Sei sulla pagina 1di 73

Alkanes are organic compounds that consist entirely of single-bonded carbon and

hydrogen atoms and lack any other functional groups. Alkanes have the general
formula CnH2n+2CnH2n+2and can be subdivided into the following three groups:
the linear straight-chain alkanes, branched alkanes, and cycloalkanes. Alkanes
are also saturated hydrocarbons. Alkanes are the simplest and least reactive
hydrocarbon species containing only carbons and hydrogens. They are commercially
very important, being the principal constituent of gasoline and lubricating oils and are
extensively employed in organic chemistry; though the role of pure alkanes (such as
hexanes) is delegated mostly to solvents. The distinguishing feature of an alkane,
making it distinct from other compounds that also exclusively contain carbon and
hydrogen, is its lack of unsaturation. That is to say, it contains no double or triple
bonds, which are highly reactive in organic chemistry. Though not totally devoid of
reactivity, their lack of reactivity under most laboratory conditions makes them a
relatively uninteresting, though very important component of organic chemistry. As you
will learn about later, the energy confined within the carbon-carbon bond and the
carbon-hydrogen bond is quite high and their rapid oxidation produces a large amount
of heat, typically in the form of fire.

 Nomenclature of Alkanes

 Properties of Alkanes

 Synthesis of Alkanes
 Reactivity of Alkanes

Nomenclature of Alkanes
The names of all alkanes end with -ane. Whether or not the carbons are linked together end-to-
end in a ring (called cyclic alkanes or cycloalkanes) or whether they contain side chains and
branches, the name of every carbon-hydrogen chain that lacks any double bonds or functional
groups will end with the suffix -ane.
Alkanes with unbranched carbon chains are simply named by the number of carbons in the chain.
The first four members of the series (in terms of number of carbon atoms) are named as follows:
1. CH4 = methane = one hydrogen-saturated carbon
2. C2H6 = ethane = two hydrogen-saturated carbons
3. C3H8 = propane = three hydrogen-saturated carbons
4. C4H10 = butane = four hydrogen-saturated carbons
Alkanes with five or more carbon atoms are named by adding the suffix -ane to the appropriate
numerical multiplier, except the terminal -a is removed from the basic numerical term. Hence,
C5H12 is called pentane, C6H14 is called hexane, C7H16 is called heptane and so forth.
Straight-chain alkanes are sometimes indicated by the prefix n- (for normal) to distinguish them
from branched-chain alkanes having the same number of carbon atoms. 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-hexane is a neurotoxin
while its branched-chain isomers are not.

IUPAC nomenclature
The IUPAC nomenclature is a system on which most organic chemists have agreed to provide
guidelines to allow them to learn from each others' works. Nomenclature, in other words,
provides a foundation of language for organic chemistry.

Number of Hydrogen to Carbons


This equation describes the relationship between the number of hydrogen and carbon
atoms in alkanes:
H = 2C + 2
where "C" and "H" are used to represent the number of carbon and hydrogen atoms
present in one molecule. If C = 2, then H = 6.
Many textbooks put this in the following format:
CnH2n+2
where "Cn" and "H2n+2" represent the number of carbon and hydrogen atoms present in
one molecule. If Cn = 3, then H2n+2 = 2(3) + 2 = 8. (For this formula look to the "n" for
the number, the "C" and the "H" letters themselves do not change.)
Progressively longer hydrocarbon chains can be made and are named systematically,
depending on the number of carbons in the longest chain.
The following table contains the systematic names for the first twenty straight
chain alkanes. It will be important to familiarize yourself with these names because they
will be the basis for naming many other organic molecules throughout your course of
study.

Drawing Hydrocarbons
Recall that when carbon makes four bonds, it adopts the tetrahedral geometry. In the
tetrahedral geometry, only two bonds can occupy a plane simultaneously. The other
two bonds point in back or in front of this plane. In order to represent the tetrahedral
geometry in two dimensions, solid wedges are used to represent bonds pointing out of
the plane of the drawing toward the viewer, and dashed wedges are used to represent
bonds pointing out of the plane of the drawing away from the viewer. Consider the
following representation of the molecule methane:

Figure 1: Two dimensional representation of methane


In the above drawing, the two hydrogens connected by solid lines, as well as the
carbon in the center of the molecule, exist in a plane (specifically, the plane of the
computer monitor / piece of paper, etc.). The hydrogen connected by a solid wedge
points out of this plane toward the viewer, and the hydrogen connected by the dashed
wedge points behind this plane and away from the viewer.
In drawing hydrocarbons, it can be time-consuming to write out each atom and bond
individually. In organic chemistry, hydrocarbons can be represented in a shorthand
notation called a skeletal structure. In a skeletal structure, only the bonds between
carbon atoms are represented. Individual carbon and hydrogen atoms are not drawn,
and bonds to hydrogen are not drawn. In the case that the molecule contains just
single bonds (sp3 bonds), these bonds are drawn in a "zig-zag" fashion. This is because
in the tetrahedral geometry all bonds point as far away from each other as possible,
and the structure is not linear. Consider the following representations of the molecule
propane:
Figure 2: Full structure of propane and skeletal structure of propane
Only the bonds between carbons have been drawn, and these have been drawn in a
"zig-zag" manner. Note that there is no representation of hydrogens in a skeletal
structure. Since, in the absence of double or triple bonds, carbon makes four bonds
total, the presence of hydrogens is implicit. Whenever an insufficient number of bonds
to a carbon atom are specified in the structure, it is assumed that the rest of the bonds
are made to hydrogens. For example, if the carbon atom makes only one explicit bond,
there are three hydrogens implicitly attached to it. If it makes two explicit bonds, there
are two hydrogens implicitly attached, etc. Note also that two lines are sufficient to
represent three carbon atoms. It is the bonds only that are being drawn out, and it is
understood that there are carbon atoms (with three hydrogens attached!) at the
terminal ends of the structure.

Alkyl Groups
Alkanes can be described by the general formula CnH2n+2. An alkyl group is formed by
removing one hydrogen from the alkane chain and is described by the formula CnH2n+1.
The removal of this hydrogen results in a stem change from -ane to -yl. Take a look at
the following examples.

The same concept can be applied to any of the straight chain alkane names provided in
the table above.

Molecular Formula Condensed Structural Fo

CH4 CH4

C2H6 CH3CH3

C3H8 CH3CH2CH3
Molecular Formula Condensed Structural Fo

C4H10 CH3(CH2)2CH3

C5H12 CH3(CH2)3CH3

C6H14 CH3(CH2)4CH3

C7H16 CH3(CH2)5CH3

C8H18 CH3(CH2)6CH3

C9H20 CH3(CH2)7CH3

C10H22 CH3(CH2)8CH3

C11H24 CH3(CH2)9CH3

C12H26 CH3(CH2)10CH3

C13H28 CH3(CH2)11CH3

C14H30 CH3(CH2)12CH3

C15H32 CH3(CH2)13CH3
Molecular Formula Condensed Structural Fo

C16H34 CH3(CH2)14CH3

C17H36 CH3(CH2)15CH3

C18H38 CH3(CH2)16CH3

C19H40 CH3(CH2)17CH3

C20H42 CH3(CH2)18CH3

Using Common Names with Branched Alkanes


Certain branched alkanes have common names that are still widely used today. These
common names make use of prefixes, such as iso-, sec-, tert-, and neo-. The
prefix iso-, which stands for isomer, is commonly given to 2-methyl alkanes. In other
words, if there is methyl group located on the second carbon of a carbon chain, we can
use the prefix iso-. The prefix will be placed in front of the alkane name that indicates
the total number of carbons. Examples:
 isopentane which is the same as 2-methylbutane
 isobutane which is the same as 2-methylpropane
To assign the prefixes sec-, which stands for secondary, and tert-, for tertiary, it is
important that we first learn how to classify carbon molecules. If a carbon is attached to
only one other carbon, it is called a primary carbon. If a carbon is attached to two
other carbons, it is called a seconday carbon. A tertiary carbon is attached to three
other carbons and last, a quaternary carbon is attached to four other carbons.
Examples:
 4-sec-butylheptane (30g)
 4-tert-butyl-5-isopropylhexane (30d); if using this example, may want to move sec/tert
after iso disc
The prefix neo- refers to a substituent whose second-to-last carbon of the chain is
trisubstituted (has three methyl groups attached to it). A neo-pentyl has five carbons
total. Examples:
 neopentane
 neoheptane

Alkoxy Groups
Alkoxides consist of an organic group bonded to a negatively charged oxygen atom. In
the general form, alkoxides are written as RO-, where R represents the organic
substituent. Similar to the alkyl groups above, the concept of naming alkoxides can be
applied to any of the straight chain alkanes provided in the table above.

Three Principles of Naming


1. Choose the longest, most substituted carbon chain containing a functional group.
2. A carbon bonded to a functional group must have the lowest possible carbon number. If
there are no functional groups, then any substitute present must have the lowest
possible number.
3. Take the alphabetical order into consideration; that is, after applying the first two rules
given above, make sure that your substitutes and/or functional groups are written in
alphabetical order.

Example 1
What is the name of the following molecule?

SOLUTION
Rule #1: Choose the longest, most substituted carbon chain containing a functional
group. This example does not contain any functional groups, so we only need to be
concerned with choosing the longest, most substituted carbon chain. The longest
carbon chain has been highlighted in red and consists of eight carbons.
Rule #2: Carbons bonded to a functional group must have the lowest possible carbon
number. If there are no functional groups, then any substitute present must have the
lowest possible number. Because this example does not contain any functional groups,
we only need to be concerned with the two substitutes present, that is, the two methyl
groups. If we begin numbering the chain from the left, the methyls would be assigned
the numbers 4 and 7, respectively. If we begin numbering the chain from the right, the
methyls would be assigned the numbers 2 and 5. Therefore, to satisfy the second rule,
numbering begins on the right side of the carbon chain as shown below. This gives the
methyl groups the lowest possible numbering.

Rule 3: In this example, there is no need to utilize the third rule. Because the two
substitutes are identical, neither takes alphabetical precedence with respect to
numbering the carbons. This concept will become clearer in the following examples.

Example 2
What is the name of the following molecule?

SOLUTION
Rule #1: Choose the longest, most substituted carbon chain containing a functional
group. This example contains two functional groups, bromine and chlorine. The longest
carbon chain has been highlighted in red and consists of seven carbons.
Rule #2: Carbons bonded to a functional group must have the lowest possible carbon
number. If there are no functional groups, then any substitute present must have the
lowest possible number. In this example, numbering the chain from the left or the right
would satisfy this rule. If we number the chain from the left, bromine and chlorine
would be assigned the second and sixth carbon positions, respectively. If we number
the chain from the right, chlorine would be assigned the second position and bromine
would be assigned the sixth position. In other words, whether we choose to number
from the left or right, the functional groups occupy the second and sixth positions in the
chain. To select the correct numbering scheme, we need to utilize the third rule.

Rule #3: After applying the first two rules, take the alphabetical order into
consideration. Alphabetically, bromine comes before chlorine. Therefore, bromine is
assigned the second carbon position, and chlorine is assigned the sixth carbon position.

Example 3
What is the name of the follow molecule?

SOLUTION
Rule #1: Choose the longest, most substituted carbon chain containing a functional
group. This example contains two functional groups, bromine and chlorine, and one
substitute, the methyl group. The longest carbon chain has been highlighted in red and
consists of seven carbons.

Rule #2: Carbons bonded to a functional group must have the lowest possible carbon
number. After taking functional groups into consideration, any substitutes present must
have the lowest possible carbon number. This particular example illustrates the point of
difference principle. If we number the chain from the left, bromine, the methyl group
and chlorine would occupy the second, fifth and sixth positions, respectively. This
concept is illustrated in the second drawing below. If we number the chain from the
right, chlorine, the methyl group and bromine would occupy the second, third and sixth
positions, respectively, which is illustrated in the first drawing below. The position of the
methyl, therefore, becomes a point of difference. In the first drawing, the methyl
occupies the third position. In the second drawing, the methyl occupies the fifth
position. To satisfy the second rule, we want to choose the numbering scheme that
provides the lowest possible numbering of this substitute. Therefore, the first of the two
carbon chains shown below is correct.

Therefore, the first numbering scheme is the appropriate one to use.


Once you have determined the correct numbering of the carbons, it is often useful to
make a list, including the functional groups, substitutes, and the name of the parent
chain.
Rule #3: After applying the first two rules, take the alphabetical order into
consideration. Alphabetically, bromine comes before chlorine. Therefore, bromine is
assigned the second carbon position, and chlorine is assigned the sixth carbon position.
Parent chain: heptane 2-Chloro 3-Methyl 6-Bromo
6-bromo-2-chloro-3-methylheptane

Problems
What is the name of the follow molecules?

Properties of Alkanes
1. Last updated
Jun 2
Alkanes are the simplest family of hydrocarbons - compounds containing carbon and hydrogen
only with only carbon-hydrogen bonds and carbon-carbon single bonds. Alkanes are not very
reactive and have little biological activity; all alkanes are colorless and odorless.
 Topic hierarchy
 Alkanes Background
This is an introductory page about alkanes, such as methane, ethane, propane, butane and the
remainder of the common alkanes. This page addresses their formulae and isomerism, their
physical properties, and an introduction to their chemical reactivity.

This is an introductory page about alkanes, such as methane, ethane, propane, butane
and the remainder of the common alkanes. This page addresses their formulae and
isomerism, their physical properties, and an introduction to their chemical reactivity.
Molecular Formulas
Alkanes are the simplest family of hydrocarbons - compounds containing carbon and
hydrogen only. Alkanes only contain carbon-hydrogen bonds and carbon-carbon single
bonds. The formula of any of the alkanes follow the general
formula CnH2n+2CnH2n+2 . The first six alkanes are tabulated below:
Table 1: The first six alkanes are
as follows:

name Formula

methane CH4

ethane C2H6

propane C3H8

butane C4H10

pentane C5H12
Table 1: The first six alkanes are
as follows:

name Formula

hexane C6H14

Isomerism
All of the alkanes containing 4 or more carbon atoms show structural isomerism,
meaning that there are two or more different structural formulae that you can draw for
each molecular formula.

Example 1: Butane or MethylPropane


C4H10 could be either of these two different molecules:

These are named butane and 2-methylpropane, respectively.

Cycloalkanes
Cycloalkanes also only contain carbon-hydrogen bonds and carbon-carbon single bonds,
but the carbon atoms are joined in a ring. The smallest cycloalkane is cyclopropane.
If you count the carbons and the hydrogens, you will see that they no longer adhere to
the general formula CnH2n+2. By joining the carbon atoms in a ring, two hydrogen atoms
are lost. The general formula for a cycloalkane is CnH2n and these are non-planar
molecules (with the exception of cyclopentane) and exist as "puckered rings".

Example 2: Cyclohexane Conformation


Cyclohexane has a ring structure that looks like this:

This is known as the "chair" form of cyclohexane because of its shape, which vaguely
resembles a chair. Cyclohexane also has a boat configuration (not shown).

Physical Properties

Boiling Points
The boiling points shown are for the "straight chain" isomers in which there are more
than one (Figure 1). Notice that the first four alkanes are gases at room temperature,
and solids do not start to appear until about C17H36.
Figure 1:Normal boiling points of first four alkanes
The temperatures cannot be more precise than those given in this chart because each
isomer has a different melting and boiling point. By the time you get 17 carbons into an
alkane, there are unbelievable numbers of isomers! Cycloalkanes have boiling points
that are approximately 10 - 20 K higher than the corresponding straight chain alkane.
There electronegativity difference between carbon and hydrogen (2.1 vs. 1.9) is small;
therefore, there is only a slight bond polarity, meaning that the only attractions
between one molecule and its neighbors will be Van der Waals dispersion forces. These
forces will be very small for a molecule like methane but will increase as the size of the
molecules increase. Therefore, the boiling points of the alkanes increase with the
molecular size.
Regarding isomers, the more branched the chain, the lower the boiling point tends to
be. Van der Waals dispersion forces are smaller for shorter molecules and only operate
over very short distances between one molecule and its neighbors. It is more difficult
for short, bulky molecules (with substantial amounts of branching) to lie close together
(compact) compared with long, thin molecules.

Example 3
The boiling points of the three isomers of C5H12 are as follows:
 pentane (309.2 K)
 2-methylbutane (301.0 K)
 2,2-dimethylpropane (282.6 K)
The slightly higher boiling points for the cycloalkanes are presumably because the
molecules can get closer together because the ring structure makes them tidier and less
"wriggly"!
Solubility
Alkanes (both normal and cycloalkanes) are virtually insoluble in water but dissolve in
organic solvents. The liquid alkanes are good solvents for many other covalent
compounds. When a molecular substance dissolves in water, the following must occur:
 breaking of the intermolecular forces within the substance. In the case of the alkanes,
these are the Van der Waals dispersion forces.
 breaking of the intermolecular forces in the water so that the substance can fit between
the water molecules. In water, the primary intermolecular attractions are hydrogen
bonds.
Breaking either of these attractions requires energy, although the amount of energy
required to break the Van der Waals dispersion forces in a compound, such as
methane, is relatively negligible; this is not true of the hydrogen bonds in water.
To simplify, a substance will dissolve if sufficient energy is released when the new
bonds are formed between the substance and the water to make up for the energy
required to break the original attractions. The only new attractions between the alkane
and the water molecules are the Van der Waals forces. These forces to do not release a
sufficient amount of energy to compensate for the energy required to break the
hydrogen bonds in water. Therefore, the alkane does not dissolve.

Solubility in organic solvents


In most organic solvents, the primary forces of attraction between the solvent
molecules are the Van der Waals forces composed of either dispersion forces or dipole-
dipole attractions. Therefore, when an alkane dissolves in an organic solvent, the Van
der Waals forces are broken and are replaced by new Van der Waals forces. The two
processes more or less cancel each other out energetically; thus, there is no barrier to
solubility.

Chemical Reactivity
Alkanes contain strong carbon-carbon single bonds and strong carbon-hydrogen bonds.
The carbon-hydrogen bonds are only very slightly polar. Therefore, there is no portion
of the molecule that carries any significant amount of positive or negative charge, which
is required for other molecules to be attracted to it. For example, many organic
reactions start because an ion or a polar molecule is attracted to a portion of an organic
molecule, which carries some positive or negative charge. This attraction does not occur
with alkanes because alkane molecules do not have this separation of charge. The net
effect is that alkanes have a fairly restricted set of reactions, including the following:
 burn them - destroying the entire molecule;
 react them with some of the halogens, breaking the carbon-hydrogen bonds;
 crack them, breaking carbon-carbon bonds.
These reactions are covered in separate pages.
Cycloalkanes are very similar to the alkanes in reactivity, except for the very small
cycloalkanes, especially cyclopropane. Cyclopropane is much more reactive than what is
expected because of the bond angles in the ring. Normally, when carbon forms four
single bonds, the bond angles are approximately 109.5°.

Example 4: Cyclopropane
In cyclopropane, the bond angles are 60°.

With the electron pairs this close together, there is a significant amount of repulsion
between the bonding pairs joining the carbon atoms, making the bonds easier to break.

Conformational Analysis of Alkanes


A conformational analysis is a study of the energetics of different spatial
arrangements of atoms relative to rotations about bonds. Conformational analyses are
assisted greatly by a representation of molecules in a manner different to skeletal
structures. A representation of a molecule in which the atoms and bonds are viewed
along the axis about which rotation occurs is called a Newman projection (Figure 2).

Figure 2: Newman projection of ethane


In a Newman projection, the molecule is viewed along an axis containing two atoms
bonded to each other and the bond between them, about which the molecule can
rotate. In a Newman projection, the "substituents" of each atom composing the bond,
be they hydrogens or functional groups, can then be viewed both in front of and behind
the carbon-carbon bond. Specifically, one can observe the angle between a substituent
on the front atom and a substituent on the back atom in the Newman projection, which
is called the dihedral angle or torsion angle.
In ethane specifically, we can imagine two possible "extreme" conformations. In one
case, the dihedral angle is 0° and the hydrogens on the first carbon line up with or
eclipse the hydrogens on the second carbon. When the dihedral angle is 0° and the
hydrogens line up perfectly, ethane has adopted the eclipsed conformation (Figure 3).
The other extreme occurs when the hydrogens on the first carbon are as far away as
possible from those on the second carbon; this occurs at a dihedral angle of 60° and is
called the staggeredconformation (Figure 3).

Figure 3: Conformations of ethane - staggered and eclipsed


The staggered conformation of ethane is a more stable, lower energy conformation
than the eclipsed conformation because the eclipsed conformation involves unfavorable
interactions between hydrogen atoms. Specifically, the negatively charged electrons in
the bonds repel each other most when the bonds line up. Thus, ethane spends most of
its time in the more stable staggered conformation.

Conformational analysis of butane


In butane, two of the substituents, one on each carbon atom being viewed, is a methyl
group. Methyl groups are much larger than hydrogen atoms. Thus, when eclipsed
conformations occur in butane, the interactions are especially unfavorable. There are
four possible "extreme" conformations of butane: 1) Fully eclipsed, when the methyl
groups eclipse each other; 2) Gauche, when the methyl groups are staggered but next
to each other; 3) Eclipsed, when the methyl groups eclipse hydrogen atoms; and
4) Anti, when the methyl groups are staggered and as far away from each other as
possible (Figure 4).

Figure 4: Conformations of butane: fully eclipsed, gauche, eclipsed, and anti


The fully eclipsed conformation is clearly the highest in energy and least favorable since
the largest groups are interacting directly with each other. As the molecule rotates, it
adopts the relatively stable gauche conformation. As it continues to rotate, it
encounters less favorable eclipsed conformation in which a methyl group eclipses a
hydrogen. As rotation continues, the molecule comes to the anti conformation, which is
the most stable since the substituents are staggered and the methyl groups are as far
away from each other as possible. An energy diagram is a graph which represents
the energy of a molecule as a function of some changing parameter. An energy diagram
can be made as a function of dihedral angle for butane (Figure 5).

Figure 5: Energy diagram for conformations of butane as a function of dihedral angle


Clearly, the anti and gauche conformations are significantly more stable than the
eclipsed and fully eclipsed conformations. Butane spends most of its time in the anti
and gauche conformations.
There is one final, very important point. At room temperature, approximately 84 kJ/mol
of thermal energy is available to molecules. Thus, if the barrier to a rotation is less than
84 kJ/mol, the molecule will rotate. In ethane and butane, the barriers to rotation are
significantly less than 84 kJ/mol. Therefore, even though the eclipsed conformations are
unfavorable, the molecules are able to adopt them. In reality, since these
conformations are not as stable, the molecules will quickly pass through them at room
temperature and return a staggered conformation. Molecules are constantly converting
between different staggered conformations all the time, quickly passing through
eclipsed conformations in between. Thus, in alkanes, no single "true" conformation
exists all the time; the molecule instead constantly converts between conformations,
spending more time in those that are more stable. This constant conversion lies in stark
contract to alkenes, which adopt the cis- or trans- (E- or Z-) conformations and retain
them at room temperature; they do not interconvert because the barrier to rotation is
too high.

Chemical Properties of Alkanes


1. Last updated
Jun 5, 2019
2.
o Alkanes Background

o Cycloalkanes
3. picture_as_pdf
LMS

Donate

Alkanes are not very reactive when compared with other chemical species. This is
because the backbone carbon atoms in alkanes have attained their octet of electrons
through forming four covalent bonds (the maximum allowed number of bonds under
the octet rule; which is why carbon's valence number is 4). These four bonds formed by
carbon in alkanes are sigma bonds, which are more stable than other types of bond
because of the greater overlap of carbon's atomic orbitals with neighboring atoms'
atomic orbitals. To make alkanes react, the input of additional energy is needed; either
through heat or radiation.
Gasoline is a mixture of the alkanes and unlike many chemicals, can be stored for long
periods and transported without problem. It is only when ignited that it has enough
energy to continue reacting. This property makes it difficult for alkanes to be converted
into other types of organic molecules. (There are only a few ways to do this). Alkanes
are also less dense than water, as one can observe, oil, an alkane, floats on water.
Alkanes are non-polar solvents. Since only C and H atoms are present, alkanes are
nonpolar. Alkanes are immiscible in water but freely miscible in other non-polar
solvents. Alkanes consisting of weak dipole dipole bonds can not break the strong
hydrogen bond between water molecules hence it is not miscible in water. The same
character is also shown by alkenes. Because alkanes contain only carbon and hydrogen,
combustion produces compounds that contain only carbon, hydrogen, and/or oxygen.
Like other hydrocarbons, combustion under most circumstances produces mainly
carbon dioxide and water. However, alkanes require more heat to combust and do not
release as much heat when they combust as other classes of hydrocarbons. Therefore,
combustion of alkanes produces higher concentrations of organic compounds containing
oxygen, such as aldehydes and ketones, when combusting at the same temperature as
other hydrocarbons.
The general formula for alkanes is CNH2N+2; the simplest possible alkane is therefore
methane, CH4. The next simplest is ethane, C2H6; the series continues indefinitely. Each
carbon atom in an alkane has sp³ hybridization.
Alkanes are also known as paraffins, or collectively as the paraffin series. These terms
are also used for alkanes whose carbon atoms form a single, unbranched chain.
Branched-chain alkanes are called isoparaffins.
Methane through Butane are very flammable gases at standard temperature and
pressure (STP). Pentane is an extremely flammable liquid boiling at 36 °C and boiling
points and melting points steadily increase from there; octadecane is the first alkane
which is solid at room temperature. Longer alkanes are waxy solids; candle wax
generally has between C20 and C25 chains. As chain length increases ultimately we reach
polyethylene, which consists of carbon chains of indefinite length, which is generally a
hard white solid.

Reactions
Alkanes react only very poorly with ionic or other polar substances. The pKa values of
all alkanes are above 50, and so they are practically inert to acids and bases. This
inertness is the source of the term paraffins (Latin para + affinis, with the meaning
here of "lacking affinity"). In crude oil the alkane molecules have remained chemically
unchanged for millions of years.
However redox reactions of alkanes, in particular with oxygen and the halogens, are
possible as the carbon atoms are in a strongly reduced condition; in the case of
methane, the lowest possible oxidation state for carbon (−4) is reached. Reaction with
oxygen leads to combustion without any smoke; with halogens, substitution. In
addition, alkanes have been shown to interact with, and bind to, certain transition metal
complexes.
Free radicals, molecules with unpaired electrons, play a large role in most reactions of
alkanes, such as cracking and reformation where long-chain alkanes are converted into
shorter-chain alkanes and straight-chain alkanes into branched-chain isomers.
In highly branched alkanes and cycloalkanes, the bond angles may differ significantly
from the optimal value (109.5°) in order to allow the different groups sufficient space.
This causes a tension in the molecule, known as steric hinderance, and can
substantially increase the reactivity. The same is preferred for alkenes too.

Cycloalkanes
1. Last updated
Jun 23, 2019

2.
o Chemical Properties of Alkanes

o Physical Properties of Cycloalkanes


3. picture_as_pdf
LMS

Donate
Cycloalkanes are cyclic hydrocarbons with the carbon atoms of the molecule are arranged in one
or more rings. Cycloalkanes are also saturated, meaning that all of the carbons atoms that make
up the ring are single bonded to other atoms (no double or triple bonds).

Physical Properties of Cycloalkanes


1. Last updated
Jun 5, 2019

2.
o Cycloalkanes

o Ring Strain and the Structure of Cycloalkanes


3. picture_as_pdf
LMS

Donate

Cycloalkanes are types of alkanes that have one or more rings of carbon atoms in their structure.
The physical properties of cycloalkanes are similar to those of alkanes, but they have higher
boiling points, melting points and higher densities due to the greater number of London forces
that they contain.

Introduction
Cycloalkanes consist of carbon and hydrogen atoms that are saturated because of the single
carbon-carbon bond (meaning that no more hydrogen atoms can be added). Cycloalkanes are
also non polar and do not have intermolecular hydrogen bonding; they are usually hydrophobic
(meaning they do not dissolve in water) and are less dense than water. Cycloalkanes can also be
used for many different purposes. These uses are typically classified by the number of carbons in
the cycloalkane ring. Many cycloalkanes are used in motor fuel, natural gas, petroleum gas,
kerosene, diesel, and many other heavy oils. There are 4 general groups of cycloalkanes:
1. Small rings (cyclopropane, cyclobutane)
2. Common rings (cyclopentane, cyclohexane, cycloheptane)
3. Medium rings (from 8-12 membered)
4. Large rings (13 membered and higher)
Cycloalkanes can be substituted and named as cycloalkyl derivatives, and disubstituted
cycloalkanes can be stereoisomers. In one isomer, if two substituents are placed on the same face
or side of the ring, they are called cis. If the two substituents are on opposite faces, they are
called trans. Substituents can also be either equitorial or axial on certain cycloalkanes, such as
cyclohexane.
Figure 1: The cis and trans isomers are stereoisomers, meaning that they have identicial connectivities but a different
arrangement of their atoms.
Generally, the melting point, the boiling point and the density of cycloalkanes increase as the
number of carbons increases. This trend occurs because of the greater number of bonds that are
in higher membered rings, thus making the bonds harder to break.

London Dispersion Forces and Cycloalkanes


Although alkanes are similar to cycloalkanes, they have higher London Dispersion forces
because the ring shape allows for a greater area of contact. Ring strain also causes certain
cylcoalkanes to be more reactive. London Dispersion Forces are the attractive or repulsive forces
between molecules or between parts of the same molecule. For cycloalkanes, London dispersion
forces refer to the repulsive forces between the molecules that cause ring strain.

Ring Strain in Cycloalkanes


Ring Strain occurs because the carbons in cycloalkanes are sp3 hybridized, which means that
they do not have the expected ideal bond angle of 109.5o ; this causes an increase in the potential
energy because of the desire for the carbons to be at an ideal 109.5o. An example of ring strain
can be seen in the diagram of cyclopropane below in which the bond angle is 60 o between the
carbons.

The reason for ring strain can be seen through the tetrahedral carbon model. The C-C-C bond
angles in cyclopropane (diagram above) (60o) and cyclobutane (90o) are much different than the
ideal bond angle of 109.5o. This bond angle causes cyclopropane and cyclobutane to have a high
ring strain. However, molecules, such as cyclohexane and cyclopentane, would have a much
lower ring strain because the bond angle between the carbons is much closer to 109.5o.
Below are some examples of cycloalkanes. Ring strain can be seen more prevalently in the
cyclopropane and cyclobutane models.

Below is a chart of cycloalkanes and their respective heats of combustion ( ΔHcomb). The
ΔHcomb value increases as the number of carbons in the cycloalkane increases (higher membered
ring), and the ΔHcomb/CH2 ratio decreases. The increase in ΔHcomb can be attributed to the greater
amount of London Dispersion forces. However, the decrease in ΔHcomb/CH2can be attributed to a
decrease in the ring strain.

How do cycloalkanes deal with ring strain?


Certain cycloalkanes, such as cyclohexane, deal with ring strain by forming conformers. A
conformer is a stereoisomer in which molecules of the same connectivity and formula exist as
different isomers, in this case, to reduce ring strain. The ring strain is reduced in conformers due
to the rotations around the sigma bonds. More about cyclohexane and its conformers can be
seen here.

Different Types of Strain


There are many different types of strain that occur with cycloalkanes. In addition to ring strain,
there is also transannular strain, eclipsing, or torsional strain and bond angle strain.Transannular
strain exists when there is steric repulsion between atoms. Eclipsing (torsional) strain exists
when a cycloalkane is unable to adopt a staggered conformation around a C-C bond, and bond
angle strain is the energy needed to distort the tetrahedral carbons enough to close the ring. The
presence of angle strain in a molecule indicates that there are bond angles in that particular
molecule that deviate from the ideal bond angles required (i.e., that molecule has conformers).

Outside links
 http://www.cem.msu.edu/~reusch/Virtu...s/cycalkan.gif
 http://img.sparknotes.com/figures/1/...6c/fig2_26.gif
 http://www.chemgapedia.de/vsengine/m...opropangif.gif

References
1. Vollhardt, K. Peter C., and Neil E. Schore. Organic Chemistry. 5th ed. New York: W.H. Freeman,
2007.
2. John E. McMurry, and Eric E. Simanek. Fundamentals of Organic Chemistry. 6th Ed. Brooks Cole,
2006.

Problems
1. Which has a higher melting point?
a. cyclopentane
b. cyclopropane
c. cycloocatne
d. they all have the same melting point.
2. Why do cycloalkanes have different physical properties from normal alkanes?
3. What is bond angle strain?
4. What are London Dispersion Forces, and how do they play a part in cycloalkanes?
5. What is the "ideal bond angle" for most cycloalkanes?

Answers
1. C; as the number of carbons increases, so does the melting point. This happens because there
are essentially more London dispersion forces acting upon the molecule which then makes it
much harder to melt, or boil.
2. Cycloalkanes have different physical properties from normal alkanes due to the greater number
of London dispersion forces that they have. Cycloalkanes also have the ability to have more
steric hindrance, thus increasing their relative energy level.
3. Bond angle strain is is the energy needed to distort the tetrahedral carbons enough to close the
ring
4. London Dispersion Forces are the attractive or repulsive forces between molecules, or in
between parts of the same molecule. They play a role in cycloalkanes because they refer to the
intramolecular forces that causes ring strain.
5. 109.5 o

Ring Strain and the Structure of


Cycloalkanes
1. Last updated
Jun 5, 2019

2.
o Physical Properties of Cycloalkanes

o Rings: cis/trans and axial/equatorial relationships


3. picture_as_pdf
LMS

Donate
Cycloalkanes are very important in components of food, pharmaceutical drugs, and much more.
However, to use cycloalkanes in such applications, we must know the effects, functions,
properties, and structures of cycloalkanes. Cycloalkanes are alkanes that are in the form of a
ring; hence, the prefix cyclo-. Stable cycloalkanes cannot be formed with carbon chains of just
any length. Recall that in alkanes, carbon adopts the sp3 tetrahedral geometry in which the angles
between bonds are 109.5°. For some cycloalkanes to form, the angle between bonds must deviate
from this ideal angle, an effect known as angle strain. Additionally, some hydrogen atoms may
come into closer proximity with each other than is desirable (become eclipsed), an effect
called torsional strain. These destabilizing effects, angle strain and torsional strain are known
together as ring strain. The smaller cycloalkanes, cyclopropane and cyclobutane, have
particularly high ring strains because their bond angles deviate substantially from 109.5° and
their hydrogens eclipse each other. Cyclopentane is a more stable molecule with a small amount
of ring strain, while cyclohexane is able to adopt the perfect geometry of a cycloalkane in which
all angles are the ideal 109.5° and no hydrogens are eclipsed; it has no ring strain at all.
Cycloalkanes larger than cyclohexane have ring strain and are not commonly encountered in
organic chemistry.
Ring Strain and the Structures of Cycloalkanes
There are many forms of cycloalkanes, such as cyclopropane, cyclobutane, cyclopentane,
cyclohexane, among others. The process of naming cycloalkanes is the same as naming alkanes
but the addition of the prefix cyclo- is required. Cyclobutane is in a form of a square, which is
highly unfavorable and unstable (this will be explained soon). There are different drawings for
cyclobutane, but they are equivalent to each other. Cyclobutane can reduce the ring string by
puckering the square cyclobutane. Cyclopentane takes the shape of a pentagon and cyclohexane
is in the shape of a hexagon.

Chair Conformation of Cyclohexane - Equitorial and Axial


There are two ways to draw cyclohexane because it can be in a hexagon shape or in a different
conformational form called the chair conformation and the boat conformation.
 The chair conformation drawing is more favored than the boat because of the energy, the steric
hindrance, and a new strain called the transannular strain.
 The boat conformation is not the favored conformation because it is less stable and has a steric
repulsion between the two H's, shown with the pink curve. This is known as the transannular
strain, which means that the strain results from steric crowding of two groups across a ring. The
boat is less stable than the chair by 6.9 kcal/mol. The boat conformation, however, is flexible,
and when we twist one of the C-C bonds, it reduces the transannular strain.
 When we twist the C-C bond in a boat, it becomes a twisted boat.

Some Conformations of Cyclohexane Rings. Image used with permission (William Reusch, MSU)
Although there are multiple ways to draw cyclohexane, the most stable and major conformer is
the chair because is has a lower activation barrier from the energy diagram.

Conformational Energy Profile of Cyclohexane. Image used with permission (William Reusch,
MSU).
The transition state structure is called a half chair. This energy diagram shows that the chair
conformation is lower in energy; therefore, it is more stable. The chair conformation is more
stable because it does not have any steric hindrance or steric repulsion between the hydrogen
bonds. By drawing cyclohexane in a chair conformation, we can see how the H's are positioned.
There are two positions for the H's in the chair conformation, which are in an axial or an
equitorial formation.
This is how a chair conformation looks, but you're probably wondering which H's are in the
equitorial and axial form. Here are more pictures to help.

These are hydrogens in the axial form.

These hydrogens are in an equitorial form. Of these two positions of the H's, the equitorial form
will be the most stable because the hydrogen atoms, or perhaps the other substituents, will not be
touching each other. This is the best time to build a chair conformation in an equitorial and an
axial form to demonstrate the stability of the equitorial form.

Ring Strain
Cycloalkanes tend to give off a very high and non-favorable energy, and the spatial orientation of
the atoms is called the ring strain. When atoms are close together, their proximity is highly
unfavorable and causes steric hindrance. The reason we do not want ring strain and steric
hindrance is because heat will be released due to an increase in energy; therefore, a lot of that
energy is stored in the bonds and molecules, causing the ring to be unstable and reactive.
Another reason we try to avoid ring strain is because it will affect the structures and the
conformational function of the smaller cycloalkanes. One way to determine the presence of ring
strain is by its heat of combustion. By comparing the heat of combustion with the value
measured for the straight chain molecule, we can determine the stability of the ring. There are
two types of strain, which are eclipsing/torsional strain and bond angle strain. Bond angle strain
causes a ring to have a poor overlap between the atoms, resulting in weak and reactive C-C
bonds. An eclipsed spatial arrangement of the atoms on the cycloalkanes results in high energy.
With so many cycloalkanes, which ones have the highest ring strain and are very unlikely to stay
in its current form? The figures below show cyclopropane, cyclobutane, and cyclopentane,
respectively. Cyclopropane is one of the cycloalkanes that has an incredibly high and
unfavorable energy, followed by cyclobutane as the next strained cycloalkane. Any ring that is
small (with three to four carbons) has a significant amount of ring strain; cyclopropane and
cyclobutane are in the category of small rings. A ring with five to seven carbons is considered to
have minimal to zero strain, and typical examples are cyclopentane, cyclohexane, and
cycloheptane. However, a ring with eight to twelve carbons is considered to have a moderate
strain, and if a ring has beyond twelve carbons, it has minimal strain.
There are different types of ring strain:
 Transannular strain isdefined as the crowding of the two groups in a ring.
 Eclipsing strain, also known as torsional strain, is intramolecular strain due to the bonding
interaction between two eclipsed atoms or groups.
 Bond angle strain is present when there is a poor overlap between the atoms. There must be an
ideal bond angle to achieve the maximum bond strength and that will allow the overlapping of
the atomic/hybrid orbitals.

Cyclohexane
Most of the time, cyclohexane adopts the fully staggered, ideal angle chair conformation. In the
chair conformation, if any carbon-carbon bond were examined, it would be found to exist with
its substituents in the staggered conformation and all bonds would be found to possess an angle
of 109.5°.

Cyclohexane in the chair conformation. Image used with permission (William Reusch, MSU).
In the chair conformation, hydrogen atoms are labeled according to their location. Those
hydrogens which exist above or below the plane of the molecule (shown with red bonds above)
are called axial. Those hydrogens which exist in the plane of the molecule (shown with blue
bonds above) are called equatorial.
Although the chair conformation is the most stable conformation that cyclohexane can adopt,
there is enough thermal energy for it to also pass through less favorable conformations before
returning to a different chair conformation. When it does so, the axial and equatorial substituents
change places. The passage of cyclohexane from one chair conformation to another, during
which the axial substituents switch places with the equatorial substituents, is called a ring flip.

Methylcyclohexane
Methylcyclohexane is cyclohexane in which one hydrogen atom is replaced with a methyl group
substituent. Methylcyclohexane can adopt two basic chair conformations: one in which the
methyl group is axial, and one in which it is equatorial. Methylcyclohexane strongly prefers the
equatorial conformation. In the axial conformation, the methyl group comes in close proximity to
the axial hydrogens, an energetically unfavorable effect known as a 1,3-diaxial
interaction (Figure 3). Thus, the equatorial conformation is preferred for the methyl group. In
most cases, if the cyclohexane ring contains a substituent, the substituent will prefer the
equatorial conformation.

Methylcyclohexane in the chair conformation. Image used with permission (William Reusch, MSU).

References
1. Bachrach, Steven M. Computational Organic Chemistry. Wiley- Interscience. (2007).
2. McMurray, John E. Organic Chemistry Sixth Edition. Brooks Cole. (2003).
3. Vollhardt and Schore. Organic Chemistry Structure and Function Fifth Edition. New York. (2007).

Problems
1. Trans- 1,2-dimethylcyclopropane is more stable than cis-1,2-dimethylcyclopropane. Why?
Drawing a picture of the two will help your explanation.
2. Out of all the cycloalkanes, which one has the most ring strain and which one is strain free?
Explain.
3. Which of these chair conformations are the most stable and why?
3.
4. What does it mean when people say "increase in heat leads to increase in energy" and how does
that statement relate to ring strains?
5. Why is that the bigger rings have lesser strains compared to smaller rings?

Answers
1. The cis isomer suffers from steric hindrance and has a larger heat of combustion.
2. Cyclopropane- ring strain. Cyclohexane chair conformation- ring strain free.
3. Top one is more stable because it is in an equitorial conformation. When assembling it with the
OChem Molecular Structure Tool Kit, equitorial formation is more spread out.
4. When there is an increase in heat there will be an increase of energy released therefore there
will be a lot of energy stored in the bond and molecule making it unstable.
5. Smaller rings are more compacts, which leads to steric hindrance and the angles for these
smaller rings are harder to get ends to meet. Bigger rings tend to have more space and that the
atoms attached to the ring won't be touching each other as much as atoms attached to the
smaller ring.

Rings: cis/trans and axial/equatorial


relationships
1. Last updated
Jun 5, 2019

2.
o Ring Strain and the Structure of Cycloalkanes

o Physical Properties of Alkanes


3. picture_as_pdf
LMS

Donate

The purpose of this page is to help organic chem students show how substituent groups are
located on ring structures. We focus here on six-membered rings (6-rings); these are among the
most common rings in organic chem (and biochem), and they suffice to raise the main issues. We
will look at how to show cis and trans relationships in simple hexagon structural formulas, and
we will look at structures showing the common "chair" conformation, focusing on axial vs
equatorial orientations. We will also discuss the relationship between cis/trans and
axial/equatorial.

Introduction
The basic approach here is to look at a series of compounds, of generally increasing complexity.
They are chosen to illustrate one new feature at a time of how to draw the structures and how to
see cis/trans and axial/equatorial features.

Cyclohexane

Figure 1: This basic 6-ring cycloalkane, without any special substituents (i.e., other than
hydrogen), is easily shown by either of the formulas in Fig 1.
Both structures clearly imply six C in a ring, with two H at each position. Fig 1A is a simple
structural formula, condensed. We say it is "condensed" because not all features are shown
explicitly. In fact, in this case, not much at all is shown explicitly: the C atoms are understood (at
each vertex), the H atoms are understood (enough at each C to give C its normal four bonds), and
the C-H bonds are understood. About all that is explicit is the basic carbon skeleton (C-C bonds),
showing that there are six C atoms in a ring. Fig 1B is very much like Fig 1A, but it adds one
new feature: Fig 1B attempts to show the conformation of the molecule. In this case, we
understand that cyclohexane and its simple derivatives tend to spend most of their time near this
"chair" conformation. Glossary entry: Conformation.
Figure 2: If we want to show the hydrogen atoms explicitly, we can do so with structural
formulas such as in Fig 2.
There is little room for confusion here, since all six C atoms are equivalent, as are all 12 H
atoms.
For notes on how to draw chairs, see the section E.2. Note: How to
draw chairs.

Chlorocyclohexane

This is an example of the next level of complexity, a mono-substituted cycloalkane. See Fig 3.

So what is new here? Not much, with the hexagon formula, Fig 3A. That type of
formula shows the basic "connectivity" of the atoms -- who is connected to whom. This
chemical has one Cl on the ring, and it does not matter where we show it. There is now
only one H on that C, but since we are not showing H explicitly here, that is not an
issue in drawing the structure. (It is an issue when you look at it and want to count H.)
With the chair formula (Fig 3B), which shows information not only about connectivity
but also about conformation, there is important new information here. In a chair, there
are two "types" of substituents: those pointing up or down, and called axial, and those
pointing "outward", and called equatorial. I have shown the chlorine atom in an
equatorial position. Why? Two reasons: it is what we would predict, and it is what is
found. Why do we predict that the Cl is equatorial? Because it is bigger than H, and
there is more room in the equatorial positions.

Helpful Hints...
If possible, examine a physical model of cyclohexane and
chlorocyclohexane, so that you can see the axial and equatorial
positions. Common ball and stick models are fine for this. It should be
easy to see that the three axial H on one side can get very near each
other.
If you do not have access to physical models, examining computer
models can also be useful. See my RasMol page for a program and
source of structures for this.
When putting substituents on chair structures, I encourage you to use
the four corner positions of the chair as much as possible. It is easier
to see the axial and equatorial relationship at the corners.
In Fig 3B I have shown the H atom that is on the same carbon as the
Cl atom. This is perhaps not necessary, since the correct number of H
atoms is understood, by counting bonds on C. But showing the H
explicitly at key C atoms helps to make the structure clearer. This may
be particularly important with hand-drawn structures. I often see
structures where I am not sure whether a particular atom is shown
axial or equatorial. But if both atoms at the position (the H as well as
the Cl) are shown, then hopefully it becomes clearer which is which. I
also encourage students who are not sure of their art work to
annotate their drawing. Say what you mean. That allows me to
distinguish whether you are unsure which direction things point or
simply unsure how to draw them.)
Again, a reminder... For notes on how to draw chairs (by hand or
using a drawing program), see the section E.2. Note: How to draw
chairs.

D. Dichlorocyclohexanes: an introduction
The next level of complexity is a di-substituted cycloalkane, "dichlorocyclohexane". The
first question we must ask is which C the two chlorine substituents are on. For now, I
want to discuss 1,3- dichlorocyclohexane. This introduces another issue: are the two Cl
on the same side of the ring, or on opposite sides? We call these "cis" (same side) and
"trans" (opposite sides). We focus on one of these, cis-1,3-dichlorocyclohexane. And for
now, we will just look at hexagon structural formulas, leaving the question of
conformation for later. Let's go through this one step at a time.

1,3-Dichlorocyclohexane
Our first attempt to draw 1,3-dichlorocyclohexane might look something like Fig 4.

The structure in Fig 4 is indeed a dichlorocyclohexane. It is even a 1,3-


dichlorocyclohexane. However, this structure provides no information about the
orientation of the two Cl atoms relative to the plane of the ring. To show a specific
isomer -- cis or trans -- we must somehow show how the two Cl atoms are oriented
relative to the plane of the ring.

cis-1,3-Dichlorocyclohexane

Fig 5 shows two common ways to show how the substituents are oriented relative to the plane of the ring. The
dichlorocyclohexane.

The basic idea in both of these is that we can imagine the ring to be planar, and then
show the groups above or below the plane of the ring. How we show this is different in
the two parts of Fig 5. In Fig 5A, we look at the ring "edge-on". The thick line for the
bottom bond is intended to convey the edge-on view (or side view); this is sometimes
omitted, especially with hand-drawn structures, but be careful then that the meaning is
clear. Once we understand that we are now looking at the ring edge-on, it is clear that
the two Cl atoms are both above the ring, hence cis. In Fig 5B, we view the ring "face-
on" (or top view), and use special bond symbols -- "stereo bonds" -- to convey up and
down: the heavy wedge -- an "up wedge" -- points upward, toward you, and the
dashed bond -- a "down bond" -- points downward, away from you. Again, both Cl are
"up", hence cis.
Notes...
For some notes on how to draw the stereo bonds, see the section E.1.
Note: How to draw stereo bonds ("up" and "down" bonds).
In discussing Fig 5, I started by saying that we imagine the ring to be
planar. Emphasize that cycloalkane rings are not really planar (except
for cyclopropane rings). As so often, the structural formula represents
the general layout of the atoms, but not the actual molecular
geometry.
Those with the Ouellette book can see examples of these two ways of
showing up/down on p 81 (top) and p 80 (middle). Most organic
chemistry books will show you this.

The conformation of cis-1,3-dichlorocyclohexane

Fig 6, at the right, is one way to show this. Fig 6 shows features of the compound that we have already noted
emphasizing what is new.

The structure shows cis-1,3-dichlorocyclohexane: a 6-ring; 2 Cl atoms, at positions 1


and 3; and cis, with both Cl on the same side of -- above -- the H that is on the same
C.
I showed the 2 Cl atoms at corner positions, and I showed the H at
the key positions explicitly. These points follow from some of
the Helpful Hints discussed earlier. It is not required that you do these
things, but they can make things easier for you -- and for anyone
reading your structures.
Now, what is new here? The conformation. We start with the notion that the
conformation of cyclohexane derivatives is based on the "chair". At each position, one
substituent is axial (loosely, perpendicular to the ring), and one is equatorial (loosely, in
the plane of the ring). There is more room in the equatorial positions (not easily seen
with these simple drawings, but ordinary ball and stick models do help with this point).
Thus we try to put the larger substituents in the equatorial positions. In this case, we
put the Cl equatorial and the H axial at each position 1 and position 3.
We are now done with this compound, cis-1,3-dichlorocyclohexane. However, we have
missed one very important concern -- because it is not an issue in this case. So let's
look at another compound.

cis-1,2-Dichlorocyclohexane

cis-1,2-Dichlorocyclohexane is like the previous compound, except that the two chloro gro

Fig 7 shows a simple structural formula for 1,2-dichlorocyclohexane, without orientation. This is analogous to Fi

Figure 8
Fig 8 shows two ways to show the cis orientation in cis-1,2-dichloro

Figures 7 and 8 above introduce no new ideas or complications. These two figures
should be straightforward.
So, what is the preferred conformation of cis-1,2-dichlorocyclohexane? This requires
careful consideration; an important lesson from this exercise is to realize that we cannot
propose a good conformation based simply on what we have learned so far.
The two guidelines we have so far for conformation of 6-rings are:
 The carbon ring is in a chair.
 Larger substituents are in equatorial positions.
Let's explore the difficulty here by looking at some things people might naively draw.
Figure 9
Fig 9 shows an attempt to draw a chair conformation of cis-1,2-dichlorocycloh

Why is Fig 9 wrong? It is the wrong chemical. The structure shown in Fig 9 is trans, not
cis. Look carefully at the 1 and 2 positions. At one of them, the H is above the Cl; at the
other, the Cl is above the H. Trans. Wrong chemical. The structure shown in Fig 9 is not
the requested chemical.

Figure 10
Fig 10 attempts to fix the problem with Fig 9. But it

Why is Fig 10 wrong? After all, it seems to address the criteria presented. It contains
both of the larger atoms (Cl) equatorial, and they are cis as desired. However, in Fig
10, the two axial groups on carbons # 1 and 2 (the two H that are shown) are both
pointing up. This is impossible. In a valid chair, the axial groups alternate up/down as
one goes around the ring; see Figure 2B, above. This follows from the tetrahedral
bonding of C. Adjacent axial groups, as relevant here, must point in opposite directions;
that condition is violated here. That is, Fig 10 is not a valid chair.
Those who find the above point new or surprising should check their
textbook. If possible, look at models of cyclohexane and simple
derivatives such as the one here. Figure 2B, above, is correct. In my
experience, many students have not yet noticed this feature of chair
conformations.
If you think you have an alternative that is better (or even
satisfactory), please show it to me. I suspect it will turn out to be
equivalent to Fig 9 or Fig 10, above. But if you think it is good, let's
discuss it.
So now what? We have a contradiction. And that really is the most important point
here. It is important to realize that we cannot draw a conformation for cis-1,2-
dichlorocyclohexane which easily fits the criteria we have used so far: a chair, with
large groups equatorial. So what do we do? Clearly, the conformation of this compound
must, in some way, involve more issues than what we have considered so far.
Instructors and books will vary in how much further explanation they want to give on
this matter. Therefore, how you proceed from here must take into account the
preferences in your course. Here is one way to proceed.
One simple way to proceed is to re-examine the two criteria we have been using, and
then state a generality about what to do in the event of a conflict. That generality is:
use the chair, and then fit the groups as best you can. That is, try to put as many of the
larger groups equatorial as you can, but realize that you may not get them all
equatorial.

Figure 11
Fig 11 shows a plausible chair conformation of cis-1,2-dichlorocyc

What have we accomplished here? First, this is the correct compound. Convince
yourself that this really is cis-1,2-dichlorocyclohexane. In particular, it is cis because at
each substituted position the Cl is "above" the H; that is, both Cl are on the same side
of the ring.
Second, this is a proper chair. Adjacent axial groups point in opposite directions. Only
two of them are shown here: the axial Cl on the upper right C is "up", and the axial H
on the lower right C is "down". Thus, these aspects satisfy the proper way to orient
things, in general, on a cyclohexane chair.
How good a conformation is the one shown in Fig 11? Actually, it is quite good, in this
case. The actual conformation has been measured, and it follows the basic ideas shown
here.
Is this the end of the story? No, but it is about enough for now. The main purpose here
was to show how one must carefully look at conformation, within the constraints of the
specific isomer one is trying to draw. Some compounds cannot be easily drawn within
the common "rules". cis-1,2-dichlorocyclohexane is one such example. In this case, we
kept the basic chair conformation, but put one larger group in the less favored axial
orientation. Measurements on many such chemicals have shown that the energetic
penalty of moving a cyclohexane ring much away from the basic chair conformation is
quite large -- certainly larger than the energetic penalty of putting one "somewhat
large" group in an axial position. Of course, with larger groups or more groups, this
might not hold.

How to draw stereo bonds ("up" and "down"


bonds)
There are various ways to show these orientations. The solid (dark) "up wedge" I used
is certainly common. Some people use an analogous "down wedge", which is light, to
indicate a down bond; unfortunately, there is no agreement as to which way the wedge
should point, and you are left relying on the lightness of the wedge to know it is
"down". The "down bond" avoids this wedge ambiguity, and just uses some kind of
light line. The down bond I used (e.g., in Figure 5B) is a dashed line; IUPAC encourages
a series of parallel lines, something like . What I did is a variation of what is
recommended by IUPAC:http://www.chem.qmul.ac.uk/iupac/stereo/intro.html.
In ISIS/Draw, the "up wedge" and "down bond" that I used, along
with other variations, are available from a tool button that may be
labeled with any of them, depending on most recent use. It is located
directly below the tool button for ordinary C-C bonds.
In Symyx Draw, the "up wedge" and "down bond", along with other
variations, are available from a tool button that may be labeled with
any of them, depending on most recent use. It is located directly
below the "Chain" tool button.
ChemSketch provides up and down wedges, but not the simple up
and down bonds discussed above. The wedges are available from the
second toolbar across the top. For an expanded discussion of using
these wedges, see the section of my ChemSketch Guide
on Stereochemistry: Wedge bonds.
As always, the information provided on these pages in intended to
help you get started. Each program has more options for drawing
bonds than discussed here. When you feel the need, look around!
How to draw chairs
Most of the structures shown on this page were drawn with the free
program ISIS/Draw. I have posted a guide to help you get started with ISIS/Draw.
ISIS/Draw provides a simple cyclohexane (6-ring) hexagon template on the toolbar
across the top. It provides templates for various 6-ring chair structures from the
Templates menu; choose Rings. There are templates for simple chairs, without
substituents (e.g., Fig 1B), and for chairs showing all the substituents (e.g., Fig 2B). In
either case, you can add, delete, or change things as you wish. Various kinds of stereo
bonds (wedges and bars) are available by clicking the left-side tool button that is just
below the regular C-C single bond button. It may have a wedge shown on it, but this
will vary depending on how it has been used. To choose a type of stereo bond, click on
the button and hold the mouse click; a new menu will appear to the right of the button.
The free drawing program Symyx Draw, the successor to ISIS/Draw, provides similar
templates and tools. A basic chair structure is provided on the default template bar that
is shown. More options are available by choosing the Rings template. See my
page Symyx Draw for a general guide for getting started with this program.
The free drawing program ChemSketch provides similar templates and tools. To find
the special templates for chairs, go to the Templates menu, choose Template
Window, and then choose "Rings" from the drop-down menu near upper left. See my
page ChemSketch for a general guide for getting started with this program.
If you want to draw chair structures by hand (and if you are going on in organic
chemistry, you should)... Be careful. The precise zigs and zags, and the angles of
substituents are all important. Your textbook may offer you some hints for how to draw
chairs. A short item in the Journal of Chemical Education offers a nice trick, showing
how the chair can be thought of as consisting of an M and a W. The article is V
Dragojlovic, A method for drawing the cyclohexane ring and its substituents. J Chem
Educ 78:923, 7/01. (I thank M Farooq Wahab, Chemistry, Univ Karachi, for suggesting
that this article be noted here.)
Aside from drawing the basic chair, the key points in adding substituents are:
 Axial groups alternate up and down, and are shown "vertical".
 Equatorial groups are approximately horizontal, but actually somewhat distorted from
that, so that the angle from the axial group is a bit more than a right angle -- reflecting
the common 109 degree bond angle.
 As cautioned before, it is usually easier to draw and see what is happening at the four
corners of the chair than at the two middle positions. Try to use the corners as much as
possible.

Other ring sizes


The main issues raised here for showing cis and trans hold for other ring sizes.
Specifically, Figure 5 (parts A and B) holds for other ring sizes. The only difference
would be the geometric figure used to show the ring. Discussion of conformation is
more complex, and must be considered for each ring size. The "chair" is a likely
conformation for 6-rings, but not for other sizes. Discussing conformations for other
ring sizes is beyond the scope of this page. 6-rings are among the most common, in
both organic and biochemistry.

Physical Properties of Alkanes


1. Last updated
Jun 5, 2019

2.
o Rings: cis/trans and axial/equatorial relationships

o Straight-Chain and Branched Alkanes


3. picture_as_pdf
LMS

Donate


 Contributed by Jim Clark
 Former Head of Chemistry and Head of Science at Truro School in Cornwall
Alkanes are not very reactive and have little biological activity; all alkanes are colorless and
odorless.

Boiling Points
The boiling points shown are for the "straight chain" isomers of which there is more than one.
The first four alkanes are gases at room temperature, and solids do not begin to appear until
about C17H36C17H36, but this is imprecise because different isomers typically have different
melting and boiling points. By the time you get 17 carbons into an alkane, there are unbelievable
numbers of isomers!
Cycloalkanes have boiling points that are approximately 20 K higher than the corresponding
straight chain alkane.
There is not a significant electronegativity difference between carbon and hydrogen, thus, there
is not any significant bond polarity. The molecules themselves also have very little polarity. A
totally symmetrical molecule like methane is completely non-polar, meaning that the only
attractions between one molecule and its neighbors will be Van der Waalsdispersion forces.
These forces will be very small for a molecule like methane but will increase as the molecules
get bigger. Therefore, the boiling points of the alkanes increase with molecular size.
Where you have isomers, the more branched the chain, the lower the boiling point tends to be.
Van der Waals dispersion forces are smaller for shorter molecules and only operate over very
short distances between one molecule and its neighbors. It is more difficult for short, fat
molecules (with lots of branching) to lie as close together as long, thin molecules.

Example 11: Structure dependent Boiling Points


For example, the boiling points of the three isomers of C5H12C5H12 are:
 pentane: 309.2 K
 2-methylbutane: 301.0 K
 2,2-dimethylpropane: 282.6 K
The slightly higher boiling points for the cycloalkanes are presumably because the molecules can
get closer together because the ring structure makes them tidier and less "wriggly"!

Solubility
Alkanes (both alkanes and cycloalkanes) are virtually insoluble in water, but dissolve in organic
solvents. However, liquid alkanes are good solvents for many other non-ionic organic
compounds.
Solubility in Water
When a molecular substance dissolves in water, the following must occur:
 break the intermolecular forces within the substance. In the case of the alkanes, these are the
Van der Waals dispersion forces.
 break the intermolecular forces in the water so that the substance can fit between the water
molecules. In water, the primary intermolecular attractions are hydrogen bonds.
Breaking either of these attractions requires energy, although the amount of energy to break the
Van der Waals dispersion forces in something like methane is relatively negligible; this is not
true of the hydrogen bonds in water.
As something of a simplification, a substance will dissolve if there is enough energy released
when new bonds are made between the substance and the water to compensate for what is used
in breaking the original attractions. The only new attractions between the alkane and the water
molecules are Van der Waals forces. These forces do not release a sufficient amount of energy to
compensate for the energy required to break the hydrogen bonds in water.; the alkane does not
dissolve.
The energy only description of solvation is an oversimplification because entropic effects are
also important when things dissolve.

Solubility in organic solvents


In most organic solvents, the primary forces of attraction between the solvent molecules are Van
der Waals - either dispersion forces or dipole-dipole attractions. Therefore, when an alkane
dissolves in an organic solvent, the Van der Waals forces are broken and are replaced by new
Van der Waals forces. The two processes more or less cancel each other out energetically; thus,
there is no barrier to solubility.

Corey-House Reaction
1. Last updated
Jun 5, 2019

2.
o Synthesis of Alkanes

o Cracking
3. picture_as_pdf
LMS

Donate
Corey-House Reaction is also called as 'coupling of alkyl halides with organo metallic
compounds'. It is a better method than Wurtz reaction. An alkyl halides and a lithium
dialkyl copper are reacted to give a higher hydrocarbon
R′−X+R2CuLi→R−R′+R−Cu+LiX(1)(1)R′−X+R2CuLi→R−R′+R−Cu+LiX

(R and R' may be same or different)

Wurtz reaction
1. Last updated
Jun 5, 2019

2.
o Cracking

o Reactivity of Alkanes
3. picture_as_pdf
LMS

Donate

Wurtz reaction is coupling of haloalkanes using sodium metal in solvent like dry ether
2R−X+2Na→R−R+2Na+X−(1)(1)2R−X+2Na→R−R+2Na+X−
The reaction consists of a halogen-metal exchange involving the free radical species R• (in a
similar fashion to the formation of a Grignard reagent and then the carbon-carbon bond
formation in a nucleophilic substitution reaction.)
One electron from the metal is transferred to the halogen to produce a metal halide and an alkyl
radical.
R−X+M→R⋅+M+X−(2)(2)R−X+M→R⋅+M+X−
The alkyl radical then accepts an electron from another metal atom to form an alkyl anion and
the metal becomes cationic. This intermediate has been isolated in a several cases.
R⋅+M→R−M+(3)(3)R⋅+M→R−M+
The nucleophilic carbon of the alkyl anion then displaces the halide in an S N2 reaction, forming a
new carbon-carbon covalent bond.
R−M++R−X→R−R+M+X−(4)(4)R−M++R−X→R−R+M+X−
Reactivity of Alkanes
1. Last updated
Jun 23, 2019

2.
o Wurtz reaction

o Alkane Heats of Combustion


3. picture_as_pdf
LMS

Donate

The alkanes and cycloalkanes, with the exception of cyclopropane, are probably the
least chemically reactive class of organic compounds. Alkanes contain strong carbon-
carbon single bonds and strong carbon-hydrogen bonds. The carbon-hydrogen bonds
are only very slightly polar; therefore, there are no portions of the molecules that carry
any significant amount of positive or negative charge that can attract other molecules
or ions. Alkanes can be burned, destroying the entire molecule (Alkane Heats of
Combustion), alkanes can react with some of the halogens, breaking carbon-hydrogen
bonds, and alkanes can crack by breaking the carbon-carbon bonds.

Alkane Heats of Combustion


1. Last updated
Jun 5, 2019

2.
o Reactivity of Alkanes

o Chlorination of Methane and the Radical Chain Mechanism


3. picture_as_pdf
LMS

Donate

 Contributed by William Reusch
 Professor Emeritus (Chemistry) at Michigan State University
The combustion of carbon compounds, especially hydrocarbons, has been the most important
source of heat energy for human civilizations throughout recorded history. The practical
importance of this reaction cannot be denied, but the massive and uncontrolled chemical changes
that take place in combustion make it difficult to deduce mechanistic paths. Using the
combustion of propane as an example, we see from the following equation that every covalent
bond in the reactants has been broken and an entirely new set of covalent bonds have formed in
the products. No other common reaction involves such a profound and pervasive change, and the
mechanism of combustion is so complex that chemists are just beginning to explore and
understand some of its elementary features.
CH3CH2CH3+5O2→3CO2+4H2O+heat(1)(1)CH3CH2CH3+5O2→3CO2+4H2O+heat
Two points concerning this reaction are important:
1. Since all the covalent bonds in the reactant molecules are broken, the quantity of heat evolved
in this reaction is related to the strength of these bonds (and, of course, the strength of the
bonds formed in the products). Precise heats of combustion measurements can provide useful
information about the structure of molecules.
2. The stoichiometry of the reactants is important. If insufficient oxygen is supplied some of the
products will consist of the less oxidized carbon monoxide COCO gas.
CH3CH2CH3+4O2→CO2+2CO+4H2O+heat(2)(2)CH3CH2CH3+4O2→CO2+2CO+4H2O+heat

Heat of Combustion
From the previous discussion, we might expect isomers to have identical heats of combustion.
However, a few simple measurements will disabuse this belief. Thus, the heat of combustion of
pentane is –782 kcal/mole, but that of its 2,2-dimethylpropane (neopentane) isomer is –777
kcal/mole. Differences such as this reflect subtle structural variations, including the greater bond
energy of 1º-C–H versus 2º-C–H bonds and steric crowding of neighboring groups. In small-ring
cyclic compounds ring strain can be a major contributor to thermodynamic stability and chemical
reactivity. The following table lists heat of combustion data for some simple cycloalkanes and
compares these with the increase per CH2unit for long chain alkanes.

Table 11: Heats of combustion of select hydrocarbons


CH2 Units ΔH25º ΔH25º
n kcal/mole per CH2 Unit

n=3 468.7 156.2

n=4 614.3 153.6

n=5 741.5 148.3

n=6 882.1 147.0

n=7 1035.4 147.9

n=8 1186.0 148.2

n=9 1335.0 148.3

n = 10 1481 148.1

m = large — 147.0

The chief source of ring strain in smaller rings is angle strain and eclipsing strain. As noted
elsewhere, cyclopropane and cyclobutane have large contributions of both strains, with angle
strain being especially severe. Changes in chemical reactivity as a consequence of angle strain
are dramatic in the case of cyclopropane, and are also evident for cyclobutane. Some examples
are shown in the following diagram. The cyclopropane reactions are additions, many of which
are initiated by electrophilic attack. The pyrolytic conversion of β-pinene to myrcene probably
takes place by an initial rupture of the 1:6 bond, giving an allylic 3º-diradical, followed
immediately by breaking of the 5:7 bond.
Figure 11: Changes in chemical reactivity as a consequence of ring strain

Complete vs. Incomplete Combustion of


Alkanes
1. Last updated
Jun 5, 2019

2.
o Chlorination of Methane and the Radical Chain Mechanism

o Cracking Alkanes
3. picture_as_pdf
LMS

Donate


 Contributed by Jim Clark
 Former Head of Chemistry and Head of Science at Truro School in Cornwall
This page deals briefly with the combustion of alkanes and cycloalkanes. In fact, there
is very little difference between the two.

Complete combustion
Complete combustion (given sufficient oxygen) of any hydrocarbon produces carbon
dioxide and water. It is quite important that you can write properly balanced equations
for these reactions, because they often come up as a part of thermochemistry
calculations. Some are easier than others. For example, with alkanes, the ones with an
even number of carbon atoms are marginally harder than those with an odd number!

Example 1: Propane Combustion


For example, with propane (C3H8), you can balance the carbons and hydrogens as you
write the equation down. Your first draft would be:
C3H8+O2→3CO2+4H2O(1)(1)C3H8+O2→3CO2+4H2O

Counting the oxygens leads directly to the final version:


C3H8+5O2→3CO2+4H2O(2)(2)C3H8+5O2→3CO2+4H2O

Example 2: Butane Combustion


With butane (C4H10), you can again balance the carbons and hydrogens as you write the
equation down.
C4H10+O2→4CO2+5H2O(3)(3)C4H10+O2→4CO2+5H2O

Counting the oxygens leads to a slight problem - with 13 on the right-hand side. The
simple trick is to allow yourself to have "six-and-a-half" O2 molecules on the left.
C4H10+612O2→4CO2+5H2O(4)(4)C4H10+612O2→4CO2+5H2O

If that offends you, double everything:


2C4H10+13O2→8CO2+10H2O(5)(5)2C4H10+13O2→8CO2+10H2O

The hydrocarbons become harder to ignite as the molecules get bigger. This is because
the bigger molecules don't vaporize so easily - the reaction is much better if the oxygen
and the hydrocarbon are well mixed as gases. If the liquid is not very volatile, only
those molecules on the surface can react with the oxygen. Bigger molecules have
greater Van der Waals attractions which makes it more difficult for them to break away
from their neighbors and turn to a gas.
Provided the combustion is complete, all the hydrocarbons will burn with a blue flame.
However, combustion tends to be less complete as the number of carbon atoms in the
molecules rises. That means that the bigger the hydrocarbon, the more likely you are to
get a yellow, smoky flame.

Incomplete combustion
Incomplete combustion (where there is not enough oxygen present) can lead to the
formation of carbon or carbon monoxide. As a simple way of thinking about it, the
hydrogen in the hydrocarbon gets the first chance at the oxygen, and the carbon gets
whatever is left over! The presence of glowing carbon particles in a flame turns it
yellow, and black carbon is often visible in the smoke. Carbon monoxide is produced as
a colorless poisonous gas.
Why carbon monoxide is poisonous: Oxygen is carried around the blood by
hemoglobin, which unfortunately binds to exactly the same site on the hemoglobin that
oxygen does. The difference is that carbon monoxide binds irreversibly (or very
strongly) - making that particular molecule of hemoglobin useless for carrying oxygen.
If you breath in enough carbon monoxide you will die from a sort of internal
suffocation.

Halogenation Alkanes
1. Last updated
Jun 5, 2019

2.
o Cracking Alkanes

o Halogenation of Alkanes
3. picture_as_pdf
LMS

Donate


 Contributed by Jim Clark
 Former Head of Chemistry and Head of Science at Truro School in Cornwall
This page describes the reactions between alkanes and cycloalkanes with the halogens
fluorine, chlorine, bromine and iodine - mainly concentrating on chlorine and bromine.
 The reaction between alkanes and fluorine: This reaction is explosive even in the
cold and dark, and you tend to get carbon and hydrogen fluoride produced. It is of no
particular interest. For example:
CH4+2F2→C+4HF(1)(1)CH4+2F2→C+4HF
 The reaction between alkanes and iodine: Iodine does not react with the alkanes
to any extent - at least, under normal lab conditions.
 The reactions between alkanes and chlorine or bromine: There is no reaction in
the dark.
Methane and Chlorine
In the presence of a flame, the reactions are rather like the fluorine one - producing a
mixture of carbon and the hydrogen halide. The violence of the reaction drops
considerably as you go from fluorine to chlorine to bromine. The interesting reactions
happen in the presence of ultra-violet light (sunlight will do). These are photochemical
reactions that happen at room temperature. We'll look at the reactions with chlorine,
although the reactions with bromine are similar, but evolve more slowly.
Substitution reactions happen in which hydrogen atoms in the methane are replaced
one at a time by chlorine atoms. You end up with a mixture of chloromethane,
dichloromethane, trichloromethane and tetrachloromethane.

The original mixture of a colorless and a green gas would produce steamy fumes of
hydrogen chloride and a mist of organic liquids. All of the organic products are liquid at
room temperature with the exception of the chloromethane which is a gas.
If you were using bromine, you could either mix methane with bromine vapor , or
bubble the methane through liquid bromine - in either case, exposed to UV light. The
original mixture of gases would, of course, be red-brown rather than green. One would
not choose to use these reactions as a means of preparing these organic compounds in
the lab because the mixture of products would be too tedious to separate. The
mechanisms for the reactions are explained on separate pages.

Larger alkanes and chlorine


You would again get a mixture of substitution products, but it is worth just looking
briefly at what happens if only one of the hydrogen atoms gets substituted
(monosubstitution) - just to show that things aren't always as straightforward as they
seem! For example, with propane, you could get one of two isomers:
If chance was the only factor, you would expect to get three times as much of the
isomer with the chlorine on the end. There are 6 hydrogens that could get replaced on
the end carbon atoms compared with only 2 in the middle. In fact, you get about the
same amount of each of the two isomers. If you use bromine instead of chlorine, the
great majority of the product is where the bromine is attached to the center carbon
atom.

Cycloalkanes
The reactions of the cycloalkanes are generally just the same as the alkanes, with the
exception of the very small ones - particularly cyclopropane. In the presence of UV
light, cyclopropane will undergo substitution reactions with chlorine or bromine just like
a non-cyclic alkane. However, it also has the ability to react in the dark. In the absence
of UV light, cyclopropane can undergo addition reactions in which the ring is broken.
For example, with bromine, cyclopropane gives 1,3-dibromopropane.

This can still happen in the presence of light - but you will get substitution reactions as
well. The ring is broken because cyclopropane suffers badly from ring strain. The bond
angles in the ring are 60° rather than the normal value of about 109.5° when the
carbon makes four single bonds. The overlap between the atomic orbitals in forming
the carbon-carbon bonds is less good than it is normally, and there is considerable
repulsion between the bonding pairs. The system becomes more stable if the ring is
broken.

Chlorination of Methane and the Radical


Chain Mechanism
1. Last updated
Jun 5, 2019

2.
o Alkane Heats of Combustion
o Complete vs. Incomplete Combustion of Alkanes
3. picture_as_pdf
LMS

Donate

Alkanes (the most basic of all organic compounds) undergo very few reactions. One of these
reactions is halogenation, or the substitution of a single hydrogen on the alkane for a single
halogen to form a haloalkane. This reaction is very important in organic chemistry because it
opens a gateway to further chemical reactions.

Introduction
While the reactions possible with alkanes are few, there are many reactions that
involve haloalkanes. In order to better understand the mechanism (a detailed look at the step by
step process through which a reaction occurs), we will closely examine the chlorination of
methane. When methane (CH4) and chlorine (Cl2) are mixed together in the absence of light at
room temperature nothing happens. However, if the conditions are changed, so that either the
reaction is taking place at high temperatures (denoted by Δ) or there is ultra violet irradiation, a
product is formed, chloromethane (CH3Cl).

Energetics
Why does this reaction occur? Is the reaction favorable? A way to answer these questions is to
look at the change in enthalpy (ΔHΔH) that occurs when the reaction takes place.
ΔH = (Energy put into reaction) – (Energy given off from reaction)
If more energy is put into a reaction than is given off, the ΔH is positive, the reaction is
endothermic and not energetically favorable. If more energy is given off in the reaction than was
put in, the ΔH is negative, the reaction is said to be exothermic and is considered favorable. The
figure below illustrates the difference between endothermic and exothermic reactions.
ΔH can also be calculated using bond dissociation energies (ΔH°):
ΔH=∑ΔH∘ of bonds broken−∑ΔH∘ of bonds formed(1)(1)ΔH=∑ΔH° of bonds broken−∑ΔH° of
bonds formed

Let’s look at our specific example of the chlorination of methane to determine if it is


endothermic or exothermic:

Since, the ΔH for the chlorination of methane is negative, the reaction is exothermic.
Energetically this reaction is favorable. In order to better understand this reaction we need to
look at the mechanism ( a detailed step by step look at the reaction showing how it occurs) by
which the reaction occurs.

Radical Chain Mechanism


The reaction proceeds through the radical chain mechanism. The radical chain mechanism is
characterized by three steps: initiation, propagation and termination. Initiation requires an
input of energy but after that the reaction is self-sustaining. The first propagation step uses up
one of the products from initiation, and the second propagation step makes another one, thus the
cycle can continue until indefinitely.

Step 1: Initiation
Initiation breaks the bond between the chlorine molecule (Cl2). For this step to occur energy must
be put in, this step is not energetically favorable. After this step, the reaction can occur
continuously (as long as reactants provide) without input of more energy. It is important to note
that this part of the mechanism cannot occur without some external energy input, through light or
heat.

Step 2: Propagation
The next two steps in the mechanism are called propagation steps. In the first propagation step, a
chlorine radical combines with a hydrogen on the methane. This gives hydrochloric acid (HCl,
the inorganic product of this reaction) and the methyl radical. In the second propagation step
more of the chlorine starting material (Cl2) is used, one of the chlorine atoms becomes a radical
and the other combines with the methyl radical.

The first propagation step is endothermic, meaning it takes in heat (requires 2 kcal/mol) and is
not energetically favorable. In contrast the second propagation step is exothermic, releasing 27
kcal/mol. Since the second propagation step is so exothermic, it occurs very quickly. The second
propagation step uses up a product from the first propagation step (the methyl radical) and
following Le Chatelier's principle, when the product of the first step is removed the equilibrium
is shifted towards it's products. This principle is what governs the unfavorable first propagation
step's occurance.
Step 3: Termination
In the termination steps, all the remaining radicals combine (in all possible manners) to form
more product (CH3Cl), more reactant (Cl2) and even combinations of the two methyl radicals to
form a side product of ethane (CH3CH3).
Problems with the Chlorination of Methane
The chlorination of methane does not necessarily stop after one chlorination. It may actually be
very hard to get a monosubstituted chloromethane. Instead di-, tri- and even tetra-
chloromethanes are formed. One way to avoid this problem is to use a much higher concentration
of methane in comparison to chloride. This reduces the chance of a chlorine radical running into
a chloromethane and starting the mechanism over again to form a dichloromethane. Through this
method of controlling product ratios one is able to have a relative amount of control over the
product.

References
1. Matyjaszewski, Krzysztof, Wojciech Jakubowski, Ke Min, Wei Tang, Jinyu Huang, Wade A.
Braunecker, and Nicolay V. Tsarevsky. "Diminishing Catalyst Concentration in Atom Transfer
Radical Polymerization with Reducing Agents." Proceedings of the National Academy of Sciences
of the United States of America 103 (2006): 15309-5314.
2. Morgan, G. T. "A State Experiment in Chemical Research." Science 72 (1930): 379-90.
3. Phillips, Francis C. "# Researches upon the Chemical Properties of Gases." Researches upon the
Chemical Properties of Gases 17 (1893): 149-236.

Outside Links
 Video of Mechanism: http://www.jbpub.com/organic-online/...s/chlormet.htm
 Wikipedia of Radical Chain Mechanism: http://en.wikipedia.org/wiki/Free_ra...l_halogenation
 Wikipedia of Le Chatelier's Principle: http://en.wikipedia.org/wiki/Le_Chat...#Concentration

Problems
Answers to these questions are in an attached slide
1. Write out the complete mechanism for the chlorination of methane.
2. Explain, in your own words, how the first propagation step can occur without input of energy if
it is energetically unfavorable.
3. Compounds other than chlorine and methane go through halogenation with the radical chain
mechanism. Write out a generalized equation for the halogenation of RH with X2including all the
different steps of the mechanism.
4. Which step of the radical chain mechanism requires outside energy? What can be used as this
energy?
5. Having learned how to calculate the change in enthalpy for the chlorination of methane apply
your knowledge and using the table provided below calculate the change in enthalpy for the
bromination of ethane.
Compound Bond Dissociation Energy
(kcal/mol)

CH3CH2-H 101

CH3CH2-Br 70

H-Br 87

Br2 46

Chlorination of Methane and the Radical


Chain Mechanism
1. Last updated
Jun 5, 2019

2.
o Alkane Heats of Combustion

o Complete vs. Incomplete Combustion of Alkanes


3. picture_as_pdf
LMS

Donate

Alkanes (the most basic of all organic compounds) undergo very few reactions. One of these
reactions is halogenation, or the substitution of a single hydrogen on the alkane for a single
halogen to form a haloalkane. This reaction is very important in organic chemistry because it
opens a gateway to further chemical reactions.

Introduction
While the reactions possible with alkanes are few, there are many reactions that
involve haloalkanes. In order to better understand the mechanism (a detailed look at the step by
step process through which a reaction occurs), we will closely examine the chlorination of
methane. When methane (CH4) and chlorine (Cl2) are mixed together in the absence of light at
room temperature nothing happens. However, if the conditions are changed, so that either the
reaction is taking place at high temperatures (denoted by Δ) or there is ultra violet irradiation, a
product is formed, chloromethane (CH3Cl).

Energetics
Why does this reaction occur? Is the reaction favorable? A way to answer these questions is to
look at the change in enthalpy (ΔHΔH) that occurs when the reaction takes place.
ΔH = (Energy put into reaction) – (Energy given off from reaction)
If more energy is put into a reaction than is given off, the ΔH is positive, the reaction is
endothermic and not energetically favorable. If more energy is given off in the reaction than was
put in, the ΔH is negative, the reaction is said to be exothermic and is considered favorable. The
figure below illustrates the difference between endothermic and exothermic reactions.

ΔH can also be calculated using bond dissociation energies (ΔH°):


ΔH=∑ΔH∘ of bonds broken−∑ΔH∘ of bonds formed(1)(1)ΔH=∑ΔH° of bonds broken−∑ΔH° of
bonds formed

Let’s look at our specific example of the chlorination of methane to determine if it is


endothermic or exothermic:
Since, the ΔH for the chlorination of methane is negative, the reaction is exothermic.
Energetically this reaction is favorable. In order to better understand this reaction we need to
look at the mechanism ( a detailed step by step look at the reaction showing how it occurs) by
which the reaction occurs.

Radical Chain Mechanism


The reaction proceeds through the radical chain mechanism. The radical chain mechanism is
characterized by three steps: initiation, propagation and termination. Initiation requires an
input of energy but after that the reaction is self-sustaining. The first propagation step uses up
one of the products from initiation, and the second propagation step makes another one, thus the
cycle can continue until indefinitely.

Step 1: Initiation
Initiation breaks the bond between the chlorine molecule (Cl2). For this step to occur energy must
be put in, this step is not energetically favorable. After this step, the reaction can occur
continuously (as long as reactants provide) without input of more energy. It is important to note
that this part of the mechanism cannot occur without some external energy input, through light or
heat.

Step 2: Propagation
The next two steps in the mechanism are called propagation steps. In the first propagation step, a
chlorine radical combines with a hydrogen on the methane. This gives hydrochloric acid (HCl,
the inorganic product of this reaction) and the methyl radical. In the second propagation step
more of the chlorine starting material (Cl2) is used, one of the chlorine atoms becomes a radical
and the other combines with the methyl radical.
The first propagation step is endothermic, meaning it takes in heat (requires 2 kcal/mol) and is
not energetically favorable. In contrast the second propagation step is exothermic, releasing 27
kcal/mol. Since the second propagation step is so exothermic, it occurs very quickly. The second
propagation step uses up a product from the first propagation step (the methyl radical) and
following Le Chatelier's principle, when the product of the first step is removed the equilibrium
is shifted towards it's products. This principle is what governs the unfavorable first propagation
step's occurance.

Step 3: Termination
In the termination steps, all the remaining radicals combine (in all possible manners) to form
more product (CH3Cl), more reactant (Cl2) and even combinations of the two methyl radicals to
form a side product of ethane (CH3CH3).
Problems with the Chlorination of Methane
The chlorination of methane does not necessarily stop after one chlorination. It may actually be
very hard to get a monosubstituted chloromethane. Instead di-, tri- and even tetra-
chloromethanes are formed. One way to avoid this problem is to use a much higher concentration
of methane in comparison to chloride. This reduces the chance of a chlorine radical running into
a chloromethane and starting the mechanism over again to form a dichloromethane. Through this
method of controlling product ratios one is able to have a relative amount of control over the
product.

References
1. Matyjaszewski, Krzysztof, Wojciech Jakubowski, Ke Min, Wei Tang, Jinyu Huang, Wade A.
Braunecker, and Nicolay V. Tsarevsky. "Diminishing Catalyst Concentration in Atom Transfer
Radical Polymerization with Reducing Agents." Proceedings of the National Academy of Sciences
of the United States of America 103 (2006): 15309-5314.
2. Morgan, G. T. "A State Experiment in Chemical Research." Science 72 (1930): 379-90.
3. Phillips, Francis C. "# Researches upon the Chemical Properties of Gases." Researches upon the
Chemical Properties of Gases 17 (1893): 149-236.

Outside Links
 Video of Mechanism: http://www.jbpub.com/organic-online/...s/chlormet.htm
 Wikipedia of Radical Chain Mechanism: http://en.wikipedia.org/wiki/Free_ra...l_halogenation
 Wikipedia of Le Chatelier's Principle: http://en.wikipedia.org/wiki/Le_Chat...#Concentration

Problems
Answers to these questions are in an attached slide
1. Write out the complete mechanism for the chlorination of methane.
2. Explain, in your own words, how the first propagation step can occur without input of energy if
it is energetically unfavorable.
3. Compounds other than chlorine and methane go through halogenation with the radical chain
mechanism. Write out a generalized equation for the halogenation of RH with X2including all the
different steps of the mechanism.
4. Which step of the radical chain mechanism requires outside energy? What can be used as this
energy?
5. Having learned how to calculate the change in enthalpy for the chlorination of methane apply
your knowledge and using the table provided below calculate the change in enthalpy for the
bromination of ethane.

Compound Bond Dissociation Energy


(kcal/mol)

CH3CH2-H 101

CH3CH2-Br 70

H-Br 87

Br2 46

Cracking Alkanes
1. Last updated
Jun 5, 2019

2.
o Complete vs. Incomplete Combustion of Alkanes

o Halogenation Alkanes
3. picture_as_pdf
LMS

Donate


 Contributed by Jim Clark
 Former Head of Chemistry and Head of Science at Truro School in Cornwall
What is cracking? Cracking is the name given to breaking up large hydrocarbon molecules into
smaller and more useful bits. This is achieved by using high pressures and temperatures without
a catalyst, or lower temperatures and pressures in the presence of a catalyst. The source of the
large hydrocarbon molecules is often the naphtha fraction or the gas oil fraction from the
fractional distillation of crude oil (petroleum). These fractions are obtained from the distillation
process as liquids, but are re-vaporized before cracking. There isn't any single unique reaction
happening in the cracker. The hydrocarbon molecules are broken up in a fairly random way to
produce mixtures of smaller hydrocarbons, some of which have carbon-carbon double bonds.
One possible reaction involving the hydrocarbon C15H32C15H32 might be:
C15H32→2C2H4ethene+C3H6propene+C8H18octane(1)(1)C15H32→2C2H4ethene+C3H6propene+C8
H18octane

Or, showing more clearly what happens to the various atoms and bonds:
This is only one way in which this particular molecule might break up. The ethene and propene
are important materials for making plastics or producing other organic chemicals. The octane is
one of the molecules found in petrol (gasoline).

Catalytic Cracking
Modern cracking uses zeolites as the catalyst. These are complex aluminosilicates, and are large
lattices of aluminium, silicon and oxygen atoms carrying a negative charge. They are, of course,
associated with positive ions such as sodium ions. You may have come across a zeolite if you
know about ion exchange resins used in water softeners. The alkane is brought into contact with
the catalyst at a temperature of about 500 °C and moderately low pressures.

The zeolites used in catalytic cracking are chosen to give high percentages of hydrocarbons with
between 5 and 10 carbon atoms - particularly useful for petrol (gasoline). It also produces high
proportions of branched alkanes and aromatic hydrocarbons like benzene. The zeolite catalyst
has sites which can remove a hydrogen from an alkane together with the two electrons which
bound it to the carbon. That leaves the carbon atom with a positive charge. Ions like this are
called carbonium ions (or carbocations). Reorganisation of these leads to the various products of
the reaction.

Thermal Cracking
In thermal cracking, high temperatures (typically in the range of 450 °C to 750 °C) and pressures
(up to about 70 atmospheres) are used to break the large hydrocarbons into smaller ones.
Thermal cracking gives mixtures of products containing high proportions of hydrocarbons with
double bonds - alkenes. This is a gross oversimplification; tn fact, there are several versions of
thermal cracking designed to produce different mixtures of products. These use completely
different sets of conditions.
Thermal cracking does not go via ionic intermediates like catalytic cracking. Instead, carbon-
carbon bonds are broken so that each carbon atom ends up with a single electron. In other words,
free radicals are formed.
Reactions of the free radicals lead to the various products.

Halogenation of Alkanes
1. Last updated
Jun 5, 2019

2.
o Halogenation Alkanes

o Alkenes
3. picture_as_pdf
LMS

Donate


 Contributed by William Reusch
 Professor Emeritus (Chemistry) at Michigan State University

Chlorination of Methane by Substitution


Halogenation is the replacement of one or more hydrogen atoms in an organic compound by a
halogen (fluorine, chlorine, bromine or iodine). Unlike the complex transformations of
combustion, the halogenation of an alkane appears to be a simple substitution reaction in which
a C-H bond is broken and a new C-X bond is formed. The chlorination of methane, shown
below, provides a simple example of this reaction.
CH4 + Cl2 + energy → CH3Cl + HCl
Since only two covalent bonds are broken (C-H & Cl-Cl) and two covalent bonds are formed (C-
Cl & H-Cl), this reaction seems to be an ideal case for mechanistic investigation and speculation.
However, one complication is that all the hydrogen atoms of an alkane may undergo substitution,
resulting in a mixture of products, as shown in the following unbalanced equation. The relative
amounts of the various products depend on the proportion of the two reactants used. In the case
of methane, a large excess of the hydrocarbon favors formation of methyl chloride as the chief
product; whereas, an excess of chlorine favors formation of chloroform and carbon tetrachloride.
CH4 + Cl2 + energy → CH3Cl + CH2Cl2 + CHCl3 + CCl4 + HCl

Empirical Considerations
The following facts must be accomodated by any reasonable mechanism for the halogenation
reaction.
1. The reactivity of the halogens decreases in the following order: F2 > Cl2 > Br2 > I2.
2. We shall confine our attention to chlorine and bromine, since fluorine is so explosively reactive
it is difficult to control, and iodine is generally unreactive.
3. Chlorinations and brominations are normally exothermic.
4. Energy input in the form of heat or light is necessary to initiate these halogenations.
5. If light is used to initiate halogenation, thousands of molecules react for each photon of light
absorbed.
6. Halogenation reactions may be conducted in either the gaseous or liquid phase.
7. In gas phase chlorinations the presence of oxygen (a radical trap) inhibits the reaction.
8. In liquid phase halogenations radical initiators such as peroxides facilitate the reaction.
The most plausible mechanism for halogenation is a chain reaction involving neutral
intermediates such as free radicals or atoms. The weakest covalent bond in the reactants is the
halogen-halogen bond (Cl-Cl = 58 kcal/mole; Br-Br = 46 kcal/mole) so the initiating step is the
homolytic cleavage of this bond by heat or light, note that chlorine and bromine both absorb
visible light (they are colored). A chain reaction mechanism for the chlorination of methane has
been described. Bromination of alkanes occurs by a similar mechanism, but is slower and more
selective because a bromine atom is a less reactive hydrogen abstraction agent than a chlorine
atom, as reflected by the higher bond energy of H-Cl than H-Br.

Selectivity
When alkanes larger than ethane are halogenated, isomeric products are formed. Thus
chlorination of propane gives both 1-chloropropane and 2-chloropropane as mono-
chlorinated products. Four constitutionally isomeric dichlorinated products are possible,
and five constitutional isomers exist for the trichlorinated propanes. Can you write
structural formulas for the four dichlorinated isomers?
CH3CH2CH3+2Cl2→FourC3H6Cl2isomers+2HCl(1)(1)CH3CH2CH3+2Cl2→FourC3H6Cl2isom
ers+2HCl

The halogenation of propane discloses an interesting feature of these reactions. All the
hydrogens in a complex alkane do not exhibit equal reactivity. For example,
propane has eight hydrogens, six of them being structurally equivalent primary, and
the other two being secondary. If all these hydrogen atoms were equally reactive,
halogenation should give a 3:1 ratio of 1-halopropane to 2-halopropane mono-
halogenated products, reflecting the primary/secondary numbers. This is not what we
observe. Light-induced gas phase chlorination at 25 ºC gives 45% 1-chloropropane and
55% 2-chloropropane.
CH3-CH2-CH3 + Cl2 → 45% CH3-CH2-CH2Cl + 55% CH3-CHCl-CH3
The results of bromination ( light-induced at 25 ºC ) are even more suprising, with 2-
bromopropane accounting for 97% of the mono-bromo product.
CH3-CH2-CH3 + Br2 → 3% CH3-CH2-CH2Br + 97% CH3-CHBr-CH3
These results suggest strongly that 2º-hydrogens are inherently more reactive than 1º-
hydrogens, by a factor of about 3:1. Further experiments showed that 3º-hydrogens
are even more reactive toward halogen atoms. Thus, light-induced chlorination of 2-
methylpropane gave predominantly (65%) 2-chloro-2-methylpropane, the substitution
product of the sole 3º-hydrogen, despite the presence of nine 1º-hydrogens in the
molecule.
(CH3)3CH + Cl2 → 65% (CH3)3CCl + 35% (CH3)2CHCH2Cl
It should be clear from a review of the two steps that make up the free radical chain
reaction for halogenation that the first step (hydrogen abstraction) is the product
determining step. Once a carbon radical is formed, subsequent bonding to a halogen
atom (in the second step) can only occur at the radical site. Consequently, an
understanding of the preference for substitution at 2º and 3º-carbon atoms must come
from an analysis of this first step.
First Step: R3CH + X· → R3C· + H-X
Second Step: R3C· + X2 → R3CX + X·
Since the H-X product is common to all possible reactions, differences in reactivity can
only be attributed to differences in C-H bond dissociation energies. In our previous
discussion of bond energy we assumed average values for all bonds of a given kind, but
now we see that this is not strictly true. In the case of carbon-hydrogen bonds, there
are significant differences, and the specific dissociation energies (energy required to
break a bond homolytically) for various kinds of C-H bonds have been measured. These
values are given in the following table.
methyl ethyl i-propyl t-butyl phenyl benzy

103 98 95 93 110 85

The difference in C-H bond dissociation energy reported for primary (1º), secondary
(2º) and tertiary (3º) sites agrees with the halogenation observations reported above,
in that we would expect weaker bonds to be broken more easily than are strong bonds.
By this reasoning we would expect benzylic and allylic sites to be exceptionally reactive
in free radical halogenation, as experiments have shown. The methyl group of toluene,
C6H5CH3, is readily chlorinated or brominated in the presence of free radical initiators
(usually peroxides), and ethylbenzene is similarly chlorinated at the benzylic location
exclusively. The hydrogens bonded to the aromatic ring (referred to as phenyl
hydrogens above) have relatively high bond dissociation energies and are not
substituted.
C6H5CH2CH3 + Cl2 → C6H5CHClCH3 + HCl

Since carbon-carbon double bonds add chlorine and bromine rapidly in liquid phase
solutions, free radical substitution reactions of alkenes by these halogens must be
carried out in the gas phase, or by other halogenating reagents. One such reagent is N-
bromosuccinimide (NBS), shown in the second equation below. By using NBS as a
brominating agent, allylic brominations are readily achieved in the liquid phase.

The covalent bond homolyses that define the bond dissociation energies listed above
may are described by the general equation:
R3C-H + energy → R3C· + H·

C-H bond dissociation energies are commonly interpreted in terms of radical


stability. However, this interpretation has been questioned by Gronert. Most
importantly, in terms of the selectivity of free radical reactions, it is the energies of the
bonds that matter, and not why they are what they are.

Complete vs. Incomplete Combustion of


Alkanes
1. Last updated
Jun 5, 2019

2.
o Chlorination of Methane and the Radical Chain Mechanism

o Cracking Alkanes
3. picture_as_pdf
LMS

Donate


 Contributed by Jim Clark
 Former Head of Chemistry and Head of Science at Truro School in Cornwall
This page deals briefly with the combustion of alkanes and cycloalkanes. In fact, there
is very little difference between the two.

Complete combustion
Complete combustion (given sufficient oxygen) of any hydrocarbon produces carbon
dioxide and water. It is quite important that you can write properly balanced equations
for these reactions, because they often come up as a part of thermochemistry
calculations. Some are easier than others. For example, with alkanes, the ones with an
even number of carbon atoms are marginally harder than those with an odd number!

Example 1: Propane Combustion


For example, with propane (C3H8), you can balance the carbons and hydrogens as you
write the equation down. Your first draft would be:
C3H8+O2→3CO2+4H2O(1)(1)C3H8+O2→3CO2+4H2O

Counting the oxygens leads directly to the final version:


C3H8+5O2→3CO2+4H2O(2)(2)C3H8+5O2→3CO2+4H2O

Example 2: Butane Combustion


With butane (C4H10), you can again balance the carbons and hydrogens as you write the
equation down.
C4H10+O2→4CO2+5H2O(3)(3)C4H10+O2→4CO2+5H2O

Counting the oxygens leads to a slight problem - with 13 on the right-hand side. The
simple trick is to allow yourself to have "six-and-a-half" O2 molecules on the left.
C4H10+612O2→4CO2+5H2O(4)(4)C4H10+612O2→4CO2+5H2O

If that offends you, double everything:


2C4H10+13O2→8CO2+10H2O(5)(5)2C4H10+13O2→8CO2+10H2O

The hydrocarbons become harder to ignite as the molecules get bigger. This is because
the bigger molecules don't vaporize so easily - the reaction is much better if the oxygen
and the hydrocarbon are well mixed as gases. If the liquid is not very volatile, only
those molecules on the surface can react with the oxygen. Bigger molecules have
greater Van der Waals attractions which makes it more difficult for them to break away
from their neighbors and turn to a gas.
Provided the combustion is complete, all the hydrocarbons will burn with a blue flame.
However, combustion tends to be less complete as the number of carbon atoms in the
molecules rises. That means that the bigger the hydrocarbon, the more likely you are to
get a yellow, smoky flame.

Incomplete combustion
Incomplete combustion (where there is not enough oxygen present) can lead to the
formation of carbon or carbon monoxide. As a simple way of thinking about it, the
hydrogen in the hydrocarbon gets the first chance at the oxygen, and the carbon gets
whatever is left over! The presence of glowing carbon particles in a flame turns it
yellow, and black carbon is often visible in the smoke. Carbon monoxide is produced as
a colorless poisonous gas.
Why carbon monoxide is poisonous: Oxygen is carried around the blood by
hemoglobin, which unfortunately binds to exactly the same site on the hemoglobin that
oxygen does. The difference is that carbon monoxide binds irreversibly (or very
strongly) - making that particular molecule of hemoglobin useless for carrying oxygen.
If you breath in enough carbon monoxide you will die from a sort of internal
suffocation.

Potrebbero piacerti anche