Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 30

c 

c 

c 
Urea


   
Carbamide, carbonyl diamide,
carbonyldiamine, diaminomethanal,
diaminomethanone

c  or    is an organic compound with the chemical formula (NH2)2CO.


The molecule has two amine (-NH2) groups joined by a carbonyl (C=O) functional
group.
Urea serves an important role in the metabolism of nitrogen-containing compounds
by animals and is the main nitrogen-containing substance in the urine of mammals.
It is solid, colourless, and odorless (although the ammonia which it gives off in the
presence of water, including water vapor in the air, has a strong odor). It is highly
soluble in water and non-toxic. Dissolved in water it is neither acidic nor alkaline.
The body uses it in many processes, most notably nitrogen excretion. Urea is
widely used in fertilizers as a convenient source of nitrogen. Urea is also an
important raw material for the chemical industry. The synthesis of this organic
compound by Friedrich Wöhler in 1828 from an inorganic precursor was an
important milestone in the development of organic chemistry, as it showed for the
first time that a molecule found in living organisms could be synthesized in the lab
without biological starting materials.

The terms {  and    are also used for a class of chemical compounds
sharing the same functional group RR'N-CO-NRR', namely a carbonyl group
attached to two organic amine residues. Examples include carbamide peroxide,
allantoin, and hydantoin. Ureas are closely related to biurets and related in
structure to amides, carbamates, carbodiimides, and thiocarbamides.

  

` 1 History
` 2 Physiology
2 2.1 In humans
2 2.2 In other species
` Ô Uses
2 Ô.1 Agriculture
2 Ô.2 Chemical industry
2 Ô.Ô Explosive
2 Ô.4 Automobile systems
2 Ô.5 Other commercial uses
2 Ô.6 Laboratory uses
2 Ô.7 Medical use
2 Ô.8 Analysis
` 4 Production
2 4.1 Industrial methods
2 4.2 Laboratory preparation
2 4.Ô Historical process
` 5 Chemical properties
2 5.1 Molecular and crystal structure
2 5.2 Reactions
` 6 Safety
` 7 References
` 8 External links

Π  

Urea was first discovered in urine in 177Ô by the French chemist Hilaire Rouelle.
In 1828, the German chemist Friedrich Wöhler obtained urea by treating silver
isocyanate with ammonium chloride.[2]

AgNCO + NH4Cl ĺ (NH2)2CO + AgCl

This was the first time an organic compound was artificially synthesized from
inorganic starting materials, without the involvement of living organisms. The
results of this experiment implicitly discredited vitalism: the theory that the
chemicals of living organisms are fundamentally different from inanimate matter.
This insight was important for the development of organic chemistry. His
discovery prompted Wöhler to write triumphantly to Berzelius: "I must tell you
that I can make urea without the use of kidneys, either man or dog. Ammonium
cyanate is urea." For this discovery, Wöhler is considered by many[’ ] the father
of organic chemistry.

Π  

Urea is synthesized in the body of many organisms as part of the urea cycle, either
from the oxidation of amino acids or from ammonia. In this cycle, amino groups
donated by ammonia and L-aspartate are converted to urea, while L-ornithine,
citrulline, L-argininosuccinate, and L-arginine act as intermediates. Urea
production occurs in the liver and is regulated by N-acetylglutamate. Urea is found
dissolved in blood (in the reference range of 2.5 to 6.7 mmol/liter) and is excreted
by the kidney as a component of urine. In addition, a small amount of urea is
excreted (along with sodium chloride and water) in sweat.

Amino acids from ingested food which are not used for the synthesis of proteins
and other biological substances are oxidized by the body, yielding urea and carbon
dioxide, as an alternative source of energy.[Ô] The oxidation pathway starts with the
removal of the amino group by a transaminase, the amino group is then fed into the
urea cycle.
Ammonia (NHÔ) is another common byproduct of the metabolism of nitrogenous
compounds. Ammonia is smaller, more volatile and more mobile than urea. If
allowed to accumulate, ammonia would raise the pH in cells to toxic levels.
Therefore many organisms convert ammonia to urea, even though this synthesis
has a net energy cost. Being practically neutral and highly soluble in water, urea is
a safe vehicle for the body to transport and excrete excess nitrogen.

In water, the amine groups undergo slow displacement by water molecules,


producing ammonia and carbonate anion. For this reason, old, stale urine has a
stronger odor than fresh urine.

Π {  

The handling of urea by the kidneys is a vital part of human metabolism. Besides
its role as carrier of waste nitrogen, urea also plays a role in the countercurrent
exchange system of the nephrons, that allows for re-absorption of water and critical
ions from the excreted urine. Urea is reabsorbed in the inner medullary collecting
ducts of the nephrons,[4] thus raising the osmolarity in the medullary interstitium
surrounding the thin ascending limb of the loop of Henle, which in turn causes
water to be reabsorbed. By action of the urea transporter 2, some of this reabsorbed
urea will eventually flow back into the thin ascending limb of the tubule, through
the collecting ducts, and into the excreted urine.

This mechanism, which is controlled by the antidiuretic hormone, allows the body
to create hyperosmotic urine, that has a higher concentration of dissolved
substances than the blood plasma. This mechanism is important to prevent the loss
of water, to maintain blood pressure, and to maintain a suitable concentration of
sodium ions in the blood plasmas.

The equivalent nitrogen content (in gram) of urea (in mmol) can be estimated by
the conversion factor 0.028 g/mmol.[5] Furthermore, 1 gram of nitrogen is roughly
equivalent to 6 gram of protein, and 1 gram of protein is roughly equivalent to
4 gram of muscle tissue. Subsequently, in situations such as muscle wasting,
1 mmol of excessive urea in the urine (as measured by urine volume in litres
multiplied by urea concentration in mmol/l) roughly corresponds to a muscle loss
of 0.67 gram.

Π    

In aquatic organisms the most common form of nitrogen waste is ammonia, while
land-dwelling organisms convert the toxic ammonia to either urea or uric acid.
Urea is found in the urine of mammals and amphibians, as well as some fish. Birds
and saurian reptiles have a different form of nitrogen metabolism, that requires less
water and leads to nitrogen being excreted in the form of uric acid. It is noteworthy
that tadpoles excrete ammonia but shift to urea production during metamorphosis.
Despite the generalization above, the urea pathway has been documented not only
in mammals and amphibians but in many other organisms as well, including birds,
invertebrates, insects, plants, yeast, fungi, and even microorganisms.[O  ]

Πc  

Π { { 

More than 90% of world production of urea is destined for use as a nitrogen-
release fertilizer. Urea has the highest nitrogen content of all solid nitrogenous
fertilizers in common use. Therefore, it has the lowest transportation costs per unit
of nitrogen nutrient. The standard crop nutrient rating of urea is .[O  ]

Many soil bacteria possess the enzyme, urease, which catalyzes the conversion of
the urea molecule to two ammonia molecules and one carbon dioxide molecule,
thus urea fertilizers are very rapidly transformed to the ammonium form in soils.
Among soil bacteria known to carry urease, some ammonia-oxidizing bacteria
(AOB), such as species of Nitrosomonas are also able to assimilate the carbon
dioxide released by the reaction to make biomass via the Calvin Cycle, and harvest
energy by oxidizing ammonia (the other product of urease) to nitrite, a process
termed nitrification.[6] Nitrite-oxidizing bacteria, especially, Nitrobacter, oxidize
nitrite to nitrate, which is extremely mobile in soils and is a major cause of water
pollution from agriculture. Ammonia and nitrate are readily absorbed by plants,
and are the dominant sources of nitrogen for plant growth. Urea is also used in
many multi-component solid fertilizer formulations. Urea is highly soluble in water
and is, therefore, also very suitable for use in fertilizer solutions (in combination
with ammonium nitrate: UAN), e.g., in 'foliar feed' fertilizers. For fertilizer use,
granules are preferred over prills because of their narrower particle size
distribution which is an advantage for mechanical application.

The most common impurity of synthetic urea is biuret, which impairs plant growth.

Urea is usually spread at rates of between 40 and Ô00 kg/ha but rates vary. Smaller
applications incur lower losses due to leaching. During summer, urea is often
spread just before, or during rain to minimize losses from volatilization (process
wherein nitrogen is lost to the atmosphere as ammonia gas). Urea is not compatible
with other fertilizers.
Because of the high nitrogen concentration in urea, it is very important to achieve
an even spread. The application equipment must be correctly calibrated and
properly used. Drilling must not occur on contact with or close to seed, due to the
risk of germination damage. Urea dissolves in water for application as a spray or
through irrigation systems.

In grain and cotton crops, urea is often applied at the time of the last cultivation
before planting. In high rainfall areas and on sandy soils (where nitrogen can be
lost through leaching) and where good in-season rainfall is expected, urea can be
side- or top-dressed during the growing season. Top-dressing is also popular on
pasture and forage crops. In cultivating sugarcane, urea is side-dressed after
planting, and applied to each ratoon crop.

In irrigated crops, urea can be applied dry to the soil, or dissolved and applied
through the irrigation water. Urea will dissolve in its own weight in water, but it
becomes increasingly difficult to dissolve as the concentration increases.
Dissolving urea in water is endothermic, causing the temperature of the solution to
fall when urea dissolves.

As a practical guide, when preparing urea solutions for fertigation (injection into
irrigation lines), dissolve no more than Ô0 kg urea per 100 L water.

In foliar sprays, urea concentrations of 0.5% ± 2.0% are often used in horticultural
crops. Low-biuret grades of urea are often indicated.

Urea absorbs moisture from the atmosphere and therefore is typically stored either
in closed/sealed bags on pallets, or, if stored in bulk, under cover with a tarpaulin.
As with most solid fertilizers, storage in a cool, dry, well-ventilated area is
recommended.

Π  { 

Urea is a raw material for the manufacture of many important chemical


compounds, such as

` üarious plastics, especially the urea-formaldehyde resins.


` üarious adhesives, such as urea-formaldehyde or the urea-melamine-
formaldehyde used in marine plywood.
` Potassium cyanate, another industrial feedstock.

Π 


Urea can be used to make urea nitrate, a high explosive which is used industrially
and as part of some improvised explosive devices.

Π{     

Urea is used in SNCR and SCR reactions to reduce the NOx pollutants in exhaust
gases from combustion from diesel, dual fuel, and lean-burn natural gas engines.
The BlueTec system, for example, injects water-based urea solution into the
exhaust system. The ammonia produced by the hydrolysis of the urea reacts with
the nitrogen oxide emissions and is converted into nitrogen and water within the
catalytic converter.

Π
   {  

This section          {  .


Please help improve this section by adding citations to reliable sources.
Unsourced material may be challenged and removed. À  

` A stabilizer in nitrocellulose explosives


` A component of animal feed, providing a relatively cheap source of nitrogen
to promote growth
` A non-corroding alternative to rock salt for road de-icing, and the
resurfacing of snowboarding halfpipes and terrain parks
` A flavor-enhancing additive for cigarettes
` A main ingredient in hair removers such as Nair and üeet
` A browning agent in factory-produced pretzels
` An ingredient in some skin cream,[7] moisturizers, hair conditioners
` A reactant in some ready-to-use cold compresses for first-aid use, due to the
endothermic reaction it creates when mixed with water
` A cloud seeding agent, along with other salts
` A flame-proofing agent, commonly used in dry chemical fire extinguisher
charges such as the urea-potassium bicarbonate mixture
` An ingredient in many tooth whitening products
` An ingredient in dish soap
` Along with ammonium phosphate, as a yeast nutrient, for fermentation of
sugars into ethanol
` A nutrient used by plankton in ocean nourishment experiments for
geoengineering purposes
` As an additive to extend the working temperature and open time of hide glue
` As a solubility-enhancing and moisture-retaining additive to dye baths for
textile dyeing or printing

Π!   {  

Urea in concentrations up to 10 M is a powerful protein denaturant as it disrupts


the noncovalent bonds in the proteins. This property can be exploited to increase
the solubility of some proteins. A mixture of urea and choline chloride is used as a
deep eutectic solvent, a type of ionic liquid.

Urea can in principle serve as a hydrogen source for subsequent power generation
in fuel cells. Urea present in urine/wastewater can be used directly (though bacteria
normally quickly degrade urea.) Producing hydrogen by electrolysis of urea
solution occurs at a lower voltage (0.Ô7ü) and thus consumes less energy than the
electrolysis of water (1.2ü).[8]

Π"{ 

Urea-containing creams are used as topical dermatological products to promote


rehydration of the skin. Urea 40% is indicated for psoriasis, xerosis,
onychomycosis, ichthyosis, eczema, keratosis, keratoderma, corns and calluses. If
covered by an occlusive dressing, 40% urea preparations may also be used for
nonsurgical debridement of nails. Urea 40% "dissolves the intercellular matrix"[9]
of the nail plate. Only diseased or dystrophic nails are removed as there is no effect
on healthy portions of the nail. This drug is also used as an earwax removal aid.

Certain types of instant cold packs (or ice packs) contain water and separated urea
crystals. Rupturing the internal water bag starts an endothermic reaction and allows
the pack to be used to reduce swelling.

Like saline, urea injection is used to perform abortions.

Urea is the main component of an alternative medicinal treatment referred to as


urine therapy.

The blood urea nitrogen (BUN) test is a measure of the amount of nitrogen in the
blood that comes from urea. It is used as a marker of renal function.

Urea labeled with carbon-14 or carbon-1Ô is used in the urea breath test, which is
used to detect the presence of the bacteria î O O   (î   ) in the
stomach and duodenum of humans, associated with peptic ulcers. The test detects
the characteristic enzyme urease, produced by î   , by a reaction that
produces ammonia from urea. This increases the pH (reduces acidity) of the
stomach environment around the bacteria. Similar bacteria species to î   can
be identified by the same test in animals such as apes, dogs, and cats (including big
cats).

Π  

Urea is readily quantified by a number of different methods, such as the diacetyl


monoxime colorimetric method, and the Berthelot reaction (after initial conversion
of urea to ammonia via urease). These methods are amenable to high throughput
instrumentation, such as automated flow injection analyzers[10] and 96-well micro-
plate spectrophotometers.[11]

Π { 

Urea is produced on a scale of some 100,000,000 tons per year worldwide.[12]

Π{    

For use in industry, urea is produced from synthetic ammonia and carbon dioxide.
Large quantities of carbon dioxide are produced during the manufacture of
ammonia from coal or from hydrocarbons such as natural gas and petroleum-
derived raw materials. Such point sources of CO2 facilitate direct synthesis of urea.

The basic process, developed in 1922, is also called the Bosch-Meiser urea process
after its discoverers. The various urea processes are characterized by the conditions
under which urea formation takes place and the way in which unconverted
reactants are further processed. The process consists of two main equilibrium
reactions, with incomplete conversion of the reactants. The first is an exothermic
reaction of liquid ammonia with dry ice to form ammonium carbamate (H2N-
COONH4):[1Ô]

2 NHÔ + CO2 ļ H2N-COONH4 ()

The second is an endothermic decomposition of ammonium carbamate into urea


and water:

H2N-COONH4 ļ (NH2)2CO + H2O

Both reactions combined are exothermic.[12]


Unconverted reactants can be used for the manufacture of other products, for
example ammonium nitrate or sulfate, or they can be recycled for complete
conversion to urea in a total-recycle process.

Urea can be produced as prills, granules, pellets, crystals, and solutions. Solid urea
is marketed as prills or granules. The advantage of prills is that, in general, they
can be produced more cheaply than granules. Properties such as impact strength,
crushing strength, and free-flowing behaviour are, in particular, important in
product handling, storage, and bulk transportation. Typical impurities in the
production are biuret and isocyanic acid:

2 NH2CONH2 ĺ H2NCONHCONH2 + NHÔ


NH2CONH2 ĺ HNCO + NHÔ

The biuret content is a serious concern because it is often toxic to the very plants
that are to be fertilized. Urea is classified on the basis of its biuret content.

Π!      

Ureas in the more general sense can be accessed in the laboratory by reaction of
phosgene with primary or secondary amines, proceeding through an isocyanate
intermediate. Non-symmetric ureas can be accessed by reaction of primary or
secondary amines with an isocyanate.

Π    

Historically urea was first noticed by Hermann Boerhaave in the early 18th century
from evaporates of urine. In 177Ô, Hilaire Rouelle obtained crystals containing
urea from dog's urine by evaporating it and treating it with alcohol in successive
filtrations. This method was aided by Carl Wilhelm Scheele's discovery that urine
treated by concentrated nitric acid precipitated crystals. Antoine François, comte
de Fourcroy and Louis Nicolas üauquelin discovered in 1799 that the nitrated
crystals were identical to Rouelle's substance and invented the term "urea."
Berzelius made further improvedments to its purification and finally William
Prout, in 1817, succeeded in obtaining and determining the chemical composition
of the pure substance. In the evolved procedure, urea was precipitated as urea
nitrate by adding strong nitric acid to urine. To purify the resulting crystals, they
were dissolved in boiling water with charcoal and filtered. After cooling pure
crystals of urea nitrate form. To reconstitute the urea from the nitrate, the crystals
are dissolved in warm water, and barium carbonate added. The water is then
evaporated and annhydrous alcohol added to extract the urea. This solution is
drained off and allowed to evaporate resulting in pure urea.

Π     

Π"{    { { 

The urea molecule is planar in the crystal structure, but the geometry around the
nitrogens is pyramidal in the gas-phase minimum-energy structure.[14] In solid urea,
the oxygen center is engaged in two N-H-O hydrogen bonds. The resulting dense
and energetically favourable hydrogen-bond network is probably established at the
cost of efficient molecular packing: The structure is quite open, the ribbons
forming tunnels with square cross-section. The carbon in urea is described as sp2
hybridized, the C-N bonds have significant double bond character, and the
carbonyl oxygen is basic compared to, say, formaldehyde. Urea's high aqueous
solubility reflects its ability to engage in extensive hydrogen bonding with water.

By virtue of its tendency to form a porous frameworks, urea has the ability to trap
many organic compounds. In these so-called clathrates, the organic "guest"
molecules are held in channels formed by interpenetrating helices comprising of
hydrogen-bonded urea molecules. This behaviour can be used to separate mixtures,
e.g. in the production of aviation fuel and lubricating oils, and in the separation of
paraffins.

As the helices are interconnected, all helices in a crystal must have the same
molecular handedness. This is determined when the crystal is nucleated and can
thus be forced by seeding. The resulting crystals have been used to separate
racemic mixtures.

Π#  

Urea reacts with alcohols to form urethanes. Urea reacts with malonic esters to
make barbituric acids.

Π$  

Urea can be irritating to skin, eyes, and the respiratory tract. Repeated or prolonged
contact with urea in fertilizer form on the skin may cause dermatitis.

High concentrations in the blood can be damaging. Ingestion of low concentrations


of urea, such as are found in typical human urine, are not dangerous with
additional water ingestion within a reasonable time-frame. Many animals (e.g.,
dogs) have a much more concentrated urine and it contains a higher urea amount
than normal human urine; this can prove dangerous as a source of liquids for
consumption in a life-threatening situation (such as in a desert).

Urea can cause algal blooms to produce toxins, and its presence in the runoff from
fertilized land may play a role in the increase of toxic blooms.[15]

The substance decomposes on heating above melting point, producing toxic gases,
and reacts violently with strong oxidants, nitrites, inorganic chlorides, chlorites and
perchlorates, causing fire and explosion.[O  ]

"  

c 
1,Ô,5-Trimethylbenzene


   
Mesitylene
a-Trimethylbenzene

   

CAS number 108-67-8

PubChem 7947

ChemSpider 7659

EC number 20Ô-604-4
AEGG C14508

Jmol-ÔD images Image 1

SMILES
Cc1cc(cc(c1)C)C
InChI
InChI=1S/C9H12/c1-7-4-8(2)6-9(Ô)5-7/h4-
6H,1-ÔHÔ
Aey: AUHZEENZYGFFBQ-
UHFFFAOYSA-N

InChI=1/C9H12/c1-7-4-8(2)6-9(Ô)5-7/h4-
6H,1-ÔHÔ
Aey: AUHZEENZYGFFBQ-
UHFFFAOYAA
   
Molecular formula C9H12
Molar mass 120.19 g/mol
Density 0.86Ô7 g/cm³ at 20 °C
Melting point -44.8 °C
Boiling point 164.7 °C
%  
Safety data from Oxford
MSDS
University
(what is this?) (verify)
Except where noted otherwise, data are given
for materials in their standard state (at 25 °C,
100 kPa)
Infobox references
"   or &'(')   % (C9H12) is an aromatic hydrocarbon with
three methyl substituents attached to the benzene ring. It is prepared by distillation
of acetone with sulfuric acid or by trimerization of propyne in sulfuric acid, which,
in both cases, acts as a catalyst and dehydrating agent. It is commonly used as a
solvent in research and industry. It is flammable and an irritant; it is a low-freezing
liquid.

In the electronics industry, mesitylene has also been used as a developer for
photopatternable silicones due to its solvent properties.

1,Ô,5-Trimethylbenzene is also a major urban volatile organic compound (üOC)


which results from combustion. It plays a significant role in aerosol and
tropospheric ozone formation as well as other reactions in atmospheric chemistry.

The    group (Mes) is a functional group found as an attachment in many


organic compounds.

The three aromatic hydrogen atoms of mesitylene are in identical chemical shift
environments. Therefore, they only give a single peak near 6.8 ppm in the 1H
NMR spectrum. For this reason, mesitylene is sometimes used as an internal
standard in NMR samples that contain aromatic protons.[1]

 {   

 {    is an inorganic compound with the formula NH4SCN. It


is the salt of the ammonium cation and the thiocyanate anion.

Πc  
Ammonium thiocyanate is used in the manufacture of herbicides, thiourea, and
transparent artificial resins; in matches; as a stabilizing agent in photography; in
various rustproofing compositions; as an adjuvant in textile dyeing and printing; as
a tracer in oil fields; in the separation of hafnium from zirconium, and in titrimetric
analyses.

Π   

Ammonium thiocyanate is made in the United States by the reaction of carbon


disulfide with aqueous ammonia. Ammonium dithiocarbamate is formed as an
intermediate in this reaction, which upon heating, decomposes to ammonium
thiocyanate and hydrogen sulfide:

CS2 + 2 NHÔ(aq) ĺ NH2C(=S)SNH4 ĺ NH4SCN + H2S

Π#  

Ammonium thiocyanate is stable in air; however, upon heating it isomerizes to


thiourea:

The equilibrium mixtures at 150 °C and 180 °C contain Ô0.Ô% and 25.Ô% (by
weight) thiourea, respectively. When heated at 200 °C, the dry powder decomposes
to ammonia, hydrogen sulfide, and carbon disulfide, leaving a residue of guanidine
thiocyanate.

NH4SCN is weakly acidic; reacts with caustic soda or caustic potash to form
sodium thiocyanate or potassium thiocyanate. It reacts with ferric salts to form a
deep-red ferric thiocyanate complex:

6 SCN- + FeÔ+ ĺ [Fe(SCN)6]Ô-

Ammonium thiocyanate reacts with several metal ions including copper, silver,
zinc, lead, and mercury, forming their thiocyanate precipitates, which may be
extracted into organic solvents.

c  {      


M  
   


 

 
 

M  
      


The c  {       is a systematic method of


naming inorganic chemical compounds, as recommended by the International
Union of Pure and Applied Chemistry (IUPAC). The rules are commonly known
as - -[1] Ideally, every inorganic compound should have a name from
which an unambiguous formula can be determined. There is also an IUPAC
nomenclature of organic chemistry.

Π$  

The names "caffeine" and "Ô,7-dihydro-1,Ô,7-trimethyl-1H-purine-2,6-dione" both


signify the same chemical. The systematic name encodes the structure and
composition of the caffeine molecule in some detail, and provides an unambiguous
reference to this compound, whereas the name "caffeine" just names it. These
advantages make the systematic name far superior to the common name when
absolute clarity and precision are required. However, for the sake of brevity, even
professional chemists will use the non-systematic name almost all of the time,
because caffeine is a well-known common chemical with a unique structure.
Similarly, H2O is most often simply called water in English, though other chemical
names do exist.

§ u



  
a
 
 
   
 V  
 
  
  
 
 

 

    !

 M 
 "
V
 
 
 V
M
 
 
# V

! 
 
 


!  
$


  M 
 V% & V%
 && '    
 
 
 a  a    
 (

 
 
   
 
 )   

  V% 
 V%  M 


 
   
* 





  a a
  
   
 +
   M 
 "   
 "#
 & 

!    a
 
a
     V   
V* 
, J  a
  
 
 
  

  
- !
!
-"
V# J     


   
 J  "
V# !  
   
!
 

Positively charged ions are called cations and negatively charged ions are called
anions. The cation is * named first. Ions can be metals or polyatomic ions.
Therefore the name of the metal or positive polyatomic ion is followed by the
name of the non-metal or negative polyatomic ion. The positive ion retains its
element name whereas for a single non-metal anion the ending is changed to -ide.

Example: sodium chloride, potassium oxide, or calcium carbonate.

When the metal has more than one possible ionic charge or oxidation number the
name becomes ambiguous. In these cases the oxidation number (the same as the
charge) of the metal ion is represented by a Roman numeral in parentheses
immediately following the metal ion name. For example in uranium(üI) fluoride
the oxidation number of uranium is 6. Another example is the iron oxides. FeO is
iron(II) oxide and Fe2OÔ is iron(III) oxide.

An older system used prefixes and suffixes to indicate the oxidation number,
according to the following scheme:

›  
  

(  .. .. 

 . . 

 . .


  ..  ..

    ..  ..


Thus the four oxyacids of chlorine are called hypochlorous acid (HOCl), chlorous
acid (HOClO), chloric acid (HOClO2) and perchloric acid (HOClOÔ), and their
respective conjugate bases are the hypochlorite, chlorite, chlorate and perchlorate
ions. This system has partially fallen out of use, but survives in the common names
of many chemical compounds: the modern literature contains few references to
"ferric chloride" (instead calling it "iron(III) chloride"), but names like "potassium
permanganate" (instead of "potassium manganate(üII)") and "sulfuric acid"
abound.

Π+   

Π,    { 

An ionic compound is named by its cation followed by its anion. See polyatomic
ions for a list of possible ions.

For cations that take on multiple charges, the charge is written using Roman
numerals in parentheses immediately following the element name) For example,
Cu(NOÔ)2 is O À
  , because the charge of two nitrate ions (NOÔ-1) is 2 ×
í1 = í2, and since the net charge of the ionic compound must be zero, the Cu ion
has a 2+ charge. This compound is therefore copper(II) nitrate. In the case of
cations with a 4+ oxidation state, the acceptable format for the Roman numeral 4 is
Iü and not IIII.

The Roman numerals in fact show the oxidation number, but in simple ionic
compounds (i.e., not metal complexes) this will always equal the ionic charge on
the metal. For a simple overview see [1], for more details see selected pages from
IUPAC rules for naming inorganic compounds.

Π 
    

Monatomic anions:

V   

u 

/#   


Polyatomic ions:

"*%


#%  

"#


"  

V    

V   

V# 


V*  


u# 

u*


u#    ! 

V#  
!
 !
!


V#
!


/*# 


/*   




/*    




V * 


V   


0##!


'*#



V*


V"
 

uV" 



1* 




Π,     

Hydrates are ionic compounds that have absorbed water. They are named as the
ionic compound followed by a numerical prefix and a  . The numerical
prefixes used are listed below (see IUPAC numerical multiplier):

§ .
 .
# .
* 
.
, 
.
2 
.
 
.
3 
.
4 
.
§ 
.

For example, CuSO4 · 5H2O is "copper(II) sulfate pentahydrate".

Π,  {  { 

Inorganic molecular compounds are named with a prefix (see list above) before
each element. The more electronegative element is written last and with an a 
suffix. For example, H2O (water) can be called    . Organic
molecules do not follow this rule. In addition, the prefix  is not used with the
first element; for example, SO2 is a    , not "monosulfur dioxide".
Sometimes prefixes are shortened when the ending vowel of the prefix "conflicts"
with a starting vowel in the compound. This makes the name easier to pronounce;
for example, CO is "carbon monoxide" (as opposed to "monooxide").

Π,  

Acids are named by the anion they form when dissolved in water. If an acid forms
an anion ending in , then its name is formed by adding the prefix   to the
anion's name and replacing the  with O. Finally the word O  is appended. For
example,  chlor OO  forms a chlor  anion. With sulfur, however, the
whole word is kept instead of the root: i.e.:  sulfur OO . Secondly, anions
with an a suffix are formed when acids with an a O suffix are dissolved, e.g.
chlor O acid (HClOÔ) dissociates into chlor anions to form salts such as sodium
chlor (NaClOÔ); anions with an a  suffix are formed when acids with an aa
suffix are dissolved in water, e.g. chloraO  (HClO2) disassociates into chlor 
anions to form salts such as sodium chlor  (NaClO2).

# -   .

From Wikipedia, the free encyclopedia


Jump to: navigation, search
M   aaaaOÀ a  


Two of the contributing structures of nitrogen dioxide

In chemistry,   or     [1] is a way of describing delocalized


electrons within certain molecules or polyatomic ions where the bonding cannot be
expressed by one single Lewis formula. A molecule or ion with such delocalized
electrons is represented by several  {  { {  [2] (also called
  { {  or   ).

Each contributing structure can be represented by a Lewis structure, with only an


integer number of covalent bonds between each pair of atoms within the
structure.[Ô] Several Lewis structures are used collectively to describe the actual
molecular structure. However these individual contributors cannot be observed in
the actual resonance-stabilized molecule; the molecule does not oscillate back and
forth between the contributing structures, as might be assumed from the word
" aO". The actual structure is an approximate intermediate between the
canonical forms, but its overall energy is lower than each of the contributors. This
intermediate form between different contributing structures is called a  
 .[4] Contributing structures differ only in the position of electrons, not in the
position of nuclei.
Resonance is a key component of valence bond theory.

Electron delocalization lowers the potential energy of the substance and thus
makes it more stable than any of the contributing structures. The difference
between the potential energy of the actual structure and that of the contributing
structure with the lowest potential energy is called the   [5] or
delocalization energy.

Resonance is distinguished from tautomerism and conformational isomerism,


which involve the formation of isomers, thus the rearrangement of the nuclear
positions.

Π/          

Molecules and ions with resonance (also called mesomerism) have the following
basic characteristics:

Contributing structures of the carbonate ion

` They can be represented by several correct Lewis formulas, called


"contributing structures", "resonance structures" or "canonical forms".
However, the real structure is not a rapid interconversion of contributing
structures. Several Lewis structures are used together, because none of them
exactly represents the actual structure. To represent the intermediate, a
resonance hybrid is used instead.
` The contributing structures are not isomers. They differ only in the position
of electrons, not in the position of nuclei.
` Each Lewis formula must have the same number of valence electrons (and
thus the same total charge), and the same number of unpaired electrons, if
any.[6]
` Bonds that have different bond orders in different contributing structures do
not have typical bond lengths. Measurements reveal intermediate bond
lengths.
` The real structure has a lower total potential energy than each of the
contributing structures would have. This means that it is more stable than
each separate contributing structure would be.
Πc   {  { {  

In Lewis formulas, covalent bonds are represented in accordance with the valence
bond theory. Each single bond is made by two valence electrons, localized between
the two bonded atoms. Each double bond has two additional localized ʌ electrons,
while each triple bond has four additional ʌ electrons (two pairs) between the
bonded atoms.

In molecules or ions that have a combination of one or more single and multiple
bonds, often the O position of the respective bonds in the Lewis formula cannot
be indicated. The ʌ electrons appear to be delocalized and the multiple bonds could
be in different positions. In those cases the molecule cannot be represented by one
single Lewis formula. To solve this problem, in valence bond theory the concept of
resonance is used, and the molecule is represented by several contributing
structures, each showing a possible distribution of single and multiple bonds. The
molecular orbital theory already includes the concept of delocalized electrons and
therefore has no need of the concept of resonance.

None of the contributing structures is considered to represent the actual structure,


since bonds that have a different bond order in different contributing structures do
not have, if measured, a bond length that is typical for a normal single or multiple
bond. Moreover, the overall energy of the actual structure is lowered with the
resonance energy.

Π#    

The actual structure of a molecule in the   a has the lowest


possible value of total energy. This structure is called the "resonance hybrid" of
that molecule. The resonance hybrid is the approximate intermediate of the
contributing structures, but the overall energy is lower than each of the
contributors, due to the resonance energy. [4]

Π"0    {  

One contributing structure may resemble the actual molecule more than another (in
the sense of energy and stability). Structures with a low value of potential energy
are more stable than those with high values and resemble the actual structure more.
The most stable contributing structures are called  O  a. Energetically
unfavourable and therefore less probable structures are   O  a.
Major contributors are generally structures
` that obey as much as possible the octet rule (8 valence electrons around each
atom rather than having deficiencies or surplus)
` that have a maximum number of covalent bonds
` that carry a minimum of charged atoms
` with negative charge, if any, on the most electronegative atoms and positive
charge, if any, on the most electropositive.

The greater the number of contributing structures, the more stable the molecule.
This is because the more states at lower energy are available to the electrons in a
particular molecule, the more stable the electrons are. Also the more volume
electrons can occupy at lower energy the more stable the molecule is.

Equivalent contributors contribute equally to the actual structure; those with low
potential energy (the major contributors) contribute more to the resonance hybrid
than the less stable minor contributors. Especially when there is more than one
major contributor, the resonance stabilization is high. High values of resonance
energy are found in aromatic molecules.

Π {  { {    

Contributing structures of the thiocyanate ion, enclosed in square brackets.

Hybrid of the Hybrid of


nitrate ion benzene.

In diagrams, contributing structures are typically separated by double-headed


arrows ( ). The arrow should not be confused with the right and left pointing
  ’ ( ). All structures together may be enclosed in large square
brackets, to indicate they picture one single molecule or ion, not different species
in a chemical equilibrium.

Alternatively to the use of resonance structures in diagrams, a hybrid diagram can


be used. In a hybrid diagram, pi bonds that are involved in resonance are usually
pictured as curves [7] or dashed lines, indicating that these are partial rather than
normal complete pi bonds. In benzene and other aromatic rings, the delocalized pi-
electrons are sometimes pictured as a solid circle.[8]

Π1 

Resonance structures of benzene

Comparing the two contributing structures of benzene, all single and double bonds
are interchanged. Bond lengths can be measured, for example using X-ray
diffraction. The average length of a C-C single bond is 154 pm; that of a C=C
double bond is 1ÔÔ pm. In localized cyclohexatriene, the carbon-carbon bonds
should be alternating 154 and 1ÔÔ pm. Instead, all carbon-carbon bonds in benzene
are found to be about 1Ô9 pm, a bond length intermediate between single and
double bond. This mixed single and double bond (or triple bond) character is
typical for all molecules in which bonds have a different bond order in different
contributing structures.

Π#  

Every structure is associated with a certain quantity of energy, which determines


the stability of the molecule or ion (the lower energy, the greater stability). A
resonance hybrid has a a O  that is intermediate between the contributing
structures; the total quantity of    , however, is lower than the
intermediate. Hybrids are therefore always more stable than any of the contributing
structures would be.[9] The molecule is sometimes said to be "stabilized by
resonance" or "resonance-stabilized," but the stabilization derives from electron
delocalization, of which "resonance" is only a description. Delocalization of the ʌ-
electrons lowers the orbital energies, imparting this stability. The difference
between the potential energy of the actual structure (the resonance hybrid) and that
of the contributing structure with the lowest potential energy is called the
"resonance energy".[5]

Π#   %

Resonance (or delocalization) energy is the amount of energy needed to convert


the true delocalized structure into that of the most stable contributing structure. The
 O  aO  can be estimated by comparing the heat of
hydrogenation of the real substance with that estimated for the contributing
structure.

The complete hydrogenation of benzene to cyclohexane via 1,Ô-cyclohexadiene


and cyclohexene is exothermic; 1 mole benzene delivers 208.4 kJ (49.8 kcal).

Hydrogenation of one double bond delivers 119.7 kJ (28.6 kcal), as can be deduced
from the last step, the hydrogenation of cyclohexene. In benzene, however, 2Ô.4 kJ
(5.6 kcal) are needed to hydrogenate one double bond. The difference, being 14Ô.1
kJ (Ô4.2 kcal), is the empirical resonance energy of benzene. Because 1,Ô-
cyclohexadiene also has a small delocalization energy (7.6 kJ or 1.8 kcal/mol) the
net resonance energy, relative to the localized cyclohexatriene, is a bit higher: 151
kJ or Ô6 kcal/mol. [10]

This measured resonance energy is also the difference between the hydrogenation
energy of three 'non-resonance' double bonds and the measured hydrogenation
energy: (Ô×119.7)í208.4=151 kJ (Ô6 kcal). [11]

, 2 The values used here are from the article of Wiberg, Nakaji, Morgan (199Ô).
üalues from other sources may differ.

Π# 3{ {    

Resonance has a deeper significance in the mathematical formalism of valence


bond theory (üB). When a molecule cannot be represented by the standard tools of
valence bond theory (promotion, hybridisation, orbital overlap, sigma and ʌ bond
formation) because no single structure predicted by üB can account for all the
properties of the molecule, one invokes the concept of resonance.

üalence bond theory gives us a model for benzene where each carbon atom makes
two sigma bonds with its neighbouring carbon atoms and one with a hydrogen
atom. But since carbon is tetravalent, it has the ability to form one more bond. In
üB it can form this extra bond with either of the neighbouring carbon atoms,
giving rise to the familiar Aekulé ring structure. But this cannot account for all
carbon-carbon bond lengths being equal in benzene. A solution is to write the
actual wavefunction of the molecule as a linear superposition of the two possible
Aekulé structures (or rather the wavefunctions representing these structures),
creating a wavefunction that is neither of its components but rather a superposition
of them.

In benzene both Aekulé structures have equal energy and are equal contributors to
the overall structure²the superposition is an equally-weighted average, or a 1:1
linear combination of the two²but this need not be the case. In general, the
superposition is written with undetermined coefficients, which are then
variationally optimized to find the lowest possible energy for the given set of basis
wavefunctions. This is taken to be the best approximation that can be made to the
real structure, though a better one may be made with addition of more structures.

Π"{   -"


. { -41.  

In molecular orbital theory, the main alternative to valence bond theory, resonance
often (but not always) translates to a delocalization of electrons in ʌ orbitals (which
are a separate concept from ʌ bonds in üB). For example, in benzene, the MO
model gives us 6 ʌ electrons completely delocalized over all 6 carbon atoms, thus
contributing something like half-bonds. This MO interpretation has inspired the
picture of the benzene ring as a hexagon with a circle inside. When describing
benzene, the üB concept of localized sigma 'bonds' and the MO concept of
'delocalized' ʌ electrons are frequently combined.

Π  

Weighting of the of resonance structures in terms of their contribution to the


overall structure can be calculated in multiple ways, using -    - methods
derived from üalence Bond theory, or else from the Natural Bond Orbitals (NBO)
approaches of Weinhold NBO5, or finally from empirical calculations based on the
Hückel method. A Hückel method-based software for teaching resonance is
available on the HuLiS Web site.
Π  

The concept of resonance was introduced into quantum mechanics by Werner


Heisenberg in 1926 in a discussion of the quantum states of the helium atom. He
compared the structure of the helium atom with the classical system of resonating
coupled harmonic oscillators. [4][12] Linus Pauling used this analogy to introduce his
resonance theory in 1928. [1Ô] In the classical system, the coupling produces two
modes, one of which is lower in frequency than either of the uncoupled vibrations;
quantum-mechanically, this lower frequency is interpreted as a lower energy. The
alternative term     popular in German and French publications with the
same meaning was introduced by Christopher Ingold in 19Ô8, but did not catch on
in the English literature. The current concept of mesomeric effect has taken on a
related but different meaning. The double headed arrow was introduced by the
German chemist Fritz Arndt who preferred the German phrase 6’ aO a or
  a.

In the Soviet Union, resonance theory ² especially as developed by Linus Pauling


² was attacked in the early 1950s as being contrary to the Marxist principles of
dialectical materialism, and in June 1951 the Soviet Academy of Sciences under
the leadership of Alexander Nesmeyanov convened a conference on the chemical
structure of organic compounds, attended by 400 physicists, chemists, and
philosophers, where "the pseudo-scientific essence of the theory of resonance was
exposed and unmasked".[14]

Due to confusion with the physical meaning of the word resonance, as no elements
actually appear to be resonating, it has been suggested that the term resonance be
abandoned in favor of % .[15] Resonance energy would become
%   and a resonance structure becomes a  { 
{ { . The double headed arrows would be replaced by commas.

Π#     

The ozone molecule is represented by two resonance structures. In reality the two
terminal oxygen atoms are equivalent and the hybrid structure is drawn on the right
with a charge of -1/2 on both oxygen atoms and partial double bonds with a full
and dashed line and bond order 1.5.[16][17]
In benzene the two cyclohexatriene A  structures first proposed by Aekulé are
taken together as contributing structures to represent the total structure. In the
hybrid structure on the right the dashed hexagon replaces three double bonds, and
represents six electrons in a set of three molecular orbitals of ʌ symmetry, with a
nodal plane in the plane of the molecule.

The allyl cation has two contributing structures with a positive charge on the
terminal carbon atoms. In the hybrid structure their charge is +1/2. The full
positive charge can also be depicted as delocalized among three carbon atoms.

In furan a lone pair of the oxygen atom interacts with the ʌ orbitals of the carbon
atoms. The curved arrows depicture the move of delocalized ʌ electrons, which
results in different contributors.

Π#     

   O O    

Often, reactive intermediates such as carbocations and free radicals have more
delocalized structure than their parent reactants, giving rise to unexpected
products. The classical example is allylic rearrangement. When 1 mole of HCl
adds to 1 mole of 1,Ô-butadiene, in addition to the ordinarily expected product Ô-
chloro-1-butene, we also find 1-chloro-2-butene. Isotope labelling experiments
have shown that what happens here is that the additional double bond shifts from
1,2 position to 2,Ô position in some of the product. This and other evidence (such
as NMR in superacid solutions) shows that the intermediate carbocation must have
a highly delocalized structure, different from its mostly classical (delocalization
exists but is small) parent molecule. This cation (an allylic cation) can be
represented using resonance, as shown above.

This observation of greater delocalization in less stable molecules is quite general.


The excited states of conjugated dienes are stabilised more by conjugation than
their ground states, causing them to become organic dyes.

A well-studied example of delocalization that does not involve ʌ electrons


(hyperconjugation) can be observed in the non-classical ion norbornyl cation.
Other examples are diborane and methanium (CH5+). These can be viewed as
containing Ô-center-2-electron bonds and are represented either by contributing
structures involving rearrangement of sigma electrons or by a special notation, a Y
that has the three nuclei at its three points.

You might also like