Sei sulla pagina 1di 44

Basic Principles of Organic Chemistry

Next: Structural Isomerism >>


Classification of Organic Compounds
a) Acyclic or open chain compounds: These contain alkanes, alkenes,
alkynes and their derivatives. These are also called aliphatic compounds.
b) Cyclic or closed chain compounds:
• Cyclic compounds whose rings are made up of only one kind of
atoms, i.e. carbon atoms are called homocyclic or cabocyclic
compounds. Aliphatic cyclic compounds are called alicyclic
compounds eg cyclopropane, cyclobutane etc

• Organic compounds containing one or more fused or isolated


benzene rings and their functinalized derivatives are called
benzenoids or aromatic compounds, eg benzene, toluene,
naphthalene, anthracene etc.

• Cyclic compounds containing one or more heteroatoms (usually O,N,


S etc) are called hetrocyclic compounds eg ethylene oxide,
tetrahydrofuran (THF), furan, pyrole etc

IUPAC Nomenclature of Organic Compounds:


a) Nomenclature of saturated hydrocarbons:
• Select the longest continuous chain of carbon atoms in the molecule.
The compound is named as a derivative of this alkane
• Number the carbon atoms in the parent chain starting from the end
which gives lowest possible sum for the numbers of the carbon atoms
carrying the substituents
• That set of locants is preferred, which when compared term by term
with other set of locants, each in order of increasing magnitude, has
the lowest term at the first point of difference. For example the set of
locants (2,7,8) is preferred over the set of locants (3,4,9) since 2
comes before 3 even though the sum of locants in the former case is
17 while in the latter case, it is 16

• The correct name is: 2,7,8 - Trimethyldecane and not 3,4,9 -


Trimethyldecane
b) Nomenclature of compounds containing functional group or
multiple bonds:
• Select the longest continuous chain containing the carbon atoms
having the functional group or those involved in the multiple bonds
• The numbering of atoms in the parent chain is done in such a way
that the carbon atom bearing the functional group or those carrying
the multiple bond gets the lowest possible number
• While writing the name of alkene (double bond) or alkyne (triple
bond), the primary suffix 'ane' of the corresponding alkane is replaced
by 'ene' and 'yne' respectively. However, if the multiple bond occurs
twice or thrice in the parent chain, the prefix di- or tri- is attached to
the primary suffix ene or yne
• In naming the organic compounds containing one functional group a
suffix known as secondary suffix is added to the primary suffix (giving
number of carbon atoms in the chain) to indicate the nature of the
functional group. A few important secondary suffixes are:

Functional Secondary Functional group Secondary


group suffix suffix

Alcohols (-OH) -ol Aldehydes (-CHO) -al

Carboxylic acids
Ketones (>C=O) -one -oic acid
(-COOH)

Acid amides
Amines (-NH2 ) -amine -amide
(-CONH2 )

Acid chlorides
-oyl chloride Esters (-COOR) -oate
(-COCL)

Nitrites (-C=N) -nitrite Thioalcohols (-SH) -thiol

Nomenclature of compounds having polyfunctional groups:


When an organic compound contains two or more functional groups, one
group is called the principal functional group while the others are called the
secondary functional groups and are treated as substituents: The order of
preference for principal group is: Carboxylic acid > acid anhydrides > esters
> acid halides > amides > nitrites > aldehydes > ketone > alcohols >
amines > double bond > triple bond
When the functional groups act as substituents, they ar named as:

Functional Functional
Prefix Prefix
group group

- COOH Carboxy -CHO Formyl

Alkoxy cabonyl or Oxo or


-COOR >CO
Carbalkoxy Kelo

-COCL Chloroformyl -OH Hydroxy

-CONH2 Carbamoyl -SH Mecaplo

-CN Cyano -NH2 Amino

-OR Akoxy =NH Imino


-X Halo -NO2 Nitro

Nomenclature of simple aromatic compounds:


a) Nuclear substituted: In these the functional group is directly attached to
the benzene ring. Most of these compounds are better known by their
common and historical names. In the IUPAC system, they are named as
derivatives of benzene.
b) Side chain substituted: In these the functional group is present in the
side chain of the benzene ring. Both in the common and IUPAC systems,
these are usually named as phenyl derivatives of the corresponding
aliphatic compounds.

Sturctural Isomerism
Componds having the same molecular formula but different
structures:
• Compounds having the same molecular formula but different
arrangement of carbon chains are called chain or nuclear isomers.
For example (i) pentane, isopentane and neopentane (ii) hexane, 2-
methulpentane, 3-methylpentane, 2,3-dimethylbutane and 2,2-
dimethulbutane etc
• Compounds having the same structure of the carbon chain but
differing in position of multiple bonds, functional groups or
substituents are called position isomers. For example, (i) but-1-ene
and but-2-ene (ii) prpan-1-ol and propan-2-ol (iii) o-,m- and p- xylenes
etc.
• Compounds having the same molecular formula but different
functional groups are calledfunctional isomers. For example, (i)
alcohols and ethers (ii) aldehydes and ketones (iii) alkynes and
dienes (iv) various amines (1o , 2o , 3o ) (v) cyanides and isocyanides
etc.
• Compounds having the same molecular formula but different number
of carbon atoms on either side of the functional group (O,S or NH)
are caled metamers. For example, (i) diethyl ether and methyl n-
propyl ether (or isopropyl methyl ether). 2-pentanone and 3-
pentanone are position isomers as well as metamers.
• Isomers obtained by 1,3-migration of a proton from one polyvalent
atom to the other within the same molecule are called tautomers. If
one tautomer contains a keto group and the other the enol group,
such a tautomerism is called keto-enol tautomerism. For example,
acetaldehyde and vinyl alcohol.
• The percentage of enol form is negligible in simple aldehydes and
ketones, ie, acetone, acetaldehyde etc. The percentage of enol form,
however, increases, if the enol form is stabilized by intramolecular H-
bonding. For example, the percentage of enol form in acetoacetic
ester is 7% while that in acetylacetone is 76%.

• Only those aldehydes and ketones which contain one or more -


hydrogens show keto-enol tautomerism.
• Primary and secondary nitroalkanes also show tautomerism but
tertiary nitroalkanes and nitroarenes do not.
• Compounds having the same molecular formula but possessing open
chain and cyclic structures are called ring chain isomers. For
example, propene and cyclopropane.

Stereoisomerism
Isomers which have the same structures but differ in the relative
arrangements of atoms or groups in space are called stereoisomers. It is of
three types, i.e. conformational isomerism, optical isomerism and
geometrical isomerism.
Geometrical Isomers: Stereoisomers which have the same structural
formula but differ in the spatial arrangement of atoms around the double
bond are called geometrical isomers. The isomer in which the two similar
atoms/groups lie on the same side of the double bond is called the cis-
isomer while that isomer in which the two similar atoms/groups lie on the
opposite side of the double bond is called the trans-isomer.
• The necessary and sufficient condition for a molecule to exhibit
geometrical isomerism is that each ot the two carbon atoms of the
double bond must have different atoms/groups which may be same of
different. Gor example , alkenes of the types abC=Cab and abC=Cde
show geometrical isomerism
• Compounds containing C=N and N=N also show geometrical
isomerism
• 1,2-1,3 and 1,4-disubstituted cycloalkanes also show geometrical
isomerism
• The dipole moment of a cis-isomer is higher than that of a trans-
isomer
• E and Z configuration of geometrical isomers
○ Give priority to the four groups attached to the double bonded
C-atoms. Higher the atomic number of the atom of the group
attached to the double bonded carbon atom higher will be its
priority. If priority cannot be decided the atoms directly attached
to double bonded C-atoms then next atoms of the group are
compared
○ If two higher priority groups are on the same side of the double
bond then it is called a Z-isomer otherwise it is called an E-
isomer
Conformational Isomerism: The infinite number of momentary
arrangement of the atoms in space which result through rotation about a
single bond are called conformations or rotational isomers
• Ethane has infinite number of conformations, of which, only two, i.e.
staggered and eclipsed are important. The staggered conformation of
ethane is more stable than the eclipsed conformation by about 3.0
kcal or 12.55 kJ mol-1
• Important conformations of n-butane are staggered (or anti), gauche
(or skew), partially eclipsed and fully eclipsed. Their relative stability
is in the order:staggered>gauche>partially eclipsed>fully eclipsed
• Cyclohexane exists in two non-planer conformations, i.e., chair and
the boat form. Both these conformations are free from angle strain
(i.e., have tetrahedral angles)
• The boat form of cyclohexane is less stable than the chair form by
about 44 kJ mol-1
• The two conformations of cyclohexane cannot be separated but ar
readily interconvertible

BASIC CONCEPTS OF CHEMISTRY


Law of Definite Proportion: We know that mass is conserved. If we were
to take a sample of a pure compound it will always contain the same
elements combined in the same proportion by mass.
Law of Multiple Proportion: If one element can combine with another to
form more than one compound then the mass ratios of the elements in the
compounds are simple whole-number ratios of each other.
Law of Definite Proportion by Volume: The volumes of reactants and
products, in chemical reactions, are related to each other by simple
whole-numbers. Of course the measurements must be taken at the same
temperature and pressure.
Law of Reciprocal Proportions: The ratio by mass, in which two elements
combine with thefixed mass of a third element, is either the same or a
simple multiple of the ratio in which they combine with each other.
Avogardo's Law: Equal volumes of all gases contain equal number of
molecules under similar conditions of temperature and pressure.
One Atomic Mass Unit (amu): It is a mass unit equal to exactly one
twelfth the mass of a carbon 12 atom.
Formula Mass of a Substance: It is the sum of the atomic masses of all
atoms in a formula unit of a compound.
Molar Mass: The mass of 1 mole of any substance.
Molar Volume (Vm): The volume of 1 mole of the substance.
Mole (m):
• It is a number which is equal to 6.023 x 1023 .
• It is the amount of substance which contains as many entities (atoms,
molecules, ions or other particles) as there are atoms in exactly 12
grams of Carbon 12 isotope.
• It is also known as Avogadro's Constant.
Number of Mole Method:
Number of moles of a substance n = w (in gms)/m
also n = volume (in lt) at NTP/22.4lt
also n = M x vol in lt (where M is Molarity = number of moles of solute per lt
of solution)
M= (w/m) x (1000/ Vm )
Number of mole method is used only in the case of balanced chemical
equations.
The coefficients of balanced equations represent the ratios in terms
of number of moles in which the reactants react or products are formed.
Number of particles(ions, atoms or molecules) in a given number of moles
= number of moles x 6.023 x 1023 .
Equivalent Weight of an Element (E): It is defined as the number of parts
by weight of the element which combine with or displace from a compund 1
part by weight of Hydrogen, 8 parts by weight of Oxygen or 35.5 parts by
weight of Chlorine.
E = M ( Molar Mass) / nf (n factor)
where:
nf = Valency in case of an atom
= Total positive or negative oxidation number of an atom in a molecule
= Basicity or Acidity.
= Change in oxidation number in case of a redox reacton.
Number of Gram Equivalents:
= weight in gms / E (Equivalent wt)
= Normality x vol in lt
Normality = (w/E) x (1000/Vm )
Number of Equivalents Method: In this method you need not balance the
chemical equation. The basic principle is that the equivalents of each
reactant which have dissapeared are equal to the number of equivalents of
each of the products formed. The working tool for this is equivalent weight.
Number of equivalents = W in gms / Eq Wt = (W x nf ) / M
Dulong and Petit's Law:
Atomic wt (approx) = 6.4/specific heat(in calories)
Curved Arrows

Additional reading recommendation: You may find Chapter 3 of Pushing


Electrons by Daniel P. Weeks (Saunders College Publishing; ISBN 0-03-0206936)
to be a useful tool for mastering the fundamentals of using curved arrows.
Discussion: Chemical reactions are a result of bonding changes. These bonding
changes are most easily described by considering the changes in electron sharing
between atoms. For example, consider the collision of two water molecules leads
to the ionization of water and the formation of hydroxide and hydronium ion.

In this reaction, the oxygen atom of one water molecule collides with a hydrogen
atom of the second water molecule. A lone pair of the oxygen atom becomes the
new O-H bond in the hydronium ion. Because a hydrogen can only be fully bond
to one other atom at a time, the old O-H bond is lost. The electron pair of the old
O-H bond becomes a lone pair, sole property of the oxygen atom of hydroxide ion.
This very simplistic step-by-step bookkeeping description of all the bonding and
electron changes in a reaction is called the reaction mechanism. Study,
understanding and prediction of reaction mechanisms is at the very heart of
reactions in organic chemistry. Mastering the fundamentals of reaction
mechanisms is a fundamental survival skill for students of organic chemistry. You
will use them every day that you study organic chemistry.
Above we used several lines of text to describe the bonding changes in a single
step of a reaction mechanism. A reaction mechanism might have ten steps or more,
making such descriptions very cumbersome. A shorthand notation that summarizes
these changes has thus been developed. This notation, called curved arrow
formalism, provides a rapid way of drawing bond and electron changes in a given
mechanism step. They are also useful to indicate electron changes between a set of
contributing resonance structures.

Each curved arrow with two barbs on the head represents the shift of one
electron pair. (Later we will encounter single-barbed curved arrows that represent
the shift of single electrons.) The curved arrows shows the direction of electron
flow. The tail shows the electron origin, and always come from an electron source,
usually a lone pair or bonding pair from a σ or π bond. The head of the arrow
indicates the electron pair destination, either as a new lone pair or a new bond. If
the arrowhead points to another atom, that atom must either have an open octet and
thus be able to accept the electron pair, or have an electron pair that can be
displaced by the incoming electron pair. Electrons never flow from atoms which
are electron-poor to atoms which are electron-rich, so a curved arrow will never
point from an atom with a positive charge to an atom with a negative charge.

New bond formed to X:

Bonding pair becomes lone pair; bond broken:


The use of curved arrows for the ionization of water are shown
below.

The tail of the curved arrow on the right starts at the oxygen lone pair, meaning this
curved arrow shows a bonding change for this lone pair. The head of the curved
arrow points to the space between the oxygen and hydrogen atoms, meaning the
electron pair ends in that space as a bond between the oxygen and the hydrogen.
The hydrogen that accepts a new bond to oxygen must give up a pair of electrons
because it cannot have more than two valence electrons at a time. Thus, the old O-
H bond is displaced by the new electron pair from the other oxygen atom. The
curved arrow on the left indicates the electron pair that was the O-H bond becomes
a lone pair on the oxygen of the hydroxide ion.
If the arrow starts at a bond between two atoms, then that bond is broken. If
the arrow ends between two atoms, then a new bond is formed between those
atoms. (If the atoms are already bonded, then a double or triple bond results.)
The process of drawing a curved arrow mechanism is also commonly called
"electron pushing."
When drawing curved arrows, the start and stopping points of the arrows are
critical. Things that make no difference are where the arrows curve up or down, or
whether they start or stop at the top or bottom of an atom. The arrows you draw
may therefore look different than the arrows shown in this tutorial.
Note also the changes in formal charge that result from the electron changes. If an
atom shares a lone pair that it used to have all to itself then its charge decreases by
one (i.e., neutral atom become +1). If an atom gains a bonding electron pair all to
itself as a lone pair then its charge increases by one unit (i.e., a neutral atom
becomes -1). The charge is conserved in this mechanism step. The total charge on
the left (zero) is the same as the total charge on the right (-1 +1 = 0). Charge is
conserved in all mechanism steps. You should make a habit of checking your work
against this point to minimize errors.
Lone pairs are often involved in reaction mechanisms, so you should be in the
habit of drawing all lone pairs. It is also important that your curved arrows be
drawn neatly and precisely, clearly showing the atomic origin of the electron pair
at the tail of the curved arrow and the electron pair destination at the head of the
arrow.

Example 1: Provide the curved arrows for the reaction of hydroxide and
hydronium ions to form two molecules of water.

Solution 1: A reasonable approach to an exercise like this is to


analyze the bond changes are then draw the corresponding
curved arrows. The oxygen of hydroxide ion has shared a lone
electron pair with the hydrogen of hydronium ion, forming a new
O-H bond. Thus we draw a curved arrow with the tail at the
hydroxide ion lone pair and ending at the hydronium ion
hydrogen. This hydrogen atom can only have one covalent bond
at a time, so it must sacrifice the bond to the hydronium ion. We
draw a curved arrow to show the bonding electron pair between
the hydrogen and oxygen of the hydronium ion moving to become
the sole property of the oxygen atom. The oxygen of the
hydroxide ion is sharing a pair of electrons that it had all to itself
before, so its formal charge drops by one unit (-1 to neutral). The
oxygen of the hydronium ion gains a lone pair of electrons that is
used to share with the hydrogen, so its formal charge decreases
by one unit (neutral to +1).

Example 2: Draw the product(s) of the following mechanism step based upon the
curved arrows.

Solution 2: The curved arrow that starts at the carbon-


carbon π bond and ends at the bromine atom indicates the π bond
electron pair has shifted to become a carbon-bromine bond. The
left-hand carbon of the π bond has lost an electron pair, so its
formal charge becomes one unit more positive (neutral to +1).
The curved arrow on the right indicates that the electron pair of
the bromine-bromine bond is shifting to reside solely on the
bromine on the right, resulting in rupture of the bond and
formation of bromine with four lone pairs and a negative charge.

Exercises:
Provide curved arrows that show how the following mechanism steps might occur.
Provide the product(s) for the following mechanism steps based upon the curved
arrows.

Electrophiles and Nucleophiles

Electrophile: A molecule or ion that accepts a pair of electrons to form a new


covalent bond (same as Lewis acid).
Leaving group: A molecule or ion that leaves with a pair of electrons that used to
be shared in a covalent bond with this fragment.

Lewis acid: A molecule or ion that accepts a pair of electrons to form a new
covalent bond (same as electrophile).

Lewis base: A molecule or ion that shares a pair of electrons to form a new
covalent bond (same as nucleophile).

Nucleophile: A molecule or ion that shares a pair of electrons to form a new


covalent bond (same as Lewis base).

Discussion: Consider the reaction of hydroxide ion (a Bronsted


base) with hydrogen chloride (a Bronsted acid).

The oxygen of hydroxide ion bears a formal charge of -1. The hydrogen of
hydrogen chloride bears a δ+ charge because chlorine is more electronegative than
hydrogen and thus the H-Cl bonding electron pair is unequally shared. We can
envision the start of the acid-base reaction between hydroxide ion and hydrogen
chloride as an electrostatic attraction between the opposite charges. As the reaction
proceeds, the oxygen atom of hydroxide ion shares a lone electron pair with the
hydrogen atom of hydrogen chloride, as shown with the curved arrow in the
reaction above. This simple reaction shares one feature in common with the
majority of reactions that you will encounter in your study of elementary organic
chemistry. One species in the reaction shares an electron pair (a Lewis base) with
another species (a Lewis acid) to make a new covalent bond. Application of this
electron sharing idea to any reaction gives you an excellent chance at starting to
figure out the mechanism for most common organic reactions.
Because of the ubiquity of electron pair sharers and acceptors in organic reactions,
we assign special and distinct terms to these species. A molecule or ion that
accepts a pair of electrons to make a new covalent bond is called
an electrophile (from the Greek for "electron loving"). An electrophile is the same
thing as aLewis acid. Any molecule, ion or atom that is electron deficient in some
way can behave as an electrophile. Electron deficiency would include a formal
positive charge (methyl carbocation), a partial positive charge (δ+), usually in
conjunction with a polar bond (such as H-Cl) or an open octet (borane). "E" or
"E+" are common abbreviations for generic electrophiles.

Typical electrophiles:
A molecule or ion that donates a pair of electrons to form a new covalent bond is
called a nucleophile (from the Greek for "nucleus loving"). A nucleophile is the
same thing as a Lewis base. Any molecule, ion or atom that has electrons that can
be shared can be a nucleophile. The most common indications that electrons are
available to be shared are formal negative charge (iodide ion), a partial negative
charge (δ-), usually in conjunction with a polar bond (methyl magnesium
bromide), aπ bond (isobutylene) or lone pairs (ammonia). "Nuc" or "Nu" are
common abbreviations for generic nucleophiles.

Typical nucleophiles:
The study of reaction mechanisms is central to the study of organic chemistry at
any level. Therefore identification of electrophiles and nucleophiles is a critical
organic chemistry survival skill. Examination of a structure for the features
discussed above is one way to identify how a molecule or ion might behave in a
reaction. Another way is by considering the curved arrows. Because electrons
flow from an electron source to a place of electron deficiency, a curved arrow
points away from a nucleophile and to an electrophile. This does not work in every
reaction, however. In some reactions, the electron flow could go in either
direction, and there are no distinct nucleophiles and electrophiles. Such reactions
are uncommon in a course of this level.

Electrons always flow from nucleophile to electrophile:


Example 1: Using the curved arrows shown below, label each reactant as a
nucleophile or electrophile.

Solution 1: A nucleophile is a molecule or ion that donates an electron pair to


form a new covalent bond. In this example, chloride ion is donating a lone pair to
form a new bond with carbon. The chloride ion is at the origin on the curved arrow
that indicates this bond change. That the chloride ion bears a formal negative
charge further suggests it should function as a nucleophile. Because nucleophile
must react with an electrophile that leaves the other molecule (a chlorosulfite ester)
to be the electrophile. The ester carbon is accepting an electron pair from the
nucleophile to form the new C-Cl bond. The other bond changes within the
chlorosulfite ester molecule are inconsequential when defining its role as a
nucleophile or electrophile. The chlorine atom that is expelled as chloride ion
accepts and electron pair from the S-Cl bond, but it does not make a new covalent
bond, so it is neither electrophile or nucleophile. In this reaction, chloride ion is
the leaving group.

Example 2: Decide if each molecule or ion shown below will react as a


nucleophile or electrophile, or both.

Solution 2: Examine each structure for the charge distribution and electronic
features discussed above.
a. Bromide ion: This atom has four lone pairs and a formal negative charge,
suggesting it is electron-rich and can therefore function as a nucleophile. It has
none of the features that would suggest it might behave as an electrophile.
b. Ammonium ion: This ion has a formal positive charge, suggesting it is electron-
poor and can therefore function as an electrophile. It has no lone pairs or areas of
negative charge, suggesting it will not function as a nucleophile.
c. Water: The oxygen atom of water has two lone pairs and a δ- charge (oxygen is
more electronegative than hydrogen). This suggests that water can behave an a
nucleophile. Each hydrogen atom bears a δ+ charge, so the molecule can behave
as an electrophile as well. Many molecules can be both nucleophiles and
electrophiles. How they behave depends upon what they react with. For example,
if water is reacted with an electrophile, the water will behave as a nucleophile.

Exercises:
Identify the nucleophiles and electrophiles in each mechanism step shown below.

Decide if each molecule or ion shown below will react as a nucleophile or


electrophile, or both.

A Brief Tutorial on Drawing Lewis Dot Structures

We will use three molecules (CO2, CO32- and NH4+) as our examples on this guided
tour of a simple method for drawing Lewis dot structures. While this algorithm
may not work in all cases, it should be adequate the vast majority of the time.

Procedure for Neutral Molecules (CO2)


1. Decide how many valence (outer shell) electrons are posessed by each atom in
the molecule.

2. If there is more than one atom type in the molecule, put the most metallic or
least electronegative atom in the center. Recall that electronegativity decreases as
atom moves further away from fluorine on the periodic chart.

Arrangement of atoms in CO2:


3. Arrange the electrons so that each atom contributes one electron to a single bond
between each atom.

4. Count the electrons around each atom: are the octets complete? If so, your Lewis
dot structure is complete.

5. If the octets are incomplete, and more electrons remain to be shared, move one
electron per bond per atom to make another bond. Note that in some structures
there will be open octets (example: the B of BF3), or atoms which have ten
electrons (example: the S of SF5).

6. Repeat steps 4 and 5 as needed until all octets are full.


7. Redraw the dots so that electrons on any given atom are in pairs wherever
possible.

Procedure for Negatively Charged Ions (CO32-)


Use the same procedure as outlined above, then as a last step add one electron per
negative charge to fill octets. Carbonate ion has a 2- charge, so we have two
electrons available to fill octets.

Using the procedure above, we arrive a this structure:


The two singly-bonded oxygen atoms each have an open octet, so we add one
electron to each so as to fill these octets. The added electrons are shown with
arrows. Don't forget to assign formal charges as well! The final Lewis structure
for carbonate ion is:

Procedure for Positively Charged Ions (NH4+)


Use the same procedure as outlined above, then remove one electron per postive
charge as needed to avoid expanded octets. When using this procedure for
positively charged ions, it may be necessary to have some atoms with expanded
octets (nitrogen in this example). For each unit of positive charge on the ion
remove on electron from these exapnded octets. If done correctly, your final
structure should have no first or second period elements with expanded octets.
Using the basic procedure outlined above, we arrive at a structure in which
nitrogen has nine valence electrons. (Electrons supplied by hydrogen are red;
electrons supplied by nitrogen are black.) Removal of one of these valence
electrons to account for the 1+ charge of ammonium ion solves this octet rule
violation.

Resonance: Vocabulary
Exercise Solutions

Contributing resonance structures: One or more alternate Lewis structures for a


molecule or ion.
Double-headed arrow: Drawn between two molecular structures to indicate that
they are contributing resonance structures.
Resonance: When a molecule can be represented by the weighted hybrid of two or
more hypothetical but reasonable Lewis structures that differ only in the
distribution of bonding and nonbonding electrons, and in which the positions of the
nuclei are constant.
Resonance hybrid structure: A structure which is the weighted average of all
contributing resonance structures for a molecule or ion. The resonance hybrid
structure is the closest representation of reality for that molecule or ion.

Resonance: Drawing Resonance Structures


Additional reading recommendation: You may find Chapter 2
of Pushing Electrons by Daniel P. Weeks (Saunders College
Publishing; ISBN 0-03-0206936) to be a useful tool for mastering
the fundamentals of resonance structures.
Discussion: Consider the Lewis structure of the carbonate ion, CO32-. The Lewis
structure for this ion has a carbon-oxygen double bond, and two carbon-oxygen
single bonds. Each of the singly bonded oxygen atoms bears a formal charge of
1-. (Review the formal charge tutorial if needed.) But which of the three oxygens
forms the double bond? There are three possibilities:

These structures are similar in that the have the same types of bonds and electron
positions, but they are not identical. The position of the carbon-oxygen double
bond makes them different. In structure A the double bond is with the top oxygen
atom, in B with the right hand oxygen atom, and C with the left hand oxygen
atom. These oxygen atoms are at different places in space, so these are different
structures. Consider this analogy: when the hands on a clock are at a 90o angle, the
time could be 3 o'clock, or 6:15. The angle between the hands stays the same, but
because they point to different places in space, they indicate a different time. The
position of the carbon-oxygen bond is like the fixed angle of the clock hands, but
pointing to different places on the clock face.
When more than one Lewis structure can be drawn, the molecule or ion is said to
have resonance. The individual Lewis structures are termed contributing
resonance structures. Resonance is a common feature of many molecules and
ions of interest in organic chemistry.
Which one of these three structures is the correct one? How could we tell? If
structure A was correct, laboratory measurements would show one shorter bond
(the carbon-oxygen double bond) and two longer bonds (the carbon-oxygen single
bonds). Measurement of structures B and C would give the same results as well.
As it turns out, laboratory measurements show that all three bonds are equal and
between single and double bond length. This suggests that none of the Lewis
structures we have drawn are correct. It further suggests that the actual structure
has three equal carbon-oxygen bonds that are intermediate between single and
double bonds.
Perhaps the three Lewis structures for carbonate ion are in rapid equilibrium. The
structures are changing so quickly that all we see can measure is an average blur
(structure D), instead of being able to detect individual structures. By analogy,
consider a camera with the shutter left open. The picture would be a blur that
looks like A, B and C all at the same time. Structures A, B and C have the same
bonds and electron distribution, the only difference is the position of the bond.
Thus the three structures have equal stability, and the three structures would occur
to the same extent at equilibrium. The blur we see would look like 1/3 A,
1/3 B and 1/3C. The bond lengths would be a blur as well; we would perceive
them as being something between single and double bonds. The charge would
shift so rapidly that we would see it on all oxygen atoms at once. Since each
oxygen atom has a 1- charge in two of the three equilibrium contributors, each
oxygen atom would appear to have, on average, a charge of 2/3-.

All laboratory experiments have failed to detect structures A, B and C. No matter


what experiments are performed, analysis has always concluded that D is the best
description for the structure of the carbonate ion. This suggests that A, B and C do
not exist, and are not adequate descriptions for carbonate ion at any time. The
actual structure is D, and not equilibrium between A, B and C.
That D appears to be a combination of A, B and C still appears to be a useful way
to determine the actual structure of carbonate ion. The only problem is
that A, B,C and D are not in equilibrium. Can they be interacting in some other
way? The answer is that the true structure of carbonate ion appears to be a
simultaneous hybrid of the three resonance contributors A, B and C.
Structure D has features derived from A, B, and C, but is never just A or just B or
just C. For example, Dhas some double bond character, and each oxygen has some
negative charge. Hybrid structure D is termed the resonance hybrid structure. It
is a much better representation of reality for carbonate ion than any of the
individual contributing resonance structure. We use a double headed arrow to
show that individual structures are related by resonance.
It is important to understand that structure D is a hybrid of A, B and C. It has
features common to A, B and C, but carbonate ion does not rapidly interconvert
between A, B and C. By analogy, consider the case of a mule. A mule is the
offspring of a horse and donkey. It has features similar to both, such as long tail
and mane, but it does not rapidly interconvert between the two. A mule is not a
horse one second, then a donkey the next. (An animal that changed would
certainly be the highlight of any zoo!)
Drawing contributing resonance structures. Resonance is an important feature of
many organic molecules. It can have a profound influence on their structure,
chemical reactions, and physical properties. Key to understanding resonance is the
ability to draw contributing resonance structures and the resonance hybrid
structure.
Resonance structures are simply alternate Lewis structures for a given ion or
molecule. Thus we can draw all resonance structures by drawing all of the
possible Lewis structures. However, it is not always easy to see what all these
Lewis structures might be. A set of Lewis structures for a give ion or molecule
must have the same number of electrons as the Lewis structures are constructed
from the same atoms. The only difference between the Lewis structures is the
placement of the electrons. The position of the atoms in space is held constant.
We can use this facts to assist us in drawing resonance structures. Because the
number of electrons is conserved, electrons taken away from one atom must appear
somewhere else in the structure. When drawing resonance structures, it is most
convenient to shift these electrons between adjacent atoms.
Electrons that can be moved between adjacent atoms in resonance structures are
lone pairs or π electrons. (Unpaired electrons present in radicals will be
considered in section 15.5 of the text.) These electrons reside on a "resonance
donor atom." The "resonance acceptor atom" must be adjacent to the donor atom.
The acceptor atoms must also have an open octet, be able to accommodate an
expanded octet (text section 1.1C; most commonly this will be chlorine, bromine,
sulfur and phosphorus) or have another electron pair (lone pair or π) that can be
displaced. An atoms with a formal positive charge can also be an resonance
acceptor atom, as long as the atom does not accept more electrons that it can
normally accommodate.
We use curved arrows as a bookkeeping tool to indicate the electron changes that
differentiate resonance structures. Although these are the same curved arrows we
use to indicate that bond changes are occurring in a reaction mechanism, they do
not mean exactly the same thing. A reaction step requires a finite amount of time
to occur, whereas a shift between resonance structures never actually occurs.
Recall that individual resonance structures do not exist, they are simply alternate
Lewis structures for the same molecule or ion. In the example below, the curved
arrow indicates that the lone pair of electrons on the left-hand carbon is moving to
a position where it can be shared between the two carbons, thus becoming the
carbon-carbon π bond. This use of curved arrows is often termed "electron
pushing."

Because electrons are shifting around, the formal charge distribution will vary
between resonance structures. The formal charge of an atom that gains a pair of
electrons through resonance becomes one unit more negative. Conversely, the
formal charge of an atom that shares a pair of electrons that it did not share
previously becomes one unit more positive.

Rules for drawing contributing resonance structures. Some rules must be


considered when drawing resonance structures.
Rule 1: All resonance structures must have the same number of valence
electrons. Electrons are not created or destroyed, nor are they lost or gained to
other molecules or ions during resonance.

The rule is violated because structure E has 12 valence electrons (four bonding
pairs and two lone pairs), whereas structure F has 14 valence electrons (five
bonding pairs and two lone pairs). Therefore these cannot be resonance structures
of the same ion. (Structure F also violates rule 2.)

Rule 2: The octet rule must be obeyed. Hydrogen may never have more than
two valence electrons. Lithium through fluorine may never have more than eight
valence electrons. (A carbon with five attachments, often called a "pentavalent
carbon" has ten valence electrons. This is forbidden because carbon does not have
the space in its orbitals to accommodate ten electrons. You should take care to
avoid this common mistake made by inexperienced organic chemistry students.)
Certain elements commonly encountered in organic chemistry may have ten
valance electrons. These elements are in periods three and higher in the periodic
table, and include chlorine, bromine, iodine, phosphorus and silicon. Of all the
atoms commonly encountered in organic chemistry, only sulfur can routinely
expand its octet to include twelve valence electrons. These atoms expand their
octets so as to improve the importance of the resonance structure.

Resonance structure G is acceptable. Structure H is not acceptable because the


carbon has ten valance electrons.

Structures I and J are both acceptable resonance contributors for bisulfate ion, the
conjugate base of sulfuric acid. The sulfur atom of structure J has 12 valence
electrons, an expanded octet. Sulfur is a third row element, so an expanded octet is
allowed. Structure J is the more important resonance structure because it
maximizes the number of covalent bonds and minimizes the number atoms with a
nonzero formal charge.

Rule 3: Nuclei do not change positions in space between resonance structures.


Resonance structures differ only in the arrangement of valence electrons.

Structures K and L are both acceptable Lewis structures, but they are not related
by resonance because the circled hydrogen atom has changed position in space.

Exercises: Draw all reasonable resonance forms for the structures shown below.
Use curved arrows to indicate electron pair changes.
Resonance: Most Important Resonance Contributor

Discussion: Many molecules or ions that participate in an organic reaction


have resonance. When writing the mechanism for the reaction, the best
representation of reality would be achieved by using the resonance hybrid
structure. However, because the resonance hybrid does not show explicitly show
electron pairs that are shared by resonance, its use in mechanisms can be unclear.
Thus we often use a single resonance contributor instead of the hybrid. When
deciding which resonance contributor to use, it makes sense to use the one that
makes the greatest contribution to the resonance hybrid. If we cannot use the
hybrid, then we should use the next closest structure. In addition, many (but not
all) reactions of molecules or ions with resonance proceed as if this most important
resonance contributor was the actual reactant. Thus we need a set of rules to
determine the most important resonance contributor. These rules are based on the
idea that if individual resonance contributors did indeed exist, the most
thermodynamically stable structures would make more significant contributions to
the resonance hybrid. Factors that enhance thermodynamic stability are
maximization of covalent bonding and minimization of charge. Resonance
increases stability by increasing the bonding between adjacent atoms and by
distributing charge over a greater number of atoms.

Preference 1: The most important contributor has the maximum number of


atoms with full octets.
This preference gets priority over the other three rules for determining the most
important resonance contributor.

The carbon of structure A has an open octet. All the atoms of structure B have full
octets. Therefore contributor B is more important than contributor A, despite the
fact that the positive charge is on the more electronegative oxygen atom instead of
the less electronegative carbon atom.
Preference 2: If a resonance contributor must have formal charge, the most
importnat contributor has these charge(s) on the atoms most willing to
accommodate them. Negative charges are best accommodated on more
electronegative atoms, whereas positive charges are best accommodated on the
least electronegative atoms.

All atoms of resonance contributors C and D have a complete octet, so we turn to


other preferences to determine the most important resonance contributor. A
negative charge is best accommodated by a more electronegative atom. Because
oxygen is more electronegative than carbon, contributor D is more important than
contributor C. (If the ion shown above was a cation, then the resonance
contributor with the positive charge on carbon would be more important than the
contributor with the positive charge on oxygen.)

Preference 3: The most significant contributor has the maximum number of


covalent bonds. Contributor B (above) is more important than
contributor Abecause B has the carbon-oxygen π bond absent in A.

Preference 4: The most significant contributor will have the least number of
formal charges.

Resonance contributor F is more significant than contributor E because F has no


atoms with formal charges, whereas E has two atoms with formal charges.
(Contributor F is also favored by Preference 3 as well.)

Preference 5: The most significant contributor has the least number of


unpaired electrons.
For example, contributors G and H each have one unpaired electron, and thus are
preferred over contributor I which has three unpaired electrons. Resonance
contributors that include avoidable unpaired electrons are rarely of any
consequence and thus should not be considered. There is one common exception:
molecular oxygen. Due to molecular orbital considerations, molecular oxygen is
best described as having two unpaired electrons and an oxygen-oxygen single bond
(contributor J) and not as lacking unpaired electrons with an oxygen-oxygen
double bond bond (contributor K).

Exercises: Determine the most significant resonance contributor for each set
of contributing resonance structures drawn previously.

Stereochemistry: Vocabulary
Exercise Solutions

Achiral: An object that is not chiral.

Chiral: An object that is not superposable with its mirror image.


Constitutional isomer: One of a set of isomers that differs in the sequence of
attachment or bonding of the atoms.

Dextrorotatory: Rotation of plane polarized light in a clockwise direction.

Diastereomer: One of a set of stereoisomers that are not enantiomers.

Enantiomer: One of a pair of stereoisomers that are nonsuperposable mirror


images.

Isomer: One of a set of compounds with identical molecular formulas.

Levorotatory: Rotation of plane polarized light in a counterclockwise direction.

Meso compound: A compound that has at least two stereocenters but is achiral.

Optically active: Rotation of plane polarized light.

Plane polarized light: Light in which the electric field component oscillates in a
single plane.

Polarimeter: A device to measure rotation of plane polarized light.

Racemic mixture: An equimolar mixture of a pair of mutual enantiomers.

Resolution: Separation of mutual enantiomers.

Specific rotation: The magnitude of rotation of plane polarized light of compound


in a 10.0 cm path length, and at concentration of 1.00g/mL or neat if the compound
is a liquid.

Stereocenter: An atom that has four different groups attached to it. (Not limited to
just carbon.)
Stereoisomer: One of a set of isomers that have the same connectivity of atoms,
but differ in the position of the atoms in space. Stereoisomers cannot be
interconverted by rotation around a single bond. (Conformational isomers are not
stereoisomers.)

Stereochemistry: Identifying Stereocenters


Discussion: A stereocenter is a carbon atom that has four
different attachments. The difference can be anything, except a
conformational change. Thus, we consider n-propyl and isopropyl
groups to be different attachments. Propyl groups that differ only
in rotation about a single bond (conformational differences) are
identical because they can rotate to adopt identical
conformations.

Figure 1. Left: This molecule has a stereocenter. The carbon


indicated by the arrow has four different attachments: CH3,
CH3CH2CH2, (CH3)2CH and Cl. Right: This molecule does not have
a stereocenter. The carbon indicated by the arrow has three
different attachments: CH3, CH3CH2CH2(two different
conformations but still the same group) and Cl.
Most instances of chiral molecules that you will encounter at this
level of organic chemistry owe their chirality to the presence of
one or more stereocenters. It is not necessary for a molecule to
have a stereocenter to be chiral. Make models of
the enantiomers of 2,3-pentadiene (1,3-dimethylallene) shown
below. Having a stereocenter does not mean a molecule will
always be chiral. Explore this point with a molecular model
of meso-tartaric acid, a compound critical to Pasteur's studies
of optical activity in organic compounds.
In the next section we explore the importance of recognizing chirality in molecular
structure. Because stereocenters are the origin of chirality of most chiral organic
molecules, it is useful to be able to recognize stereocenters within a molecule.
Example 1: In each molecule shown below, identify the stereocenter(s).

Solution 1:
a. When looking for stereocenters, pay careful attention to the "four different
attachments" criterion. Recall that hydrogens of "stick structures" are usually not
drawn. Also note attachments that differ only in conformation do not make a
stereocenter. The only carbon of 2-butanol that bears 4 different attachments is the
alcohol carbon.

b. 2-Methyl-3-pentanol has only one stereocenter: the alcohol carbon. Don't


neglect the hydrogen on this carbon just because it isn't shown. The other carbon
bearing wedge and broken line bonds is not a stereocenter, because it has two of
the same thing (methyl groups) attached. We often use the wedge and broken line
notation to imply a three dimensional arrangement of atoms, but its does not
always occur on a stereocenter.

Exercises: Locate all stereocenters in each structure.


Stereochemistry: Determining Molecular Chirality
Discussion: Chiral objects are not superposable with their mirror images. An
excellent example of this is your hands. Hold your hands out in front of you, with
the palms facing together. Neglecting unnatural additions such as jewelry, note that
your hands are mirror images. Now turn your hands so that both palms face the
same direction. Note that the thumbs now point in opposite directions. When the
thumbs point in the same direction, the palms are opposite. Your hands are mirror
images, but not superposable. Each hand is therefore chiral.
Achiral objects may be superposed on their mirror image. Examine two sheets of
blank paper in the same way as you experimented with your hands. Notice the
sheets of paper are mirror images, but superposable. The sheets of paper are
therefore achiral.
All objects can be classified as chiral or achiral, including molecules. If our hands
were molecules, they would be a pair of enantiomers.
We know from basic biology that interaction of molecules, such as the docking of a
substrate to an enzyme, is vital to living organisms. Because enzymes and their
substrates may be chiral, it is useful to understand how achiral and chiral
molecules can interact. (enzymes are constructed from a group of about 20 small
molecules called amino acids. All amino acids except one are chiral, so the
enzymes they make are chiral as well.) The way a hand slips into a glove provides
a useful way to model this effect. The glove is the enzyme, and the hand is the
substrate that must fit properly into the enzyme pocket for the enzyme to be able to
act upon the substrate. (Verify that a pair of gloves are chiral in the same way you
explored the chirality of your hands.) Your right hand fits nicely into the right
handed glove, but does not fit well into the left-handed glove. Likewise, your left
hand fits well into the left-handed glove, but the right hand does not. Imagine the
glove represents an enzyme and your hand the substrate. The left-handed
enzyme/glove would accept the left-handed substrate/hand readily, and would be
able to act upon the substrate. The left-handed enzyme/glove cannot readily accept
the right hand/substrate, as so this enzyme cannot readily act upon this substrate.
This simple model implies that an enzyme will act on one enantiomer more readily
than another. Thus, enantiomers of drugs can have different effects in the body,
because they are acted upon differently by enzymes, despite the fact that they have
the same set of functional groups.
That enantiomers of drugs can have different biological effects has been
demonstrated in many instances, but perhaps none so dramatically as the in the
case of the drug thalidomide. In the late 1950s, the racemic form of this drug was
prescribed as a sedative or hypnotic for pregnant women. Some women who took
the drug delivered children with severe birth defects. A substance that causes fetal
abnormalities is called a teratogen. Further research revealed that one enantiomer
of thalidomide has the desired sedative effects, while the other enantiomer was
teratogenic. The enantiomers of thalidomide were acting differently in the body,
because they interacted differently with chiral biomolecules such as enzymes. The
drug was quickly removed from the market.

Chiral molecules are nonsuperposable with their mirror images. This can be tested
on paper or with molecular models using the two methods described below.
Internal mirror plane. We can look for a plane of symmetry in the molecule.
Imagine this plane as a mirror through the middle of the molecule. If one half of
the molecule is reflected into the other half, then the molecule is achiral. If no such
mirror plane exist, the molecule is usually chiral. (There are symmetry elements
other than a mirror plane that may render a molecule achiral, but these are rarely
encountered and thus beyond the scope of an introductory organic chemistry
course.) Molecular models can be used in the same way.
Example 1: Using the method of an internal mirror plane, determine if
cyclohexanol is chiral or achiral.
Solution 1: To determine if cyclohexanol is chiral using the internal mirror
method, draw a mirror plane through the middle of molecule. If there are any
unique functional groups or atoms within the molecule, these must lie within the
mirror plane, so that one half of the atom or functional group is reflected into the
other half. In this case, there is only one alcohol functional group, so it must be
contained in the mirror plane. Figure 1 shows that the mirror plane bisects the
molecule into two equivalent halves, so cyclohexanol is achiral.

Figure 1. Two views of cyclohexanol showing the internal mirror plane. The
mirror plane is indicated by the dashed line.

Figure 2. Cyclohexanol molecular model. A vertical mirror plane


bisects the molecule through the middle of the picture.
Superposable models. To determine if a molecule is chiral using
the superposability requirement, build a molecular model of the
molecule in question, then a build a mirror image of this model.
Now try to superpose the models by aligning them so that all the
atoms match up. The models may be manipulated in any way,
such as rotation around single bonds (changing molecular
conformation) or changing perspective, but bonds cannot be
broken. If all the atoms can be made to line up, the molecule is
achiral. If they cannot be aligned, the molecule is chiral.

Example 2: Using the method of superposable molecular models, determine if


cyclopentanol and 2-chlorobutane are chiral or achiral.

Solution 2:

Figure 3. Left: Molecular model of cyclopentanol and its mirror image. Right: A
top view showing that these cyclopentanol models can be aligned (superposed), so
cyclopentanol is achiral.

Figure 4. Left: Molecular model of 2-chlorobutane and its mirror image. Right:
The same models stacked. The Cl-C-H portions of the models can be made to
superpose, but at the same time the methyl and ethyl groups do not. The 2-
chlorobutane models cannot be superposed, so the molecule is chiral.
Exercises: Using either method discussed above, determine if the molecules shown
below are chiral or achiral.

Stereochemistry: Determining Molecular Chirality


Exercise Solutions

a. 3-Methylhexane:
This molecule is chiral because the mirror images are not superposable. The case
is very much like 2-chlorobutane, except the molecule has a propyl group instead
of a chlorine atom. Each molecule has a single stereocenter. Examination of other
molecules with single stereocenters should rapidly convince you that molecules
with a single stereocenter are always chiral.

b. Benzene:
Benzene is achiral because it has many internal mirror planes. Perhaps the most
obvious of these is the molecular plane that contains all twelve atoms.

c. Cyclohexane:
Cyclohexane can exist in many different conformations (chair, boat, etc.). In cases
of conformationally fluxional molecules, if there is one conformation that is
achiral, then the molecule as a whole is achiral. The chair conformation has a
mirror plane through the middle of the ring, containing C1 and C4, so this
conformation is achiral. Because the chair conformation is achiral, we consider
cyclohexane to be an achiral molecule.

This molecule contains no mirror planes, and mirror image molecular models are
not superposable. Therefore this molecule is chiral.

This molecule has two stereocenters, so at first guess, it might appear to be chiral.
However, it has an internal mirror plane:

This is a meso compound, and is achiral, as are all meso compounds. Note that for
a compound to be meso, the pairs of stereocenters must have the same groups
attached to them.

At first glance, this compound appears to be chiral because it contains no apparent


internal mirror plane. As with the cyclohexane case, we need to examine all
conformations to see if any are achiral. A good way to do this in a molecule with
multiple stereocenters is to hold one stereocenter still (the left hand one in this
case), and rotate around the single bonds to try and make the stereocenters align.
Such a rotation in this molecule reveals an internal mirror plane:
This molecule is therefore a conformational isomer of the molecule in (e), and is
thus achiral.

At first glance, we might suspect this to be a meso compound because the two
stereocenters have the same attachments. However, rotation around the bond that
attaches the stereocenters reveals that they are not reflected by an internal mirror
plane:

Careful examination will reveal that this molecule has no conformations with an
internal mirror plane and so it is chiral.

In the quest for an internal mirror plane, rotation around the bond connecting the
two stereocenters reveals the molecule to be meso and therefore achiral:
Stereochemistry: Drawing Enantiomers and Diastereomers
Discussion: For the same reasons as it is important to recognize and
classify stereoisomers it is valuable to be able to
draw enantiomers and diastereomers as well.
Drawing enantiomers: Recall that enantiomers are molecules which are
nonsuperposable mirror images. Because these are mirror image molecules, and
because stereoisomers can only have two absolute configurations, all
the stereocenters of one enantiomer will be the mirror image of the other
enantiomer. A mirror image stereocenter can be draw by switching the place of two
groups attached to that stereocenter. We can also draw the mirror image of the
enantiomer, using an imaginary mirror.
Example 1: Draw the enantiomer of (R)-2-chlorobutane shown below.

Solution 1: Let's switch the places of the methyl and ethyl groups. This gives us
(S)-2-chlorobutane. (You should label the stereocenter as R or S to verify this.)

Example 2: Draw the enantiomer of (2R,3R)-tartaric acid, shown below.

Solution 2: Let's use the mirror image technique for this example. Once again, you
may wish to verify the answer by labeling the stereocenters of the mirror image
molecule as R or S.
Drawing diastereomers: Recall that stereoisomers differ in the position of the
atoms in space, and that diastereomers are stereoisomers that are not enantiomers.
Some thought on this will suggest that we can draw the diastereomer of a given
structure by inverting one or more, but not all of the stereocenters.

Example 3: Draw all the diastereomers of (2R,3R)-tartaric acid (structure shown


above).
Solution 3: How many diastereomers can there be? Recall that for a molecule with
n stereocenters, the molecule can have 2n stereoisomers (section 4.4 of the text).
Thus, (2R,3R)-tartaric acid, which has two stereocenters, can have at most 22 = 4
stereoisomers. The (2S,3S) stereoisomer cannot be a diastereomer of (2R,3R)
because both stereocenters have been inverted. This leaves two possibilities, each
of which has one stereocenter the same as the given molecule: (2R,3S) and
(2S,3R). The structures can be drawn by inverting the stereocenter that is altered.

(In this particular case, the new diastereomers are meso compounds, and are
identical. This happens when the two stereocenters have the same attachments.)

Exercises: Draw the enantiomer for each compound.


Draw a diastereomer for each structure.

Stereochemistry: Specific Rotation and Related


Calculations
Discussion: With one exception, the chemical and physical properties
of enantiomers are identical in the absence of other chiral materials. The
exception is the rotation of plane polarized light. Because plane polarized light
is effected differently by enantiomers, it can be a useful laboratory tool to quantify
the amount of an enantiomer or enantiomers present in a sample. Recall that equal
amounts of mutual enantiomers rotate plane polarized light to an equal extent but
in opposite directions. Rotation of plane polarized light occurs at the molecular
level, so the extent of the rotation is proportional to the amount of optically
active material that is present as expressed in the following equation.
[α]λT = α/lc
Where:
[α]λT = specific rotation in degrees (The correct units are deg
cm2 g-1, but are usually just given as degrees). λ is the
wavelength of light used for the observation (usually 589 nm, the
D line of a sodium lamp unless otherwise specified. This
wavelength is responsible for the orange-yellow color of the
common sodium vapor street light.), and T is the temperature
in oC. This value is characteristic for a given compound, just like
the melting point. Concentration and solvent data is included if
relevant. Example: (c = 10, CH3OH) after the rotation means the
specific rotation was determined at a concentration of 10 g ml-1 in
methanol.
α = observed rotation in degrees.
l = cell path length in decimeters (1 decimeter = 1 dm = 10
cm. A standard polarimeter tube is 1.00 dm in length.)
c = concentration in g ml-1 (the density if the sample is a pure
liquid).

Armed with these facts, and some basic concepts from general chemistry, we can
perform just about any sort of simple calculation involving optical activity and
enantiomeric composition of a sample.

Example 1: A sample of pure (S)-2-butanol was placed in a 10.0 cm polarimeter


tube. Using the D line of a sodium lamp, the observed rotation at 20oC was +104o.
The density of this compound is 0.805 g ml-1. What is the specific rotation of (S)-
2-butanol?
Solution: Plugging the numbers into [α]λT = α/lc we get: [α]λT = (+104o) / (1.00
dm) (0.805 g ml-1) = +129o. Thus we would write [α]D20 = +129o (neat). ("Neat"
refers to a liquid that has not been diluted.)

Example 2: Calculate the observed rotation of a solution of 0.5245g of (S)-1-


amino-1-phenylethane diluted to a volume of 10.0 mL with methanol at 20oC,
using the D line of a sodium lamp and a 1.00 dm tube. Specific rotation of this
material: [α]D23 = -30.0o.
Solution: Solving the specific rotation equation for observed rotation, we get α =
[α]λTlc. The sample size is 0.5245 g, but this has been diluted to 10.0 mL, so the
sample concentration is 0.5245 g/10.0 mL = 0.05245 g ml-1. Plugging in the
numbers, we get α = (-30.0o) (1.00 dm) (0.05245 g ml-1). Solving, we find α =
-1.57o.

Exercises:
a. Calculate the specific rotation of (2R, 3R)-tartaric acid based on the following
observation: A 0.856 g sample of the pure acid was diluted to 10.0 mL with water
and placed in a 1.00 dm polarimeter tube. The observed rotation using the 589 nm
line of a sodium lamp at 20.0 oC was + 1.06o.
b. What is the expected observed rotation of a 1.0 x 10-4 M methanol solution of
the potent anticancer drug paclitaxel (also called taxol)? [α]D20 = -49o (c = 1,
CH3OH). Paclitaxel has a molecular weight of 853.93 g mole-1.
c. A certain compound has a specific rotation of -43.2o (c = 5, toluene). What is the
observed rotation of a sample of 1.24 g of the enantiomer of this compound when
diluted to a concentration of 1.00 g ml-1 in the same solvent?
d. In his classic studies of stereochemistry and optical activity in organic
compounds, Pasteur measured the optical activity of many solutions. For the
naturally occurring enantiomer of tartaric acid, [α]D20 = +12.4o (c = 20, H2O). What
can be concluded about the ratio of tartaric acid enantiomers present in the solution
if the observed rotation is (i) -6.0o, or (ii) 0o?

Stereochemistry: Specific Rotation and Related Calculations


Exercise Solutions
a. 0.856 g in 10.0 mL of solution = 0.0856 g ml-1.
[α]λT = α/lc
So: [α]D20.0 = +1.06o / (1.00 dm) (0.0856 g ml-1) = +12.38o

b. [α]λT = α/lc
c = 0.0001 M = 0.0001 moles L-1 = (0.0001) (853.93 g mole-1) (0.001 L) = 0.00085
g ml-1
Assume l = 1.0 dm (standard polarimeter tube)
Solving for α: α = [α]λTlc = (-49o) (1.0 dm) (0.00085 g ml-1) = -0.042o.
c. Solutions of enantiomers of equal concentrations rotate plane polarized light to
an equal extent but in opposite directions. If a certain compound has a specific
rotation of -43.2o (c = 5, toluene) then the enantiomer has a specific rotation of
+43.2o (c = 5, toluene). From the previous question,
α = [α]λTlc = (+43.2o) (1.00 dm) (1.00 g ml-1) = +43.2o. We could have anticipated
this result without doing a calculation, as the rotation is observed under the
"standard conditions" of 1.00 dm tube and 1.00 g ml-1 concentration.
d. i) If the sample is levorotatory, then the concentration of the (-) enantiomer is
greater than the concentration of the (+) enantiomer. We cannot calculate the exact
concentration, as we do not know how much (-) enantiomer is being canceled by
the (+) enantiomer. For example, the sample might contain 1.1 g ml-1 of the (-)
enantiomer and 1.0 g ml-1 of the (+) enantiomer. This solution would have the
same rotation as a 0.1 g ml-1 solution of the (-) enantiomer alone.
ii) Enantiomers have equal but opposite rotations. An observed rotation of zero
implies the enantiomers are present in equal concentrations (a racemic mixture).

Potrebbero piacerti anche