Acetic acid: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
m Reverting possible vandalism by 24.36.1.172 to version by DePiep. False positive? Report it. Thanks, ClueBot NG. (1640082) (Bot)
No edit summary
Line 125: Line 125:
}}
}}


'''Acetic acid''' {{IPAc-en|ə|ˈ|s|iː|t|ɨ|k}} (systematically named '''ethanoic acid''' {{IPAc-en|ˌ|ɛ|θ|ə|ˈ|n|oʊ|ɨ|k}}) is an [[organic compound]] with the [[chemical formula]] CH<sub>3</sub>COOH (also written as CH<sub>3</sub>CO<sub>2</sub>H or C<sub>2</sub>H<sub>4</sub>O<sub>2</sub>). It is a colourless liquid that when undiluted is also called ''glacial acetic acid''. Acetic acid is the main component of [[vinegar]] (apart from water; vinegar is roughly 8% acetic acid by volume), and has a distinctive sour taste and pungent smell. Besides its production as household vinegar, it is mainly produced as a precursor to polyvinylacetate and cellulose acetate. Although it is classified as a [[acid strength|weak acid]], concentrated acetic acid is corrosive and attacks the skin.
'''Acetic acid''' {{IPAc-en|ə|ˈ|s|iː|t|ɨ|k}} (systematically named '''ethanoic acid''' {{IPAc-en|ˌ|ɛ|θ|ə|ˈ|n|oʊ|ɨ|k}}) is an [[organic compound]] with the [[chemical formula]] CH<sub>3</sub>COOH (also written as CH<sub>3</sub>CO<sub>2</sub>H or C<sub>2</sub>H<sub>4</sub>O<sub>2</sub>). It is a colourless liquid that when undiluted is also called ''glacial acetic acid''. Acetic acid is the main component of [[vinegar]] (apart from water; vinegar is roughly 8% acetic acid by volume), and smells like fart, or methane. Besides its production as household vinegar, it is mainly produced as a precursor fart fart fart fart fart fart fart to polyvinylacetate and cellulose acetate. Although it is classified as a [[acid strength|weak acid]], concentrated acetic acid is corrosive and attacks the skin.


Acetic acid is one of the simplest [[carboxylic acid]]s. It is an important [[reagent|chemical reagent]] and industrial chemical, mainly used in the production of [[cellulose acetate]] for [[photographic film]] and [[polyvinyl acetate]] for wood [[Adhesive|glue]], as well as synthetic fibers and fabrics. In households, diluted acetic acid is often used in [[descaling agent]]s. In the [[food industry]], acetic acid is used under the [[E number|food additive code]] E260 as an [[acidity regulator]] and as a condiment. As a [[food additive]] it is approved for usage in many countries, including Canada,<ref name=CDJ2013>{{cite web|title=Food and Drug Regulations (C.R.C., c. 870)|url=http://laws-lois.justice.gc.ca/eng/regulations/C.R.C.,_c._870/FullText.html|work=Consolidated Regulations|publisher=Canadian Department of Justice|accessdate=21 July 2013|date=31 May 2013}}</ref> the EU,<ref>UK Food Standards Agency: {{cite web |url=http://www.food.gov.uk/safereating/chemsafe/additivesbranch/enumberlist |title=Current EU approved additives and their E Numbers |accessdate=27 October 2011}}</ref> USA<ref>US Food and Drug Administration: {{cite web |url=http://www.fda.gov/Food/FoodIngredientsPackaging/FoodAdditives/FoodAdditiveListings/ucm091048.htm |title=Listing of Food Additives Status Part I |accessdate=27 October 2011}}</ref> and Australia and New Zealand.<ref>Australia New Zealand Food Standards Code{{cite web |url=http://www.comlaw.gov.au/Details/F2011C00827 |title=Standard 1.2.4 - Labeling of ingredients |accessdate=27 October 2011}}</ref>
Acetic acid is one of the simplest [[carboxylic acid]]s. It is an important [[reagent|chemical reagent]] and industrial chemical, mainly used in the production of [[cellulose acetate]] for [[photographic film]] and [[polyvinyl acetate]] for wood [[Adhesive|glue]], as well as synthetic fibers and fabrics. In households, diluted acetic acid is often used in [[descaling agent]]s. In the [[food industry]], acetic acid is used under the [[E number|food additive code]] E260 as an [[acidity regulator]] and as a condiment. As a [[food additive]] it is approved for usage in many countries, including Canada,<ref name=CDJ2013>{{cite web|title=Food and Drug Regulations (C.R.C., c. 870)|url=http://laws-lois.justice.gc.ca/eng/regulations/C.R.C.,_c._870/FullText.html|work=Consolidated Regulations|publisher=Canadian Department of Justice|accessdate=21 July 2013|date=31 May 2013}}</ref> the EU,<ref>UK Food Standards Agency: {{cite web |url=http://www.food.gov.uk/safereating/chemsafe/additivesbranch/enumberlist |title=Current EU approved additives and their E Numbers |accessdate=27 October 2011}}</ref> USA<ref>US Food and Drug Administration: {{cite web |url=http://www.fda.gov/Food/FoodIngredientsPackaging/FoodAdditives/FoodAdditiveListings/ucm091048.htm |title=Listing of Food Additives Status Part I |accessdate=27 October 2011}}</ref> and Australia and New Zealand.<ref>Australia New Zealand Food Standards Code{{cite web |url=http://www.comlaw.gov.au/Details/F2011C00827 |title=Standard 1.2.4 - Labeling of ingredients |accessdate=27 October 2011}}</ref>

Revision as of 22:09, 5 January 2014

Acetic acid
Skeletal formula of acetic acid
Skeletal formula of acetic acid
Spacefill model of acetic acid
Spacefill model of acetic acid
Skeletal formula of acetic acid with all explicit hydogens added
Skeletal formula of acetic acid with all explicit hydogens added
Ball and stick model of acetic acid
Ball and stick model of acetic acid
Sample of acetic acid in a reagent bottle
Names
IUPAC name
Acetic acid[3][4]
Systematic IUPAC name
Ethanoic acid[5]
Other names
Methanecarboxylic acid[1][2]
Identifiers
3D model (JSmol)
3DMet
Abbreviations AcOH
506007
ChEBI
ChEMBL
ChemSpider
DrugBank
ECHA InfoCard 100.000.528 Edit this at Wikidata
EC Number
  • 200-580-7
E number E260 (preservatives)
1380
KEGG
MeSH Acetic+acid
RTECS number
  • AF1225000
UNII
UN number 2789
  • InChI=1S/C2H4O2/c1-2(3)4/h1H3,(H,3,4) checkY
    Key: QTBSBXVTEAMEQO-UHFFFAOYSA-N checkY
  • CC(O)=O
Properties
C2H4O2
Molar mass 60.052 g·mol−1
Appearance Colourless liquid
Odor vinegar-like
Density 1.049 g cm-3
Melting point 16 °C; 61 °F; 289 K
Boiling point 118 °C; 244 °F; 391 K
Miscible
log P -0.322
Acidity (pKa) 4.76
Basicity (pKb) 9.198
1.371
Viscosity 1.22 mPa s
1.74 D
Thermochemistry
123.1 J K-1 mol-1
158.0 J K-1 mol-1
-483.88--483.16 kJ mol-1
-875.50--874.82 kJ mol-1
Hazards
GHS labelling:
GHS02: Flammable GHS05: Corrosive
Danger
H226, H314
P280, P305+P351+P338, P310
NFPA 704 (fire diamond)
NFPA 704 four-colored diamondHealth 3: Short exposure could cause serious temporary or residual injury. E.g. chlorine gasFlammability 2: Must be moderately heated or exposed to relatively high ambient temperature before ignition can occur. Flash point between 38 and 93 °C (100 and 200 °F). E.g. diesel fuelInstability 0: Normally stable, even under fire exposure conditions, and is not reactive with water. E.g. liquid nitrogenSpecial hazards (white): no code
3
2
0
Flash point 40 °C (104 °F; 313 K)
Explosive limits 4-16%
Lethal dose or concentration (LD, LC):
3.31 g kg-1, oral (rat)
Related compounds
Supplementary data page
Acetic acid (data page)
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).
☒N verify (what is checkY☒N ?)

Acetic acid /əˈst[invalid input: 'ɨ']k/ (systematically named ethanoic acid /ˌɛθəˈn[invalid input: 'ɨ']k/) is an organic compound with the chemical formula CH3COOH (also written as CH3CO2H or C2H4O2). It is a colourless liquid that when undiluted is also called glacial acetic acid. Acetic acid is the main component of vinegar (apart from water; vinegar is roughly 8% acetic acid by volume), and smells like fart, or methane. Besides its production as household vinegar, it is mainly produced as a precursor fart fart fart fart fart fart fart to polyvinylacetate and cellulose acetate. Although it is classified as a weak acid, concentrated acetic acid is corrosive and attacks the skin.

Acetic acid is one of the simplest carboxylic acids. It is an important chemical reagent and industrial chemical, mainly used in the production of cellulose acetate for photographic film and polyvinyl acetate for wood glue, as well as synthetic fibers and fabrics. In households, diluted acetic acid is often used in descaling agents. In the food industry, acetic acid is used under the food additive code E260 as an acidity regulator and as a condiment. As a food additive it is approved for usage in many countries, including Canada,[6] the EU,[7] USA[8] and Australia and New Zealand.[9]

The global demand of acetic acid is around 6.5 million tonnes per year (Mt/a), of which approximately 1.5 Mt/a is met by recycling; the remainder is manufactured from petrochemical feedstock.[10] As a chemical reagent, biological sources of acetic acid are of interest but generally uncompetitive. Vinegar is dilute acetic acid, often produced by fermentation and subsequent oxidation of ethanol.

Nomenclature

The trivial name acetic acid is the most commonly used and preferred IUPAC name. The systematic name ethanoic acid, a valid IUPAC name, is constructed according to the substitutive nomenclature.[5] The name acetic acid derives from acetum, the Latin word for vinegar, and is related to the word acid itself.

Glacial acetic acid is a trivial name for water-free (anhydrous) acetic acid. Similar to the German name Eisessig (ice-vinegar), the name comes from the ice-like crystals that form slightly below room temperature at 16.6 °C (61.9 °F) (the presence of 0.1% water lowers its melting point by 0.2 °C).[11]

A common abbreviation for acetic acid is AcOH, where Ac stands for the acetyl group CH3−C(=O)−. Acetate (CH3COO) is abbreviated AcO. The Ac is not to be confused with the abbreviation for the chemical element actinium.[12] To better reflect its structure, acetic acid is often written as CH3–C(O)OH, CH3–C(=O)OH, CH3COOH, and CH3CO2H. In the context of acid-base reactions, the abbreviation HAc is sometimes used,[13] where Ac instead stands for acetate. Acetate is the ion resulting from loss of H+ from acetic acid. The name acetate can also refer to a salt containing this anion, or an ester of acetic acid.[14]

History

Vinegar was known early in civilization as the natural result of exposure of beer and wine to air, because acetic acid-producing bacteria are present globally. The use of acetic acid in alchemy extends into the 3rd century BC, when the Greek philosopher Theophrastus described how vinegar acted on metals to produce pigments useful in art, including white lead (lead carbonate) and verdigris, a green mixture of copper salts including copper(II) acetate. Ancient Romans boiled soured wine to produce a highly sweet syrup called sapa. Sapa that was produced in lead pots was rich in lead acetate, a sweet substance also called sugar of lead or sugar of Saturn, which contributed to lead poisoning among the Roman aristocracy.[15]

In the Renaissance, glacial acetic acid was prepared through the dry distillation of certain metal acetates (the most noticeable one being copper(II) acetate). The 16th-century German alchemist Andreas Libavius described such a procedure, and he compared the glacial acetic acid produced by this means to vinegar. The presence of water in vinegar has such a profound effect on acetic acid's properties that for centuries chemists believed that glacial acetic acid and the acid found in vinegar were two different substances. French chemist Pierre Adet proved them identical.[15][16]

glass beaker of crystallized acetic acid
Crystallised acetic acid.

In 1845 German chemist, Hermann Kolbe synthesized acetic acid from inorganic compounds for the first time. This reaction sequence consisted of chlorination of carbon disulfide to carbon tetrachloride, followed by pyrolysis to tetrachloroethylene and aqueous chlorination to trichloroacetic acid, and concluded with electrolytic reduction to acetic acid.[17]

By 1910, most glacial acetic acid was obtained from the "pyroligneous liquor" from distillation of wood. The acetic acid was isolated from this by treatment with milk of lime, and the resulting calcium acetate was then acidified with sulfuric acid to recover acetic acid. At that time, Germany was producing 10,000 tons of glacial acetic acid, around 30% of which was used for the manufacture of indigo dye.[15][18]

Because both methanol and carbon monoxide are commodity raw materials, methanol carbonylation long appeared to be attractive precursors to acetic acid. Henri Dreyfus at British Celanese developed a methanol carbonylation pilot plant as early as 1925.[19] However, a lack of practical materials that could contain the corrosive reaction mixture at the high pressures needed (200 atm or more) discouraged commercialization of these routes. The first commercial methanol carbonylation process, which used a cobalt catalyst, was developed by German chemical company BASF in 1963. In 1968, a rhodium-based catalyst (cis−[Rh(CO)2I2]) was discovered that could operate efficiently at lower pressure with almost no by-products. US chemical company Monsanto Company built the first plant using this catalyst in 1970, and rhodium-catalyzed methanol carbonylation became the dominant method of acetic acid production (see Monsanto process). In the late 1990s, the chemicals company BP Chemicals commercialized the Cativa catalyst ([Ir(CO)2I2]), which is promoted by ruthenium for greater efficiency. This iridium-catalyzed Cativa process is greener and more efficient[20] and has largely supplanted the Monsanto process, often in the same production plants.

In the interstellar medium

Acetic acid was discovered in interstellar medium in 1996 by a team led by David Mehringer[21] who detected it using the former Berkeley-Illinois-Maryland Association array at the Hat Creek Radio Observatory and the former Millimeter Array located at the Owens Valley Radio Observatory. It was first detected in the Sagittarius B2 North molecular cloud (also known as the Sgr B2 Large Molecule Heimat source). Acetic acid has the distinction of being the first molecule discovered in the interstellar medium using solely radio interferometers; in all previous ISM molecular discoveries made in the millimeter and centimeter wavelength regimes, single dish radio telescopes were at least partly responsible for the detections.[21]

Chemical properties

Acetic acid crystals

Acidity

The hydrogen center in the carboxyl group (−COOH) in carboxylic acids such as acetic acid can separate from the molecule by ionization:

CH3CO2H → CH3CO2- + H+

Because of this release of the proton (H+), acetic acid has acidic character. Acetic acid is a weak monoprotic acid. In aqueous solution, it has a pKa value of 4.76.[22] Its conjugate base is acetate (CH3COO). A 1.0 M solution (about the concentration of domestic vinegar) has a pH of 2.4, indicating that merely 0.4% of the acetic acid molecules are dissociated.[23]

Deprotonation equilibrium of acetic acid in water

Cyclic dimer of acetic acid; dashed green lines represent hydrogen bonds

Structure

In solid acetic acid, the molecules form pairs (dimers), being connected by hydrogen bonds.[24] The dimers can also be detected in the vapour at 120 °C (248 °F). Dimers also occur in the liquid phase in dilute solutions in non-hydrogen-bonding solvents, and a certain extent in pure acetic acid,[25] but are disrupted by hydrogen-bonding solvents. The dissociation enthalpy of the dimer is estimated at 65.0–66.0 kJ/mol, and the dissociation entropy at 154–157 J mol−1 K−1.[26] Other lower carboxylic acids dimerize in a similar fashion.[27]

Solvent properties

Liquid acetic acid is a hydrophilic (polar) protic solvent, similar to ethanol and water. With a moderate relative static permittivity (dielectric constant) of 6.2, it dissolves not only polar compounds such as inorganic salts and sugars, but also non-polar compounds such as oils and elements such as sulfur and iodine. It readily mixes with other polar and non-polar solvents such as water, chloroform, and hexane. With higher alkanes (starting with octane), acetic acid is not completely miscible anymore, and its miscibility continues to decline with longer n-alkanes.[28] This dissolving property and miscibility of acetic acid makes it a widely used industrial chemical, for example, as a solvent in the production of dimethyl terephthalate.[10]

Chemical reactions

Organic chemistry

Two typical organic reactions of acetic acid

Acetic acid undergoes the typical chemical reactions of a carboxylic acid. Upon treatment with a standard base, it converts to metal acetate and water. With strong bases (e.g., organolithium reagents), it can be doubly deprotonated to give LiCH2CO2Li. Reduction of acetic acid gives ethanol. The OH group is the main site of reaction, as illustrated by the conversion of acetic acid to acetyl chloride. Other substitution derivatives include acetic anhydride; this anhydride is produced by loss of water from two molecules of acetic acid. Esters of acetic acid can likewise be formed via Fischer esterification, and amides can be formed. When heated above 440 °C (824 °F), acetic acid decomposes to produce carbon dioxide and methane, or to produce ketene and water:[29][30][31]

CH3COOH → CH4 + CO2
CH3COOH → CH2CO + H2O

Reactions with inorganic compounds

Acetic acid is mildly corrosive to metals including iron, magnesium, and zinc, forming hydrogen gas and salts called acetates:

Mg + 2 CH3COOH → (CH3COO)2Mg + H2

Because aluminium forms a passivating acid-resistant film of aluminium oxide, aluminium tanks are used to transport acetic acid. Metal acetates can also be prepared from acetic acid and an appropriate base, as in the popular "baking soda + vinegar" reaction:

NaHCO3 + CH3COOH → CH3COONa + CO2 + H2O

A color reaction for salts of acetic acid is iron(III) chloride solution, which results in a deeply red color that disappears after acidification.[32] A more sensitive test uses lanthanum nitrate with iodine and ammonia to give a blue solution.[33] Acetates when heated with arsenic trioxide form cacodyl oxide, which can be detected by its malodorous vapors.[34]

Biochemistry

At physiological pHs, acetic acid is usually fully ionized to acetate. In biochemistry, acetate and acetic acid are equivalent.

The acetyl group, derived from acetic acid, is fundamental to all forms of life. When bound to coenzyme A, it is central to the metabolism of carbohydrates and fats. Unlike longer-chain carboxylic acids (the fatty acids), acetic acid does not occur in natural triglycerides. However, the artificial triglyceride triacetin (glycerine triacetate) is a common food additive and is found in cosmetics and topical medicines.[35]

Acetic acid is produced and excreted by acetic acid bacteria, notable ones being the Acetobacter genus and Clostridium acetobutylicum. These bacteria are found universally in foodstuffs, water, and soil, and acetic acid is produced naturally as fruits and other foods spoil. Acetic acid is also a component of the vaginal lubrication of humans and other primates, where it appears to serve as a mild antibacterial agent.[36]

Production

Purification and concentration plant for acetic acid in 1884

Acetic acid is produced industrially both synthetically and by bacterial fermentation. About 75% of acetic acid made for use in the chemical industry is made by the carbonylation of methanol, explained below.[10] Alternative methods account for the rest. The biological route accounts for only about 10% of world production, but it remains important for the production of vinegar, as many food purity laws stipulate that vinegar used in foods must be of biological origin. As of 2003–2005, total worldwide production of virgin acetic acid was estimated at 5 Mt/a (million tonnes per year), approximately half of which was then produced in the United States. European production stood at approximately 1 Mt/a and was declining, and 0.7 Mt/a were produced in Japan. Another 1.5 Mt were recycled each year, bringing the total world market to 6.5 Mt/a.[37][38] Since then the global production has increased to 10.7 Mt/a (in 2010), and further, however, slowing increase in production is predicted.[39] The two biggest producers of virgin acetic acid are Celanese and BP Chemicals. Other major producers include Millennium Chemicals, Sterling Chemicals, Samsung, Eastman, and Svensk Etanolkemi.[40]

Methanol carbonylation

Most acetic acid is produced by methanol carbonylation. In this process, methanol and carbon monoxide react to produce acetic acid according to the equation:

CH3OH + CO → CH3COOH

The process involves iodomethane as an intermediate, and occurs in three steps. A catalyst, metal carbonyl, is needed for the carbonylation (step 2).[41]

  1. CH3OH + HI → CH3I + H2O
  2. CH3I + CO → CH3COI
  3. CH3COI + H2O → CH3COOH + HI

Two related processes for the carbonylation of methanol: the rhodium-catalyzed Monsanto process, and the iridium-catalyzed Cativa process. The latter process is greener and more efficient[20] and has largely supplanted the former process, often in the same production plants. Catalytic amounts of water are used in both processes, but the Cativa process requires less, so the water-gas shift reaction is suppressed, and less byproducts are formed.

By altering the process conditions, acetic anhydride may also be produced on the same plant using the rhodium catalysts.[42]

Acetaldehyde oxidation

Prior to the commercialization of the Monsanto process, most acetic acid was produced by oxidation of acetaldehyde. This remains the second-most-important manufacturing method, although it is usually uncompetitive with the carbonylation of methanol.

The acetaldehyde may be produced via oxidation of butane or light naphtha, or by hydration of ethylene. When butane or light naphtha is heated with air in the presence of various metal ions, including those of manganese, cobalt, and chromium, peroxides form and then decompose to produce acetic acid according to the chemical equation

2 C4H10 + 5 O2 → 4 CH3COOH + 2 H2O

The typical reaction is conducted at temperatures and pressures designed to be as hot as possible while still keeping the butane a liquid. Typical reaction conditions are 150 °C (302 °F) and 55 atm.[43] Side-products may also form, including butanone, ethyl acetate, formic acid, and propionic acid. These side-products are also commercially valuable, and the reaction conditions may be altered to produce more of them where needed. However, the separation of acetic acid from these by-products adds to the cost of the process.[44]

Under similar conditions and using similar catalysts as are used for butane oxidation, the oxygen in air to produce acetic acid can oxidize acetaldehyde.[44]

2 CH3CHO + O2 → 2 CH3COOH

Using modern catalysts, this reaction can have an acetic acid yield greater than 95%. The major side-products are ethyl acetate, formic acid, and formaldehyde, all of which have lower boiling points than acetic acid and are readily separated by distillation.[44]

Ethylene oxidation

Acetaldehyde may be prepared from ethylene via the Wacker process, and then oxidized as above. In more recent times, chemical company Showa Denko, which opened an ethylene oxidation plant in Ōita, Japan, in 1997, commercialized a cheaper single-stage conversion of ethylene to acetic acid.[45] The process is catalyzed by a palladium metal catalyst supported on a heteropoly acid such as tungstosilicic acid. It is thought to be competitive with methanol carbonylation for smaller plants (100–250 kt/a), depending on the local price of ethylene. The approach will be based on utilizing a novel selective photocatalytic oxidation technology for the selective oxidation of ethylene and ethane to acetic acid. Unlike traditional oxidation catalysts, the selective oxidation process will use UV light to produce acetic acid at ambient temperatures and pressure.

Oxidative fermentation

For most of human history, acetic acid bacteria of the genus Acetobacter have made acetic acid, in the form of vinegar. Given sufficient oxygen, these bacteria can produce vinegar from a variety of alcoholic foodstuffs. Commonly used feeds include apple cider, wine, and fermented grain, malt, rice, or potato mashes. The overall chemical reaction facilitated by these bacteria is:

C2H5OH + O2 → CH3COOH + H2O

A dilute alcohol solution inoculated with Acetobacter and kept in a warm, airy place will become vinegar over the course of a few months. Industrial vinegar-making methods accelerate this process by improving the supply of oxygen to the bacteria.[46]

The first batches of vinegar produced by fermentation probably followed errors in the winemaking process. If must is fermented at too high a temperature, acetobacter will overwhelm the yeast naturally occurring on the grapes. As the demand for vinegar for culinary, medical, and sanitary purposes increased, vintners quickly learned to use other organic materials to produce vinegar in the hot summer months before the grapes were ripe and ready for processing into wine. This method was slow, however, and not always successful, as the vintners did not understand the process.[47]

One of the first modern commercial processes was the "fast method" or "German method", first practised in Germany in 1823. In this process, fermentation takes place in a tower packed with wood shavings or charcoal. The alcohol-containing feed is trickled into the top of the tower, and fresh air supplied from the bottom by either natural or forced convection. The improved air supply in this process cut the time to prepare vinegar from months to weeks.[48]

Nowadays, most vinegar is made in submerged tank culture, first described in 1949 by Otto Hromatka and Heinrich Ebner.[49] In this method, alcohol is fermented to vinegar in a continuously stirred tank, and oxygen is supplied by bubbling air through the solution. Using modern applications of this method, vinegar of 15% acetic acid can be prepared in only 24 hours in batch process, even 20% in 60-hour fed-batch process.[47]

The oxygen supply should be kept optimum. Lack of O2 will kill the bacteria because they are extremely sensitive to overcome this problem hence we have to use efficient aeration. Efficient aeration can be achieved with the used of compressed air and proper mechanical device. Consider shear stress imparted to the fluid and the microorganisms. The efficiency depends on the ratio between the energy inputs necessary per unit weight of O2 transferred to the culture.[citation needed]

When there is over-oxidation, acetic acid will convert to CO2 and H2O, this will decrease acetic acid production. Acetic acid concentrations above 6% of the total culture and avoiding the total depletion of ethanol will prevent overoxidization.[citation needed]

Anaerobic fermentation

Species of anaerobic bacteria, including members of the genus Clostridium or Acetobacterium can convert sugars to acetic acid directly, without using ethanol as an intermediate. The overall chemical reaction conducted by these bacteria may be represented as:

C6H12O6 → 3 CH3COOH

These acetogenic bacteria produce acetic acid from one-carbon compounds, including methanol, carbon monoxide, or a mixture of carbon dioxide and hydrogen:

2 CO2 + 4 H2 → CH3COOH + 2 H2O

This ability of Clostridium to utilize sugars directly, or to produce acetic acid from less costly inputs, means that these bacteria could potentially produce acetic acid more efficiently than ethanol-oxidizers like Acetobacter. However, Clostridium bacteria are less acid-tolerant than Acetobacter. Even the most acid-tolerant Clostridium strains can produce vinegar of only a few per cent acetic acid, compared to Acetobacter strains that can produce vinegar of up to 20% acetic acid. At present, it remains more cost-effective to produce vinegar using Acetobacter than to produce it using Clostridium and then concentrate it. As a result, although acetogenic bacteria have been known since 1940, their industrial use remains confined to a few niche applications.[50]

Uses

Acetic acid is a chemical reagent for the production of chemical compounds. The largest single use of acetic acid is in the production of vinyl acetate monomer, closely followed by acetic anhydride and ester production. The volume of acetic acid used in vinegar is comparatively small.[10][38]

Vinyl acetate monomer

The major use of acetic acid is for the production of vinyl acetate monomer (VAM). In 2008, this application was estimated to consume one third of the world's production of acetic acid.[10] The reaction is of ethylene and acetic acid with oxygen over a palladium catalyst, conducted in the gas phase.[51]

2 H3C–COOH + 2 C2H4 + O2 → 2 H3C–CO–O–CH=CH2 + 2 H2O

Vinyl acetate can be polymerized to polyvinyl acetate or to other polymers, which are components in paints and adhesives.[51]

Ester production

The major esters of acetic acid are commonly used solvents for inks, paints and coatings. The esters include ethyl acetate, n-butyl acetate, isobutyl acetate, and propyl acetate. They are typically produced by catalyzed reaction from acetic acid and the corresponding alcohol:

H3C-COOH + HO-R → H3C-CO-O-R + H2O, (R = a general alkyl group)

Most acetate esters, however, are produced from acetaldehyde using the Tishchenko reaction. In addition, ether acetates are used as solvents for nitrocellulose, acrylic lacquers, varnish removers, and wood stains. First, glycol monoethers are produced from ethylene oxide or propylene oxide with alcohol, which are then esterified with acetic acid. The three major products are ethylene glycol monoethyl ether acetate (EEA), ethylene glycol monobutyl ether acetate (EBA), and propylene glycol monomethyl ether acetate (PMA, more commonly known as PGMEA in semiconductor manufacturing processes, where it is used as a resist solvent). This application consumes about 15% to 20% of worldwide acetic acid. Ether acetates, for example EEA, have been shown to be harmful to human reproduction.[38]

Acetic anhydride

The product of the condensation of two molecules of acetic acid is acetic anhydride. The worldwide production of acetic anhydride is a major application, and uses approximately 25% to 30% of the global production of acetic acid. The main process involves dehydration of acetic acid to give ketene at 700-750 °C. Ketene is thereafter reacted with acetic acid to obtain the anhydride:[52]

CH3CO2H → CH2=C=O + H2O
CH3CO2H + CH2=C=O → (CH3CO)2O

Acetic anhydride is an acetylation agent. As such, its major application is for cellulose acetate, a synthetic textile also used for photographic film. Acetic anhydride is also a reagent for the production of heroin and other compounds.[52]

Use as solvent

Glacial acetic acid is an excellent polar protic solvent, as noted above. It is frequently used as a solvent for recrystallization to purify organic compounds. Acetic acid is used as a solvent in the production of terephthalic acid (TPA), the raw material for polyethylene terephthalate (PET). In 2006, about 20% of acetic acid was used for TPA production.[38]

Acetic acid is often used as a solvent for reactions involving carbocations, such as Friedel-Crafts alkylation. For example, one stage in the commercial manufacture of synthetic camphor involves a Wagner-Meerwein rearrangement of camphene to isobornyl acetate; here acetic acid acts both as a solvent and as a nucleophile to trap the rearranged carbocation.[53]

Glacial acetic acid is used in analytical chemistry for the estimation of weakly alkaline substances such as organic amides. Glacial acetic acid is a much weaker base than water, so the amide behaves as a strong base in this medium. It then can be titrated using a solution in glacial acetic acid of a very strong acid, such as perchloric acid.[54]

Medical use

Diluted acetic acid is used in physical therapy using iontophoresis.[55]

Vinegar

Vinegar is typically 4-18% acetic acid by mass. Vinegar is used directly as a condiment, and in the pickling of vegetables and other foods. Table vinegar tends to be more diluted (4% to 8% acetic acid), while commercial food pickling employs solutions that are more concentrated. The amount of acetic acid used as vinegar on a worldwide scale is not large, but is by far the oldest and best-known application.[56]

Other derivatives

Organic or inorganic salts are produced from acetic acid, including:

Substituted acetic acids produced include:

Amounts of acetic acid used in these other applications together (apart from TPA) account for another 5–10% of acetic acid use worldwide. These applications are, however, not expected to grow as much as TPA production.[38]

Health effects and Safety

Concentrated acetic acid is corrosive to skin and must, therefore, be handled with appropriate care, since it can cause skin burns, permanent eye damage, and irritation to the mucous membranes.[57][58] These burns or blisters may not appear until hours after exposure. Latex gloves offer no protection, so specially resistant gloves, such as those made of nitrile rubber, are worn when handling the compound. Concentrated acetic acid can be ignited with difficulty in the laboratory. It becomes a flammable risk if the ambient temperature exceeds 39 °C (102 °F), and can form explosive mixtures with air above this temperature (explosive limits: 5.4–16%). Acetic acid is a strong eye, skin, and mucous membrane irritant. Prolonged skin contact with glacial acetic acid may result in tissue destruction. Inhalation exposure (8 hours) to acetic acid vapours at 10 ppm could produce some irritation of eyes, nose, and throat; at 100 ppm marked lung irritation and possible damage to lungs, eyes, and skin might result. Vapour concentrations of 1,000 ppm cause marked irritation of eyes, nose and upper respiratory tract and cannot be tolerated. These predictions were based on animal experiments and industrial exposure. Skin sensitization to acetic acid is rare, but has occurred. It has been reported that, 12 workers exposed for two or more years to an estimated mean acetic acid airborne concentration of 51 ppm, there were symptoms of conjunctive irritation, upper respiratory tract irritation, and hyperkeratotic dermatitis. Exposure to 50 ppm or more is intolerable to most persons and results in intensive lacrimation and irritation of the eyes, nose, and throat, with pharyngeal oedema and chronic bronchitis. Un acclimatized humans experience extreme eye and nasal irritation at concentrations in excess of 25 ppm, and conjunctivitis from concentrations below 10 ppm has been reported. In a study of 5 workers exposed for 7 to 12 years to concentrations of 80 to 200 ppm at peaks, the principal findings were blackening and hyperkeratosis of the skin of the hands, conjunctivitis (but no corneal damage), bronchitis and pharyngitis, and erosion of the exposed teeth (incisors and canines).[59]

The hazards of solutions of acetic acid depend on the concentration; solutions of 10-25 w/. The following table lists the EU classification of acetic acid solutions:[60]

Concentration
by weight
Molarity Classification R-Phrases
10–25% 1.67–4.16 mol/L Irritant (Xi) Template:R36/38
25–90% 4.16–14.99 mol/L Corrosive (C) Template:R34
>90% >14.99 mol/L Corrosive (C) Flammable (F) Template:R10, Template:R35

Solutions at more than 25% acetic acid are handled in a fume hood because of the pungent, corrosive vapor. Dilute acetic acid, in the form of vinegar, is harmless. However, ingestion of stronger solutions is dangerous to human and animal life. It can cause severe damage to the digestive system, and a potentially lethal change in the acidity of the blood.

Due to incompatibilities, it is recommended to keep acetic acid away from chromic acid, ethylene glycol, nitric acid, perchloric acid, permanganates, peroxides and hydroxyls.[61]

See also

References

  1. ^ Scientific literature reviews on generally recognized as safe (GRAS) food ingredients. National Technical Information Service. 1974. p. 1.
  2. ^ "Chemistry", volume 5, Encyclopedia Britannica, 1961, page 374
  3. ^ IUPAC, Commission on Nomenclature of Organic Chemistry (1993). "Table 28(a) Carboxylic acids and related groups. Unsubstituted parent structures". A Guide to IUPAC Nomenclature of Organic Compounds (Recommendations 1993). Blackwell Scientific publications. ISBN 0-632-03488-2.
  4. ^ "Acetic Acid - PubChem Public Chemical Database". The PubChem Project. USA: National Center for Biotechnology Information.
  5. ^ a b IUPAC Provisional Recommendations 2004 Chapter P-12.1; page 4
  6. ^ "Food and Drug Regulations (C.R.C., c. 870)". Consolidated Regulations. Canadian Department of Justice. 31 May 2013. Retrieved 21 July 2013.
  7. ^ UK Food Standards Agency: "Current EU approved additives and their E Numbers". Retrieved 27 October 2011.
  8. ^ US Food and Drug Administration: "Listing of Food Additives Status Part I". Retrieved 27 October 2011.
  9. ^ Australia New Zealand Food Standards Code"Standard 1.2.4 - Labeling of ingredients". Retrieved 27 October 2011.
  10. ^ a b c d e Hosea Cheung; Robin S. Tanke; G. Paul Torrence. "Acetic Acid". Ullmann's Encyclopedia of Industrial Chemistry. Weinheim: Wiley-VCH. doi:10.1002/14356007.a01_045.pub2. ISBN 978-3527306732.
  11. ^ Armarego,W.L.F. and Chai,Christina (2009). Purification of Laboratory Chemicals, 6th edition. Butterworth-Heinemann. ISBN 1-85617-567-7.{{cite book}}: CS1 maint: multiple names: authors list (link)
  12. ^ Cooper, Caroline (9 August 2010). Organic Chemist's Desk Reference (2 ed.). CRC Press. pp. 102–104. ISBN 1-4398-1166-0.
  13. ^ DeSousa, Luís R. (1995). Common Medical Abbreviations. Cengage Learning. p. 97. ISBN 0-8273-6643-4.
  14. ^ Hendrickson, James B.; Cram, Donald J.; Hammond, George S. (1970). Organic Chemistry (3 ed.). Tokyo: McGraw Hill Kogakusha. p. 135.
  15. ^ a b c Martin, Geoffrey (1917). Industrial and Manufacturing Chemistry (Part 1, Organic ed.). London: Crosby Lockwood. pp. 330–331.
  16. ^ P. A. Adet (1798). "Mémoire sur l'acide acétique (Memoir on acetic acid)". Annales de Chemie. 27: 299–319.
  17. ^ Goldwhite, Harold (2003). "Short summary of the career of the German organic chemist, Hermann Kolbe" (PDF). New Haven Section Bulletin American Chemical Society. 20 (3). {{cite journal}}: Unknown parameter |month= ignored (help)
  18. ^ Schweppe, Helmut (1979). "Identification of dyes on old textiles". Journal of the American Institute for Conservation. 19 (1/3): 14–23. doi:10.2307/3179569. JSTOR 3179569.
  19. ^ Wagner, Frank S. (1978). "Acetic acid". In Grayson, Martin (ed.). Kirk-Othmer Encyclopedia of Chemical Technology (3rd ed.). New York: John Wiley & Sons.
  20. ^ a b Lancaster, Mike (2002). Green Chemistry, an Introductory Text. Cambridge: Royal Society of Chemistry. pp. 262–266. ISBN 0-85404-620-8.
  21. ^ a b Mehringer, David M.; et al. (1997). "Detection and Confirmation of Interstellar Acetic Acid". Astrophysical Journal Letters. 480: L71. Bibcode:1997ApJ...480L..71M. doi:10.1086/310612Template:Inconsistent citations{{cite journal}}: CS1 maint: postscript (link)
  22. ^ Goldberg, R.; Kishore, N.; Lennen, R. (2002). "Thermodynamic Quantities for the Ionization Reactions of Buffers" (PDF). Journal of Physical and Chemical Reference Data. 31 (2): 231–370. Bibcode:1999JPCRD..31..231G. doi:10.1063/1.1416902.
  23. ^ [H3O+] = 10−2.4 = 0.4 %
  24. ^ Jones, R.E.; Templeton, D.H. (1958). "The crystal structure of acetic acid". Acta Crystallographica. 11 (7): 484–487. doi:10.1107/S0365110X58001341.
  25. ^ Briggs, James M.; Toan B. Nguyen; William L. Jorgensen (1991). "Monte Carlo simulations of liquid acetic acid and methyl acetate with the OPLS potential functions". Journal of Physical Chemistry. 95 (8): 3315–3322. doi:10.1021/j100161a065.
  26. ^ Togeas, James B. (2005). "Acetic Acid Vapor: 2. A Statistical Mechanical Critique of Vapor Density Experiments". Journal of Physical Chemistry A. 109 (24): 5438–5444. doi:10.1021/jp058004j. PMID 16839071.
  27. ^ McMurry, John (2000). Organic Chemistry (5 ed.). Brooks/Cole. p. 818. ISBN 0-534-37366-6.
  28. ^ Zieborak, K.; K. Olszewski (1958). Bulletin de L'Academie Polonaise des Sciences-Serie des Sciences Chimiques Geologiques et Geographiques. 6 (2): 3315–3322. {{cite journal}}: Missing or empty |title= (help)
  29. ^ P. G. Blake; G. E. Jackson (1968). "The thermal decomposition of acetic acid". Journal of the Chemical Society B Physical Organic: 1153–1155. doi:10.1039/J29680001153.
  30. ^ C. H. Bamford; M. J. S. Dewar (1949). "608. The thermal decomposition of acetic acid". Journal of the Chemical Society: 2877. doi:10.1039/JR9490002877.
  31. ^ Duan, Xiaofeng; Michael Page (1995). "Theoretical Investigation of Competing Mechanisms in the Thermal Unimolecular Decomposition of Acetic Acid and the Hydration Reaction of Ketene". Journal of the American Chemical Society. 117 (18): 5114–5119. doi:10.1021/ja00123a013. ISSN 0002-7863.
  32. ^ Charlot, G.; Murray, R. G. (1954). Qualitative Inorganic Analysis (4 ed.). CUP Archive. p. 110.
  33. ^ Neelakantam, K.; L Ramachangra Row (1940). "The Lanthanum Nitrate Test for Acetatein Inorganic Qualitative Analysis" (PDF). Retrieved 5 June 2013.
  34. ^ Brantley, L. R.; T. M. Cromwell; J. F. Mead (1947). "Detection of acetate ion by the reaction with arsenious oxide to form cacodyl oxide". Journal of Chemical Education. 24 (7): 353. Bibcode:1947JChEd..24..353B. doi:10.1021/ed024p353. ISSN 0021-9584.
  35. ^ Fiume, M. Z. (June 2003). "Final report on the safety assessment of triacetin". International Journal of Toxicology. 22 (Suppl 2): 1–10. doi:10.1177/1091581803022S203. PMID 14555416.
  36. ^ executive ed.: J. Buckingham (1996). Dictionary of Organic Compounds. Vol. 1 (6th ed.). London: Chapman & Hall. ISBN 0-412-54090-8.
  37. ^ "Production report". Chemical & Engineering News: 67–76. 11 July 2005.
  38. ^ a b c d e Suresh, Bala (2003). "Acetic Acid". Chemicals Economic Handbook. SRI International. p. 602.5000.
  39. ^ Acetic Acid :: Petrochemicals :: World Petrochemicals :: SRI Consulting. http://chemical.ihs.com/WP/Public/Reports/acetic_acid/ (accessed Dec 18, 2011).
  40. ^ "Reportlinker Adds Global Acetic Acid Market Analysis and Forecasts". Market Research Database. March 2009. p. contents. Retrieved 6 June 2013.
  41. ^ Yoneda, N.; Kusano, S.; Yasui, M.; Pujado, P.; Wilcher, S. (2001). "Recent advances in processes and catalysts for the production of acetic acid". Applied Catalysis A, General. 221 (1–2): 253–265. doi:10.1016/S0926-860X(01)00800-6.
  42. ^ Zoeller, J. R.; Agreda, V. H.; Cook, S. L.; Lafferty, N. L.; Polichnowski, S. W.; Pond, D. M. (1992). "Eastman Chemical Company Acetic Anhydride Process". Catalysis Today. 13 (1): 73–91. doi:10.1016/0920-5861(92)80188-S.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  43. ^ Chenier, Philip J. (2002). Survey of Industrial Chemistry (3 ed.). Springer. p. 151. ISBN 0-306-47246-5.
  44. ^ a b c Sano, Ken‐ichi; Hiroshi Uchida; Syoichirou Wakabayashi (1999). "A new process for acetic acid production by direct oxidation of ethylene". Catalysis Surveys from Japan. 3 (1): 55–60. doi:10.1023/A:1019003230537. ISSN 1384-6574.
  45. ^ Sano, Ken-ichi; Uchida, Hiroshi; Wakabayashi, Syoichirou (1999). A new process for acetic acid production by direct oxidation of ethylene. Vol. 3. pp. 66–60. doi:10.1023/A:1019003230537. {{cite book}}: |journal= ignored (help)
  46. ^ Chotani, Gopal K.; Gaertner, Alfred L.; Arbige, Michael V.; Timothy C. Dodge (2007). "Industrial Biotechnology: Discovery to Delivery". Kent and Riegel's Handbook of Industrial Chemistry and Biotechnology. Springer. pp. 32–34. ISBN 978-0-387-27842-1.
  47. ^ a b Otto Hromatka and Heinrich Ebner (1959). "Vinegar by Submerged Oxidative Fermentation". Industrial & Engineering Chemistry. 51 (10): 1279–1280. doi:10.1021/ie50598a033.
  48. ^ Everett P. Partridge (1931). "Acetic Acid and Cellulose Acetate in the United States A General Survey of Economic and Technical Developments". Industrial & Engineering Chemistry. 23 (5): 482–498. doi:10.1021/ie50257a005.
  49. ^ O Hromatka, H Ebner (1949). "Investigations on vinegar fermentation: Generator for vinegar fermentation and aeration procedures". Enzymologia. 13: 369.
  50. ^ Jia Huey Sim, Azlina Harun Kamaruddin, Wei Sing Long and Ghasem Najafpour (2007). "Clostridium aceticum—A potential organism in catalyzing carbon monoxide to acetic acid: Application of response surface methodology". Enzyme and Microbial Technology. 40 (5): 1234–1243. doi:10.1016/j.enzmictec.2006.09.017.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  51. ^ a b Günter Roscher. "VInyl Esters". Ullmann's Encyclopedia of Industrial Chemistry. Weinheim: Wiley-VCH. doi:10.1002/14356007.a27_419. ISBN 978-3527306732.
  52. ^ a b Heimo Held; Alfred Rengstl; Dieter Mayer. "Acetic Anhydride and Mixed Fatty Acid Anhydrides". Ullmann's Encyclopedia of Industrial Chemistry. Weinheim: Wiley-VCH. doi:10.1002/14356007.a01_065. ISBN 978-3527306732.
  53. ^ Sell, Charles S. (2006). "4.2.15 Bicyclic Monoterpenoids". The Chemistry of Fragrances: From Perfumer to Consumer. RSC Paperbacks Series. Vol. 38 (2 ed.). Great Britain: Royal Society of Chemistry. p. 80. ISBN 0-85404-824-3.
  54. ^ Felgner, Andrea. "Titration in Non-Aqueous Media". Sigma-Aldrich. Retrieved 5 June 2013.
  55. ^ Kolt, Gregory S.; Snyder-Mackler, Lynn (2007). Physical Therapies in Sport and Exercise. Elsevier Health Sciences. p. 223. ISBN 978-0-443-10351-3. Retrieved 7 June 2013.
  56. ^ Bernthsen, A.; Sudborough, J. J. (1922). Organic Chemistry. London: Blackie and Son. p. 155.
  57. ^ "ICSC 0363 – ACETIC ACID". International Programme on Chemical Safety. 5 June 2010.
  58. ^ "Occupational Safety and Health Guideline for Acetic Acid" (PDF). Centers for Disease Control and Prevention. Retrieved 8 May 2013.
  59. ^ Virginia Department of Health DIVISION OF HEALTH HAZARDS CONTROLPREPARED BY: PETER C. SHERERTZ, Ph.D. (Toxicologist) June 1, 1994
  60. ^ Yee, Allan (10 May 2013). "HSIS Consolidated List – Alphabetical Index". Safe Work Australia. Retrieved 11 June 2013.
  61. ^ "Acetic acid MSDS". 21 May 2013. Retrieved 7 June 2013.

External links

Template:Link FA Template:Link FA Template:Link FA Template:Link FA Template:Link GA