Alkane: Difference between revisions

Source: Wikipedia, the free encyclopedia.
Content deleted Content added
JKlug81 (talk | contribs)
→‎Further reading: Adding a new external link that talks about alkanes as an interesting chemical
Tag: Reverted
Line 586: Line 586:
* [https://www2.chemistry.msu.edu/faculty/reusch/VirtTxtJml/intro1.htm Virtual Textbook of Organic Chemistry]
* [https://www2.chemistry.msu.edu/faculty/reusch/VirtTxtJml/intro1.htm Virtual Textbook of Organic Chemistry]
* [https://log-web.de/chemie/Start.htm?name=ethaneCryst&lang=en A visualization of the crystal structures of alkanes up to nonan]
* [https://log-web.de/chemie/Start.htm?name=ethaneCryst&lang=en A visualization of the crystal structures of alkanes up to nonan]
* [https://www.wolframscience.com/nks/notes-12-11--interesting-chemicals/ Interesting Chemicals - Alkanes (paraffins)]


{{Hydrocarbons}}
{{Hydrocarbons}}

Revision as of 20:25, 9 March 2021

Chemical structure of methane, the simplest alkane

In

pentacontane (C50H102) or 6-ethyl-2-methyl-5-(1-methylethyl) octane, an isomer of tetradecane
(C14H30).

The International Union of Pure and Applied Chemistry (IUPAC) defines alkanes as "acyclic branched or unbranched hydrocarbons having the general formula Template:CarbonnTemplate:Hydrogen2n+2, and therefore consisting entirely of hydrogen atoms and saturated carbon atoms". However, some sources use the term to denote any saturated hydrocarbon, including those that are either monocyclic (i.e. the cycloalkanes) or polycyclic,[2] despite their having a distinct general formula (i.e. cycloalkanes are CnH2n).

In an alkane, each carbon atom is sp3-hybridized with 4 sigma bonds (either C–C or C–H), and each hydrogen atom is joined to one of the carbon atoms (in a C–H bond). The longest series of linked carbon atoms in a molecule is known as its carbon skeleton or carbon backbone. The number of carbon atoms may be considered as the size of the alkane.

One group of the

standard ambient temperature and pressure
(SATP), for which the number of carbon atoms in the carbon backbone is greater than about 17. With their repeated –CH2 units, the alkanes constitute a homologous series of organic compounds in which the members differ in molecular mass by multiples of 14.03 u (the total mass of each such methylene-bridge unit, which comprises a single carbon atom of mass 12.01 u and two hydrogen atoms of mass ~1.01 u each).

Methane is produced by methanogenic bacteria and some long-chain alkanes function as pheromones in certain animal species or as protective waxes in plants and fungi. Nevertheless, most alkanes do not have much biological activity. They can be viewed as molecular trees upon which can be hung the more active/reactive functional groups of biological molecules.

The alkanes have two main commercial sources: petroleum (crude oil)[3] and natural gas.

An

alkyl
group is an alkane-based molecular fragment that bears one open valence for bonding. They are generally abbreviated with the symbol for any organyl group, R, although Alk is sometimes used to specifically symbolize an alkyl group (as opposed to an alkenyl group or aryl group).

Structure and classification

Saturated hydrocarbons are hydrocarbons having only single covalent bonds between their carbons. They can be:

  • linear (general formula C
    n
    H
    2n+2
    ) wherein the carbon atoms are joined in a snake-like structure;
  • branched (general formula C
    n
    H
    2n+2
    , n > 3) wherein the carbon backbone splits off in one or more directions; the simplest branched alkane is 2-methyl-propane, or more commonly known as isobutane, which has three carbon atoms as backbone, and another carbon (a methyl group) linked to the second atom of the backbone, so as to form a "T" shaped molecule. The 'iso-' prefix comes from the fact that a hydrogen atom misses from the central carbon atom of the isopropyl radical (most known for its isopropyl alcohol compound, or rubbing alcohol);
  • cyclic (general formula C
    n
    H
    2n
    , n > 3) wherein the carbon backbone is linked so as to form a loop.

According to the definition by

acyclic
(loopless) ones, corresponding to k = 0.

Isomerism

n-butane and isobutane are the two C4H10 isomers; cyclobutane and methylcyclopropane are the two C4H8 alkane isomers.
Bicyclo[1.1.0]butane is the only C4H6 alkane and has no alkane isomer; tetrahedrane
(below) is the only C4H4 alkane and so has no alkane isomer.
Tetrahedrane

Alkanes with more than three carbon atoms can be arranged in various ways, forming structural isomers. The simplest isomer of an alkane is the one in which the carbon atoms are arranged in a single chain with no branches. This isomer is sometimes called the n-isomer (n for "normal", although it is not necessarily the most common). However the chain of carbon atoms may also be branched at one or more points. The number of possible isomers increases rapidly with the number of carbon atoms. For example, for acyclic alkanes:[4]

Branched alkanes can be

molecular formula. Cyclobutane and methylcyclopropane are isomers of each other, but are not isomers of butane.[citation needed
]

Nomenclature

The IUPAC nomenclature (systematic way of naming compounds) for alkanes is based on identifying hydrocarbon chains. Unbranched, saturated hydrocarbon chains are named systematically with a Greek numerical prefix denoting the number of carbons and the suffix "-ane".[5]

In 1866,

hydrocarbons CnH2n+2, CnH2n, CnH2n−2, CnH2n−4, CnH2n−6.[6] Now, the first three name hydrocarbons with single, double and triple bonds;[7] "-one" represents a ketone; "-ol" represents an alcohol or OH group; "-oxy-" means an ether and refers to oxygen between two carbons, so that methoxymethane is the IUPAC name for dimethyl ether.[8]

It is difficult or impossible to find compounds with more than one IUPAC name. This is because shorter chains attached to longer chains are prefixes and the convention includes brackets. Numbers in the name, referring to which carbon a group is attached to, should be as low as possible so that 1- is implied and usually omitted from names of organic compounds with only one side-group. Symmetric compounds will have two ways of arriving at the same name.[citation needed]

Linear alkanes

Straight-chain alkanes are sometimes indicated by the prefix "n-" or "n-"(for "normal") where a non-linear isomer exists. Although this is not strictly necessary, the usage is still common in cases where there is an important difference in properties between the straight-chain and branched-chain isomers, e.g., "n-butane" rather than simply "butane" to distinguish it from isobutane. Alternative names for this group are: linear paraffins or n-paraffins.[citation needed]

The first six members of the series (in terms of number of carbon atoms) are named as follows:

methane
CH4 – one carbon and 4 hydrogen
ethane
C2H6 – two carbon and 6 hydrogen
propane
C3H8 – three carbon and 8 hydrogen
butane
C4H10 – four carbon and 10 hydrogen
pentane
C5H12 – five carbon and 12 hydrogen
hexane
C6H14 – six carbon and 14 hydrogen

The first four names were

Latin prefix non-. For a more complete list, see list of straight-chain alkanes
.

Branched alkanes

Ball-and-stick model of isopentane (common name) or 2-methylbutane (IUPAC systematic name)

Simple branched alkanes often have a common name using a prefix to distinguish them from linear alkanes, for example n-pentane, isopentane, and neopentane.

IUPAC naming conventions can be used to produce a systematic name.

The key steps in the naming of more complicated branched alkanes are as follows:[10]

  • Identify the longest continuous chain of carbon atoms
  • Name this longest root chain using standard naming rules
  • Name each side chain by changing the suffix of the name of the alkane from "-ane" to "-yl"
  • Number the longest continuous chain in order to give the lowest possible numbers for the side-chains[11]
  • Number and name the side chains before the name of the root chain
  • If there are multiple side chains of the same type, use prefixes such as "di-" and "tri-" to indicate it as such, and number each one.
  • Add side chain names in alphabetical (disregarding "di-" etc. prefixes) order in front of the name of the root chain
Comparison of nomenclatures for three isomers of C5H12
Common name n-pentane isopentane neopentane
IUPAC name pentane 2-methylbutane 2,2-dimethylpropane
Structure

Saturated cyclic hydrocarbons

Though technically distinct from the alkanes, this class of hydrocarbons is referred to by some as the "cyclic alkanes." As their description implies, they contain one or more rings.[citation needed]

Simple cycloalkanes have a prefix "cyclo-" to distinguish them from alkanes. Cycloalkanes are named as per their acyclic counterparts with respect to the number of carbon atoms in their backbones, e.g., cyclopentane (C5H10) is a cycloalkane with 5 carbon atoms just like pentane (C5H12), but they are joined up in a five-membered ring. In a similar manner, propane and cyclopropane, butane and cyclobutane, etc.[citation needed]

Substituted cycloalkanes are named similarly to substituted alkanes – the cycloalkane ring is stated, and the substituents are according to their position on the ring, with the numbering decided by the Cahn–Ingold–Prelog priority rules.[9]

Trivial/common names

The trivial (non-

systematic) name for alkanes is 'paraffins'. Together, alkanes are known as the 'paraffin series'. Trivial names for compounds are usually historical artifacts. They were coined before the development of systematic names, and have been retained due to familiar usage in industry. Cycloalkanes are also called naphthenes.[citation needed
]

It is almost certain that the term 'paraffin' stems from the petrochemical industry. Branched-chain alkanes are called isoparaffins. The use of the term "paraffin" is a general term and often does not distinguish between pure compounds and mixtures of isomers, i.e., compounds of the same chemical formula, e.g., pentane and isopentane.[citation needed]

In IUPAC

The following trivial names are retained in the IUPAC system:

Non-IUPAC

Some non-IUPAC trivial names are occasionally used:

Physical properties

All alkanes are colorless.[13][14] Alkanes with the lowest molecular weights are gasses, those of intermediate molecular weight are liquids, and the heaviest are waxy solids.[citation needed]

Table of alkanes

Alkane Formula Boiling point [°C] Melting point [°C] Density [kg/m3] (at 20 °C) Isomers
Methane CH4 -162 −182 0.656 (gas) 1
Ethane C2H6 −89 −183 1.26 (gas) 1
Propane C3H8 −42 −188 2.01 (gas) 1
Butane C4H10 0 −138 2.48 (gas) 2
Pentane C5H12 36 −130 626 (liquid) 3
Hexane C6H14 69 −95 659 (liquid) 5
Heptane C7H16 98 −91 684 (liquid) 9
Octane C8H18 126 −57 703 (liquid) 18
Nonane C9H20 151 −54 718 (liquid) 35
Decane C10H22 174 −30 730 (liquid) 75
Undecane C11H24 196 −26 740 (liquid) 159
Dodecane C12H26 216 −10 749 (liquid) 355
Tridecane C13H28 235 -5.4 756 (liquid) 802
Tetradecane C14H30 253 5.9 763 (liquid) 1858
Pentadecane C15H32 270 10 769 (liquid)
Hexadecane C16H34 287 18 773 (liquid)
Heptadecane C17H36 303 22 777 (solid)
Octadecane C18H38 317 28 781 (solid)
Nonadecane C19H40 330 32 785 (solid)
Icosane C20H42 343 37 789 (solid)
Triacontane
C30H62 450 66 810 (solid)
Tetracontane
C40H82 525 82 817 (solid)
Pentacontane
C50H102 575 91 824 (solid)
Hexacontane
C60H122 625 100 829 (solid)
Heptacontane
C70H142 653 109 869 (solid)

Boiling point

Melting (blue) and boiling (orange) points of the first 16 n-alkanes in °C.

Alkanes experience intermolecular van der Waals forces. Stronger intermolecular van der Waals forces give rise to greater boiling points of alkanes.[15]

There are two determinants for the strength of the van der Waals forces:

  • the number of electrons surrounding the molecule, which increases with the alkane's molecular weight
  • the surface area of the molecule

Under

molecular weight) of the molecule. As a rule of thumb, the boiling point rises 20–30 °C for each carbon added to the chain; this rule applies to other homologous series.[15]

A straight-chain alkane will have a boiling point higher than a branched-chain alkane due to the greater surface area in contact, thus the greater van der Waals forces, between adjacent molecules. For example, compare

n-butane (butane), which boil at −12 and 0 °C, and 2,2-dimethylbutane and 2,3-dimethylbutane which boil at 50 and 58 °C, respectively.[15] For the latter case, two molecules 2,3-dimethylbutane can "lock" into each other better than the cross-shaped 2,2-dimethylbutane, hence the greater van der Waals forces.[citation needed
]

On the other hand, cycloalkanes tend to have higher boiling points than their linear counterparts due to the locked conformations of the molecules, which give a plane of intermolecular contact.[16]

Melting points

The

boiling points for the same reason as outlined above. That is, (all other things being equal) the larger the molecule the higher the melting point. There is one significant difference between boiling points and melting points. Solids have more rigid and fixed structure than liquids. This rigid structure requires energy to break down. Thus the better put together solid structures will require more energy to break apart. For alkanes, this can be seen from the graph above (i.e., the blue line). The odd-numbered alkanes have a lower trend in melting points than even numbered alkanes. This is because even numbered alkanes pack well in the solid phase, forming a well-organized structure, which requires more energy to break apart. The odd-numbered alkanes pack less well and so the "looser" organized solid packing structure requires less energy to break apart.[17] For a visualization of the crystal structures see.[18]

The melting points of branched-chain alkanes can be either higher or lower than those of the corresponding straight-chain alkanes, again depending on the ability of the alkane in question to pack well in the solid phase: This is particularly true for isoalkanes (2-methyl isomers), which often have melting points higher than those of the linear analogues.[citation needed]

Conductivity and solubility

Alkanes do not conduct electricity in any way, nor are they substantially

polarized by an electric field. For this reason, they do not form hydrogen bonds and are insoluble in polar solvents such as water. Since the hydrogen bonds between individual water molecules are aligned away from an alkane molecule, the coexistence of an alkane and water leads to an increase in molecular order (a reduction in entropy). As there is no significant bonding between water molecules and alkane molecules, the second law of thermodynamics suggests that this reduction in entropy should be minimized by minimizing the contact between alkane and water: Alkanes are said to be hydrophobic as they repel water.[citation needed
]

Their solubility in nonpolar solvents is relatively high, a property that is called lipophilicity. Alkanes are, for example, miscible in all proportions among themselves.[citation needed]

The density of the alkanes usually increases with the number of carbon atoms but remains less than that of water. Hence, alkanes form the upper layer in an alkane–water mixture.[citation needed]

Molecular geometry

sp3-hybridization in methane.

The molecular structure of the alkanes directly affects their physical and chemical characteristics. It is derived from the electron configuration of carbon, which has four valence electrons. The carbon atoms in alkanes are always sp3-hybridized, that is to say that the valence electrons are said to be in four equivalent orbitals derived from the combination of the 2s orbital and the three 2p orbitals. These orbitals, which have identical energies, are arranged spatially in the form of a tetrahedron, the angle of cos−1(−1/3) ≈ 109.47° between them.[citation needed]

Bond lengths and bond angles

An alkane has only C–H and C–C single bonds. The former result from the overlap of an sp3 orbital of carbon with the 1s orbital of a hydrogen; the latter by the overlap of two sp3 orbitals on adjacent carbon atoms. The bond lengths amount to 1.09 × 10−10 m for a C–H bond and 1.54 × 10−10 m for a C–C bond.[citation needed]

The tetrahedral structure of methane.

The spatial arrangement of the bonds is similar to that of the four sp3 orbitals—they are tetrahedrally arranged, with an angle of 109.47° between them. Structural formulae that represent the bonds as being at right angles to one another, while both common and useful, do not correspond with the reality.[citation needed]

Conformation

The structural formula and the

torsion angle between the atoms or groups bound to the atoms at each end of the bond. The spatial arrangement described by the torsion angles of the molecule is known as its conformation
.

Newman projections of the two conformations of ethane: eclipsed on the left, staggered on the right.
Ball-and-stick models of the two rotamers of ethane

eclipsed conformation and staggered conformation
.

The two conformations differ in energy: the staggered conformation is 12.6 kJ/mol lower in energy (more stable) than the eclipsed conformation (the least stable).

This difference in energy between the two conformations, known as the

torsion energy
, is low compared to the thermal energy of an ethane molecule at ambient temperature. There is constant rotation about the C–C bond. The time taken for an ethane molecule to pass from one staggered conformation to the next, equivalent to the rotation of one CH3 group by 120° relative to the other, is of the order of 10−11 seconds.

The case of

higher alkanes
is more complex but based on similar principles, with the antiperiplanar conformation always being the most favored around each carbon–carbon bond. For this reason, alkanes are usually shown in a zigzag arrangement in diagrams or in models. The actual structure will always differ somewhat from these idealized forms, as the differences in energy between the conformations are small compared to the thermal energy of the molecules: Alkane molecules have no fixed structural form, whatever the models may suggest.

Spectroscopic properties

Virtually all organic compounds contain carbon–carbon, and carbon–hydrogen bonds, and so show some of the features of alkanes in their spectra. Alkanes are notable for having no other groups, and therefore for the absence of other characteristic spectroscopic features of a functional group like –OH, –CHO, –COOH etc.

Infrared spectroscopy

The carbon–hydrogen stretching mode gives a strong absorption between 2850 and 2960 cm−1, while the carbon–carbon stretching mode absorbs between 800 and 1300 cm−1. The carbon–hydrogen bending modes depend on the nature of the group: methyl groups show bands at 1450 cm−1 and 1375 cm−1, while methylene groups show bands at 1465 cm−1 and 1450 cm−1. Carbon chains with more than four carbon atoms show a weak absorption at around 725 cm−1.

NMR spectroscopy

The proton resonances of alkanes are usually found at

relaxation time
, and can be missed in weak samples, or samples that have not been run for a sufficiently long time.

Mass spectrometry

Alkanes have a high

free radicals
. The fragment resulting from the loss of a single methyl group (M − 15) is often absent, and other fragments are often spaced by intervals of fourteen mass units, corresponding to sequential loss of CH2 groups.

Chemical properties

Alkanes are only weakly reactive with most chemical compounds. The

crude oil the alkane molecules have remained chemically unchanged for millions of years.[citation needed
]

]

In highly branched alkanes, the bond angle may differ significantly from the optimal value (109.5°) to accommodate bulky groups. Such distortions introduce a tension in the molecule, known as

steric hindrance
or strain. Strain substantially increases reactivity.

However, in general and perhaps surprisingly, when branching is not extensive enough to make highly disfavorable 1,2- and 1,3-alkyl–alkyl steric interactions (worth ~3.1 kcal/mol and ~3.7 kcal/mol in the case of the eclipsing conformations of butane and pentane, respectively) unavoidable, the branched alkanes are actually more thermodynamically stable than their linear (or less branched) isomers. For example, the highly branched 2,2,3,3-tetramethylbutane is about 1.9 kcal/mol more stable than its linear isomer, n-octane.[19] Due to the subtlety of this effect, the exact reasons for this rule have been vigorously debated in the chemical literature and is yet unsettled. Several explanations, including stabilization of branched alkanes by electron correlation,[20] destabilization of linear alkanes by steric repulsion,[21] stabilization by neutral hyperconjugation,[22][19] and/or electrostatic effects[23] have been advanced as possibilities. The controversy is related to the question of whether the traditional explanation of hyperconjugation is the primary factor governing the stability of alkyl radicals.[24][21]

Reactions with oxygen (combustion reaction)

All alkanes react with oxygen in a combustion reaction, although they become increasingly difficult to ignite as the number of carbon atoms increases. The general equation for complete combustion is:

CnH2n+2 + (3/2n + 1/2) O2 → (n + 1) H2O + n CO2
or CnH2n+2 + (3n + 1/2) O2 → (n + 1) H2O + n CO2

In the absence of sufficient oxygen, carbon monoxide or even soot can be formed, as shown below:

CnH2n+2 + (n + 1/2O2 → (n + 1) H2O + n CO
CnH2n+2 + (1/2n + 1/2O2 → (n + 1) H2O + n C

For example, methane:

2 CH4 + 3 O2 → 2 CO + 4 H2O
CH4 + 3/2 O2 → CO + 2 H2O

See the

alkane heat of formation table
for detailed data. The
standard enthalpy change of combustion
, ΔcH, for alkanes increases by about 650 kJ/mol per CH2 group. Branched-chain alkanes have lower values of ΔcH than straight-chain alkanes of the same number of carbon atoms, and so can be seen to be somewhat more stable.

Reactions with halogens

Alkanes react with

Free radicals are the reactive species that participate in the reaction, which usually leads to a mixture of products. The reaction is highly exothermic, and can lead to an explosion.[citation needed
]

These reactions are an important industrial route to halogenated hydrocarbons. There are three steps:

  • Initiation the halogen radicals form by homolysis. Usually, energy in the form of heat or light is required.
  • Chain reaction or Propagation then takes place—the halogen radical abstracts a hydrogen from the alkane to give an alkyl radical. This reacts further.
  • Chain termination where the radicals recombine.

Experiments have shown that all halogenation produces a mixture of all possible isomers, indicating that all hydrogen atoms are susceptible to reaction. The mixture produced, however, is not a statistical mixture: Secondary and tertiary hydrogen atoms are preferentially replaced due to the greater stability of secondary and tertiary free-radicals. An example can be seen in the monobromination of propane:[15]

Monobromination of propane
Monobromination of propane

Cracking

Cracking breaks larger molecules into smaller ones. This can be done with a thermal or catalytic method. The thermal cracking process follows a

intermolecular hydrogen transfer or hydride transfer. In both types of processes, the corresponding reactive intermediates (radicals, ions) are permanently regenerated, and thus they proceed by a self-propagating chain mechanism. The chain of reactions is eventually terminated by radical or ion recombination.[citation needed
]

Isomerization and reformation

Dragan and his colleague were the first to report about isomerization in alkanes.

octane number of the substance. Butane is the most common alkane that is put under the process of isomerization, as it makes many branched alkanes with high octane numbers.[25]

Other reactions

Alkanes will react with

fluorosulfonic acid (HSO3F), called magic acid, can protonate alkanes.[26]

Occurrence

Occurrence of alkanes in the Universe

Methane and ethane make up a tiny proportion of Jupiter's atmosphere
Extraction of oil, which contains many distinct hydrocarbons including alkanes

Alkanes form a small portion of the

ppm ethane), Saturn (0.2% methane, 5 ppm ethane), Uranus (1.99% methane, 2.5 ppm ethane) and Neptune (1.5% methane, 1.5 ppm ethane). Titan (1.6% methane), a satellite of Saturn, was examined by the Huygens probe, which indicated that Titan's atmosphere periodically rains liquid methane onto the moon's surface.[27] Also on Titan the Cassini mission has imaged seasonal methane/ethane lakes near the polar regions of Titan. Methane and ethane have also been detected in the tail of the comet Hyakutake. Chemical analysis showed that the abundances of ethane and methane were roughly equal, which is thought to imply that its ices formed in interstellar space, away from the Sun, which would have evaporated these volatile molecules.[28] Alkanes have also been detected in meteorites such as carbonaceous chondrites
.

Occurrence of alkanes on Earth

Traces of methane gas (about 0.0002% or 1745 ppb) occur in the Earth's atmosphere, produced primarily by methanogenic microorganisms, such as Archaea in the gut of ruminants.[29]

The most important commercial sources for alkanes are natural gas and

hydrocarbons. These hydrocarbons were formed when marine animals and plants (zooplankton and phytoplankton) died and sank to the bottom of ancient seas and were covered with sediments in an anoxic
environment and converted over many millions of years at high temperatures and high pressure to their current form. Natural gas resulted thereby for example from the following reaction:

C6H12O6 → 3 CH4 + 3 CO2

These hydrocarbon deposits, collected in porous rocks trapped beneath impermeable cap rocks, comprise commercial

oil fields. They have formed over millions of years and once exhausted cannot be readily replaced. The depletion of these hydrocarbons reserves is the basis for what is known as the energy crisis
.

Methane is also present in what is called

renewable energy source
.

Alkanes have a low solubility in water, so the content in the oceans is negligible; however, at high pressures and low temperatures (such as at the bottom of the oceans), methane can co-crystallize with water to form a solid methane clathrate (methane hydrate). Although this cannot be commercially exploited at the present time, the amount of combustible energy of the known methane clathrate fields exceeds the energy content of all the natural gas and oil deposits put together. Methane extracted from methane clathrate is, therefore, a candidate for future fuels.

Biological occurrence

Acyclic alkanes occur in nature in various ways.

Bacteria and archaea
Methanogenic archaea in the gut of this cow are responsible for some of the methane in Earth's atmosphere.

Certain types of bacteria can metabolize alkanes: they prefer even-numbered carbon chains as they are easier to degrade than odd-numbered chains.[30]

On the other hand, certain

oxidized organic compounds. The energy is released by the oxidation of hydrogen
:

CO2 + 4 H2 → CH4 + 2 H2O

Methanogens are also the producers of

intestines of humans. Methanogenic archaea are, hence, at the end of the carbon cycle, with carbon being released back into the atmosphere after having been fixed by photosynthesis. It is probable that our current deposits of natural gas were formed in a similar way.[33]

Fungi and plants

Alkanes also play a role, if a minor role, in the biology of the three eukaryotic groups of organisms: fungi, plants and animals. Some specialized yeasts, e.g., Candida tropicale, Pichia sp., Rhodotorula sp., can use alkanes as a source of carbon or energy. The fungus Amorphotheca resinae prefers the longer-chain alkanes in aviation fuel, and can cause serious problems for aircraft in tropical regions.[34]

In plants, the solid long-chain alkanes are found in the plant cuticle and epicuticular wax of many species, but are only rarely major constituents.[35] They protect the plant against water loss, prevent the leaching of important minerals by the rain, and protect against bacteria, fungi, and harmful insects. The carbon chains in plant alkanes are usually odd-numbered, between 27 and 33 carbon atoms in length[35] and are made by the plants by decarboxylation of even-numbered fatty acids. The exact composition of the layer of wax is not only species-dependent but changes also with the season and such environmental factors as lighting conditions, temperature or humidity.[35]

More volatile short-chain alkanes are also produced by and found in plant tissues. The

Jeffrey pine is noted for producing exceptionally high levels of n-heptane in its resin, for which reason its distillate was designated as the zero point for one octane rating. Floral scents have also long been known to contain volatile alkane components, and n-nonane is a significant component in the scent of some roses.[36] Emission of gaseous and volatile alkanes such as ethane, pentane, and hexane by plants has also been documented at low levels, though they are not generally considered to be a major component of biogenic air pollution.[37]

Edible vegetable oils also typically contain small fractions of biogenic alkanes with a wide spectrum of carbon numbers, mainly 8 to 35, usually peaking in the low to upper 20s, with concentrations up to dozens of milligrams per kilogram (parts per million by weight) and sometimes over a hundred for the total alkane fraction.[38]

Animals

Alkanes are found in animal products, although they are less important than unsaturated hydrocarbons. One example is the shark liver oil, which is approximately 14%

pentacosane (C25H52), 3-methylpentaicosane (C26H54) and 9-methylpentaicosane (C26H54) are transferred by body contact. With others like the tsetse fly Glossina morsitans morsitans, the pheromone contains the four alkanes 2-methylheptadecane (C18H38), 17,21-dimethylheptatriacontane (C39H80), 15,19-dimethylheptatriacontane (C39H80) and 15,19,23-trimethylheptatriacontane (C40H82), and acts by smell over longer distances. Waggle-dancing honey bees produce and release two alkanes, tricosane and pentacosane.[39]

Ecological relations

Early spider orchid (Ophrys sphegodes)

One example, in which both plant and animal alkanes play a role, is the ecological relationship between the

heptacosane
(C27H56) in the ratio 3:3:1, and males are attracted by specifically this odor. The orchid takes advantage of this mating arrangement to get the male bee to collect and disseminate its pollen; parts of its flower not only resemble the appearance of sand bees but also produce large quantities of the three alkanes in the same ratio as female sand bees. As a result, numerous males are lured to the blooms and attempt to copulate with their imaginary partner: although this endeavor is not crowned with success for the bee, it allows the orchid to transfer its pollen, which will be dispersed after the departure of the frustrated male to other blooms.

Production

Petroleum refining

An oil refinery at Martinez, California.

As stated earlier, the most important source of alkanes is natural gas and

crude oil.[15] Alkanes are separated in an oil refinery by fractional distillation
and processed into many products.

Fischer–Tropsch

The

petroleum distillates
.

Laboratory preparation

There is usually little need for alkanes to be synthesized in the laboratory, since they are usually commercially available. Also, alkanes are generally unreactive chemically or biologically, and do not undergo

functional group interconversions cleanly. When alkanes are produced in the laboratory, it is often a side-product of a reaction. For example, the use of n-butyllithium as a strong base
gives the conjugate acid, n-butane as a side-product:

C4H9Li + H2O → C4H10 + LiOH

However, at times it may be desirable to make a section of a molecule into an alkane-like functionality (

hydroxy group, it gives ethanol, which is not an alkane. To do so, the best-known methods are hydrogenation of alkenes
:

RCH=CH2 + H2 → RCH2CH3     (R =
alkyl
)

Alkanes or alkyl groups can also be prepared directly from

removes hydroxyl groups from alcohols e.g.

Barton–McCombie deoxygenation scheme

and the Clemmensen reduction[42][43][44][45] removes carbonyl groups from aldehydes and ketones to form alkanes or alkyl-substituted compounds e.g.:

Clemmensen Reduction

Preparation of alkanes from other organic compounds

Alkanes can be prepared from a variety of organic compounds.These include alkenes, alkynes, haloalkanes, alcohols, aldehydes and ketones and carboxylic acids.

From alkenes and alkynes

When alkenes and alkynes are subjected to hydrogenation reaction by treating them with hydrogen in the presence of palladium or platinum or nickel catalyst, they produce alkanes. In this reaction powdered catalyst is preferred to increase the surface area so that adsorption of hydrogen on the catalyst increases. In this reaction the hydrogen gets attached on the catalyst to form a hydrogen-catalyst bond which leads to weakening of H-H bond, thereby leading to the addition of hydrogen on alkenes and alkynes. The reaction is exothermic because the product alkane is stable as it has more sigma bonds than the reactant alkenes and alkynes due to conversion of pi bond to sigma bonds.[46]

From haloalkanes

Alkanes can be produced from haloalkanes using different methods.

Wurtz reaction

When haloalkane is treated with sodium in dry ether, alkane with double the number of carbon atoms is obtained. This reaction proceeds through free radical intermediate and has possibility of alkene formation in case of tertiary haloalkanes and vicinal dihalides.

Corey-House-Synthesis

When haloalkane is treated with dialkyl lithium cuprite, which is otherwise known as Gilman's reagent, any higher alkane is obtained.

Reaction with metal hydride

When haloalkanes are treated with metal hydride ,eg, sodium hydride and lithium aluminium hydride.

Frankland reaction

When haloalkane is treated with zinc in ester, alkane is obtained.

Fittig reaction

When aryl halide is treated with sodium in dry ether, it forms biphenyl.

Ullmann biaryl synthesis

When aryl halide is treated with copper, it forms biphenyl.

Wurtz-Fittig reaction

When aryl halide is treated with haloalkane, we get alkyl benzene.

Applications

The applications of alkanes depend on the number of carbon atoms. The first four alkanes are used mainly for heating and cooking purposes, and in some countries for electricity generation. Methane and ethane are the main components of natural gas; they are normally stored as gases under pressure. It is, however, easier to transport them as liquids: This requires both compression and cooling of the gas.

aerosol sprays
.

From pentane to octane the alkanes are highly volatile liquids. They are used as fuels in internal combustion engines, as they vaporize easily on entry into the combustion chamber without forming droplets, which would impair the uniformity of the combustion. Branched-chain alkanes are preferred as they are much less prone to premature ignition, which causes knocking, than their straight-chain homologues. This propensity to premature ignition is measured by the octane rating of the fuel, where 2,2,4-trimethylpentane (isooctane) has an arbitrary value of 100, and heptane has a value of zero. Apart from their use as fuels, the middle alkanes are also good solvents for nonpolar substances.

Alkanes from nonane to, for instance, hexadecane (an alkane with sixteen carbon atoms) are liquids of higher viscosity, less and less suitable for use in gasoline. They form instead the major part of diesel and aviation fuel. Diesel fuels are characterized by their cetane number, cetane being an old name for hexadecane. However, the higher melting points of these alkanes can cause problems at low temperatures and in polar regions, where the fuel becomes too thick to flow correctly.

Alkanes from hexadecane upwards form the most important components of

lubricating oil. In the latter function, they work at the same time as anti-corrosive agents, as their hydrophobic nature means that water cannot reach the metal surface. Many solid alkanes find use as paraffin wax, for example, in candles. This should not be confused however with true wax, which consists primarily of esters
.

Alkanes with a chain length of approximately 35 or more carbon atoms are found in bitumen, used, for example, in road surfacing. However, the higher alkanes have little value and are usually split into lower alkanes by cracking.

Some synthetic polymers such as polyethylene and polypropylene are alkanes with chains containing hundreds or thousands of carbon atoms. These materials are used in innumerable applications, and billions of kilograms of these materials are made and used each year.

Environmental transformations

Alkanes are chemically very inert apolar molecules which are not very reactive as organic compounds. This inertness yields serious ecological issues if they are released into the environment. Due to their lack of functional groups and low water solubility, alkanes show poor bioavailability for microorganisms.[48]

There are, however, some microorganisms possessing the metabolic capacity to utilize n-alkanes as both carbon and energy sources.[49] Some bacterial species are highly specialised in degrading alkanes; these are referred to as hydrocarbonoclastic bacteria.[50]

Hazards

Methane is flammable, explosive and dangerous to inhale; because it is a colorless, odorless gas, special caution must be taken around methane.[51] Ethane is also extremely flammable, explosive, and dangerous to inhale.[52] Both of them may cause suffocation.[51][52] Propane, too, is flammable and explosive,[53] and may cause drowsiness or unconsciousness if inhaled.[53] Butane presents the same hazards as propane.[54]

Alkanes also pose a threat to the environment. Branched alkanes have a lower biodegradability than unbranched alkanes.[55] Methane is considered to be the greenhouse gas that is most dangerous to the environment,[56] although the amount of methane in the atmosphere is relatively low.[56]

See also

References

  1. ^
  2. ^ "Alkanes". Chemistry LibreTexts. 28 November 2016.
  3. .
  4. ^ On-Line Encyclopedia of Integer Sequences (sequence A000602 in the OEIS)
  5. . Retrieved 12 February 2007.
  6. ^ Alkane Nomenclature Archived 2 February 2012 at the Wayback Machine
  7. ^ Thus, the ending "-diene" is applied in some cases where von Hofmann had "-ine"
  8. .
  9. ^ a b William Reusch. "Nomenclature – Alkanes". Virtual Textbook of Organic Chemistry. Archived from the original on 21 May 2016. Retrieved 5 April 2007.
  10. ^ William Reusch. "Examples of the IUPAC Rules in Practice". Virtual Textbook of Organic Chemistry. Archived from the original on 21 May 2016. Retrieved 5 April 2007.
  11. ^ "IUPAC Rules". www.chem.uiuc.edu. Retrieved 13 August 2018.
  12. , p. 206
  13. ^ "Archived copy" (PDF). Archived from the original (PDF) on 29 October 2013. Retrieved 17 February 2014.{{cite web}}: CS1 maint: archived copy as title (link)
  14. ^ "13. Hydrocarbons | Textbooks". textbook.s-anand.net. Archived from the original on 8 May 2011. Retrieved 3 October 2014.
  15. ^ .
  16. ^ "Physical Properties of Cycloalkanes". Chemistry LibreTexts. 29 November 2015. Archived from the original on 2 February 2018. Retrieved 2 February 2018.
  17. .
  18. ^ visualization of crystal structures.
  19. ^
    OCLC 957525299.{{cite book}}: CS1 maint: location missing publisher (link
    )
  20. .
  21. ^ .
  22. .
  23. .
  24. .
  25. ^ a b c Asinger, Friedrich (1967). Paraffins; Chemistry and Technology. Oxford: Pergamon Press.
  26. .
  27. ^ Emily Lakdawalla. "Titan: Arizona in an Icebox?". Archived from the original on 6 April 2008. Retrieved 21 January 2004.
  28. S2CID 27362518
    .
  29. .
  30. ^ "Metabolism of Alkanes and Fatty Acids – eQuilibrator 0.2 beta documentation". equilibrator.weizmann.ac.il. Retrieved 11 April 2018.
  31. ^ TodayIFoundOut.com, Matt Blitz -. "Do Cow Farts Actually Contribute to Global Warming?". Gizmodo. Retrieved 11 April 2018.
  32. PMID 28168033
    .
  33. ^ Society, National Geographic (24 July 2012). "natural gas". National Geographic Society. Retrieved 11 April 2018.
  34. .
  35. ^
  36. .
  37. S2CID 94021819. Archived from the original
    (PDF) on 13 March 2013.
  38. .
  39. PMID 17713987.{{cite journal}}: CS1 maint: unflagged free DOI (link
    )
  40. .
  41. .
  42. .
  43. ^ Buchanan, J. G. St. C.; Woodgate, P. D. Quart. Rev. 1969, 23, 522, (Review).
  44. .
  45. ^ Yamamura, S.; Nishiyama, S. Compr. Org. Synth. 1991, 8, 309–313, (Review).
  46. ^ Harikiran.B, K E Mannanam, Kottayam
  47. ^ "Using propane as a fuel" (PDF). Archived from the original (PDF) on 12 October 2013. Retrieved 27 November 2012.
  48. .
  49. .
  50. .
  51. ^ a b "CDC - METHANE - International Chemical Safety Cards - NIOSH". www.cdc.gov. Retrieved 19 September 2017.
  52. ^ a b "CDC - ETHANE - International Chemical Safety Cards - NIOSH". www.cdc.gov. Retrieved 19 September 2017.
  53. ^ a b "CDC - PROPANE - International Chemical Safety Cards - NIOSH". www.cdc.gov. Archived from the original on 23 November 2017. Retrieved 19 September 2017.
  54. ^ "CDC - BUTANE - International Chemical Safety Cards - NIOSH". www.cdc.gov. Retrieved 19 September 2017.
  55. .
  56. ^ a b "Pollutant Fact Sheet". apps.sepa.org.uk. Retrieved 19 September 2017.

Further reading