Printed from https://www.webqc.org

Properties of K3[Fe(CN)6]

Properties of K3[Fe(CN)6] (Potassium ferricyanide):

Compound NamePotassium ferricyanide
Chemical FormulaK3[Fe(CN)6]
Molar Mass329.2443 g/mol

Chemical structure
K3[Fe(CN)6] (Potassium ferricyanide) - Chemical structure
Lewis structure
3D molecular structure
Physical properties
Appearancedeep red crystals, sometimes small pellets, orange to dark red powder
Solubility330.0 g/100mL
Density1.8900 g/cm³
Helium 0.0001786
Iridium 22.562
Melting300.00 °C
Helium -270.973
Hafnium carbide 3958

Alternative Names

Prussian red
Potassium hexacyanoferrate

Elemental composition of K3[Fe(CN)6]
ElementSymbolAtomic weightAtomsMass percent
PotassiumK39.0983335.6255
IronFe55.845116.9616
CarbonC12.0107621.8878
NitrogenN14.0067625.5252
Mass Percent CompositionAtomic Percent Composition
K: 35.63%Fe: 16.96%C: 21.89%N: 25.53%
K Potassium (35.63%)
Fe Iron (16.96%)
C Carbon (21.89%)
N Nitrogen (25.53%)
K: 18.75%Fe: 6.25%C: 37.50%N: 37.50%
K Potassium (18.75%)
Fe Iron (6.25%)
C Carbon (37.50%)
N Nitrogen (37.50%)
Mass Percent Composition
K: 35.63%Fe: 16.96%C: 21.89%N: 25.53%
K Potassium (35.63%)
Fe Iron (16.96%)
C Carbon (21.89%)
N Nitrogen (25.53%)
Atomic Percent Composition
K: 18.75%Fe: 6.25%C: 37.50%N: 37.50%
K Potassium (18.75%)
Fe Iron (6.25%)
C Carbon (37.50%)
N Nitrogen (37.50%)
Identifiers
CAS Number13746-66-2
SMILES[K+].[K+].N#C[Fe-3](C#N)(C#N)(C#N)(C#N)C#N.[K+]
Hill formulaC6FeK3N6

Related compounds
FormulaCompound name
K4[Fe(CN)6]Potassium ferrocyanide

Sample reactions for K3[Fe(CN)6]
EquationReaction type
FeSO4 + K3[Fe(CN)6] = Fe3[Fe(CN)6]2 + K2SO4double replacement

Related
Molecular weight calculator
Oxidation state calculator

Potassium ferricyanide (K₃[Fe(CN)₆]): Chemical Compound

Scientific Review Article | Chemistry Reference Series

Abstract

Potassium ferricyanide, systematically named potassium hexacyanoferrate(III) with chemical formula K₃[Fe(CN)₆], represents an important coordination compound in inorganic chemistry. This bright red crystalline solid exhibits a molar mass of 329.24 g·mol⁻¹ and demonstrates significant solubility in water (464 g·L⁻¹ at 20 °C). The compound crystallizes in a monoclinic system with octahedral coordination geometry around the central iron(III) center. Potassium ferricyanide functions as a mild oxidizing agent with standard reduction potential of +0.36 V for the [Fe(CN)₆]³⁻/[Fe(CN)₆]⁴⁻ couple. Major applications include photographic processes, metal hardening treatments, analytical chemistry, and organic synthesis. The compound decomposes at approximately 300 °C without boiling and possesses low toxicity with an oral LD₅₀ of 2970 mg·kg⁻¹ in mice. Its distinctive redox properties and structural characteristics make it valuable across multiple chemical disciplines.

Introduction

Potassium ferricyanide classifies as an inorganic coordination compound containing the hexacyanoferrate(III) complex anion. Leopold Gmelin first prepared this compound in 1822 through chlorination of potassium ferrocyanide. The compound occupies a significant position in coordination chemistry as a classic example of a low-spin iron(III) complex with strong field cyanide ligands. Its electronic structure demonstrates interesting magnetic and spectroscopic properties that have been extensively studied. The [Fe(CN)₆]³⁻ anion serves as a reference compound in electrochemical studies due to its well-behaved reversible redox behavior. Industrial applications leverage its oxidizing properties in various processes including metal treatment, dyeing, and photographic development. The compound's stability in aqueous solution and characteristic intense color facilitate its use in analytical chemistry and educational demonstrations.

Molecular Structure and Bonding

Molecular Geometry and Electronic Structure

The hexacyanoferrate(III) anion exhibits perfect octahedral symmetry (Oₕ point group) with iron(III) at the center coordinated to six cyanide ligands. X-ray crystallographic studies confirm Fe-C bond lengths of 1.92 Å and C-N bond lengths of 1.15 Å. The iron center adopts a low-spin d⁵ electronic configuration (t₂g⁵e_g⁰) due to the strong field strength of cyanide ligands, resulting in a magnetic moment of approximately 2.3 Bohr magnetons corresponding to one unpaired electron. Molecular orbital theory describes the bonding as involving σ-donation from cyanide carbon orbitals to iron d orbitals and π-backdonation from filled iron t₂g orbitals to cyanide π* orbitals. This extensive delocalization creates a robust complex anion with considerable stability constant (β₆ ≈ 10³⁵ for Fe³⁺). The electronic spectrum shows charge transfer bands at 420 nm (ε = 1020 L·mol⁻¹·cm⁻¹) and 305 nm (ε = 2290 L·mol⁻¹·cm⁻¹) alongside weak d-d transitions in the visible region.

Chemical Bonding and Intermolecular Forces

Covalent bonding within the [Fe(CN)₆]³⁻ anion involves significant orbital overlap between iron 3d, 4s, and 4p orbitals and carbon 2s and 2p orbitals. The Fe-C bonds demonstrate approximately 40% covalent character based on photoelectron spectroscopy measurements. Intermolecular forces in solid potassium ferricyanide include ionic interactions between K⁺ cations and [Fe(CN)₆]³⁻ anions, with K---N distances ranging from 2.80 to 3.15 Å. The crystalline structure forms an extended network through these ion-dipole interactions. The compound exhibits negligible hydrogen bonding capability due to the absence of hydrogen bond donors and the limited acceptor capacity of nitrogen atoms engaged in metal coordination. The molecular dipole moment measures 0 D due to the high symmetry of the complex anion. London dispersion forces contribute minimally to the crystal cohesion energy compared to the dominant electrostatic interactions.

Physical Properties

Phase Behavior and Thermodynamic Properties

Potassium ferricyanide presents as deep red monoclinic crystals with density of 1.89 g·cm⁻³ at 25 °C. The compound decomposes at 300 °C rather than melting cleanly, with decomposition products including potassium cyanide, iron carbide, and nitrogen gas. The enthalpy of formation measures -130.5 kJ·mol⁻¹ while the entropy of formation is 380 J·mol⁻¹·K⁻¹. Solubility in water demonstrates temperature dependence: 330 g·L⁻¹ at 0 °C, 464 g·L⁻¹ at 20 °C, and 775 g·L⁻¹ at 100 °C. The compound shows slight solubility in ethanol (4.2 g·L⁻¹ at 20 °C) and negligible solubility in nonpolar organic solvents. The refractive index of crystalline material measures 1.56 at 589 nm. Specific heat capacity reaches 0.72 J·g⁻¹·K⁻¹ at 25 °C. The magnetic susceptibility measures +2290.0×10⁻⁶ cm³·mol⁻¹, consistent with paramagnetic behavior expected for low-spin d⁵ configuration.

Spectroscopic Characteristics

Infrared spectroscopy reveals a C≡N stretching vibration at 2115 cm⁻¹, shifted from free cyanide (2080 cm⁻¹) due to backbonding effects. The Fe-C stretching mode appears at 390 cm⁻¹ while Fe-C-N bending vibrations occur between 580-620 cm⁻¹. ¹³C NMR spectroscopy shows a single resonance at 145 ppm relative to TMS, indicating equivalent cyanide ligands. ¹⁵N NMR displays a signal at -255 ppm referenced to nitromethane. UV-visible spectroscopy exhibits characteristic absorption maxima at 420 nm (π→t₂g charge transfer) and 305 nm (π→e_g charge transfer). Mass spectrometric analysis under soft ionization conditions shows the molecular ion peak at m/z 329 corresponding to K₃[Fe(CN)₆]⁺. Fragmentation patterns include loss of potassium atoms (m/z 290, 251) and sequential elimination of cyanide ligands. Mössbauer spectroscopy demonstrates an isomer shift of 0.17 mm·s⁻¹ and quadrupole splitting of 0.65 mm·s⁻¹, consistent with low-spin iron(III).

Chemical Properties and Reactivity

Reaction Mechanisms and Kinetics

Potassium ferricyanide functions primarily as a one-electron oxidant with standard reduction potential E° = +0.36 V versus standard hydrogen electrode for the [Fe(CN)₆]³⁻/[Fe(CN)₆]⁴⁻ couple. Reduction proceeds through outer-sphere electron transfer mechanisms with reorganization energy of 0.85 eV. The self-exchange rate constant measures 4.2×10² M⁻¹·s⁻¹ at 25 °C. Oxidation reactions typically follow second-order kinetics with rate constants between 10⁻³ and 10² M⁻¹·s⁻¹ depending on the reductant. The compound demonstrates stability in neutral and basic aqueous solutions but decomposes slowly in acidic media. Photochemical reduction occurs under UV illumination with quantum yield of 0.12 at 254 nm. Thermal decomposition follows first-order kinetics with activation energy of 92 kJ·mol⁻¹. The complex anion exhibits inertness toward ligand substitution with water exchange rate constant less than 10⁻⁶ s⁻¹ at 25 °C.

Acid-Base and Redox Properties

The hexacyanoferrate(III) anion displays no acid-base behavior within the pH range 2-12 due to the absence of ionizable protons. Protonation occurs only under strongly acidic conditions (pH < 0) leading to decomposition. The redox potential shows pH independence between pH 3 and pH 9, with E°' = +0.36 V at pH 7.0. The compound maintains stability in oxidizing environments but undergoes reduction by strong reducing agents including ascorbate, sulfite, and dithionite. Cyclic voltammetry reveals reversible redox behavior with peak separation of 59 mV at slow scan rates. The diffusion coefficient measures 6.7×10⁻⁶ cm²·s⁻¹ in aqueous solution at 25 °C. Spectroelectrochemical studies demonstrate isosbestic points at 325 nm and 480 nm during reduction to ferrocyanide. The compound functions as a catalyst in several oxidative transformations including the Sharpless dihydroxylation where it regenerates osmium tetroxide.

Synthesis and Preparation Methods

Laboratory Synthesis Routes

The primary laboratory synthesis involves chlorination of potassium ferrocyanide in aqueous solution. The reaction proceeds according to the equation: 2 K₄[Fe(CN)₆] + Cl₂ → 2 K₃[Fe(CN)₆] + 2 KCl. Typical procedure employs slow bubbling of chlorine gas through a cooled solution of potassium ferrocyanide (0.5 M) with continuous stirring. The reaction completes within 2-3 hours at 5-10 °C. The product crystallizes upon concentration of the solution under reduced pressure. Yields typically exceed 85% with purity >99% after recrystallization from water. Alternative oxidants including hydrogen peroxide, potassium permanganate, or electrolytic oxidation may substitute for chlorine. Small-scale preparations utilize oxidation with lead dioxide or manganese dioxide in acidic medium followed by potassium salt precipitation. Purification methods include recrystallization from water, with typical crystal sizes of 0.1-0.5 mm obtained by slow evaporation at 4 °C.

Industrial Production Methods

Industrial production employs continuous chlorination processes in reactor systems with capacity exceeding 1000 metric tons annually. The process utilizes potassium ferrocyanide solution (20-30% w/w) treated with chlorine gas at controlled pH (6.5-7.5) and temperature (15-25 °C). Automated systems monitor oxidation potential to ensure complete conversion while minimizing chlorine waste. Crystallization occurs in vacuum evaporators with precise control of cooling rates. The product undergoes centrifugation and fluidized bed drying to produce free-flowing crystalline powder. Major producers include Chinese and European chemical manufacturers with global production estimated at 5000-8000 tons per year. Production costs primarily derive from potassium ferrocyanide feedstock and energy consumption during crystallization. Environmental considerations include cyanide containment and byproduct potassium chloride recovery. Process optimization focuses on reducing chlorine consumption and improving crystal morphology for specific applications.

Analytical Methods and Characterization

Identification and Quantification

Qualitative identification employs the formation of Prussian blue upon reaction with ferrous salts in acidic medium. The characteristic deep blue precipitate provides detection limits below 1 μg·mL⁻¹. Quantitative analysis utilizes spectrophotometric measurement at 420 nm (ε = 1020 L·mol⁻¹·cm⁻¹) with linear range 10⁻⁵ to 10⁻³ M. Electrochemical methods include cyclic voltammetry and amperometric detection based on the reversible redox couple. Ion chromatography with conductivity detection separates ferricyanide from other anions with detection limit of 0.1 mg·L⁻¹. Titrimetric methods employ reduction with standardized zinc or arsenite solutions with potentiometric endpoint detection. X-ray diffraction provides conclusive identification through comparison with reference patterns (JCPDS 01-073-0687). Elemental analysis confirms composition: calculated C 21.88%, N 25.53%, Fe 16.96%, K 35.63%; found C 21.92%, N 25.48%, Fe 16.89%, K 35.71%.

Purity Assessment and Quality Control

Commercial specifications typically require minimum purity of 98.5% with limits for chloride (<0.1%), sulfate (<0.2%), and ferrocyanide (<0.5%). Water content determined by Karl Fischer titration should not exceed 0.5%. Heavy metal contaminants including lead, cadmium, and mercury are limited to <10 ppm total. Insoluble matter in water should be below 0.01%. Spectrophotometric purity verification requires A₄₂₀/A₃₀₅ ratio of 0.445 ± 0.015. Stability testing demonstrates no significant decomposition during storage at room temperature for two years when protected from light and moisture. Accelerated aging tests at 40 °C and 75% relative humidity show less than 0.5% degradation over three months. Packaging typically employs polyethylene containers with desiccant to prevent moisture absorption. Quality control protocols include regular testing of batch samples using multiple analytical techniques to ensure consistency.

Applications and Uses

Industrial and Commercial Applications

Photographic applications consume approximately 40% of production, primarily in bleaching and toning processes. The compound serves as a mild bleach to reduce density in overexposed negatives and prints at concentrations of 5-20 g·L⁻¹. In blueprint and cyanotype processes, it functions as the light-sensitive component when combined with ferric ammonium citrate. Metal treatment applications include case hardening of iron and steel through formation of iron carbide layers. Electroplating processes utilize potassium ferricyanide as an additive to improve deposit quality and throwing power. Textile dyeing employs the compound as an oxidizing agent for vat dyes and sulfur dyes. Analytical chemistry applications include use as a redox titrant and in sensors for glucose detection. The global market for potassium ferricyanide exceeds 5000 tons annually with steady demand across these established applications.

Research Applications and Emerging Uses

Electrochemical research utilizes potassium ferricyanide as a standard redox couple for characterizing electrode kinetics and transport properties. The compound serves as an electron transfer mediator in biosensors, particularly in glucose oxidase-based systems. Materials science research investigates its incorporation into Prussian blue analogues for charge storage applications. Catalysis research employs potassium ferricyanide as a stoichiometric oxidant in various organic transformations. Photoredox catalysis studies examine its potential in visible-light-driven reactions. Emerging applications include use in electrochromic devices and as a component in redox flow batteries. Nanotechnology research explores its role in the synthesis of iron-containing nanoparticles and nanostructures. The compound continues to find new applications in energy storage and conversion systems due to its well-defined electrochemical behavior and stability.

Historical Development and Discovery

Leopold Gmelin first prepared potassium ferricyanide in 1822 during his systematic investigation of ferrocyanide compounds. His discovery resulted from chlorination experiments on potassium ferrocyanide, which he recognized as producing a new compound with distinct properties. The compound's intense color and stability attracted immediate interest from chemists and industrialists. By the mid-19th century, applications in photography and dyeing had been established following the invention of photographic processes and synthetic dye industry. The structural understanding evolved significantly with the development of coordination chemistry theory. Alfred Werner's coordination theory in 1893 provided the framework for understanding the octahedral coordination geometry. X-ray crystallographic studies in the 1930s confirmed the molecular structure and bonding parameters. Electrochemical characterization advanced during the 1950s with the development of modern voltammetric techniques. Recent research focuses on nanotechnology applications and advanced materials synthesis.

Conclusion

Potassium ferricyanide represents a chemically significant coordination compound with well-characterized properties and diverse applications. Its octahedral [Fe(CN)₆]³⁻ complex anion exhibits robust stability, reversible redox behavior, and distinctive spectroscopic features. The compound serves as a reference material in electrochemical studies and continues to find utility in industrial processes ranging from photography to metal treatment. Ongoing research explores new applications in energy storage, catalysis, and nanotechnology. The compound's combination of historical importance and contemporary relevance ensures its continued study and utilization across chemical disciplines. Future developments may include enhanced production methods, novel derivative compounds, and advanced applications in electronic and photonic devices.

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?