Synthetic colorant

Last updated
Tassels dyed with BASF dyes 1901 The aniline colours of the Badische Anilin- & Soda-Fabrik, Ludwigshafen on Rhine and their application on wool, cotton, silk and other textile fibres, page 295.jpg
Tassels dyed with BASF dyes 1901


A colorant is any substance that changes the spectral transmittance or reflectance of a material. [1] Synthetic colorants are those created in a laboratory or industrial setting. The production and improvement of colorants was a driver of the early synthetic chemical industry, in fact many of today's largest chemical producers started as dye-works in the late 19th or early 20th centuries, including Bayer AG(1863). [2] Synthetics are extremely attractive for industrial and aesthetic purposes as they have they often achieve higher intensity and color fastness than comparable natural pigments and dyes used since ancient times. Market viable large scale production of dyes occurred nearly simultaneously in the early major producing countries Britain (1857), France (1858), Germany (1858), and Switzerland (1859), and expansion of associated chemical industries followed. [3] The mid-nineteenth century through WWII saw an incredible expansion of the variety and scale of manufacture of synthetic colorants. Synthetic colorants quickly became ubiquitous in everyday life, from clothing to food. This stems from the invention of industrial research and development laboratories in the 1870s, and the new awareness of empirical chemical formulas as targets for synthesis by academic chemists. The dye industry became one of the first instances where directed scientific research lead to new products, and the first where this occurred regularly.

Contents

Dyes versus pigments

Colorants can be divided into pigments and dyes. Broadly, dyes are soluble and become fixed to a substrate via impregnation, while pigments are insoluble and require a binding agent to adhere to a substrate. Dyes, therefore, must have an affinity for the substance they are intended to color. [4] Chemically speaking, for now, pigments can be organic or inorganic, while dyes are only organic. Furthermore, organic white pigments do not exist, despite the fact that the majority of purified crystalline organic products are white in appearance. [5] This story is complicated somewhat by lake pigments, or lakes, which are dyes modified with a chemical process to form an insoluble pigment. Typically this involves precipitating the natural extracts as salts in alkaline conditions. [6] The historical importance of both pigments and dyes is closely related, as the markets for both, as well as the types and variety available, have always been closely tied.

History

Early colorants date to prehistoric times. Human beings were already relying on natural substances, primarily from vegetables, but also from animals, to color their homes and artifacts. Cave drawings like those in Altamira or Lascaux were made in the Ice Age 15,000 to 30,000 years ago. Using pigments for coloration is among the oldest cultural activities of mankind. [5] The important substrates of pre-industrial societies were generally naturally occurring (cotton, silk, wool, leather, paper) and therefore share similarities, since they are primarily saccharide or peptide polymers. [7]

The nineteenth and twentieth century in particular saw an expansion in colorant use and production, yielding many pigments and dyes in use today. The availability of strong acidic or alkaline environments like sulphuric acid and synthetic sodium carbonate was crucial in this process. These conditions became possible due to price drops in reagents due to new industrial preparations like the LeBlanc process, where potassium carbonate formerly obtained from ashes was replaced by sodium carbonate. [6] However, many early colorants are no longer produced due to economics, or high toxicity, for example Schweinfurt green (cupric acetate arsenite), Scheele's green (copper(II) arsenite), and Naples yellow (lead antimonate). [5]

The late 1850s saw the introduction of the first modern synthetic dyes, which brought more color and variety of color to Europe. In addition to being multi-varied and extraordinarily intense, these new dyes were notoriously unstable, rapidly fading and turning when exposed to sunlight, washing, and other chemical or physical agents. This led to new systems of categorization and study of colorants, which in turn lead to the synthesis of more color-fast modern colorants. Synthetic colors found themselves in not only dyes and paints but also inks and foodstuffs, permeating consumer culture. [8]

Natural products

In ancient cave paintings natural manganese oxide and charcoal were used for black shades and iron oxides for yellow, orange, and red color tones. [5] Examples of similar earth pigments that persisted to more modern times are the red pigment vermilion (mercury sulphide), the yellow orpiment (arsenic trisulphide), the green malachite (basic copper carbonate) and the blue lapis lazuli (natural ultramarine). Natural sources of white pigments include chalk and kaolin, while black pigments are often obtained as charcoal and as soot. [7]

Early production and syntheses

Example of the Virgin Mary painted wearing a blue robe Intercession of Charles Borromeo supported by the Virgin Mary - Detail Rottmayr Fresco - Karlskirche - Vienna.JPG
Example of the Virgin Mary painted wearing a blue robe

In ancient times, through the Industrial Revolution, various inorganic pigments like Egyptian Blue were synthesized, many with toxic chemicals like arsenic and antimony. These toxic pigments were used for cosmetics and painting. In ancient Egypt, blue was considered the color of the divine. As a result, the early synthetic compound Egyptian Blue, became an incredibly important pigment. It was used for the depiction of eyes, hair and decoration in the graphic representation of pharaohs. [5] Blue, particularly ultramarine pigment made from ground lapis lazuli remained significant for depictions of the divine through the Renaissance. Pre-industrial revolution painters in Europe used ultramarine almost exclusively for the robes of Mary because of the pigment's great expense, until the work of Jean-Baptiste Guimet and Christian Gmelin made it commercially available in larger, cheaper quantities.

At the beginning of the eighteenth century, the first products of the fledgling color industry were Prussian blue and Naples yellow. The first synthetically produced white pigment was white lead (lead carbonate). It was known in Roman times. Around 1800, more inorganic white pigments were developed including zinc white (zinc oxide) was developed, followed by antimony white (antimony oxide) and zinc sulfide. [5] The printers and dyers at that time had access to lead acetate, alum, copper acetate, nitric acid, ammonia and ammonium chloride, potassium carbonate, potassium tartrate, gallic acid, gums, bleaching lyes, hydrochloric acid, sulfuric acid, carbonates, sulfates, and acetates. [9] Small scale workshops evolved into ever larger and larger manufactories.

Other inorganic pigments developed in the nineteenth century were cobalt blue, Scheele's green, and chrome yellow. The availability of sulphuric and sulfurous acids facilitated further experiments, leading to the isolation of alizarin and purpurin in 1826. Madder based pigments such as Brown Madder (obtained in 1840) were developed due to research by British and German chemists into Turkey red, also known as Rouge d’Andrinopole. [6]

First "scientific" syntheses: aniline dyes 1858 – 1870

In the mid nineteenth century, the coal tar industry, particularly in England, produced the precursors needed for a large amount of organic syntheses, in large quantities. [10] For the first eight years after the first marketable synthetic dye, Mauveine, until the middle of the 1860s, British and French firms were the major dye producers. The second half of the 1860s saw German dye works surpassing their competition in both capacity and market share. During 1870, German firms were responsible for roughly half of the world's production of dyes and pigments. Aniline dyes were produced at scale, in part because of many advances in the synthesis of their precursors. [11] Antione Bechamp described a process for reducing nitrobenzene to aniline in 1854, known as the Bechamp Process, making the production of aniline easy. [12] Widespread isolation of phenol from coal tar, made its nitration more economical, generally the path of the synthesis flowed: coal tar → nitrobenzene → aniline → dyes. [13] According to Henry Perkin himself "This industry holds an[ sic ] unique position in the history of chemical industries, as it was entirely the outcome of scientific research." [10]

First scientific synthetic dye: picric acid

The first synthetic dye was picric acid. It was prepared in a laboratory in 1771, and commercially produced by M. Guinon in Lyon in 1845. [13] It dyed silk fabric yellow; however the color fastness properties were not good, thus it had very limited commercial success. [7] [14] It was, however, purchased in limited amounts by French dyers. [15]

A letter with a sample of mauveine dyed silk Mauv2.jpg
A letter with a sample of mauveine dyed silk

William Henry Perkin’s mauveine

In 1856, 18 year old William Perkin accidentally discovered a dye he called mauve while trying to make quinine from the oxidation of allyl toluene in his home lab for his academic advisor and boss August Wilhelm von Hoffman. [16] Hoffman reportedly referred to aniline, a major step in the synthesis, as his "first love," and was excited to have Perkin working with it. [10] Perkin communicated with the textile industry, including Pullars of Perth, and John Hyde Christie, the chemist and general manager of John Orr Ewing and Co. about how to best market and produce his dye. [14] He started production of aniline purple near London at the end of 1857 and remained the only producer for at least a few months. Perkin began making the intermediates for his dyes in-house, for example, nitro-benzene, expanding the scale of operations. [3] By the summer of 1859, according to a satirical magazine Punch, London had fallen ill with 'the mauve measles'. [17]

An illustration of an aniline dye-works in a journal of the period Journal (1882) (14799745243).jpg
An illustration of an aniline dye-works in a journal of the period

Rapid expansion

By the end of 1858 there were already eight firms producing aniline dyes. [13] By 1861 there were twenty-nine British patents on coloring matters from aniline. By 1864 68 firms were producing dyes. [3] This was driven by the textile industry, which employed new designs requiring the colorful aniline dyes. Even Hofmann, who had at first criticized his student for leaving his academic research of quinine, later synthesized his own aniline dye, rosaniline. [17] In 1858 the German chemist Johann Peter Griess obtained a yellow dye by reacting nitrous acid with aniline. It didn't last commercially, but it created even more interest in aniline as precursor for colorful compounds. [7] French chemist François-Emmanuel Verguin reacted aniline with stannic chloride to yield fuchsine, a rose colored dye, the first of the triphenylmethane dyes. Further work by Hoffman [18] along with the discovery of benzene’s structure (1858) and carbon’s tetravalency(1865), this science built the groundwork for modern organic chemistry. [19]

In the late 1860s many companies began offering a full spectrum of colors, and were already outcompeting many natural dyes for market share. Prices continually fell, and new colors and products regularly entered the market. On January 1, 1868, there were 52 producers of aniline dyes. [3] Members of enlightened scientific societies from all over Europe competed for expertise and authority with dyers and printers in factories and workshops. [9] Many soluble salts of acid dyes synthesized for textile-related purposes were transformed into insoluble salts or lake pigments by reaction with water-soluble salts of calcium, barium or lead, whereas basic dyes were treated with tannins or antimony potassium tartrate to yield pigments. [4]

Synthetic coal tar alizarin dye samples, 1908. They are described systematically with their properties, such as fastness to light Sample Book, The Coal Tar Colour of Meister Lucius and Bruning, Ltd. . . . Applied in Calico Printing, Synthetic Red Dyes- Mordant Alizarins, plates 122-23, and Alizarin Turkey Reds, plate 169, 1908 (CH 68766109).jpg
Synthetic coal tar alizarin dye samples, 1908. They are described systematically with their properties, such as fastness to light

Synthetic alizarin 1868 – 1873

The development of synthetic alizarin opened up a huge market that was formerly served by natural dye makers. Alizarin was the first dye whose structure chemist determined, and they quickly set it as a target of synthesis, succeeding by 1868. [3] Other chemical components of natural madder were identified and applied by the mid-nineteenth century, including purpurin, which produced a delicate lilac colour, and green alizarin, which was patented in Britain and famously displayed at the 1867 Paris International Exhibition. [14] Similar to aniline dyes, the precursors for Synthetic Alizarin were easily obtainable from coal tar. Germany dominated the synthetic alizarin market, however foreign competition was not non-extant, for example the British Alizarine Company Ltd. [14]

Azo-dyes from coupling reactions 1878 – 1885

In 1858 Peter Griess passed ‘nitrous fumes’ () into a solution of 2-amino-4,6-dinitrophenol (picramic acid) and isolated a product belonging to a new class of compounds: Azo dyes. Later, a new class of azo dyes that were based on "coupling" reactions entered the market. The new azo dyes were easy to make and assumed a vast variety of incredibly intense colors based on the chosen precursors. [7] The chemists Z. Roussin, H. Caro, O. Witt, and P. Griess all put azo dyes on the market, and attempted to keep the syntheses as industrial secrets, Hoffman, however, determined the structure of their dyes and published his findings. [13]

4-hydroxyphenylazobenzene, a representative yellow azo dye compound 4-hydroxyphenylazobenzene.png
4-hydroxyphenylazobenzene, a representative yellow azo dye compound

This caused another rapid expansion, particularly in Germany. Between 1877 and 1887, 130 German patents for azo dyes were filed and 105 new dyes made it to market. [13] It also lead to a difference in how chemical companies interacted with consumers. German dye firms developed in-house marketing and distribution capabilities coordinated directly with their research and development departments. [3] Paul Schützenberger, in response to what he had seen at the 1878 Universal Exposition commented, "The abundance, the variety of combinations is such that we do not know whether to be more amazed by their multiplicity or by the imagination required to name them. Indeed, it is by the thousands that dyers create, every season, new colors for their sample cards." [8] Professional societies based on the synthetic dye industries began to form. [20] By the First World War, the largest number of dyes sold in the market fell into the class of azo dyes. [3] 1885, an azo-naphthol, Para-red, became the first water-insoluble organic pigment not containing acidic or basic groups. [4]

New dyes and larger markets 1900 – 1913

Historical BASF chemical products, including dyes Historische Farbstoffsammlung DSC00363.JPG
Historical BASF chemical products, including dyes

The twentieth century was again characterized by increases in scope and scale of chemical production. Pigments like cadmium selenide, manganese blue, molybdenum red, and bismuth vanadate were synthesized. High purity titanium dioxide and zinc oxide were produced for the first time on an industrial scale and introduced synthetic white pigments. [5] The first insoluble organic pigments, the red naphthols, containing neither acid nor basic groups, were produced and sold. [21] Furthermore, the quality of the new dyes increased. Chemist Rene Bohn developed a brilliant blue vat dye, indanthrone, with excellent color fastness in 1901. BASF(Badische Anilin und Soda Fabrik), the largest manufacturer of vat dyes, sold it as Indanthren Blue RS, along with the synthetic indigo they placed on the market in 1897. [19] Allegedly James Morton, a leader in England's textile industry, was out walking when he saw some tapestries he produced using aniline dyes had already faded, despite only recently being put on display. He was so dismayed that he began to have dye samples exposed to the sun to check for light-fastness. He then employed a Scottish chemist named John Christie to synthesize dyes based on the chemical structures that were more stable to sunlight, and began to market the dyes in his products as fast dyes, or sundour, which can translate to "hard to move" in Scots. [17]

Synthetic dyes were now produced in Britain, Germany, France, the US, Switzerland, Russia, the Austrian Empire, the Netherlands, Belgium, and Italy. At the end of this period, this grew to include Rumania (one firm), Greece (one firm), and Canada (two firms). [3] The scale of the chemical plants also grew, for instance the Bayer company in 1907 had a reactor to make azo dye with a capacity of 20,000 liters. [3] From 1900 to the first World War German firms controlled around 75% of the dye market. [3] The concentration of chemical producers in Germany was perturbed by WW1, however, and the chemical industry of the United States of America in particular expanded rapidly, although Germany always remained a major player.

WWI and the American dye industry 1913 – 1930

Through 1914, the US dye market was dominated by German imports, there were only a few small companies and German subsidiaries. With WW1, however, German dye factories now had to switch to making explosives and German shipping was cut off by British blockades. Prices quickly went up and U. S. companies built plants to meet demand. [22] American pharmaceutical giants, even at that time, like Dow, DuPont, and others began to produce dyes and were extremely successful with simple sulphur and vat dyes. Dow Chemical developed a synthetic process for indigo in 1915, and American industry and universities worked together to reverse engineer German chemical production secrets. After the war some American munitions factories converted to dye-works, intuiting that if the reverse was possible for the German chemical industry during the war, then it ought to be feasible. [23]

1881 Renoir painting featuring prominent use of bright red pigments Renoir, Pierre-Auguste - The Two Sisters, On the Terrace.jpg
1881 Renoir painting featuring prominent use of bright red pigments

Artistic use

Synthetic colorants gained popularity as quickly with artists as with industry. The painters of the impressionist school in particular were famous early adopters. Critical reviews of Impressionists’ blues made comparisons to laundresses’ tubs, in particular the practice of laundry bluing, and to chemical waste dumped into the Seine by dye factories. [8] One critic accused Edgar Degas, known for experiments in aquatint, pastel and oil painting as having an obsession with "chemistry," evoking a laboratory in description of his studio. Interestingly, Degas was known to be in correspondence with chemist Marcellin Berthelot, considered the father of organic synthetic chemistry in France. Pierre-Auguste Renoir’s later paintings relied heavily on alizarin crimson. He also employed cobalt blue or a mixture of ultramarine and cobalt blue, a synthetic pigment. [8] New pigments and dyes were not limited to the artists of Europe, even Japanese printmakers were using dyes like rosaniline as early as 1863. [24]

Colorants

Prussian Blue pigment applied to canvas with oil paint Prussian blue.jpg
Prussian Blue pigment applied to canvas with oil paint

Prussian Blue

Prussian Blue, also known as Berlin Blue, Paris Blue, or Turnbull's Blue, is an inorganic pigment, produced in large quantities for both artistic purposes and textiles. It has the chemical formula . With a history dating back to the early eighteenth century, Prussian blue remains a popular artistic pigment. Studies of Prussian Blue lead to discoveries about hydrogen cyanide. It is an antidote for heavy metal poisoning, and is famed for being used to color the uniforms of the Prussian army in the eighteenth century. [25]

Mauveine

Mauveine was discovered when Henry Perkin was trying to convert an artificial base into the natural alkaloid quinine. He tried adding aniline – a different base with a simpler construction. This created a black product. After purification, drying and washing with alcohol, Perkin had a mauve dye. Perkin filed his patent in August 1856 and a new dye industry was born. He at first called his discovery Tyrian Purple evoking the value of the ancient, highly expensive, pigment. Other names include aniline purple and Perkin's mauve. [7] Rather than one homogenous molecule, the original mauvine was primarily a mix of four major compounds, mauveine A, mauveine B, mauveine C, and mauveine B2, although there were other mauvine and pseudo mauveines in the dye product. [26]

The structure of alizarin Alizarin.svg
The structure of alizarin

Synthetic alizarin

Natural Alizarin was the first colorant to have its structure determined, making it one of the first targets for synthesis. The first synthesis of alizarin was patented by Carl Graebe and Carl Liebermann in 1868. It entailed the dibromination of anthraquinone, followed by fusion with sodium hydroxide. The second, much cheaper, synthetic path was developed in 1869 by Graebe, Liebermann and Heinrich Caro. It entailed the treatment of anthraquinone with fuming sulphuric acid, followed by a treatment with sodium hydroxide and potassium chlorate. Perkin submitted his own patent for a nearly identical process just a day later, and was awarded the patent in England. [6] [13]

Science

Colorants function through selective electromagnetic absorbance in the visible spectrum. A given pigment or dye molecule absorbs different wavelengths of electromagnetic radiation according to its atomic structure, and local chemical environment. The quantum behavior of a chemical typically results in distinct resonant frequencies of chemical bonds, which can be excited best by discrete wavelengths—meaning broad spectrum radiation has its spectra changed via absorption upon interaction. The physical shape, size, organization and concentration of dyes and pigments can also drastically affect observed color. Pigments are particularly susceptible to altered appearances based on physical properties.

Most modern synthetic dye molecules contain two components. The first part is an aromatic benzene ring or system of benzene rings, often substituted. The second is a chromophore, a conjugated double bond system with unsaturated groups. When exposed to visible light, this is the part that absorbs or reflects color. [16] [27] Other components of colorant molecules can tune intensity, color, solubility and substrate affinity.

Dyes and pigments can be categorized according to their synthetic or chemical properties. British chemist Edward Chambers Nicholson showed that pure aniline produced no dye. Hofmann showed that toluidine must be present to produce these dyes. Aniline dyes, including mauve, are prepared from aniline containing amounts of toluidine. [19] One can also classify dyes based on chemical formulas, azo-dyes from coupling, or diazonation—reactions that all have a characteristic azo group. [27]

Related Research Articles

<span class="mw-page-title-main">Dye</span> Soluble chemical substance or natural material which can impart color to other materials

A dye is a colored substance that chemically bonds to the substrate to which it is being applied. This distinguishes dyes from pigments which do not chemically bind to the material they color. Dye is generally applied in an aqueous solution and may require a mordant to improve the fastness of the dye on the fiber.

<span class="mw-page-title-main">Violet (color)</span> Color

Violet is the color of light at the short wavelength end of the visible spectrum, between blue and invisible ultraviolet. It is one of the seven colors that Isaac Newton labeled when dividing the spectrum of visible light in 1672. Violet light has a wavelength between approximately 380 and 435 nanometers. The color's name is derived from the violet flower.

<span class="mw-page-title-main">Pigment</span> Colored material

A pigment is a colored material that is completely or nearly insoluble in water. In contrast, dyes are typically soluble, at least at some stage in their use. Generally dyes are often organic compounds whereas pigments are often inorganic compounds. Pigments of prehistoric and historic value include ochre, charcoal, and lapis lazuli.

<span class="mw-page-title-main">Indigo dye</span> Chemical compound, food additive and dye

Indigo dye is an organic compound with a distinctive blue color. Historically, indigo was a natural dye extracted from the leaves of some plants of the Indigofera genus, in particular Indigofera tinctoria; dye-bearing Indigofera plants were commonly grown and used throughout the world, in Asia in particular, as an important crop, with the production of indigo dyestuff economically important due to the previous rarity of some blue dyestuffs historically.

<span class="mw-page-title-main">William Henry Perkin</span> British chemist known for his accidental discovery of the first synthetic dye

Sir William Henry Perkin was a British chemist and entrepreneur best known for his serendipitous discovery of the first commercial synthetic organic dye, mauveine, made from aniline. Though he failed in trying to synthesise quinine for the treatment of malaria, he became successful in the field of dyes after his first discovery at the age of 18.

<span class="mw-page-title-main">Food coloring</span> Substance used to color to food or drink

Food coloring, or color additive, is any dye, pigment, or substance that imparts color when it is added to food or drink. They come in many forms consisting of liquids, powders, gels, and pastes. Food coloring is used in both commercial food production and domestic cooking. Food colorants are also used in a variety of non-food applications, including cosmetics, pharmaceuticals, home craft projects, and medical devices.

<span class="mw-page-title-main">Alizarin</span> Chemical compound and histologic stain

Alizarin is an organic compound with formula C
14
H
8
O
4
that has been used throughout history as a prominent red dye, principally for dyeing textile fabrics. Historically it was derived from the roots of plants of the madder genus. In 1869, it became the first natural dye to be produced synthetically.

<span class="mw-page-title-main">Aniline</span> Organic compound (C₆H₅NH₂); simplest aromatic amine

Aniline is an organic compound with the formula C6H5NH2. Consisting of a phenyl group attached to an amino group, aniline is the simplest aromatic amine. It is an industrially significant commodity chemical, as well as a versatile starting material for fine chemical synthesis. Its main use is in the manufacture of precursors to polyurethane, dyes, and other industrial chemicals. Like most volatile amines, it has the odor of rotten fish. It ignites readily, burning with a smoky flame characteristic of aromatic compounds. It is toxic to humans.

<span class="mw-page-title-main">Mauveine</span>

Mauveine, also known as aniline purple and Perkin's mauve, was one of the first synthetic dyes. It was discovered serendipitously by William Henry Perkin in 1856 while he was attempting to synthesise the phytochemical quinine for the treatment of malaria. It is also among the first chemical dyes to have been mass-produced.

A lake pigment is a pigment made by precipitating a dye with an inert binder, or mordant, usually a metallic salt. Unlike vermilion, ultramarine, and other pigments made from ground minerals, lake pigments are organic. Manufacturers and suppliers to artists and industry frequently omit the lake designation in the name. Many lake pigments are fugitive because the dyes involved are not lightfast. Red lakes were particularly important in Renaissance and Baroque paintings; they were often used as translucent glazes to portray the colors of rich fabrics and draperies.

<span class="mw-page-title-main">Quinacridone</span> Organic compound used as a pigment

Quinacridone is an organic compound used as a pigment. Numerous derivatives constitute the quinacridone pigment family, which finds extensive use in industrial colorant applications such as robust outdoor paints, inkjet printer ink, tattoo inks, artists' watercolor paints, and color laser printer toner. As pigments, the quinacridones are insoluble. The development of this family of pigments supplanted the alizarin dyes.

<span class="mw-page-title-main">Brilliant blue FCF</span> Chemical compound

Brilliant blue FCF is a synthetic organic compound used primarily as a blue colorant for processed foods, medications, dietary supplements, and cosmetics. It is classified as a triarylmethane dye and is known under various names, such as FD&C Blue No. 1 or acid blue 9. It is denoted by E number E133 and has a color index of 42090. It has the appearance of a blue powder and is soluble in water and glycerol, with a maximum absorption at about 628 nanometers. It is one of the oldest FDA-approved color additives and is generally considered nontoxic and safe.

<span class="mw-page-title-main">Heinrich Caro</span> German chemist

Heinrich Caro, was a German Jewish chemist.

<span class="mw-page-title-main">Azo dye</span> Class of organic compounds used as dye

Azo dyes are organic compounds bearing the functional group R−N=N−R′, in which R and R′ are usually aryl and substituted aryl groups. They are a commercially important family of azo compounds, i.e. compounds containing the C-N=N-C linkage. Azo dyes are synthetic dyes and do not occur naturally. Most azo dyes contain only one azo group, but some dyes called "disazo dyes" contain two azo groups, some dyes called "trisazo dyes" contain three azo groups and are or more. Azo dyes comprise 60-70% of all dyes used in food and textile industries. Azo dyes are widely used to treat textiles, leather articles, and some foods. Chemically related derivatives of azo dyes include azo pigments, which are insoluble in water and other solvents.

<span class="mw-page-title-main">Diazonium compound</span> Diazonium salts of formula R-N≡N+

Diazonium compounds or diazonium salts are a group of organic compounds sharing a common functional group [R−N+≡N]X where R can be any organic group, such as an alkyl or an aryl, and X is an inorganic or organic anion, such as a halide.

In organic chemistry, an azo coupling is an organic reaction between a diazonium compound and another aromatic compound that produces an azo compound. In this electrophilic aromatic substitution reaction, the aryldiazonium cation is the electrophile and the activated arene is a nucleophile. In most cases, including the examples below, the diazonium compound is also aromatic.

<span class="mw-page-title-main">Rose madder</span> Red paint made from the madder plant

Rose madder is a red paint made from the pigment madder lake, a traditional lake pigment extracted from the common madder plant Rubia tinctorum.

<span class="mw-page-title-main">Anthraquinone dyes</span>

Anthraquinone dyes are an abundant group of dyes comprising a anthraquinone unit as the shared structural element. Anthraquinone itself is colourless, but red to blue dyes are obtained by introducing electron donor groups such as hydroxy or amino groups in the 1-, 4-, 5- or 8-position. Anthraquinone dyestuffs are structurally related to indigo dyestuffs and are classified together with these in the group of carbonyl dyes.

<span class="mw-page-title-main">Glossary of dyeing terms</span>

Dyeing is the craft of imparting colors to textiles in loose fiber, yarn, cloth or garment form by treatment with a dye. Archaeologists have found evidence of textile dyeing with natural dyes dating back to the Neolithic period. In China, dyeing with plants, barks and insects has been traced back more than 5,000 years. Natural insect dyes such as Tyrian purple and kermes and plant-based dyes such as woad, indigo and madder were important elements of the economies of Asia and Europe until the discovery of man-made synthetic dyes in the mid-19th century. Synthetic dyes quickly superseded natural dyes for the large-scale commercial textile production enabled by the industrial revolution, but natural dyes remained in use by traditional cultures around the world.

Rudolf Hugo Nietzki was a German chemist who specialized in industrial dyes derived from coal tar. While a professor at the University of Basel in Switzerland he initiated the university's association with to the local chemical industry.

References

  1. Malacara, Daniel (2011-08-04). Color Vision and Colorimetry: Theory and Applications, Second Edition. doi: 10.1117/3.881172 . ISBN   9780819483980.
  2. MacLeod, Roy (2003). "Toward a New SynthesisErnst Homburg;, Anthony S. Travis;, Harm G. Schröter (Editors). The Chemical Industry in Europe, 1850–1914: Industrial Growth, Pollution, and Professionalization. (Chemists and Chemistry, 17.) viii + 344 pp., illus., tables, index. Dordrecht, Netherlands: Kluwer Academic Publishers, 1998. $155, Nlg 285". Isis. 94 (1): 114–116. doi:10.1086/376104. ISSN   0021-1753. S2CID   146874448.
  3. 1 2 3 4 5 6 7 8 9 10 Murmann, Johann Peter; Homburg, Ernst (2002). "Comparing evolutionary dynamics across different national settings: the case of the synthetic dye industry, 1857–1914". Journal of Evolutionary Economics. 11 (2): 177–205. doi:10.1007/pl00003863. ISSN   0936-9937. S2CID   12992663.
  4. 1 2 3 Centeno, Silvia A.; Buisan, Virginia Lladó; Ropret, Polonca (2006). "Raman study of synthetic organic pigments and dyes in early lithographic inks (1890–1920)". Journal of Raman Spectroscopy. 37 (10): 1111–1118. Bibcode:2006JRSp...37.1111C. doi:10.1002/jrs.1594. ISSN   0377-0486.
  5. 1 2 3 4 5 6 7 Pfaff, Gerhard (2017-01-11). Inorganic Pigments. Berlin, Boston: De Gruyter. doi:10.1515/9783110484519. ISBN   978-3-11-048451-9.
  6. 1 2 3 4 Degano, Ilaria; Tognotti, Pietro; Kunzelman, Diane; Modugno, Francesca (2017-02-13). "HPLC-DAD and HPLC-ESI-Q-ToF characterisation of early 20th century lake and organic pigments from Lefranc archives". Heritage Science. 5 (1). doi: 10.1186/s40494-017-0120-y . ISSN   2050-7445.
  7. 1 2 3 4 5 6 Kanetkar, Vinod R. (2010). "Colour: History and advancements". Resonance. 15 (9): 794–803. doi:10.1007/s12045-010-0089-2. ISSN   0971-8044. S2CID   121668638.
  8. 1 2 3 4 Kalba, Laura Anne (2017). Color in the age of impressionism : commerce, technology, and art. University Park, Pennsylvania. ISBN   978-0-271-07980-6. OCLC   1052514699.
  9. 1 2 Nieto-Galan, Agusí (2010), "Between Craft Routines and Academic Rules", Materials and Expertise in Early Modern Europe, University of Chicago Press, pp. 321–354, doi:10.7208/chicago/9780226439709.003.0012, ISBN   978-0-226-43968-6 , retrieved 2020-02-14
  10. 1 2 3 Perkin, W. H. (1896). "The origin of the coal-tar colour industry, and the contributions of Hofmann and his pupils". Journal of the Chemical Society, Transactions. 69: 596. doi:10.1039/ct8966900596. ISSN   0368-1645.
  11. Reinhardt, Carsten; Travis, Anthony S. (2001), "Aspects of Paper Tools in the Industrial-Academic Context: Constitutions and Structures of Aniline Dyes, 1860–1880", Tools and Modes of Representation in the Laboratory Sciences, Springer Netherlands, pp. 79–94, doi:10.1007/978-94-015-9737-1_6, ISBN   978-90-481-5859-1 , retrieved 2020-02-14
  12. Sonawane, Shirish; Setty, Y. Pydi; Sapavatu, Srinu Naik, eds. (2015-04-15). Chemical and Bioprocess Engineering. Apple Academic Press. doi:10.1201/b18402. ISBN   978-0-429-15974-9.
  13. 1 2 3 4 5 6 Travis, Anthony S.; Hornix, Willem J.; Bud, Robert; Hornix, Willem J. (1992). "From process to plant: innovation in the early artificial dye industry". The British Journal for the History of Science. 25 (1): 65–90. doi:10.1017/s0007087400045337. ISSN   0007-0874. S2CID   144110681.
  14. 1 2 3 4 Tuckett, Sally; Nenadic, Stana (2012). "Colouring the Nation: A New In-Depth Study of the Turkey Red Pattern Books in the National Museums Scotland". Textile History. 43 (2): 161–182. doi:10.1179/0040496912z.00000000016. hdl: 20.500.11820/eacbb713-62e2-46f6-9945-212414398d71 . ISSN   0040-4969. S2CID   191476924.
  15. Thompson, Lewis (1850). The Chemistry Of Gas Lighting; FROM VOLS II. AND III. OF "THE JOURNAL, OF GAS LIGHTING.". London: J. Clatton.
  16. 1 2 Bernard, James P. (2018-07-16). "A Brief History of Synthetic Dyes | First Source Worldwide, LLC". First Source Worldwide. Retrieved 2020-02-14.
  17. 1 2 3 "The colourful chemistry of artificial dyes". Science Museum. Retrieved 2020-02-14.
  18. Travis, Anthony S. (1998), "Theory from Practice: Portraying the Constitution of Synthetic Dyestuffs in the 1860s", The Invisible Industrialist, Palgrave Macmillan UK, pp. 122–142, doi:10.1007/978-1-349-26443-8_5, ISBN   978-1-349-26445-2 , retrieved 2020-02-14
  19. 1 2 3 "Dye - Development of synthetic dyes". Encyclopedia Britannica. Retrieved 2020-02-14.
  20. "Home Page". SDC | Society of Dyers and Colourists. Retrieved 2020-02-14.
  21. Gabrieli, F.; Doherty, B.; Miliani, C.; Degano, I.; Modugno, F.; Uldank, D.; Kunzelman, D.; Buzzegoli, E.; Patti, M.; Rosi, F. (2016-10-17). "Micro-Raman and SER spectroscopy to unfold Lefranc's early organic pigment formulations". Journal of Raman Spectroscopy. 47 (12): 1505–1513. Bibcode:2016JRSp...47.1505G. doi:10.1002/jrs.5052. ISSN   0377-0486.
  22. "Colorants Industry History". www.colorantshistory.org. Archived from the original on December 11, 2004. Retrieved 2020-02-14.{{cite web}}: CS1 maint: unfit URL (link)
  23. Mock, Gary N (2002). "The textile dye industry in the United States". Review of Progress in Coloration and Related Topics. 32 (1): 80–87. doi:10.1111/j.1478-4408.2002.tb00252.x.
  24. Cesaratto, Anna; Luo, Yan-Bing; Smith, Henry D.; Leona, Marco (2018-04-03). "A timeline for the introduction of synthetic dyestuffs in Japan during the late Edo and Meiji periods". Heritage Science. 6 (1). doi: 10.1186/s40494-018-0187-0 . ISSN   2050-7445.
  25. Haythornthwaite, Philip J. (28 November 1991). Frederick the Great's army. Osprey. ISBN   1-85532-160-2. OCLC   24796374.
  26. Cova, Tânia F. G. G.; Pais, Alberto A. C. C.; Seixas de Melo, J. Sérgio (2017-07-28). "Reconstructing the historical synthesis of mauveine from Perkin and Caro: procedure and details". Scientific Reports. 7 (1): 6806. Bibcode:2017NatSR...7.6806C. doi: 10.1038/s41598-017-07239-z . ISSN   2045-2322. PMC   5533699 . PMID   28754965.
  27. 1 2 Lichtarowicz, Marek. "Colorants". www.essentialchemicalindustry.org. Retrieved 2020-02-14.

Further reading