. High-performance liquid chromatography with ultraviolet detection provides specific quantification using reversed-phase C18 columns with mobile phases consisting of methanol-water-acetic acid (30:69:1 v/v) at flow rates of 1.0 mL/min. Retention times typically range from 6.5-7.2 minutes with detection limits of 0.1 mg/L. Gas chromatography-mass spectrometry requires derivatization with BSTFA (N,O-bis(trimethylsilyl)trifluoroacetamide) to improve volatility, producing characteristic fragment ions at m/z 355 [M+], 340 [M-CH₃]⁺, and 282 [M-COOTMS]⁺.

Capillary electrophoresis methods employ borate buffers at pH 9.2 with detection at 214 nm, achieving separation efficiencies exceeding 200,000 theoretical plates. Spectrophotometric quantification utilizes the intense UV absorption at 270 nm with molar absorptivity of 9200 M⁻¹cm⁻¹. Titrimetric methods employ potentiometric endpoint detection with 0.1 M sodium hydroxide, though these lack specificity in complex matrices. The combination of chromatographic separation with mass spectrometric detection provides definitive identification with quantification limits of 0.01 mg/kg in food matrices.

Purity Assessment and Quality Control

Pharmaceutical-grade saccharin must conform to strict purity specifications outlined in various pharmacopeias. The United States Pharmacopeia requires identification by infrared spectroscopy matching the reference spectrum, with loss on drying not exceeding 1.0% and residue on ignition below 0.1%. Heavy metal content must not exceed 10 ppm, while related substances determined by HPLC should show no individual impurity exceeding 0.1% and total impurities below 0.5%.

Common impurities include ortho-toluenesulfonamide (0.05-0.2%), ortho-sulfobenzoic acid (0.01-0.1%), and various oxidation byproducts. Water content determined by Karl Fischer titration must be less than 0.5% for the free acid form. The sodium salt specification includes sodium content between 10.8-11.2% measured by atomic absorption spectroscopy. Microbiological testing requires absence of Escherichia coli and Salmonella species in 10 g samples, with total aerobic microbial count below 1000 CFU/g.

Applications and Uses

Industrial and Commercial Applications

Saccharin serves primarily as a high-intensity sweetener in food and beverage products, with estimated annual consumption of 20,000 metric tons worldwide. The compound's exceptional stability under thermal processing conditions (up to 150°C) and wide pH range (2-8) makes it particularly suitable for baked goods, canned fruits, and carbonated beverages. Typical usage levels range from 100-500 mg/kg in finished products, providing equivalent sweetness to 3-15% sucrose solutions.

Non-food applications include use in electroplating baths as a brightening agent for nickel and copper deposition, where it functions as a leveling agent at concentrations of 0.5-2.0 g/L. Pharmaceutical formulations employ saccharin as a sweetening excipient in liquid medications, chewable tablets, and oral care products at concentrations up to 1200 mg/L. Agricultural applications include use as an animal feed additive to improve palatability, particularly for swine and poultry feeds at 50-200 mg/kg concentrations.

Research Applications and Emerging Uses

Recent research applications exploit saccharin's unique chemical structure for novel purposes. The compound serves as a building block in organic synthesis, particularly for preparation of N-substituted derivatives with potential biological activity. Coordination chemistry utilizes saccharin as a ligand for various metal ions, forming complexes with interesting catalytic properties. Materials science investigations explore saccharin-based ionic liquids for specialized extraction processes.

Emerging applications include use as a template molecule in molecular imprinting polymers for sensor development. Electrochemical studies investigate saccharin's corrosion inhibition properties on mild steel in acidic environments, showing efficiency exceeding 85% at 500 ppm concentration. Photocatalytic degradation studies employ saccharin as a model compound for advanced oxidation process development. The compound's fluorescence properties in certain solvents suggest potential applications in analytical sensing methods.

Historical Development and Discovery

The discovery of saccharin in 1879 by Constantin Fahlberg represents a classic example of serendipitous scientific discovery. While working under Ira Remsen at Johns Hopkins University investigating oxidation products of coal tar derivatives, Fahlberg noticed a sweet taste on his hands during evening meals. Systematic investigation identified the source as ortho-benzoic sulfimide, which had been prepared earlier that day from toluene sulfonation products. Fahlberg and Remsen published their findings in 1879, though controversy soon emerged regarding proper credit allocation.

Commercial production began in 1884 in Germany, with initial applications limited by cost and production scale. Wartime sugar shortages during World War I dramatically increased demand, leading to expanded production facilities in Europe and North America. The 1960s and 1970s witnessed renewed interest as dieting trends created markets for low-calorie sweeteners. Regulatory challenges emerged in the 1970s based on rodent carcinogenicity studies, though subsequent epidemiological investigations cleared the compound of human cancer risk. Modern production methods have optimized yields and purity while reducing environmental impact through improved waste treatment technologies.

Conclusion

Saccharin remains a chemically fascinating compound with significant industrial importance more than a century after its discovery. Its unique molecular structure combining benzisothiazole and sulfimide functionalities creates distinctive physical and chemical properties, particularly exceptional thermal stability and intense sweetness perception. The compound's behavior in various chemical environments reflects the complex electronic interactions between its functional groups. Modern synthetic methods produce high-purity material efficiently while minimizing environmental impact. Ongoing research continues to reveal new applications beyond sweetening, particularly in materials science and coordination chemistry. The historical development of saccharin from laboratory curiosity to industrial commodity illustrates the interplay between scientific discovery, technological innovation, and regulatory science.

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?