Вы находитесь на странице: 1из 130

Chirality (chemistry)

From Wikipedia, the free encyclopedia

"L-form" redirects here. For the bacterial strains, see L-form bacteria.

Two enantiomers of a generic amino acid

(S)-Alanine (left) and (R)-alanine (right) in zwitterionic form at neutral pH

A chiral molecule /karl/ is a type of molecule that has a non-superposable mirror image. The presence of
an asymmetric carbon atom is often the feature that causes chirality in molecules.[1][2][3][4]
Achiral objects, such as atoms, are symmetrical, identical to their mirror image.
Human hands are perhaps the most universally recognized example of chirality: the left hand is a nonsuperposable mirror image of the right hand; no matter how the two hands are oriented, it is impossible for all
the major features of both hands to coincide. This difference in symmetry becomes obvious if a left-handed
glove is placed on a right hand. The term chirality is derived from the Greek word for hand, (kheir). It is a
mathematical approach to the concept of "handedness".
In chemistry, chirality usually refers to molecules. Two mirror images of a chiral molecule are
called enantiomers or optical isomers. Pairs of enantiomers are often designated as "right-" and "left-handed".
Molecular chirality is of interest because of its application to stereochemistry in inorganic chemistry, organic
chemistry, physical chemistry, biochemistry, and supramolecular chemistry.

Contents
[hide]

1 History

2 Symmetry

3 Naming conventions

3.1 By configuration: R- and S-

3.2 By optical activity: (+)- and ()- or d- and l-

3.3 By configuration: D- and L-

4 Nomenclature

5 Stereogenic centers
5.1 The identity of the stereogenic atom

6 Properties of enantiomers

7 In biology
7.1 D-Amino Acid Natural Abundance

8 Inorganic chemistry

9 Chirality of compounds with a stereogenic "lone pair"

10 See also

11 References

12 External links

History[edit]

The term optical activity is derived from the interaction of chiral materials with polarized light. In solution, the
()-form, or levorotary form, of an optical isomer rotates the plane of a beam ofpolarized light counterclockwise.
The (+)-form, or dextrorotatory form, does the opposite. The property was first observed by Jean-Baptiste
Biot in 1815,[5] and gained considerable importance in the sugar industry, analytical chemistry, and
pharmaceuticals. Louis Pasteur deduced in 1848 that this phenomenon has a molecular basis. [6] Artificial
composite materials displaying the analog of optical activity but in the microwave region were introduced
by J.C. Bose in 1898,[7] and gained considerable attention from the mid-1980s. [8] The term chirality itself was
coined by Lord Kelvin in 1894.[9] Different enantiomers or diastereomers of a compound were formerly
called optical isomers due to their different optical properties.[10]

Symmetry[edit]
The symmetry of a molecule (or any other object) determines whether it is chiral. A molecule is achiral (not
chiral) when an improper rotation, that is a combination of a rotation and a reflection in a plane, perpendicular
to the axis of rotation, results in the same molecule - see chirality (mathematics). For tetrahedral molecules, the
molecule is chiral if all four substituents are different.
A chiral molecule is not necessarily asymmetric (devoid of any symmetry element), as it can have, for
example, rotational symmetry.

Naming conventions[edit]
By configuration: R- and S-[edit]
For chemists, the R / S system is the most important nomenclature system for denoting enantiomers, which
does not involve a reference molecule such as glyceraldehyde. It labels each chiral center R or S according to
a system by which its substituents are each assigned a priority, according to the CahnIngoldPrelog priority
rules (CIP), based on atomic number. If the center is oriented so that the lowest-priority of the four is pointed
away from a viewer, the viewer will then see two possibilities: If the priority of the remaining three substituents
decreases in clockwise direction, it is labeled R (for Rectus, Latin for right), if it decreases in counterclockwise
direction, it is S (for Sinister, Latin for left).[11]
This system labels each chiral center in a molecule (and also has an extension to chiral molecules not involving
chiral centers). Thus, it has greater generality than the D/L system, and can label, for example, an (R,R) isomer
versus an (R,S) diastereomers.
The R / S system has no fixed relation to the (+)/() system. An R isomer can be either dextrorotatory or
levorotatory, depending on its exact substituents.
The R / S system also has no fixed relation to the D/L system. For example, the side-chain one
of serine contains a hydroxyl group, -OH. If a thiol group, -SH, were swapped in for it, the D/Llabeling would, by

its definition, not be affected by the substitution. But this substitution would invert the molecule's R / S labeling,
because the CIP priority of CH2OH is lower than that for CO2H but the CIP priority of CH2SH is higher than that
for CO2H.
For this reason, the D/L system remains in common use in certain areas of biochemistry, such as amino acid
and carbohydrate chemistry, because it is convenient to have the same chiral label for all of the commonly
occurring structures of a given type of structure in higher organisms. In the D/L system, they are nearly all
consistent - naturally occurring amino acids are all L, while naturally occurring carbohydrates are nearly all D. In
the R / S system, they are mostly S, but there are some common exceptions.

By optical activity: (+)- and ()- or d- and l-[edit]


An enantiomer can be named by the direction in which it rotates the plane of polarized light. If it rotates the light
clockwise (as seen by a viewer towards whom the light is traveling), that enantiomer is labeled (+). Its mirrorimage is labeled (). The (+) and () isomers have also been termed d- and l-, respectively
(for dextrorotatory and levorotatory). Naming with d- and l- is easy to confuse with D- and L- labeling and is
therefore strongly discouraged by IUPAC.[12]

By configuration: D- and L-[edit]


An optical isomer can be named by the spatial configuration of its atoms. The D/L system, not to be confused
with the d- and l-system, see above, does this by relating the molecule toglyceraldehyde. Glyceraldehyde is
chiral itself, and its two isomers are labeled

and L (typically typeset in small caps in published work). Certain

chemical manipulations can be performed on glyceraldehyde without affecting its configuration, and its
historical use for this purpose (possibly combined with its convenience as one of the smallest commonly used
chiral molecules) has resulted in its use for nomenclature. In this system, compounds are named by analogy to
glyceraldehyde, which, in general, produces unambiguous designations, but is easiest to see in the small
biomolecules similar to glyceraldehyde. One example is the amino acid alanine, which has two optical isomers,
and they are labeled according to which isomer of glyceraldehyde they come from. On the other hand, glycine,
the amino acid derived from glyceraldehyde, has no optical activity, as it is not chiral (achiral). Alanine,
however, is chiral.
The D/L labeling is unrelated to (+)/(); it does not indicate which enantiomer is dextrorotatory and which is
levorotatory. Rather, it says that the compound's stereochemistry is related to that of
the dextrorotatory or levorotatory enantiomer of glyceraldehydethe dextrorotatory isomer of glyceraldehyde
is, in fact, the D- isomer. Nine of the nineteen L-amino acids commonly found in proteins are dextrorotatory (at a
wavelength of 589 nm), and D-fructose is also referred to as levulose because it is levorotatory.
A rule of thumb for determining the D/L isomeric form of an amino acid is the "CORN" rule. The groups:
COOH, R, NH2 and H (where R is the side-chain)

are arranged around the chiral center carbon atom. With the hydrogen atom away from the viewer, if the
arrangement of the CORN groups around the carbon atom is center is clockwise, then it is the
[13]

If the arrangement is counter-clockwise, it is the

proteins. For most amino acids, the

form.

form. The L form is the usual one found in natural

form corresponds to an S absolute stereochemistry, but is R instead

for certain side-chains.

Nomenclature[edit]

Any non-racemic chiral substance is called scalemic.[14]

A chiral substance is enantiopure or homochiral when only one of two possible enantiomers is
present.

A chiral substance is enantioenriched or heterochiral when an excess of one enantiomer is present


but not to the exclusion of the other.

Enantiomeric excess or ee is a measure for how much of one enantiomer is present compared to the
other. For example, in a sample with 40% ee in R, the remaining 60% is racemic with 30% of R and
30% of S, so that the total amount of R is 70%.

Stereogenic centers[edit]
Main article: Stereogenic center
In general, chiral molecules have point chirality at a single stereogenic atom, which has four different
substituents. The two enantiomers of such compounds are said to have different absolute
configurations at this center. This center is thus stereogenic (i.e., a grouping within a molecular entity that
may be considered a focus of stereoisomerism).
Normally when an atom has four different substituents, it is chiral. However in rare cases, two of the
ligands differ from each other by being mirror images of each other. When this happens, the mirror image
of the molecule is identical to the original, and the molecule is achiral. This is called pseudochirality.
A molecule can have multiple stereogenic centers without being chiral overall if there is a symmetry
between the two (or more) stereocenters themselves. Such a molecule is called a meso compound.
It is also possible for a molecule to be chiral without having actual point chirality. Common examples
include 1,1'-bi-2-naphthol (BINOL), 1,3-dichloro-allene, and BINAP, which have axial chirality, (E)cyclooctene, which has planar chirality, and certain calixarenes and fullerenes, which have inherent
chirality.

A form of point chirality can also occur if a molecule contains a tetrahedral subunit which cannot easily
rearrange, for instance 1-bromo-1-chloro-1-fluoroadamantane and methylethylphenyltetrahedrane.
It is important to keep in mind that molecules have considerable flexibility and thus, depending on the
medium, may adopt a variety of different conformations. These various conformations are themselves
almost always chiral. When assessing chirality, a time-averaged structure is considered and for routine
compounds, one should refer to the most symmetric possible conformation.
When the optical rotation for an enantiomer is too low for practical measurement, it is said to
exhibit cryptochirality.
Even isotopic differences must be considered when examining chirality. Replacing one of the two 1H atoms
at the CH2 position of benzyl alcohol with a deuterium (2H) makes that carbon a stereocenter. The resulting
benzyl--d alcohol exists as two distinct enantiomers, which can be assigned by the usual stereochemical
naming conventions. The S enantiomer has []D = +0.715.[15]

The identity of the stereogenic atom[edit]


The stereogenic atom in chiral molecules is usually carbon, as in many biological molecules. However, it
may also be a metal atom (as in many chiral coordination compounds), nitrogen, phosphorus, or sulfur.

The
chiral
atom

Carbon

1
stereog Serine,glycera
enic
ldehyde
center

Phospho Phospho
Nitro
rus
rus
gen (phosph (phosph
ates)
ines)

Sarin, V
X

2
Adenosi
Trge
stereog Threonine, iso
ne
DIPAM
r's
enic
leucine
triphosp P
base
centers
hate

3 or
more
Metstereog enkephalin, le
enic
u-enkephalin
centers

DNA

Properties of enantiomers[edit]

Sulfur

Metal (type of metal)

Tris(bipyridine)ruthenium(II) (ruthenium),
Esomeprazole,ar cismodafinil
Dichlorobis(ethylenediamine)cobalt(III) (c
obalt), hexol (cobalt)

Dithionous acid

Normally, the two enantiomers of a molecule behave identically to each other. For example, they will
migrate with identical Rf in thin layer chromatography and have identical retention time inHPLC.
Their NMR and IR spectra are identical. However, enantiomers behave differently in the presence of other
chiral molecules or objects. For example, enantiomers do not migrate identically on chiral chromatographic
media, such as quartz or standard media that have been chirally modified. The NMR spectra of
enantiomers are affected differently by single-enantiomer chiral additives such as EuFOD.
Chiral compounds rotate plane polarized light. Each enantiomer will rotate the light in a different sense,
clockwise or counterclockwise. Molecules that do this are said to be optically active.
Characteristically, different enantiomers of chiral compounds often taste and smell differently and have
different effects as drugs see below. These effects reflect the chirality inherent in biological systems.
One chiral 'object' that interacts differently with the two enantiomers of a chiral compound is circularly
polarised light: An enantiomer will absorb left- and right-circularly polarised light to differing degrees. This
is the basis of circular dichroism (CD) spectroscopy. Usually the difference in absorptivity is relatively small
(parts per thousand). CD spectroscopy [16] is a powerful analytical technique for investigating the secondary
structure of proteins and for determining the absolute configurations of chiral compounds, in particular,
transition metal complexes. CD spectroscopy is replacing polarimetry as a method for characterising chiral
compounds, although the latter is still popular with sugar chemists.

In biology[edit]
Many biologically active molecules are chiral, including the naturally occurring amino acids (the building
blocks of proteins) and sugars. In biological systems, most of these compounds are of the same chirality:
most amino acids are L and sugars are D. Typical naturally occurring proteins, made of

amino acids, are

known as left-handed proteins, whereas D amino acids produce right-handed proteins.


The origin of this homochirality in biology is the subject of much debate.[17] Most scientists believe that
Earth life's "choice" of chirality was purely random, and that if carbon-based life forms exist elsewhere in
the universe, their chemistry could theoretically have opposite chirality. However, there is some suggestion
that early amino acids could have formed in comet dust. In this case, circularly polarised radiation (which
makes up 17% of stellar radiation) could have caused the selective destruction of one chirality of amino
acids, leading to a selection bias which ultimately resulted in all life on Earth being homochiral. [18]
Enzymes, which are chiral, often distinguish between the two enantiomers of a chiral substrate. Imagine
an enzyme as having a glove-like cavity that binds a substrate. If this glove is right-handed, then one
enantiomer will fit inside and be bound, whereas the other enantiomer will have a poor fit and is unlikely to
bind.

D-form

amino acids tend to taste sweet, whereas L-forms are usually tasteless.[19] Spearmint leaves

and caraway seeds, respectively, contain R-()-carvone and S-(+)-carvone - enantiomers of carvone.
[20]

These smell different to most people because our olfactory receptors also contain chiral molecules that

behave differently in the presence of different enantiomers.


Chirality is important in context of ordered phases as well, for example the addition of a small amount of an
optically active molecule to a nematic phase (a phase that has long range orientational order of molecules)
transforms that phase to a chiral nematic phase (or cholesteric phase). Chirality in context of such phases
in polymeric fluids has also been studied in this context. [21]

D-Amino

Acid Natural Abundance[edit]

The relative abundances of each of the different D-isomers of several amino acids have recently been
quantified by collecting experimentally reported data from the proteome across all organisms in the SwissProt database. The D-isomers observed experimentally were found to occur very rarely as shown in the
following table in the database of protein sequences containing over 187 million amino acids. [22]
D-amino

acid

# of Times Experimentally Observed

D-alanine

664

D-serine

114

D-methionine

19

D-phenylalanine

15

D-valine

D-tryptophan

D-leucine

D-asparagine

D-threonine

However, the D-isomers are not uncommon as free amino acids. Humans have special enzymes to
process then, D-amino acid oxidase and D-aspartate oxidase. D-glutamic acid, D-glutamin, andD-alanine
are also extremely common at a part of the peptidoglycan layer in the bacterial cell wall. In addition, Dserine is a neurotransmitter, and produced in humans by serine racemase.

Inorganic chemistry[edit]

Delta-ruthenium-tris(bipyridine) cation

Main article: Complex (chemistry): Isomerism


Many coordination compounds are chiral. At one time, chirality was only associated with organic chemistry,
but this misconception was overthrown by the resolution of a purely inorganic compound, hexol, by Alfred
Werner. A famous example is tris(bipyridine)ruthenium(II) complex in which the three bipyridine ligands
adopt a chiral propeller-like arrangement.[23] In this case, the Ru atom is the stereogenic center. The two
enantiomers of complexes such as [Ru(2,2-bipyridine) 3]2+ may be designated as (capital lambda, the
Greek version of "L") for a left-handed twist of the propeller described by the ligands, and (capital delta,
Greek "D") for a right-handed twist pictured.

Chirality of compounds with a stereogenic "lone pair"[edit]


When a nonbonding pair of electrons, a lone pair, occupies space, chirality can result. The effect is
pervasive in certain amines, phosphines,[24]sulfonium and oxonium ions, sulfoxides, and even carbanions.

The main requirement is that aside from the lone pair, the other three substituents differ mutually. Chiral
phosphine ligands are useful in asymmetric synthesis.

Geometric inversion among the lone pair and three bonded groups on a tetrahedral amine

Chiral amines are special in the sense that the enantiomers can rarely be separated. The energy barrier
for nitrogen inversion of the stereocenter is generally only about 30 kJ/mol, which means that the two
stereoisomers rapidly interconvert at room temperature. As a result, such chiral amines cannot be resolved
into individual enantiomers unless some of the substituents are constrained in cyclic structures as
in Trger's base.

See also[edit]

Stereochemistry for overview of stereochemistry in general

Supramolecular chirality

Chirality (physics)

Chirality (mathematics)

Pfeiffer Effect

Chemical chirality in popular fiction

BROWSE > HOME / ALPHABETICAL SEARCH / C-D / CHIRAL MOLECULE

Print This Page

Chiral Molecule

A chiral molecule is a molecule that is not superimposable on its mirror image.


eg. 1:

Molecule 1 is not superimposable on its mirror image and, therefore, is chiral.


eg. 2:

Molecule 2 is not superimposable on its mirror image and, therefore, is chiral.


An achiral molecule is a molecule that is superimposable on its mirror image.
eg. 1:

Molecule 3 is superimposable on its mirror image and, therefore, is achiral.


eg. 2:

Molecule 4 is superimposable on its mirror image and, therefore, is achiral.


Alternatively, an achiral molecule is a molecule that has at least one plane of
symmetry.
eg. 1:

The vertical plane that bisects the bromine atom and the methyl group, which is the
plane of the screen, is a plane of symmetry. Thus, 3 is achiral.
eg. 2:

The vertical plane that bisects the molecule perpendicular to the plane of the screen
is a plane of symmetry. Thus, 4 is achiral.
A chiral molecule has no plane of symmetry.
eg. 1

1 is chiral and has no plane of symmetry.

eg. 2:

2 is chiral and has no plane of symmetry.


Although relatively rare, molecules do exist that have no plane of symmetry but is
achiral.
eg:

Thus, presence of a plane of symmetry is not a foolproof method to determine


whether a molecule is chiral or achiral.

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 lefthanded 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 lefthanded 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.

Link to:

Definitions: Chiral

A molecule is chiral if it is not superimposable on its mirror image. Most chiral molecules can be
identified by their lack of a plane of symmetry or a center of symmetry. Your hand is a chiral object, as it
does not have either of these types of symmetry.

The molecule on the left has a plane of symmetry through the center carbon. This is a mirror plane; in

other words, one half of the molecule is a perfect reflection of the other half of the molecule. This
molecule is not chiral because of its mirror plane.
Molecules which are identical (superimposable) with their mirror image geometries are always optically
inactive (achiral); whereas the non-superimposability of a structure with its mirror image results
in chirality (optical activity, see below). A simple, but not always accurate test whether a molecule is
achiral or not is the presence of a mirror plane (equal to a plane of reflection or plane of symmetry,
symmetry element ) in the structure of a molecule. Compounds possessing mirror symmetry are always
optically inactive, such as, for example, cis- andtrans-1,4-disubstituted cyclohexane derivatives, or
symmetrically cis-1,2-disubstituted cyclohexanes (see also the 3D structures at 'Cycloalkanes').

In addition, compounds which posses a center of inversion (equal to a center of symmetry, symmetry
element i) are also always achiral (see the above example of -truxillic acid which is of Ci symmetry, but
does not have a mirror plane of symmetry).
However, there are molecules featuring neither a mirror plane nor a center of inversion i, but which are
still achiral. Most accurately, all molecules which have a n-fold alternating axis of symmetry (equal
to an improper rotation axis or a rotary-reflection axis, symmetry element Sn) are achiral (and thus
superimposable with their mirror images). A Snaxis is composed of two successive transformations,
first a rotation through 360/n, followed by a reflection through a plane perpendicular to that axis; neither
operation alone (rotation or reflection) is a valid symmetry operation for these molecules, but only the
combination of both. Note, that a S1 axis is identical to a simple mirror plane , and a S2 axis is equivalent
to a center of inversion i.

Molecules which do not posses a mirror plane or a center of inversion i, but a S4 axis are not very
common, the examples given below are 1,3,5,7-tetrabromo-2,4,6,8-tetramethyl-cyclooctane and 2,3,7,8tetramethyl-spiro[4.4]nonane (both S4 symmetry only). Combinations of S4 axis with other symmetry
elements are common, e.g. methane CH4 possess three S4axis along with six mirror planes , four C3 and
three C2 axis; in total the over-all symmetry of methane is described by the Td point group.

On some rare occasions molecules possessing Sn axis with n > 4 are found, the one example above is
provided by [6.5]coronane (low-energy solid-state conformation of point group S6) which also has the
symmetry elements i and C3.
NOTE: In many cases the modern drawings of chemical formulas may erroneously suggest a
compound to be of different (i.e. wrong) symmetry than it really is. In all cases, the actually
prevailing geometry (very often, but not necessarily the low-energy conformation) of a molecule
must be considered when establishing the symmetry of a molecule. Below examples are given, in
which chirality results from conformational effects, whereas chemical formulas at first sight
suggest planar conformations of molecules (see 'Helical Chirality' below). Chemical formulas are
very often helpful, but not always accurate. For example, the above formula of [6.5]coronane
virtually implies a D3d point group (symmetry elements i, three mirror planes , one C3, three C2,
and one S6 axis), but the solid-state conformation actually is of S6 symmetry (still achiral) only.
Chiral Compounds - 3D Structures
The ultimate criterion for chirality (handedness) of a molecule is the non-superimposability of a structure
with its mirror image geometry through pure translation and/or rotation only. Chiral molecules related to
each other as mutual mirror images may be separated into two enantiomers (reflection isomers, mirror
images) with identical chemical (stability and reactivity in achiral environments) and physical (scalar)
properties (melting and boiling point, spectroscopic data, etc.), except for their specific optical rotation (the
optical activity of enantiomers is of equal absolute magnitude, but of opposite sign).
Chirality of molecules may originate from configurational or conformational effects of structures.
This differentiation of configurational and conformational stereoisomers is not allways
unambiguous, but generally conformational isomers may interconvert in each other through
rotations about C-C single bonds only (this will not interconvert configurational isomers). Below,
numerous examples for the different origins of chirality of organic compounds are given.
The formal maxmimum number of configurational stereoisomers (including E/Z-isomers for double bonds)
of any compound may be calculated from the number of stereocenters (see'Asymmetric Substituted
Atoms' below) and the number of stereogenic double bonds (double bonds carrying different substitutents
at either end, E/Z-isomerism) present in the molecule:
Maxmimum number of configurational stereoisomers: Nmax = 2(n + m)
where n = the number of stereocenters and m = the number of stereogenic double bonds.
This includes E/Z-isomers of alkenes which may be regared as configurational isomers (not
interconvertible through rotation about C-C single bonds), and which are not superimposable to each
other. These isomers are not related to each other as mirror images, and thus they are in
fact diastereomers (see also below 'Two or More Asymmetric Substituted Atoms'). For
each stereocenter and stereogenic double bond present in a molecule a pair of stereoisomers may be

generated (enantiomers or diastereomers).


The actual number of different stereocenter may be smaller than the formal maximum number Nmax as
defined above if constitutional symmetry is present in the molecule (see below 'Two or More Asymmetric
Substituted Atoms' for meso-compounds). Steric strain and geometrical limitations may also reduce the
number of possible stereoisomers (e.g. double bonds in small and normal rings may only adopt Zconfiguration, or bridged and bicyclic ring systems may require certain rigid linkage geometries and
relative configurations of stereocenters - see also below for 'Substituted Adamantane Derivatives').
On the other hand, this maximum number may be exceeded if hindered rotation about C-C single bonds
results in additional stereoisomers (see below 'Biphenyls and Binaphthyls').
Chiral Compounds - Asymmetric Substituted Atoms - 3D Structures
Any molecule with a single chirality center (any atom holding a set of ligands in a spatial arrangement
which is not superimposable on its mirror image) must be chiral. This is the generalized extension of the
traditional concept of the asymmetrically substituted carbon atom (van't Hoff): any tetrahedral carbon
atom that is attached to four different entities (atoms or groups, e.g. CR 1R2R3R4) acts as a chirality center,
and the corresponding compound may be separated into enantiomers. Thus 2-bromo-butane is chiral
(four different substituents -Br, -C2H5, -CH3, and -H at C-2). This rule applies no matter how slight the
differences between the four groups are, including isotopic substitution: 1-butanol-1-d is also a chiral
compound.

The above defintion is not only restricted to tetrahedral carbon atoms, but to any other type of central
atom with an appropriate set of bound groups or ligands. Numerous sulfur derivatives exhibit pyramidal
bonding where the non-bonded electron pair located at the sulfur atom acts as a fourth ligand. In many
cases these compounds are configurationally sufficiently stable to be separated into enantiomers.

The same would apply to nitrogen derivatives (tertiary amines), but usually these compounds rapidly
interconvert through an trigonal planar transition state (pyramidal inversion) and thus prevent separation
into enantiomers. Ammonia interconverts 2*1011 times per second, and although this process is slower for
substituted amines it is still very fast at room temperature. Exceptions are provided by nitrogen atoms in
small rings such as aziridines (for which the trigonal planar transition state of inversion builds up strain
energy in the ring), or nitrogen bonded to other atoms with non-bonded electron pairs (such as oxygen).
Compounds of these types may be resolved into optically pure enantiomers.

Other examples of configurationally stable amines are bicyclic ring systems with nitrogen located at the

bridgehead positions. The geometrical restrictions operative in these ring systems may also prevent
inversion. As an typical example, Trger's base has been separated into enantiomers (see above).
Phosphorous inverts less rapid than nitrogen and arsenic still more slowly. The above rules are not only
restricted to tetrahedral centers, but also apply to octahedral and other coordination geometries of
appropriate substitution, including metal complexes and inorganic structures.

The above example of a chiral substituted ferrocene may also be classified as being planar chiral (see
below 'Planar Chirality').
Chiral Compounds - Two or More Asymmetric Substituted Atoms - 3D Structures
Compunds with two or more asymmetrically substituted atoms (chiral centers) may be optically active.
Typical examples are (2S,3S)- and (2R,3R)-2,3-dibromo-butane, or (2S,3S)- and (2R,3R)-tartaric acid
(german: Weinsure). However, if pairs of equivalent stereocenters of opposite configuration are present
in the molecule, the compounds are optically inactive (achiralmeso-compounds) as both chiral centers
neutralize each other (internal compensation). For example, (2S,3R)- and (2R,3S)-2,3-dibromo-butane
are identical (meso) as both structures can be superimposed through simple rotations and translations
only. The same applies to meso-tartaric acid.

Stereoisomers of compunds which are not not related as mirror images are called diastereoisomers.
Structures of this type cannot be interconverted into each other through translation, rotation, and mirror
symmetry operations. In contrast to enantiomers, diastereoisomers have different chemical (towards
achiral as well as chiral reagents) and physical properties (including totally independent values for their
specific optical rotation).
trans-1,2-Disubstituted cyclohexanes of C2 symmetry are chiral compounds, whereas symmetrically cis1,2-disubstituted cyclohexanes are optically inactive as they posses a mirror plane of symmetry.
However, the latter type compounds exist in chiral conformations which interconvert rapidly into each
other through chair-antichair inversions of the cyclohexane ring (see also the 3D structures at 'Achiral
Compounds' and the chapter 'Ring Pseudorotation' in the MolArch+ - Movies section of this web-site).

Chiral Compounds - Substituted Adamantane Derivatives - 3D Structures


Adamantane derivates of suitable substitution may also be chiral compounds. In these structures, not a
single atom, but an entire group (the adamantyl residue) holds four substituents (adamantane derivative
on the left below) in a spatial arrangement that causes the compound to be non-superimposable with its
mirror image geometry. The central adamantyl residue may be regarded as an extended tetrahedron.

This example demonstrates that the stereogenic center of a moleucle needs not to be located on a
specific atom. In this case, the stereogenic center is located in the center of the adamantyl cage.
Disubstituted adamantanes (note the different substitution positions and pattern) are also chiral
compounds - they do not feature a chiral center but an axis of chirality(see below 'Axial Chirality').
In fact, the above (left) adamantane derivative features four asymmetric substituted carbon atoms (the
four tertiary carbon atoms of adamantane backbone). As derived above, the formal maximum number of
stereoisomers (see abvoe 'Chiral Compounds') would be Nmax = 24 = 16. However, in this case only two
stereoisomers actually exist, as the steric strain and geometrical limitations of the adamantane resdiue
require all substituent to point towards the outside of the molecule (in fact, once the first stereocenter is
defined, the remaining three are determined by their relative configuration, and thus the number of
observed stereoisomers is 2).
Similar limitations on the number of stereoisomers are observed in almost all bi- and polycyclic systems
with small rings. Twistanes (even the unsubstituted parent hydrocarbon) are chiral. As another example,
camphor yields two stereoisomers only, although it features two stereocenters in position 1 and 4 (but the
relative configuration of both centers is constrained in the bicyclic ring system):

Chiral Compounds - Axial Chirality - 3D Structures

Some compounds which do not have asymmetrically substituted carbon atoms (or any other atom type)
may still be chiral if they feature two perpendicular planes which are not symmetry planes. If
these disymmetric (chiral) planes cannot freely rotate against each other, the corresponding compounds
are chiral. Compounds of this type are said to be axially chiral (they feture an axis of chirality instead
a center of chirality)
Typical examples are allenes in which the central atom is sp-hybridized, and the planes containing the
substitutents on either end of the double bonds are aligned perpendicular to each other (in contrast to
simple olefines where all substituents are contained in a single plane of the -bond; note the difference
to E/Z-isomerism of alkenes). For an even number of double bonds in the allene, and if neither side of it is
symmetrically substituted, these compounds are optically active and thus chiral.

Very similar arrangements are observed in chiral spiro-annelated ring systems and compounds with
exocyclic double-bonds (see examples below).

As demonstrated above, suitably disubstituted adamantane derivatives also show axial chirality (see
above 'Substituted Adamantane Derivatives').
Chiral Compounds - Biphenyls and Binaphthyls - 3D Structures
The same rules as outlined above for the allenes and spiranes apply also to biphenyl- and binaphtylderivatives. If both aromatic ring systems are asymmetrically substituted, the compounds are chiral. As
the chirality of these structures originates not from an asymmetrically substituted atom center, but from an
asymmetric axis around which rotation is hindered, these enantiomers are also called atropisomers. In the
biphenyls, the ortho-substitutents must be large enough to prevent rotation around the central single
bond; if hydrogen atoms are present in these positions the barrier of rotation may be too small to prevent
interconversion of the enantiomeric forms at room temperature and the two structures may not be
separated, or may racemize slowly on standing.

Please note, that for the biphenyls and binaphtyls derivatives the formal maximum number of
stereoisomers (see above 'Chiral Compounds') is exceeded. As these moleucles do neither feature
a stereocenter or a stereogenic double bond, no stereoisomers would be expected. Only the hindered

rotation about the central C-C single bond leads to the stereoisomerism of these compounds. Therefore,
biphenyl- and binaphtyl-derivatives are conformational stereoisomers (not configurational stereoisomers).
In particular the enentiomerically pure binaphthyl derivatives are of great use in asymmetric catalysis. For
some animations of the dynamic behavior of biphenyls and binaphthyls and the process of racemization
see the chapter 'Atropsiomers' in the MolArch+ - Movies section of this web-site.
Chiral Compounds - Helical Chirality - 3D Structures
Helices are chiral as they can exist in enantiomeric left- or right-handed forms. Typical examples for
helical strutures are provided by the helicenes (benzologues of phenanthrene). With four or more rings,
steric hinderance at both ends of these molecule prevents the formation of planar conformations, and
helicenes rather adopt non-planar, but helical and enantiomeric structures with C2 symmetry (see also the
3D structures at 'Helicenes' and the chapter 'Racemization of [8]Helicene' in the MolArch+ Movies section of this web-site).

Other examples of chiral helical structures are provided by trans-cyclooctene and suitably substituted
heptalenes. Heptalene is not planar, and its twisted structure results in chirality. Although these
conformations generally interconvert rapidly, bulky substituents may sufficiently slow down this process to
optically resolve and separate these compounds.

The above example of the chirality of (E)-cyclooctene is also example for 'Planar Chirality' (see below).
Chiral Compounds - Planar Chirality - 3D Structures
Planar chirality may arise if an appropriately substituted planar group of atoms or ring system is bridged
by a linker-chain extending into the space above or below of this plane. Commmon examples are the
planar chirality of cyclophanes or alkenes as shown below. Even (E)-cyclooctene is a planar chiral
compound (see also above 'Helical Chirality'):

The above mentioned suitably substituted ferrocenes are also planar chiral compounds (see 'Asymmetric
Substituted Atoms').
Chiral Compounds - Other Examples - 3D Structures
There are many more examples known for which chirality of molecules results from hindered rotation of
groups or spatial arrangements of chemical moieties, a few examples are listed below:

Even catenanes and molecular knots made up from achiral molecules are chiral.

For more informations on other research topics, please refer to the complete list of
publications and to the gallery of graphics and animations.
Molecules which are identical (superimposable) with their mirror image geometries are always optically
inactive (achiral); whereas the non-superimposability of a structure with its mirror image results
in chirality (optical activity, see below). A simple, but not always accurate test whether a molecule is
achiral or not is the presence of a mirror plane (equal to a plane of reflection or plane of symmetry,
symmetry element ) in the structure of a molecule. Compounds possessing mirror symmetry are always
optically inactive, such as, for example, cis- andtrans-1,4-disubstituted cyclohexane derivatives, or
symmetrically cis-1,2-disubstituted cyclohexanes (see also the 3D structures at 'Cycloalkanes').

In addition, compounds which posses a center of inversion (equal to a center of symmetry, symmetry
element i) are also always achiral (see the above example of -truxillic acid which is of Ci symmetry, but
does not have a mirror plane of symmetry).
However, there are molecules featuring neither a mirror plane nor a center of inversion i, but which are
still achiral. Most accurately, all molecules which have a n-fold alternating axis of symmetry (equal
to an improper rotation axis or a rotary-reflection axis, symmetry element Sn) are achiral (and thus
superimposable with their mirror images). A Snaxis is composed of two successive transformations,
first a rotation through 360/n, followed by a reflection through a plane perpendicular to that axis; neither
operation alone (rotation or reflection) is a valid symmetry operation for these molecules, but only the
combination of both. Note, that a S1 axis is identical to a simple mirror plane , and a S2 axis is equivalent
to a center of inversion i.

Molecules which do not posses a mirror plane or a center of inversion i, but a S4 axis are not very
common, the examples given below are 1,3,5,7-tetrabromo-2,4,6,8-tetramethyl-cyclooctane and 2,3,7,8tetramethyl-spiro[4.4]nonane (both S4 symmetry only). Combinations of S4 axis with other symmetry
elements are common, e.g. methane CH4 possess three S4axis along with six mirror planes , four C3 and
three C2 axis; in total the over-all symmetry of methane is described by the Td point group.

On some rare occasions molecules possessing Sn axis with n > 4 are found, the one example above is
provided by [6.5]coronane (low-energy solid-state conformation of point group S6) which also has the
symmetry elements i and C3.
NOTE: In many cases the modern drawings of chemical formulas may erroneously suggest a
compound to be of different (i.e. wrong) symmetry than it really is. In all cases, the actually
prevailing geometry (very often, but not necessarily the low-energy conformation) of a molecule
must be considered when establishing the symmetry of a molecule. Below examples are given, in
which chirality results from conformational effects, whereas chemical formulas at first sight

suggest planar conformations of molecules (see 'Helical Chirality' below). Chemical formulas are
very often helpful, but not always accurate. For example, the above formula of [6.5]coronane
virtually implies a D3d point group (symmetry elements i, three mirror planes , one C3, three C2,
and one S6 axis), but the solid-state conformation actually is of S6 symmetry (still achiral) only.
Chiral Compounds - 3D Structures
The ultimate criterion for chirality (handedness) of a molecule is the non-superimposability of a structure
with its mirror image geometry through pure translation and/or rotation only. Chiral molecules related to
each other as mutual mirror images may be separated into two enantiomers (reflection isomers, mirror
images) with identical chemical (stability and reactivity in achiral environments) and physical (scalar)
properties (melting and boiling point, spectroscopic data, etc.), except for their specific optical rotation (the
optical activity of enantiomers is of equal absolute magnitude, but of opposite sign).
Chirality of molecules may originate from configurational or conformational effects of structures.
This differentiation of configurational and conformational stereoisomers is not allways
unambiguous, but generally conformational isomers may interconvert in each other through
rotations about C-C single bonds only (this will not interconvert configurational isomers). Below,
numerous examples for the different origins of chirality of organic compounds are given.
The formal maxmimum number of configurational stereoisomers (including E/Z-isomers for double bonds)
of any compound may be calculated from the number of stereocenters (see'Asymmetric Substituted
Atoms' below) and the number of stereogenic double bonds (double bonds carrying different substitutents
at either end, E/Z-isomerism) present in the molecule:
Maxmimum number of configurational stereoisomers: Nmax = 2(n + m)
where n = the number of stereocenters and m = the number of stereogenic double bonds.
This includes E/Z-isomers of alkenes which may be regared as configurational isomers (not
interconvertible through rotation about C-C single bonds), and which are not superimposable to each
other. These isomers are not related to each other as mirror images, and thus they are in
fact diastereomers (see also below 'Two or More Asymmetric Substituted Atoms'). For
each stereocenter and stereogenic double bond present in a molecule a pair of stereoisomers may be
generated (enantiomers or diastereomers).
The actual number of different stereocenter may be smaller than the formal maximum number Nmax as
defined above if constitutional symmetry is present in the molecule (see below 'Two or More Asymmetric
Substituted Atoms' for meso-compounds). Steric strain and geometrical limitations may also reduce the
number of possible stereoisomers (e.g. double bonds in small and normal rings may only adopt Zconfiguration, or bridged and bicyclic ring systems may require certain rigid linkage geometries and
relative configurations of stereocenters - see also below for 'Substituted Adamantane Derivatives').
On the other hand, this maximum number may be exceeded if hindered rotation about C-C single bonds
results in additional stereoisomers (see below 'Biphenyls and Binaphthyls').
Chiral Compounds - Asymmetric Substituted Atoms - 3D Structures
Any molecule with a single chirality center (any atom holding a set of ligands in a spatial arrangement
which is not superimposable on its mirror image) must be chiral. This is the generalized extension of the
traditional concept of the asymmetrically substituted carbon atom (van't Hoff): any tetrahedral carbon
atom that is attached to four different entities (atoms or groups, e.g. CR 1R2R3R4) acts as a chirality center,
and the corresponding compound may be separated into enantiomers. Thus 2-bromo-butane is chiral

(four different substituents -Br, -C2H5, -CH3, and -H at C-2). This rule applies no matter how slight the
differences between the four groups are, including isotopic substitution: 1-butanol-1-d is also a chiral
compound.

The above defintion is not only restricted to tetrahedral carbon atoms, but to any other type of central
atom with an appropriate set of bound groups or ligands. Numerous sulfur derivatives exhibit pyramidal
bonding where the non-bonded electron pair located at the sulfur atom acts as a fourth ligand. In many
cases these compounds are configurationally sufficiently stable to be separated into enantiomers.

The same would apply to nitrogen derivatives (tertiary amines), but usually these compounds rapidly
interconvert through an trigonal planar transition state (pyramidal inversion) and thus prevent separation
into enantiomers. Ammonia interconverts 2*1011 times per second, and although this process is slower for
substituted amines it is still very fast at room temperature. Exceptions are provided by nitrogen atoms in
small rings such as aziridines (for which the trigonal planar transition state of inversion builds up strain
energy in the ring), or nitrogen bonded to other atoms with non-bonded electron pairs (such as oxygen).
Compounds of these types may be resolved into optically pure enantiomers.

Other examples of configurationally stable amines are bicyclic ring systems with nitrogen located at the
bridgehead positions. The geometrical restrictions operative in these ring systems may also prevent
inversion. As an typical example, Trger's base has been separated into enantiomers (see above).
Phosphorous inverts less rapid than nitrogen and arsenic still more slowly. The above rules are not only
restricted to tetrahedral centers, but also apply to octahedral and other coordination geometries of
appropriate substitution, including metal complexes and inorganic structures.

The above example of a chiral substituted ferrocene may also be classified as being planar chiral (see
below 'Planar Chirality').
Chiral Compounds - Two or More Asymmetric Substituted Atoms - 3D Structures

Compunds with two or more asymmetrically substituted atoms (chiral centers) may be optically active.
Typical examples are (2S,3S)- and (2R,3R)-2,3-dibromo-butane, or (2S,3S)- and (2R,3R)-tartaric acid
(german: Weinsure). However, if pairs of equivalent stereocenters of opposite configuration are present
in the molecule, the compounds are optically inactive (achiralmeso-compounds) as both chiral centers
neutralize each other (internal compensation). For example, (2S,3R)- and (2R,3S)-2,3-dibromo-butane
are identical (meso) as both structures can be superimposed through simple rotations and translations
only. The same applies to meso-tartaric acid.

Stereoisomers of compunds which are not not related as mirror images are called diastereoisomers.
Structures of this type cannot be interconverted into each other through translation, rotation, and mirror
symmetry operations. In contrast to enantiomers, diastereoisomers have different chemical (towards
achiral as well as chiral reagents) and physical properties (including totally independent values for their
specific optical rotation).
trans-1,2-Disubstituted cyclohexanes of C2 symmetry are chiral compounds, whereas symmetrically cis1,2-disubstituted cyclohexanes are optically inactive as they posses a mirror plane of symmetry.
However, the latter type compounds exist in chiral conformations which interconvert rapidly into each
other through chair-antichair inversions of the cyclohexane ring (see also the 3D structures at 'Achiral
Compounds' and the chapter 'Ring Pseudorotation' in the MolArch+ - Movies section of this web-site).

Chiral Compounds - Substituted Adamantane Derivatives - 3D Structures


Adamantane derivates of suitable substitution may also be chiral compounds. In these structures, not a
single atom, but an entire group (the adamantyl residue) holds four substituents (adamantane derivative
on the left below) in a spatial arrangement that causes the compound to be non-superimposable with its
mirror image geometry. The central adamantyl residue may be regarded as an extended tetrahedron.

This example demonstrates that the stereogenic center of a moleucle needs not to be located on a
specific atom. In this case, the stereogenic center is located in the center of the adamantyl cage.
Disubstituted adamantanes (note the different substitution positions and pattern) are also chiral
compounds - they do not feature a chiral center but an axis of chirality(see below 'Axial Chirality').
In fact, the above (left) adamantane derivative features four asymmetric substituted carbon atoms (the
four tertiary carbon atoms of adamantane backbone). As derived above, the formal maximum number of
stereoisomers (see abvoe 'Chiral Compounds') would be Nmax = 24 = 16. However, in this case only two
stereoisomers actually exist, as the steric strain and geometrical limitations of the adamantane resdiue
require all substituent to point towards the outside of the molecule (in fact, once the first stereocenter is
defined, the remaining three are determined by their relative configuration, and thus the number of
observed stereoisomers is 2).
Similar limitations on the number of stereoisomers are observed in almost all bi- and polycyclic systems
with small rings. Twistanes (even the unsubstituted parent hydrocarbon) are chiral. As another example,
camphor yields two stereoisomers only, although it features two stereocenters in position 1 and 4 (but the
relative configuration of both centers is constrained in the bicyclic ring system):

Chiral Compounds - Axial Chirality - 3D Structures


Some compounds which do not have asymmetrically substituted carbon atoms (or any other atom type)
may still be chiral if they feature two perpendicular planes which are not symmetry planes. If
these disymmetric (chiral) planes cannot freely rotate against each other, the corresponding compounds
are chiral. Compounds of this type are said to be axially chiral (they feture an axis of chirality instead
a center of chirality)
Typical examples are allenes in which the central atom is sp-hybridized, and the planes containing the
substitutents on either end of the double bonds are aligned perpendicular to each other (in contrast to
simple olefines where all substituents are contained in a single plane of the -bond; note the difference
to E/Z-isomerism of alkenes). For an even number of double bonds in the allene, and if neither side of it is
symmetrically substituted, these compounds are optically active and thus chiral.

Very similar arrangements are observed in chiral spiro-annelated ring systems and compounds with
exocyclic double-bonds (see examples below).

As demonstrated above, suitably disubstituted adamantane derivatives also show axial chirality (see
above 'Substituted Adamantane Derivatives').
Chiral Compounds - Biphenyls and Binaphthyls - 3D Structures
The same rules as outlined above for the allenes and spiranes apply also to biphenyl- and binaphtylderivatives. If both aromatic ring systems are asymmetrically substituted, the compounds are chiral. As
the chirality of these structures originates not from an asymmetrically substituted atom center, but from an
asymmetric axis around which rotation is hindered, these enantiomers are also called atropisomers. In the
biphenyls, the ortho-substitutents must be large enough to prevent rotation around the central single
bond; if hydrogen atoms are present in these positions the barrier of rotation may be too small to prevent
interconversion of the enantiomeric forms at room temperature and the two structures may not be
separated, or may racemize slowly on standing.

Please note, that for the biphenyls and binaphtyls derivatives the formal maximum number of
stereoisomers (see above 'Chiral Compounds') is exceeded. As these moleucles do neither feature
a stereocenter or a stereogenic double bond, no stereoisomers would be expected. Only the hindered
rotation about the central C-C single bond leads to the stereoisomerism of these compounds. Therefore,
biphenyl- and binaphtyl-derivatives are conformational stereoisomers (not configurational stereoisomers).
In particular the enentiomerically pure binaphthyl derivatives are of great use in asymmetric catalysis. For
some animations of the dynamic behavior of biphenyls and binaphthyls and the process of racemization
see the chapter 'Atropsiomers' in the MolArch+ - Movies section of this web-site.
Chiral Compounds - Helical Chirality - 3D Structures
Helices are chiral as they can exist in enantiomeric left- or right-handed forms. Typical examples for
helical strutures are provided by the helicenes (benzologues of phenanthrene). With four or more rings,
steric hinderance at both ends of these molecule prevents the formation of planar conformations, and
helicenes rather adopt non-planar, but helical and enantiomeric structures with C2 symmetry (see also the
3D structures at 'Helicenes' and the chapter 'Racemization of [8]Helicene' in the MolArch+ -

Movies section of this web-site).

Other examples of chiral helical structures are provided by trans-cyclooctene and suitably substituted
heptalenes. Heptalene is not planar, and its twisted structure results in chirality. Although these
conformations generally interconvert rapidly, bulky substituents may sufficiently slow down this process to
optically resolve and separate these compounds.

The above example of the chirality of (E)-cyclooctene is also example for 'Planar Chirality' (see below).
Chiral Compounds - Planar Chirality - 3D Structures
Planar chirality may arise if an appropriately substituted planar group of atoms or ring system is bridged
by a linker-chain extending into the space above or below of this plane. Commmon examples are the
planar chirality of cyclophanes or alkenes as shown below. Even (E)-cyclooctene is a planar chiral
compound (see also above 'Helical Chirality'):

The above mentioned suitably substituted ferrocenes are also planar chiral compounds (see 'Asymmetric
Substituted Atoms').
Chiral Compounds - Other Examples - 3D Structures
There are many more examples known for which chirality of molecules results from hindered rotation of
groups or spatial arrangements of chemical moieties, a few examples are listed below:

Even catenanes and molecular knots made up from achiral molecules are chiral.

For more informations on other research topics, please refer to the complete list of
publications and to the gallery of graphics and animations.
Molecules which are identical (superimposable) with their mirror image geometries are always optically
inactive (achiral); whereas the non-superimposability of a structure with its mirror image results
in chirality (optical activity, see below). A simple, but not always accurate test whether a molecule is
achiral or not is the presence of a mirror plane (equal to a plane of reflection or plane of symmetry,
symmetry element ) in the structure of a molecule. Compounds possessing mirror symmetry are always
optically inactive, such as, for example, cis- andtrans-1,4-disubstituted cyclohexane derivatives, or
symmetrically cis-1,2-disubstituted cyclohexanes (see also the 3D structures at 'Cycloalkanes').

In addition, compounds which posses a center of inversion (equal to a center of symmetry, symmetry
element i) are also always achiral (see the above example of -truxillic acid which is of Ci symmetry, but
does not have a mirror plane of symmetry).
However, there are molecules featuring neither a mirror plane nor a center of inversion i, but which are
still achiral. Most accurately, all molecules which have a n-fold alternating axis of symmetry (equal
to an improper rotation axis or a rotary-reflection axis, symmetry element Sn) are achiral (and thus
superimposable with their mirror images). A Snaxis is composed of two successive transformations,
first a rotation through 360/n, followed by a reflection through a plane perpendicular to that axis; neither
operation alone (rotation or reflection) is a valid symmetry operation for these molecules, but only the
combination of both. Note, that a S1 axis is identical to a simple mirror plane , and a S2 axis is equivalent

to a center of inversion i.

Molecules which do not posses a mirror plane or a center of inversion i, but a S4 axis are not very
common, the examples given below are 1,3,5,7-tetrabromo-2,4,6,8-tetramethyl-cyclooctane and 2,3,7,8tetramethyl-spiro[4.4]nonane (both S4 symmetry only). Combinations of S4 axis with other symmetry
elements are common, e.g. methane CH4 possess three S4axis along with six mirror planes , four C3 and
three C2 axis; in total the over-all symmetry of methane is described by the Td point group.

On some rare occasions molecules possessing Sn axis with n > 4 are found, the one example above is
provided by [6.5]coronane (low-energy solid-state conformation of point group S6) which also has the
symmetry elements i and C3.
NOTE: In many cases the modern drawings of chemical formulas may erroneously suggest a
compound to be of different (i.e. wrong) symmetry than it really is. In all cases, the actually
prevailing geometry (very often, but not necessarily the low-energy conformation) of a molecule
must be considered when establishing the symmetry of a molecule. Below examples are given, in
which chirality results from conformational effects, whereas chemical formulas at first sight
suggest planar conformations of molecules (see 'Helical Chirality' below). Chemical formulas are
very often helpful, but not always accurate. For example, the above formula of [6.5]coronane
virtually implies a D3d point group (symmetry elements i, three mirror planes , one C3, three C2,
and one S6 axis), but the solid-state conformation actually is of S6 symmetry (still achiral) only.
Chiral Compounds - 3D Structures
The ultimate criterion for chirality (handedness) of a molecule is the non-superimposability of a structure
with its mirror image geometry through pure translation and/or rotation only. Chiral molecules related to
each other as mutual mirror images may be separated into two enantiomers (reflection isomers, mirror
images) with identical chemical (stability and reactivity in achiral environments) and physical (scalar)
properties (melting and boiling point, spectroscopic data, etc.), except for their specific optical rotation (the
optical activity of enantiomers is of equal absolute magnitude, but of opposite sign).
Chirality of molecules may originate from configurational or conformational effects of structures.
This differentiation of configurational and conformational stereoisomers is not allways
unambiguous, but generally conformational isomers may interconvert in each other through
rotations about C-C single bonds only (this will not interconvert configurational isomers). Below,

numerous examples for the different origins of chirality of organic compounds are given.
The formal maxmimum number of configurational stereoisomers (including E/Z-isomers for double bonds)
of any compound may be calculated from the number of stereocenters (see'Asymmetric Substituted
Atoms' below) and the number of stereogenic double bonds (double bonds carrying different substitutents
at either end, E/Z-isomerism) present in the molecule:
Maxmimum number of configurational stereoisomers: Nmax = 2(n + m)
where n = the number of stereocenters and m = the number of stereogenic double bonds.
This includes E/Z-isomers of alkenes which may be regared as configurational isomers (not
interconvertible through rotation about C-C single bonds), and which are not superimposable to each
other. These isomers are not related to each other as mirror images, and thus they are in
fact diastereomers (see also below 'Two or More Asymmetric Substituted Atoms'). For
each stereocenter and stereogenic double bond present in a molecule a pair of stereoisomers may be
generated (enantiomers or diastereomers).
The actual number of different stereocenter may be smaller than the formal maximum number Nmax as
defined above if constitutional symmetry is present in the molecule (see below 'Two or More Asymmetric
Substituted Atoms' for meso-compounds). Steric strain and geometrical limitations may also reduce the
number of possible stereoisomers (e.g. double bonds in small and normal rings may only adopt Zconfiguration, or bridged and bicyclic ring systems may require certain rigid linkage geometries and
relative configurations of stereocenters - see also below for 'Substituted Adamantane Derivatives').
On the other hand, this maximum number may be exceeded if hindered rotation about C-C single bonds
results in additional stereoisomers (see below 'Biphenyls and Binaphthyls').
Chiral Compounds - Asymmetric Substituted Atoms - 3D Structures
Any molecule with a single chirality center (any atom holding a set of ligands in a spatial arrangement
which is not superimposable on its mirror image) must be chiral. This is the generalized extension of the
traditional concept of the asymmetrically substituted carbon atom (van't Hoff): any tetrahedral carbon
atom that is attached to four different entities (atoms or groups, e.g. CR 1R2R3R4) acts as a chirality center,
and the corresponding compound may be separated into enantiomers. Thus 2-bromo-butane is chiral
(four different substituents -Br, -C2H5, -CH3, and -H at C-2). This rule applies no matter how slight the
differences between the four groups are, including isotopic substitution: 1-butanol-1-d is also a chiral
compound.

The above defintion is not only restricted to tetrahedral carbon atoms, but to any other type of central
atom with an appropriate set of bound groups or ligands. Numerous sulfur derivatives exhibit pyramidal
bonding where the non-bonded electron pair located at the sulfur atom acts as a fourth ligand. In many
cases these compounds are configurationally sufficiently stable to be separated into enantiomers.

The same would apply to nitrogen derivatives (tertiary amines), but usually these compounds rapidly
interconvert through an trigonal planar transition state (pyramidal inversion) and thus prevent separation
into enantiomers. Ammonia interconverts 2*1011 times per second, and although this process is slower for
substituted amines it is still very fast at room temperature. Exceptions are provided by nitrogen atoms in
small rings such as aziridines (for which the trigonal planar transition state of inversion builds up strain
energy in the ring), or nitrogen bonded to other atoms with non-bonded electron pairs (such as oxygen).
Compounds of these types may be resolved into optically pure enantiomers.

Other examples of configurationally stable amines are bicyclic ring systems with nitrogen located at the
bridgehead positions. The geometrical restrictions operative in these ring systems may also prevent
inversion. As an typical example, Trger's base has been separated into enantiomers (see above).
Phosphorous inverts less rapid than nitrogen and arsenic still more slowly. The above rules are not only
restricted to tetrahedral centers, but also apply to octahedral and other coordination geometries of
appropriate substitution, including metal complexes and inorganic structures.

The above example of a chiral substituted ferrocene may also be classified as being planar chiral (see
below 'Planar Chirality').
Chiral Compounds - Two or More Asymmetric Substituted Atoms - 3D Structures
Compunds with two or more asymmetrically substituted atoms (chiral centers) may be optically active.
Typical examples are (2S,3S)- and (2R,3R)-2,3-dibromo-butane, or (2S,3S)- and (2R,3R)-tartaric acid
(german: Weinsure). However, if pairs of equivalent stereocenters of opposite configuration are present
in the molecule, the compounds are optically inactive (achiralmeso-compounds) as both chiral centers
neutralize each other (internal compensation). For example, (2S,3R)- and (2R,3S)-2,3-dibromo-butane
are identical (meso) as both structures can be superimposed through simple rotations and translations
only. The same applies to meso-tartaric acid.

Stereoisomers of compunds which are not not related as mirror images are called diastereoisomers.
Structures of this type cannot be interconverted into each other through translation, rotation, and mirror
symmetry operations. In contrast to enantiomers, diastereoisomers have different chemical (towards
achiral as well as chiral reagents) and physical properties (including totally independent values for their
specific optical rotation).
trans-1,2-Disubstituted cyclohexanes of C2 symmetry are chiral compounds, whereas symmetrically cis1,2-disubstituted cyclohexanes are optically inactive as they posses a mirror plane of symmetry.
However, the latter type compounds exist in chiral conformations which interconvert rapidly into each
other through chair-antichair inversions of the cyclohexane ring (see also the 3D structures at 'Achiral
Compounds' and the chapter 'Ring Pseudorotation' in the MolArch+ - Movies section of this web-site).

Chiral Compounds - Substituted Adamantane Derivatives - 3D Structures


Adamantane derivates of suitable substitution may also be chiral compounds. In these structures, not a
single atom, but an entire group (the adamantyl residue) holds four substituents (adamantane derivative
on the left below) in a spatial arrangement that causes the compound to be non-superimposable with its
mirror image geometry. The central adamantyl residue may be regarded as an extended tetrahedron.

This example demonstrates that the stereogenic center of a moleucle needs not to be located on a
specific atom. In this case, the stereogenic center is located in the center of the adamantyl cage.
Disubstituted adamantanes (note the different substitution positions and pattern) are also chiral
compounds - they do not feature a chiral center but an axis of chirality(see below 'Axial Chirality').
In fact, the above (left) adamantane derivative features four asymmetric substituted carbon atoms (the
four tertiary carbon atoms of adamantane backbone). As derived above, the formal maximum number of
stereoisomers (see abvoe 'Chiral Compounds') would be Nmax = 24 = 16. However, in this case only two
stereoisomers actually exist, as the steric strain and geometrical limitations of the adamantane resdiue
require all substituent to point towards the outside of the molecule (in fact, once the first stereocenter is
defined, the remaining three are determined by their relative configuration, and thus the number of
observed stereoisomers is 2).
Similar limitations on the number of stereoisomers are observed in almost all bi- and polycyclic systems
with small rings. Twistanes (even the unsubstituted parent hydrocarbon) are chiral. As another example,

camphor yields two stereoisomers only, although it features two stereocenters in position 1 and 4 (but the
relative configuration of both centers is constrained in the bicyclic ring system):

Chiral Compounds - Axial Chirality - 3D Structures


Some compounds which do not have asymmetrically substituted carbon atoms (or any other atom type)
may still be chiral if they feature two perpendicular planes which are not symmetry planes. If
these disymmetric (chiral) planes cannot freely rotate against each other, the corresponding compounds
are chiral. Compounds of this type are said to be axially chiral (they feture an axis of chirality instead
a center of chirality)
Typical examples are allenes in which the central atom is sp-hybridized, and the planes containing the
substitutents on either end of the double bonds are aligned perpendicular to each other (in contrast to
simple olefines where all substituents are contained in a single plane of the -bond; note the difference
to E/Z-isomerism of alkenes). For an even number of double bonds in the allene, and if neither side of it is
symmetrically substituted, these compounds are optically active and thus chiral.

Very similar arrangements are observed in chiral spiro-annelated ring systems and compounds with
exocyclic double-bonds (see examples below).

As demonstrated above, suitably disubstituted adamantane derivatives also show axial chirality (see
above 'Substituted Adamantane Derivatives').
Chiral Compounds - Biphenyls and Binaphthyls - 3D Structures
The same rules as outlined above for the allenes and spiranes apply also to biphenyl- and binaphtylderivatives. If both aromatic ring systems are asymmetrically substituted, the compounds are chiral. As
the chirality of these structures originates not from an asymmetrically substituted atom center, but from an
asymmetric axis around which rotation is hindered, these enantiomers are also called atropisomers. In the
biphenyls, the ortho-substitutents must be large enough to prevent rotation around the central single
bond; if hydrogen atoms are present in these positions the barrier of rotation may be too small to prevent
interconversion of the enantiomeric forms at room temperature and the two structures may not be

separated, or may racemize slowly on standing.

Please note, that for the biphenyls and binaphtyls derivatives the formal maximum number of
stereoisomers (see above 'Chiral Compounds') is exceeded. As these moleucles do neither feature
a stereocenter or a stereogenic double bond, no stereoisomers would be expected. Only the hindered
rotation about the central C-C single bond leads to the stereoisomerism of these compounds. Therefore,
biphenyl- and binaphtyl-derivatives are conformational stereoisomers (not configurational stereoisomers).
In particular the enentiomerically pure binaphthyl derivatives are of great use in asymmetric catalysis. For
some animations of the dynamic behavior of biphenyls and binaphthyls and the process of racemization
see the chapter 'Atropsiomers' in the MolArch+ - Movies section of this web-site.
Chiral Compounds - Helical Chirality - 3D Structures
Helices are chiral as they can exist in enantiomeric left- or right-handed forms. Typical examples for
helical strutures are provided by the helicenes (benzologues of phenanthrene). With four or more rings,
steric hinderance at both ends of these molecule prevents the formation of planar conformations, and
helicenes rather adopt non-planar, but helical and enantiomeric structures with C2 symmetry (see also the
3D structures at 'Helicenes' and the chapter 'Racemization of [8]Helicene' in the MolArch+ Movies section of this web-site).

Other examples of chiral helical structures are provided by trans-cyclooctene and suitably substituted
heptalenes. Heptalene is not planar, and its twisted structure results in chirality. Although these
conformations generally interconvert rapidly, bulky substituents may sufficiently slow down this process to
optically resolve and separate these compounds.

The above example of the chirality of (E)-cyclooctene is also example for 'Planar Chirality' (see below).
Chiral Compounds - Planar Chirality - 3D Structures
Planar chirality may arise if an appropriately substituted planar group of atoms or ring system is bridged

by a linker-chain extending into the space above or below of this plane. Commmon examples are the
planar chirality of cyclophanes or alkenes as shown below. Even (E)-cyclooctene is a planar chiral
compound (see also above 'Helical Chirality'):

The above mentioned suitably substituted ferrocenes are also planar chiral compounds (see 'Asymmetric
Substituted Atoms').
Chiral Compounds - Other Examples - 3D Structures
There are many more examples known for which chirality of molecules results from hindered rotation of
groups or spatial arrangements of chemical moieties, a few examples are listed below:

Even catenanes and molecular knots made up from achiral molecules are chiral.

For more informations on other research topics, please refer to the complete list of
publications and to the gallery of graphics and animations.
Compunds with two or more asymmetrically substituted atoms (chiral centers) may be optically active.
Typical examples are (2S,3S)- and (2R,3R)-2,3-dibromo-butane, or (2S,3S)- and (2R,3R)-tartaric acid

(german: Weinsure). However, if pairs of equivalent stereocenters of opposite configuration are present
in the molecule, the compounds are optically inactive (achiralmeso-compounds) as both chiral centers
neutralize each other (internal compensation). For example, (2S,3R)- and (2R,3S)-2,3-dibromo-butane
are identical (meso) as both structures can be superimposed through simple rotations and translations
only. The same applies to meso-tartaric acid.

Stereoisomers of compunds which are not not related as mirror images are called diastereoisomers.
Structures of this type cannot be interconverted into each other through translation, rotation, and mirror
symmetry operations. In contrast to enantiomers, diastereoisomers have different chemical (towards
achiral as well as chiral reagents) and physical properties (including totally independent values for their
specific optical rotation).
trans-1,2-Disubstituted cyclohexanes of C2 symmetry are chiral compounds, whereas symmetrically cis1,2-disubstituted cyclohexanes are optically inactive as they posses a mirror plane of symmetry.
However, the latter type compounds exist in chiral conformations which interconvert rapidly into each
other through chair-antichair inversions of the cyclohexane ring (see also the 3D structures at 'Achiral
Compounds' and the chapter 'Ring Pseudorotation' in the MolArch+ - Movies section of this web-site).

Chiral Compounds - Substituted Adamantane Derivatives - 3D Structures


Adamantane derivates of suitable substitution may also be chiral compounds. In these structures, not a
single atom, but an entire group (the adamantyl residue) holds four substituents (adamantane derivative
on the left below) in a spatial arrangement that causes the compound to be non-superimposable with its
mirror image geometry. The central adamantyl residue may be regarded as an extended tetrahedron.

This example demonstrates that the stereogenic center of a moleucle needs not to be located on a
specific atom. In this case, the stereogenic center is located in the center of the adamantyl cage.
Disubstituted adamantanes (note the different substitution positions and pattern) are also chiral
compounds - they do not feature a chiral center but an axis of chirality(see below 'Axial Chirality').
In fact, the above (left) adamantane derivative features four asymmetric substituted carbon atoms (the
four tertiary carbon atoms of adamantane backbone). As derived above, the formal maximum number of
stereoisomers (see abvoe 'Chiral Compounds') would be Nmax = 24 = 16. However, in this case only two
stereoisomers actually exist, as the steric strain and geometrical limitations of the adamantane resdiue
require all substituent to point towards the outside of the molecule (in fact, once the first stereocenter is
defined, the remaining three are determined by their relative configuration, and thus the number of
observed stereoisomers is 2).
Similar limitations on the number of stereoisomers are observed in almost all bi- and polycyclic systems
with small rings. Twistanes (even the unsubstituted parent hydrocarbon) are chiral. As another example,
camphor yields two stereoisomers only, although it features two stereocenters in position 1 and 4 (but the
relative configuration of both centers is constrained in the bicyclic ring system):

Chiral Compounds - Axial Chirality - 3D Structures


Some compounds which do not have asymmetrically substituted carbon atoms (or any other atom type)
may still be chiral if they feature two perpendicular planes which are not symmetry planes. If
these disymmetric (chiral) planes cannot freely rotate against each other, the corresponding compounds
are chiral. Compounds of this type are said to be axially chiral (they feture an axis of chirality instead
a center of chirality)
Typical examples are allenes in which the central atom is sp-hybridized, and the planes containing the
substitutents on either end of the double bonds are aligned perpendicular to each other (in contrast to
simple olefines where all substituents are contained in a single plane of the -bond; note the difference
to E/Z-isomerism of alkenes). For an even number of double bonds in the allene, and if neither side of it is
symmetrically substituted, these compounds are optically active and thus chiral.

Very similar arrangements are observed in chiral spiro-annelated ring systems and compounds with
exocyclic double-bonds (see examples below).

As demonstrated above, suitably disubstituted adamantane derivatives also show axial chirality (see
above 'Substituted Adamantane Derivatives').
Chiral Compounds - Biphenyls and Binaphthyls - 3D Structures
The same rules as outlined above for the allenes and spiranes apply also to biphenyl- and binaphtylderivatives. If both aromatic ring systems are asymmetrically substituted, the compounds are chiral. As
the chirality of these structures originates not from an asymmetrically substituted atom center, but from an
asymmetric axis around which rotation is hindered, these enantiomers are also called atropisomers. In the
biphenyls, the ortho-substitutents must be large enough to prevent rotation around the central single
bond; if hydrogen atoms are present in these positions the barrier of rotation may be too small to prevent
interconversion of the enantiomeric forms at room temperature and the two structures may not be
separated, or may racemize slowly on standing.

Please note, that for the biphenyls and binaphtyls derivatives the formal maximum number of
stereoisomers (see above 'Chiral Compounds') is exceeded. As these moleucles do neither feature
a stereocenter or a stereogenic double bond, no stereoisomers would be expected. Only the hindered
rotation about the central C-C single bond leads to the stereoisomerism of these compounds. Therefore,
biphenyl- and binaphtyl-derivatives are conformational stereoisomers (not configurational stereoisomers).
In particular the enentiomerically pure binaphthyl derivatives are of great use in asymmetric catalysis. For
some animations of the dynamic behavior of biphenyls and binaphthyls and the process of racemization
see the chapter 'Atropsiomers' in the MolArch+ - Movies section of this web-site.
Chiral Compounds - Helical Chirality - 3D Structures
Helices are chiral as they can exist in enantiomeric left- or right-handed forms. Typical examples for
helical strutures are provided by the helicenes (benzologues of phenanthrene). With four or more rings,
steric hinderance at both ends of these molecule prevents the formation of planar conformations, and
helicenes rather adopt non-planar, but helical and enantiomeric structures with C2 symmetry (see also the
3D structures at 'Helicenes' and the chapter 'Racemization of [8]Helicene' in the MolArch+ Movies section of this web-site).

Other examples of chiral helical structures are provided by trans-cyclooctene and suitably substituted
heptalenes. Heptalene is not planar, and its twisted structure results in chirality. Although these
conformations generally interconvert rapidly, bulky substituents may sufficiently slow down this process to
optically resolve and separate these compounds.

The above example of the chirality of (E)-cyclooctene is also example for 'Planar Chirality' (see below).
Chiral Compounds - Planar Chirality - 3D Structures
Planar chirality may arise if an appropriately substituted planar group of atoms or ring system is bridged
by a linker-chain extending into the space above or below of this plane. Commmon examples are the
planar chirality of cyclophanes or alkenes as shown below. Even (E)-cyclooctene is a planar chiral
compound (see also above 'Helical Chirality'):

The above mentioned suitably substituted ferrocenes are also planar chiral compounds (see 'Asymmetric
Substituted Atoms').
Chiral Compounds - Other Examples - 3D Structures
There are many more examples known for which chirality of molecules results from hindered rotation of
groups or spatial arrangements of chemical moieties, a few examples are listed below:

Even catenanes and molecular knots made up from achiral molecules are chiral.

For more informations on other research topics, please refer to the complete list of
publications and to the gallery of graphics and animations.
Compunds with two or more asymmetrically substituted atoms (chiral centers) may be optically active.
Typical examples are (2S,3S)- and (2R,3R)-2,3-dibromo-butane, or (2S,3S)- and (2R,3R)-tartaric acid
(german: Weinsure). However, if pairs of equivalent stereocenters of opposite configuration are present
in the molecule, the compounds are optically inactive (achiralmeso-compounds) as both chiral centers
neutralize each other (internal compensation). For example, (2S,3R)- and (2R,3S)-2,3-dibromo-butane
are identical (meso) as both structures can be superimposed through simple rotations and translations
only. The same applies to meso-tartaric acid.

Stereoisomers of compunds which are not not related as mirror images are called diastereoisomers.
Structures of this type cannot be interconverted into each other through translation, rotation, and mirror
symmetry operations. In contrast to enantiomers, diastereoisomers have different chemical (towards
achiral as well as chiral reagents) and physical properties (including totally independent values for their
specific optical rotation).
trans-1,2-Disubstituted cyclohexanes of C2 symmetry are chiral compounds, whereas symmetrically cis1,2-disubstituted cyclohexanes are optically inactive as they posses a mirror plane of symmetry.
However, the latter type compounds exist in chiral conformations which interconvert rapidly into each
other through chair-antichair inversions of the cyclohexane ring (see also the 3D structures at 'Achiral

Compounds' and the chapter 'Ring Pseudorotation' in the MolArch+ - Movies section of this web-site).

Chiral Compounds - Substituted Adamantane Derivatives - 3D Structures


Adamantane derivates of suitable substitution may also be chiral compounds. In these structures, not a
single atom, but an entire group (the adamantyl residue) holds four substituents (adamantane derivative
on the left below) in a spatial arrangement that causes the compound to be non-superimposable with its
mirror image geometry. The central adamantyl residue may be regarded as an extended tetrahedron.

This example demonstrates that the stereogenic center of a moleucle needs not to be located on a
specific atom. In this case, the stereogenic center is located in the center of the adamantyl cage.
Disubstituted adamantanes (note the different substitution positions and pattern) are also chiral
compounds - they do not feature a chiral center but an axis of chirality(see below 'Axial Chirality').
In fact, the above (left) adamantane derivative features four asymmetric substituted carbon atoms (the
four tertiary carbon atoms of adamantane backbone). As derived above, the formal maximum number of
stereoisomers (see abvoe 'Chiral Compounds') would be Nmax = 24 = 16. However, in this case only two
stereoisomers actually exist, as the steric strain and geometrical limitations of the adamantane resdiue
require all substituent to point towards the outside of the molecule (in fact, once the first stereocenter is
defined, the remaining three are determined by their relative configuration, and thus the number of
observed stereoisomers is 2).
Similar limitations on the number of stereoisomers are observed in almost all bi- and polycyclic systems
with small rings. Twistanes (even the unsubstituted parent hydrocarbon) are chiral. As another example,
camphor yields two stereoisomers only, although it features two stereocenters in position 1 and 4 (but the
relative configuration of both centers is constrained in the bicyclic ring system):

Chiral Compounds - Axial Chirality - 3D Structures


Some compounds which do not have asymmetrically substituted carbon atoms (or any other atom type)
may still be chiral if they feature two perpendicular planes which are not symmetry planes. If
these disymmetric (chiral) planes cannot freely rotate against each other, the corresponding compounds
are chiral. Compounds of this type are said to be axially chiral (they feture an axis of chirality instead
a center of chirality)
Typical examples are allenes in which the central atom is sp-hybridized, and the planes containing the
substitutents on either end of the double bonds are aligned perpendicular to each other (in contrast to
simple olefines where all substituents are contained in a single plane of the -bond; note the difference
to E/Z-isomerism of alkenes). For an even number of double bonds in the allene, and if neither side of it is
symmetrically substituted, these compounds are optically active and thus chiral.

Very similar arrangements are observed in chiral spiro-annelated ring systems and compounds with
exocyclic double-bonds (see examples below).

As demonstrated above, suitably disubstituted adamantane derivatives also show axial chirality (see
above 'Substituted Adamantane Derivatives').
Chiral Compounds - Biphenyls and Binaphthyls - 3D Structures
The same rules as outlined above for the allenes and spiranes apply also to biphenyl- and binaphtylderivatives. If both aromatic ring systems are asymmetrically substituted, the compounds are chiral. As
the chirality of these structures originates not from an asymmetrically substituted atom center, but from an
asymmetric axis around which rotation is hindered, these enantiomers are also called atropisomers. In the
biphenyls, the ortho-substitutents must be large enough to prevent rotation around the central single
bond; if hydrogen atoms are present in these positions the barrier of rotation may be too small to prevent
interconversion of the enantiomeric forms at room temperature and the two structures may not be
separated, or may racemize slowly on standing.

Please note, that for the biphenyls and binaphtyls derivatives the formal maximum number of
stereoisomers (see above 'Chiral Compounds') is exceeded. As these moleucles do neither feature
a stereocenter or a stereogenic double bond, no stereoisomers would be expected. Only the hindered
rotation about the central C-C single bond leads to the stereoisomerism of these compounds. Therefore,
biphenyl- and binaphtyl-derivatives are conformational stereoisomers (not configurational stereoisomers).
In particular the enentiomerically pure binaphthyl derivatives are of great use in asymmetric catalysis. For
some animations of the dynamic behavior of biphenyls and binaphthyls and the process of racemization
see the chapter 'Atropsiomers' in the MolArch+ - Movies section of this web-site.
Chiral Compounds - Helical Chirality - 3D Structures
Helices are chiral as they can exist in enantiomeric left- or right-handed forms. Typical examples for
helical strutures are provided by the helicenes (benzologues of phenanthrene). With four or more rings,
steric hinderance at both ends of these molecule prevents the formation of planar conformations, and
helicenes rather adopt non-planar, but helical and enantiomeric structures with C2 symmetry (see also the
3D structures at 'Helicenes' and the chapter 'Racemization of [8]Helicene' in the MolArch+ Movies section of this web-site).

Other examples of chiral helical structures are provided by trans-cyclooctene and suitably substituted
heptalenes. Heptalene is not planar, and its twisted structure results in chirality. Although these
conformations generally interconvert rapidly, bulky substituents may sufficiently slow down this process to
optically resolve and separate these compounds.

The above example of the chirality of (E)-cyclooctene is also example for 'Planar Chirality' (see below).
Chiral Compounds - Planar Chirality - 3D Structures
Planar chirality may arise if an appropriately substituted planar group of atoms or ring system is bridged
by a linker-chain extending into the space above or below of this plane. Commmon examples are the

planar chirality of cyclophanes or alkenes as shown below. Even (E)-cyclooctene is a planar chiral
compound (see also above 'Helical Chirality'):

The above mentioned suitably substituted ferrocenes are also planar chiral compounds (see 'Asymmetric
Substituted Atoms').
Chiral Compounds - Other Examples - 3D Structures
There are many more examples known for which chirality of molecules results from hindered rotation of
groups or spatial arrangements of chemical moieties, a few examples are listed below:

Even catenanes and molecular knots made up from achiral molecules are chiral.

For more informations on other research topics, please refer to the complete list of
publications and to the gallery of graphics and animations.
Compunds with two or more asymmetrically substituted atoms (chiral centers) may be optically active.
Typical examples are (2S,3S)- and (2R,3R)-2,3-dibromo-butane, or (2S,3S)- and (2R,3R)-tartaric acid
(german: Weinsure). However, if pairs of equivalent stereocenters of opposite configuration are present

in the molecule, the compounds are optically inactive (achiralmeso-compounds) as both chiral centers
neutralize each other (internal compensation). For example, (2S,3R)- and (2R,3S)-2,3-dibromo-butane
are identical (meso) as both structures can be superimposed through simple rotations and translations
only. The same applies to meso-tartaric acid.

Stereoisomers of compunds which are not not related as mirror images are called diastereoisomers.
Structures of this type cannot be interconverted into each other through translation, rotation, and mirror
symmetry operations. In contrast to enantiomers, diastereoisomers have different chemical (towards
achiral as well as chiral reagents) and physical properties (including totally independent values for their
specific optical rotation).
trans-1,2-Disubstituted cyclohexanes of C2 symmetry are chiral compounds, whereas symmetrically cis1,2-disubstituted cyclohexanes are optically inactive as they posses a mirror plane of symmetry.
However, the latter type compounds exist in chiral conformations which interconvert rapidly into each
other through chair-antichair inversions of the cyclohexane ring (see also the 3D structures at 'Achiral
Compounds' and the chapter 'Ring Pseudorotation' in the MolArch+ - Movies section of this web-site).

Chiral Compounds - Substituted Adamantane Derivatives - 3D Structures


Adamantane derivates of suitable substitution may also be chiral compounds. In these structures, not a
single atom, but an entire group (the adamantyl residue) holds four substituents (adamantane derivative
on the left below) in a spatial arrangement that causes the compound to be non-superimposable with its
mirror image geometry. The central adamantyl residue may be regarded as an extended tetrahedron.

This example demonstrates that the stereogenic center of a moleucle needs not to be located on a
specific atom. In this case, the stereogenic center is located in the center of the adamantyl cage.
Disubstituted adamantanes (note the different substitution positions and pattern) are also chiral
compounds - they do not feature a chiral center but an axis of chirality(see below 'Axial Chirality').
In fact, the above (left) adamantane derivative features four asymmetric substituted carbon atoms (the
four tertiary carbon atoms of adamantane backbone). As derived above, the formal maximum number of
stereoisomers (see abvoe 'Chiral Compounds') would be Nmax = 24 = 16. However, in this case only two
stereoisomers actually exist, as the steric strain and geometrical limitations of the adamantane resdiue
require all substituent to point towards the outside of the molecule (in fact, once the first stereocenter is
defined, the remaining three are determined by their relative configuration, and thus the number of
observed stereoisomers is 2).
Similar limitations on the number of stereoisomers are observed in almost all bi- and polycyclic systems
with small rings. Twistanes (even the unsubstituted parent hydrocarbon) are chiral. As another example,
camphor yields two stereoisomers only, although it features two stereocenters in position 1 and 4 (but the
relative configuration of both centers is constrained in the bicyclic ring system):

Chiral Compounds - Axial Chirality - 3D Structures


Some compounds which do not have asymmetrically substituted carbon atoms (or any other atom type)
may still be chiral if they feature two perpendicular planes which are not symmetry planes. If
these disymmetric (chiral) planes cannot freely rotate against each other, the corresponding compounds
are chiral. Compounds of this type are said to be axially chiral (they feture an axis of chirality instead
a center of chirality)
Typical examples are allenes in which the central atom is sp-hybridized, and the planes containing the
substitutents on either end of the double bonds are aligned perpendicular to each other (in contrast to
simple olefines where all substituents are contained in a single plane of the -bond; note the difference
to E/Z-isomerism of alkenes). For an even number of double bonds in the allene, and if neither side of it is
symmetrically substituted, these compounds are optically active and thus chiral.

Very similar arrangements are observed in chiral spiro-annelated ring systems and compounds with
exocyclic double-bonds (see examples below).

As demonstrated above, suitably disubstituted adamantane derivatives also show axial chirality (see
above 'Substituted Adamantane Derivatives').
Chiral Compounds - Biphenyls and Binaphthyls - 3D Structures
The same rules as outlined above for the allenes and spiranes apply also to biphenyl- and binaphtylderivatives. If both aromatic ring systems are asymmetrically substituted, the compounds are chiral. As
the chirality of these structures originates not from an asymmetrically substituted atom center, but from an
asymmetric axis around which rotation is hindered, these enantiomers are also called atropisomers. In the
biphenyls, the ortho-substitutents must be large enough to prevent rotation around the central single
bond; if hydrogen atoms are present in these positions the barrier of rotation may be too small to prevent
interconversion of the enantiomeric forms at room temperature and the two structures may not be
separated, or may racemize slowly on standing.

Please note, that for the biphenyls and binaphtyls derivatives the formal maximum number of
stereoisomers (see above 'Chiral Compounds') is exceeded. As these moleucles do neither feature
a stereocenter or a stereogenic double bond, no stereoisomers would be expected. Only the hindered
rotation about the central C-C single bond leads to the stereoisomerism of these compounds. Therefore,
biphenyl- and binaphtyl-derivatives are conformational stereoisomers (not configurational stereoisomers).
In particular the enentiomerically pure binaphthyl derivatives are of great use in asymmetric catalysis. For
some animations of the dynamic behavior of biphenyls and binaphthyls and the process of racemization
see the chapter 'Atropsiomers' in the MolArch+ - Movies section of this web-site.
Chiral Compounds - Helical Chirality - 3D Structures
Helices are chiral as they can exist in enantiomeric left- or right-handed forms. Typical examples for
helical strutures are provided by the helicenes (benzologues of phenanthrene). With four or more rings,
steric hinderance at both ends of these molecule prevents the formation of planar conformations, and
helicenes rather adopt non-planar, but helical and enantiomeric structures with C2 symmetry (see also the
3D structures at 'Helicenes' and the chapter 'Racemization of [8]Helicene' in the MolArch+ Movies section of this web-site).

Other examples of chiral helical structures are provided by trans-cyclooctene and suitably substituted
heptalenes. Heptalene is not planar, and its twisted structure results in chirality. Although these
conformations generally interconvert rapidly, bulky substituents may sufficiently slow down this process to
optically resolve and separate these compounds.

The above example of the chirality of (E)-cyclooctene is also example for 'Planar Chirality' (see below).
Chiral Compounds - Planar Chirality - 3D Structures
Planar chirality may arise if an appropriately substituted planar group of atoms or ring system is bridged
by a linker-chain extending into the space above or below of this plane. Commmon examples are the
planar chirality of cyclophanes or alkenes as shown below. Even (E)-cyclooctene is a planar chiral
compound (see also above 'Helical Chirality'):

The above mentioned suitably substituted ferrocenes are also planar chiral compounds (see 'Asymmetric
Substituted Atoms').
Chiral Compounds - Other Examples - 3D Structures
There are many more examples known for which chirality of molecules results from hindered rotation of
groups or spatial arrangements of chemical moieties, a few examples are listed below:

Even catenanes and molecular knots made up from achiral molecules are chiral.

For more informations on other research topics, please refer to the complete list of
publications and to the gallery of graphics and animations.
Compunds with two or more asymmetrically substituted atoms (chiral centers) may be optically active.
Typical examples are (2S,3S)- and (2R,3R)-2,3-dibromo-butane, or (2S,3S)- and (2R,3R)-tartaric acid
(german: Weinsure). However, if pairs of equivalent stereocenters of opposite configuration are present
in the molecule, the compounds are optically inactive (achiralmeso-compounds) as both chiral centers
neutralize each other (internal compensation). For example, (2S,3R)- and (2R,3S)-2,3-dibromo-butane
are identical (meso) as both structures can be superimposed through simple rotations and translations
only. The same applies to meso-tartaric acid.

Stereoisomers of compunds which are not not related as mirror images are called diastereoisomers.
Structures of this type cannot be interconverted into each other through translation, rotation, and mirror
symmetry operations. In contrast to enantiomers, diastereoisomers have different chemical (towards
achiral as well as chiral reagents) and physical properties (including totally independent values for their
specific optical rotation).
trans-1,2-Disubstituted cyclohexanes of C2 symmetry are chiral compounds, whereas symmetrically cis1,2-disubstituted cyclohexanes are optically inactive as they posses a mirror plane of symmetry.
However, the latter type compounds exist in chiral conformations which interconvert rapidly into each
other through chair-antichair inversions of the cyclohexane ring (see also the 3D structures at 'Achiral

Compounds' and the chapter 'Ring Pseudorotation' in the MolArch+ - Movies section of this web-site).

Chiral Compounds - Substituted Adamantane Derivatives - 3D Structures


Adamantane derivates of suitable substitution may also be chiral compounds. In these structures, not a
single atom, but an entire group (the adamantyl residue) holds four substituents (adamantane derivative
on the left below) in a spatial arrangement that causes the compound to be non-superimposable with its
mirror image geometry. The central adamantyl residue may be regarded as an extended tetrahedron.

This example demonstrates that the stereogenic center of a moleucle needs not to be located on a
specific atom. In this case, the stereogenic center is located in the center of the adamantyl cage.
Disubstituted adamantanes (note the different substitution positions and pattern) are also chiral
compounds - they do not feature a chiral center but an axis of chirality(see below 'Axial Chirality').
In fact, the above (left) adamantane derivative features four asymmetric substituted carbon atoms (the
four tertiary carbon atoms of adamantane backbone). As derived above, the formal maximum number of
stereoisomers (see abvoe 'Chiral Compounds') would be Nmax = 24 = 16. However, in this case only two
stereoisomers actually exist, as the steric strain and geometrical limitations of the adamantane resdiue
require all substituent to point towards the outside of the molecule (in fact, once the first stereocenter is
defined, the remaining three are determined by their relative configuration, and thus the number of
observed stereoisomers is 2).
Similar limitations on the number of stereoisomers are observed in almost all bi- and polycyclic systems
with small rings. Twistanes (even the unsubstituted parent hydrocarbon) are chiral. As another example,
camphor yields two stereoisomers only, although it features two stereocenters in position 1 and 4 (but the
relative configuration of both centers is constrained in the bicyclic ring system):

Chiral Compounds - Axial Chirality - 3D Structures


Some compounds which do not have asymmetrically substituted carbon atoms (or any other atom type)
may still be chiral if they feature two perpendicular planes which are not symmetry planes. If
these disymmetric (chiral) planes cannot freely rotate against each other, the corresponding compounds
are chiral. Compounds of this type are said to be axially chiral (they feture an axis of chirality instead
a center of chirality)
Typical examples are allenes in which the central atom is sp-hybridized, and the planes containing the
substitutents on either end of the double bonds are aligned perpendicular to each other (in contrast to
simple olefines where all substituents are contained in a single plane of the -bond; note the difference
to E/Z-isomerism of alkenes). For an even number of double bonds in the allene, and if neither side of it is
symmetrically substituted, these compounds are optically active and thus chiral.

Very similar arrangements are observed in chiral spiro-annelated ring systems and compounds with
exocyclic double-bonds (see examples below).

As demonstrated above, suitably disubstituted adamantane derivatives also show axial chirality (see
above 'Substituted Adamantane Derivatives').
Chiral Compounds - Biphenyls and Binaphthyls - 3D Structures
The same rules as outlined above for the allenes and spiranes apply also to biphenyl- and binaphtylderivatives. If both aromatic ring systems are asymmetrically substituted, the compounds are chiral. As
the chirality of these structures originates not from an asymmetrically substituted atom center, but from an
asymmetric axis around which rotation is hindered, these enantiomers are also called atropisomers. In the
biphenyls, the ortho-substitutents must be large enough to prevent rotation around the central single
bond; if hydrogen atoms are present in these positions the barrier of rotation may be too small to prevent
interconversion of the enantiomeric forms at room temperature and the two structures may not be
separated, or may racemize slowly on standing.

Please note, that for the biphenyls and binaphtyls derivatives the formal maximum number of
stereoisomers (see above 'Chiral Compounds') is exceeded. As these moleucles do neither feature
a stereocenter or a stereogenic double bond, no stereoisomers would be expected. Only the hindered
rotation about the central C-C single bond leads to the stereoisomerism of these compounds. Therefore,
biphenyl- and binaphtyl-derivatives are conformational stereoisomers (not configurational stereoisomers).
In particular the enentiomerically pure binaphthyl derivatives are of great use in asymmetric catalysis. For
some animations of the dynamic behavior of biphenyls and binaphthyls and the process of racemization
see the chapter 'Atropsiomers' in the MolArch+ - Movies section of this web-site.
Chiral Compounds - Helical Chirality - 3D Structures
Helices are chiral as they can exist in enantiomeric left- or right-handed forms. Typical examples for
helical strutures are provided by the helicenes (benzologues of phenanthrene). With four or more rings,
steric hinderance at both ends of these molecule prevents the formation of planar conformations, and
helicenes rather adopt non-planar, but helical and enantiomeric structures with C2 symmetry (see also the
3D structures at 'Helicenes' and the chapter 'Racemization of [8]Helicene' in the MolArch+ Movies section of this web-site).

Other examples of chiral helical structures are provided by trans-cyclooctene and suitably substituted
heptalenes. Heptalene is not planar, and its twisted structure results in chirality. Although these
conformations generally interconvert rapidly, bulky substituents may sufficiently slow down this process to
optically resolve and separate these compounds.

The above example of the chirality of (E)-cyclooctene is also example for 'Planar Chirality' (see below).
Chiral Compounds - Planar Chirality - 3D Structures
Planar chirality may arise if an appropriately substituted planar group of atoms or ring system is bridged
by a linker-chain extending into the space above or below of this plane. Commmon examples are the

planar chirality of cyclophanes or alkenes as shown below. Even (E)-cyclooctene is a planar chiral
compound (see also above 'Helical Chirality'):

The above mentioned suitably substituted ferrocenes are also planar chiral compounds (see 'Asymmetric
Substituted Atoms').
Chiral Compounds - Other Examples - 3D Structures
There are many more examples known for which chirality of molecules results from hindered rotation of
groups or spatial arrangements of chemical moieties, a few examples are listed below:

Even catenanes and molecular knots made up from achiral molecules are chiral.

For more informations on other research topics, please refer to the complete list of
publications and to the gallery of graphics and animations.
Compunds with two or more asymmetrically substituted atoms (chiral centers) may be optically active.
Typical examples are (2S,3S)- and (2R,3R)-2,3-dibromo-butane, or (2S,3S)- and (2R,3R)-tartaric acid
(german: Weinsure). However, if pairs of equivalent stereocenters of opposite configuration are present

in the molecule, the compounds are optically inactive (achiralmeso-compounds) as both chiral centers
neutralize each other (internal compensation). For example, (2S,3R)- and (2R,3S)-2,3-dibromo-butane
are identical (meso) as both structures can be superimposed through simple rotations and translations
only. The same applies to meso-tartaric acid.

Stereoisomers of compunds which are not not related as mirror images are called diastereoisomers.
Structures of this type cannot be interconverted into each other through translation, rotation, and mirror
symmetry operations. In contrast to enantiomers, diastereoisomers have different chemical (towards
achiral as well as chiral reagents) and physical properties (including totally independent values for their
specific optical rotation).
trans-1,2-Disubstituted cyclohexanes of C2 symmetry are chiral compounds, whereas symmetrically cis1,2-disubstituted cyclohexanes are optically inactive as they posses a mirror plane of symmetry.
However, the latter type compounds exist in chiral conformations which interconvert rapidly into each
other through chair-antichair inversions of the cyclohexane ring (see also the 3D structures at 'Achiral
Compounds' and the chapter 'Ring Pseudorotation' in the MolArch+ - Movies section of this web-site).

Chiral Compounds - Substituted Adamantane Derivatives - 3D Structures


Adamantane derivates of suitable substitution may also be chiral compounds. In these structures, not a
single atom, but an entire group (the adamantyl residue) holds four substituents (adamantane derivative
on the left below) in a spatial arrangement that causes the compound to be non-superimposable with its
mirror image geometry. The central adamantyl residue may be regarded as an extended tetrahedron.

This example demonstrates that the stereogenic center of a moleucle needs not to be located on a
specific atom. In this case, the stereogenic center is located in the center of the adamantyl cage.
Disubstituted adamantanes (note the different substitution positions and pattern) are also chiral
compounds - they do not feature a chiral center but an axis of chirality(see below 'Axial Chirality').
In fact, the above (left) adamantane derivative features four asymmetric substituted carbon atoms (the
four tertiary carbon atoms of adamantane backbone). As derived above, the formal maximum number of
stereoisomers (see abvoe 'Chiral Compounds') would be Nmax = 24 = 16. However, in this case only two
stereoisomers actually exist, as the steric strain and geometrical limitations of the adamantane resdiue
require all substituent to point towards the outside of the molecule (in fact, once the first stereocenter is
defined, the remaining three are determined by their relative configuration, and thus the number of
observed stereoisomers is 2).
Similar limitations on the number of stereoisomers are observed in almost all bi- and polycyclic systems
with small rings. Twistanes (even the unsubstituted parent hydrocarbon) are chiral. As another example,
camphor yields two stereoisomers only, although it features two stereocenters in position 1 and 4 (but the
relative configuration of both centers is constrained in the bicyclic ring system):

Chiral Compounds - Axial Chirality - 3D Structures


Some compounds which do not have asymmetrically substituted carbon atoms (or any other atom type)
may still be chiral if they feature two perpendicular planes which are not symmetry planes. If
these disymmetric (chiral) planes cannot freely rotate against each other, the corresponding compounds
are chiral. Compounds of this type are said to be axially chiral (they feture an axis of chirality instead
a center of chirality)
Typical examples are allenes in which the central atom is sp-hybridized, and the planes containing the
substitutents on either end of the double bonds are aligned perpendicular to each other (in contrast to
simple olefines where all substituents are contained in a single plane of the -bond; note the difference
to E/Z-isomerism of alkenes). For an even number of double bonds in the allene, and if neither side of it is
symmetrically substituted, these compounds are optically active and thus chiral.

Very similar arrangements are observed in chiral spiro-annelated ring systems and compounds with
exocyclic double-bonds (see examples below).

As demonstrated above, suitably disubstituted adamantane derivatives also show axial chirality (see
above 'Substituted Adamantane Derivatives').
Chiral Compounds - Biphenyls and Binaphthyls - 3D Structures
The same rules as outlined above for the allenes and spiranes apply also to biphenyl- and binaphtylderivatives. If both aromatic ring systems are asymmetrically substituted, the compounds are chiral. As
the chirality of these structures originates not from an asymmetrically substituted atom center, but from an
asymmetric axis around which rotation is hindered, these enantiomers are also called atropisomers. In the
biphenyls, the ortho-substitutents must be large enough to prevent rotation around the central single
bond; if hydrogen atoms are present in these positions the barrier of rotation may be too small to prevent
interconversion of the enantiomeric forms at room temperature and the two structures may not be
separated, or may racemize slowly on standing.

Please note, that for the biphenyls and binaphtyls derivatives the formal maximum number of
stereoisomers (see above 'Chiral Compounds') is exceeded. As these moleucles do neither feature
a stereocenter or a stereogenic double bond, no stereoisomers would be expected. Only the hindered
rotation about the central C-C single bond leads to the stereoisomerism of these compounds. Therefore,
biphenyl- and binaphtyl-derivatives are conformational stereoisomers (not configurational stereoisomers).
In particular the enentiomerically pure binaphthyl derivatives are of great use in asymmetric catalysis. For
some animations of the dynamic behavior of biphenyls and binaphthyls and the process of racemization
see the chapter 'Atropsiomers' in the MolArch+ - Movies section of this web-site.
Chiral Compounds - Helical Chirality - 3D Structures
Helices are chiral as they can exist in enantiomeric left- or right-handed forms. Typical examples for
helical strutures are provided by the helicenes (benzologues of phenanthrene). With four or more rings,
steric hinderance at both ends of these molecule prevents the formation of planar conformations, and
helicenes rather adopt non-planar, but helical and enantiomeric structures with C2 symmetry (see also the
3D structures at 'Helicenes' and the chapter 'Racemization of [8]Helicene' in the MolArch+ Movies section of this web-site).

Other examples of chiral helical structures are provided by trans-cyclooctene and suitably substituted
heptalenes. Heptalene is not planar, and its twisted structure results in chirality. Although these
conformations generally interconvert rapidly, bulky substituents may sufficiently slow down this process to
optically resolve and separate these compounds.

The above example of the chirality of (E)-cyclooctene is also example for 'Planar Chirality' (see below).
Chiral Compounds - Planar Chirality - 3D Structures
Planar chirality may arise if an appropriately substituted planar group of atoms or ring system is bridged
by a linker-chain extending into the space above or below of this plane. Commmon examples are the
planar chirality of cyclophanes or alkenes as shown below. Even (E)-cyclooctene is a planar chiral
compound (see also above 'Helical Chirality'):

The above mentioned suitably substituted ferrocenes are also planar chiral compounds (see 'Asymmetric
Substituted Atoms').
Chiral Compounds - Other Examples - 3D Structures
There are many more examples known for which chirality of molecules results from hindered rotation of
groups or spatial arrangements of chemical moieties, a few examples are listed below:

Even catenanes and molecular knots made up from achiral molecules are chiral.

For more informations on other research topics, please refer to the complete list of
publications and to the gallery of graphics and animations.
Compunds with two or more asymmetrically substituted atoms (chiral centers) may be optically active.
Typical examples are (2S,3S)- and (2R,3R)-2,3-dibromo-butane, or (2S,3S)- and (2R,3R)-tartaric acid
(german: Weinsure). However, if pairs of equivalent stereocenters of opposite configuration are present
in the molecule, the compounds are optically inactive (achiralmeso-compounds) as both chiral centers
neutralize each other (internal compensation). For example, (2S,3R)- and (2R,3S)-2,3-dibromo-butane
are identical (meso) as both structures can be superimposed through simple rotations and translations
only. The same applies to meso-tartaric acid.

Stereoisomers of compunds which are not not related as mirror images are called diastereoisomers.
Structures of this type cannot be interconverted into each other through translation, rotation, and mirror
symmetry operations. In contrast to enantiomers, diastereoisomers have different chemical (towards
achiral as well as chiral reagents) and physical properties (including totally independent values for their
specific optical rotation).
trans-1,2-Disubstituted cyclohexanes of C2 symmetry are chiral compounds, whereas symmetrically cis1,2-disubstituted cyclohexanes are optically inactive as they posses a mirror plane of symmetry.
However, the latter type compounds exist in chiral conformations which interconvert rapidly into each
other through chair-antichair inversions of the cyclohexane ring (see also the 3D structures at 'Achiral

Compounds' and the chapter 'Ring Pseudorotation' in the MolArch+ - Movies section of this web-site).

Chiral Compounds - Substituted Adamantane Derivatives - 3D Structures


Adamantane derivates of suitable substitution may also be chiral compounds. In these structures, not a
single atom, but an entire group (the adamantyl residue) holds four substituents (adamantane derivative
on the left below) in a spatial arrangement that causes the compound to be non-superimposable with its
mirror image geometry. The central adamantyl residue may be regarded as an extended tetrahedron.

This example demonstrates that the stereogenic center of a moleucle needs not to be located on a
specific atom. In this case, the stereogenic center is located in the center of the adamantyl cage.
Disubstituted adamantanes (note the different substitution positions and pattern) are also chiral
compounds - they do not feature a chiral center but an axis of chirality(see below 'Axial Chirality').
In fact, the above (left) adamantane derivative features four asymmetric substituted carbon atoms (the
four tertiary carbon atoms of adamantane backbone). As derived above, the formal maximum number of
stereoisomers (see abvoe 'Chiral Compounds') would be Nmax = 24 = 16. However, in this case only two
stereoisomers actually exist, as the steric strain and geometrical limitations of the adamantane resdiue
require all substituent to point towards the outside of the molecule (in fact, once the first stereocenter is
defined, the remaining three are determined by their relative configuration, and thus the number of
observed stereoisomers is 2).
Similar limitations on the number of stereoisomers are observed in almost all bi- and polycyclic systems
with small rings. Twistanes (even the unsubstituted parent hydrocarbon) are chiral. As another example,
camphor yields two stereoisomers only, although it features two stereocenters in position 1 and 4 (but the
relative configuration of both centers is constrained in the bicyclic ring system):

Chiral Compounds - Axial Chirality - 3D Structures


Some compounds which do not have asymmetrically substituted carbon atoms (or any other atom type)
may still be chiral if they feature two perpendicular planes which are not symmetry planes. If
these disymmetric (chiral) planes cannot freely rotate against each other, the corresponding compounds
are chiral. Compounds of this type are said to be axially chiral (they feture an axis of chirality instead
a center of chirality)
Typical examples are allenes in which the central atom is sp-hybridized, and the planes containing the
substitutents on either end of the double bonds are aligned perpendicular to each other (in contrast to
simple olefines where all substituents are contained in a single plane of the -bond; note the difference
to E/Z-isomerism of alkenes). For an even number of double bonds in the allene, and if neither side of it is
symmetrically substituted, these compounds are optically active and thus chiral.

Very similar arrangements are observed in chiral spiro-annelated ring systems and compounds with
exocyclic double-bonds (see examples below).

As demonstrated above, suitably disubstituted adamantane derivatives also show axial chirality (see
above 'Substituted Adamantane Derivatives').
Chiral Compounds - Biphenyls and Binaphthyls - 3D Structures
The same rules as outlined above for the allenes and spiranes apply also to biphenyl- and binaphtylderivatives. If both aromatic ring systems are asymmetrically substituted, the compounds are chiral. As
the chirality of these structures originates not from an asymmetrically substituted atom center, but from an
asymmetric axis around which rotation is hindered, these enantiomers are also called atropisomers. In the
biphenyls, the ortho-substitutents must be large enough to prevent rotation around the central single
bond; if hydrogen atoms are present in these positions the barrier of rotation may be too small to prevent
interconversion of the enantiomeric forms at room temperature and the two structures may not be
separated, or may racemize slowly on standing.

Please note, that for the biphenyls and binaphtyls derivatives the formal maximum number of
stereoisomers (see above 'Chiral Compounds') is exceeded. As these moleucles do neither feature
a stereocenter or a stereogenic double bond, no stereoisomers would be expected. Only the hindered
rotation about the central C-C single bond leads to the stereoisomerism of these compounds. Therefore,
biphenyl- and binaphtyl-derivatives are conformational stereoisomers (not configurational stereoisomers).
In particular the enentiomerically pure binaphthyl derivatives are of great use in asymmetric catalysis. For
some animations of the dynamic behavior of biphenyls and binaphthyls and the process of racemization
see the chapter 'Atropsiomers' in the MolArch+ - Movies section of this web-site.
Chiral Compounds - Helical Chirality - 3D Structures
Helices are chiral as they can exist in enantiomeric left- or right-handed forms. Typical examples for
helical strutures are provided by the helicenes (benzologues of phenanthrene). With four or more rings,
steric hinderance at both ends of these molecule prevents the formation of planar conformations, and
helicenes rather adopt non-planar, but helical and enantiomeric structures with C2 symmetry (see also the
3D structures at 'Helicenes' and the chapter 'Racemization of [8]Helicene' in the MolArch+ Movies section of this web-site).

Other examples of chiral helical structures are provided by trans-cyclooctene and suitably substituted
heptalenes. Heptalene is not planar, and its twisted structure results in chirality. Although these
conformations generally interconvert rapidly, bulky substituents may sufficiently slow down this process to
optically resolve and separate these compounds.

The above example of the chirality of (E)-cyclooctene is also example for 'Planar Chirality' (see below).
Chiral Compounds - Planar Chirality - 3D Structures
Planar chirality may arise if an appropriately substituted planar group of atoms or ring system is bridged
by a linker-chain extending into the space above or below of this plane. Commmon examples are the

planar chirality of cyclophanes or alkenes as shown below. Even (E)-cyclooctene is a planar chiral
compound (see also above 'Helical Chirality'):

The above mentioned suitably substituted ferrocenes are also planar chiral compounds (see 'Asymmetric
Substituted Atoms').
Chiral Compounds - Other Examples - 3D Structures
There are many more examples known for which chirality of molecules results from hindered rotation of
groups or spatial arrangements of chemical moieties, a few examples are listed below:

Even catenanes and molecular knots made up from achiral molecules are chiral.

For more informations on other research topics, please refer to the complete list of
publications and to the gallery of graphics and animations.

Chirality and Symmetry

All objects may be classified with respect to a property we call chirality (from the Greek cheir meaning hand).
A chiral object is not identical in all respects (i.e. superimposable) with its mirror image. An achiral object is
identical with (superimposable on) its mirror image. Chiral objects have a "handedness", for example, golf clubs,
scissors, shoes and a corkscrew. Thus, one can buy right or left-handed golf clubs and scissors. Likewise, gloves and
shoes come in pairs, a right and a left. Achiral objects do not have a handedness, for example, a baseball bat (no
writing or logos on it), a plain round ball, a pencil, a T-shirt and a nail. The chirality of an object is related to its
symmetry, and to this end it is useful to recognize certain symmetry elements that may be associated with a given
object. A symmetry element is a plane, a line or a point in or through an object, about which a rotation or reflection
leaves the object in an orientation indistinguishable from the original. Some examples of symmetry elements are
shown below.

The face playing card provides an example of a center or point of symmetry. Starting from such a point, a line drawn
in any direction encounters the same structural features as the opposite (180) line. Four random lines of this kind are
shown in green. An example of a molecular configuration having a point of symmetry is (E)-1,2-dichloroethene.
Another way of describing a point of symmetry is to note that any point in the object is reproduced by reflection
through the center onto the other side. In these two cases the point of symmetry is colored magenta.
The boat conformation of cyclohexane shows an axis of symmetry (labeled C 2 here) and two intersecting planes of
symmetry (labeled ). The notation for a symmetry axis is C n, where n is an integer chosen so that rotation about the
axis by 360/n returns the object to a position indistinguishable from where it started. In this case the rotation is by
180, so n=2. A plane of symmetry divides the object in such a way that the points on one side of the plane are
equivalent to the points on the other side by reflection through the plane. In addition to the point of symmetry noted
earlier, (E)-1,2-dichloroethene also has a plane of symmetry (the plane defined by the six atoms), and a C 2axis,
passing through the center perpendicular to the plane. The existence of a reflective symmetry element (a point or
plane of symmetry) is sufficient to assure that the object having that element is achiral. Chiral objects, therefore, do
not have any reflective symmetry elements, but may have rotational symmetry axes, since these elements do not
require reflection to operate. In addition to the chiral vs achiral distinction, there are two other terms often used to
refer to the symmetry of an object. These are:
1.
2.

Dissymmetry: The absence of reflective symmetry elements. All dissymmetric objects are chiral.
Asymmetry: The absence of all symmetry elements. All asymmetric objects are chiral.
Models of some additional three-dimensional examples are provided on the interactive symmetry page. The
symmetry elements of a structure provide insight concerning the structural equivalence or nonequivalence of similar
component atoms or groups Examples of this symmetry analysis may be viewed by Clicking Here.

External Links
Molecular Symmetry and Chirality
Introduction and Overview
Note: viewing the structures on these pages requires use of the MDL Chime Plug-In.

The symmetry of a molecule describes how its different parts relate to one another
geometrically. symmetry plays an important role in many areas of chemistry, with effects on:

physical properties: e.g. dipole moment, chirality

spectroscopic properties: transition intensities, geometric equivalence of groups or nuclei

bonding interactions: bonds require overlap of atomic orbitals of correct symmetry

Terms and Definitions


Symmetry Operation: a spatial manipulation performed on a molecule that leaves it in an
configuration identical to and superimposable upon the original configuration.
Symmetry Element: an axis, plane, or point about which a symmetry operation is performed.
Point Group: a symbol that identifies all the symmetry elements present in a molecule.
Symmetry Elements and Operations
Element

none

proper
rotation axis

mirror plane

inversion
centre

improper
rotation axis

Effects on Molecular Properties


Chirality:
To be chiral, a molecule must lack both i and .*
Molecules belonging to groups Cn (including E) and Dn are chiral.
* This is a small oversimplification. The most precise requirement for chirality is the lack of
any Sn element, but because hardly anybody really knows what those look like, there are a
series of increasingly precise shortcuts used to Spotting a Chiral Compound. Or, you can just
skip to the summary at the end.
Polarity:
To be polar, a molecule must lack all of i, C2' axes, and h.
Molecules belonging to groups Cs, Cn (incl E) and Cnv (including Cv) may be polar.

This page is maintained and copyright by W. Stephen McNeil at UBC Okanagan.


Last updated January 12, 2010. Introduction:

Chirality is a property of individual molecules. If a molecule is chiral then we know


that the molecule has two enantiomeric forms (which are almost identical but are
actually different molecules). I like to think of this as the 'evil twin' property. If a
molecule is chiral then we know that it has an 'evil twin' (an enantiomer). If a
molecule is not chiral (achiral) then it is unique and does not have an 'evil twin'.

How to:
There are three different common tests for chirality. You can use which ever test is
easiest to apply for a particular molecule. (Note: that they should all give the same
answer for a particular molecule).
Test 1: Draw the mirror image of the molecule and see if the two molecules are the
same or different. If they are different, then the molecule is chiral. If they are the
same, then it is not chiral.
Molecule 1 and its mirror image are different. Therefore, molecule 1 is chiral. (The
mirror image is drawn from the perspective of having the mirror behind molecule 1,
which flips all the chiral centers.)

Molecule 2 and its mirror image are the same. Therefore, molecule 2 is not chiral.
(This is a little difficult to see and so it is probably easier to use test 3 for this molecule.
There is a plane of symmetry in molecule 2 that cuts throught the NH 2 and OH groups.)
Test 2: If the molecule has only one chiral center, then the molecule is chiral. If the
molecule has more than one chiral center, it is most likely chiral. The exceptions
are meso-compounds, which have chiral centers but are not chiral due to the
presence of a place of symmetry.

Molecule 3 has a single chiral center (carbon 2). Therefore, this molecule is chiral. (Note
that in this depiction, we have not specified if the NH2 group is up or down. This means

usually means that it is racemic mixture of the two. However, the molecule can still be
chiral even if it is a racemate.)
Molecule 4 has two chiral centers. Therefore, it is most likely chiral. If we use test 1 or
2, we find that it is in fact chiral. (However, if the two methyl groups were cis, then the
molecule would still have two chiral centers but not be chiral and would be a meso
compound.)
Test 3: If a molecule has a plane of symmetry, then the molecule is not chiral
(achiral).

Molecule 2 has a plane of symmetry. Therefore, the molecule is not chiral. (If the NH2
and OH groups were cis, there still would be a plane of symmetry and the molecule would
still be not chiral.)
Molecule 4 does not have a plane of symmetry. Therefore, the molecule is most likely
chiral. (The only exception would be if the molecule had a point of symmetry. Then it
would not be chiral but point symmetry is rare.)
Molecule 3 does not have a plane of symmetry. Therefore, the molecule is most likely
chiral. (It may look like there is a plane of symmetry in the plane of the page. But we
have to remember that the NH2 group is either up or down which breaks the plane of
symmetry).

ntroduction:
Chiral centers are tetrahedral atoms (usually carbons) that have four different
substituents. Each chiral center in a molecule will be either R or S. As noted above,
molecules with a single chiral center are chiral. Molecules with more than one chiral
center are usually chiral. The exception are meso-compounds.

How to:
Step 1: Eliminate the atoms that cannot be chiral centers. These include CH2 groups,
CH3 groups, oxygens, halogens, and any atom that is part of a double or triple bond.
For molecule 1, we can eliminate every atom as a possible chiral centers except for one
(which is highlighted with a red arrow).

Step 2: For the remaining atoms, list out the groups (substituents) attached to that
atom. If there are four different groups, then it is a chiral center. (Note that two
substituents can appear to be the same if you look only at the first attached atom but
you have to keep going to check if they are really the same or are different.)

The four groups attached to this carbon are:


-Br, -H, -CH2, -C=. These are all different.
Therefore, this is a chiral center.
The four groups attached to this carbon are:
-H, -methyl, -ethyl, -propyl. These are all different. Therefore, this is a chiral center.
(Note that if you stopped at just the first attached atom then it would look like that the
two CH2 groups were the same. However, if you keep going down the chain, you find that
they are different.)
Groups attached to this carbon are: -H, -NH2,
-CH2CH2CHOH, -CH2CH2CHOH. The last two are the same. Therefore, this is NOT a chiral
center. (Note, that we had to go all the way to the other side of the molecule to
convince ourselves that they really are the same.)
Groups attached to this carbon are: -H, -NH2,
-CH2CH2, -CH2CHOH. There are four different groups attached to this carbon so this is a
chiral center. (Note that we had to go around the ring to see if the two CH 2's were the
same or different. We can stop at the first point where we find that they are different.
HINT: The most common type of chiral atom in organic molecules are carbons
because they can be SP3 hybridized and can form four bonds. The other types of
atoms that can be chiral are quaternary nitrogens, tertiary nitrogens that are at the
bridgehead (junction) of a bicyclic ring system, hypervalent phosphorous (with more
than three bonds), and hypervalent sulfur (with more than two bonds). However,
these other types of chiral atoms are very unusual.

Introd
uction:
If a molecule has two or more chiral centers, it is usually chiral. The exception are
meso-molecules, which are not chiral. These are molecules that due to symmetry
have chiral centers that cancel each other out.

How to:
A meso-molecule must satisfy BOTH of the following criteria
1) Has two or more chiral centers.
2) Has a plane of symmetry.
A person is meso (NOT CHIRAL) even though they have chiral elements (hands and feet).
There is a plane of symmetry down the middle of a person, which makes a person the
same as their mirror image.
Examples:

This molecule is meso (NOT CHIRAL). It has two chiral centers and a plane of symmetry.
This molecule is not meso (CHIRAL). It has two chiral centers but no plane of symmetry.
This molecule is not meso (NOT CHIRAL). It has a plane of symmetry but no chiral
centers. The carbons attached to the NH2 groups may look like chiral centers but the are
not.
ntroduction:
When finding the relationship between two molecules, one of the most difficult
decisions is whether two molecules are the same or different. Often the two
molecules will be shown in different orientations. Therefore, to compare the two
molecules, it is helpful to change the orientation of one of the molecules by rotating
it and/or flipping it. You can do this by making a model, but it is faster to learn how
do this without making a model.

How to:
Rotating. When rotating a molecule in the plane of the page, the up and down groups
should remain the same. In both examples below, the OH group points up and will
stay pointing up after rotation.
Examples:

Hint: You can rotate the paper to check that the two molecules that you have drawn
are the same. Make sure the the position (left and right) of the groups did not get
switched.

Flipping. When flipping the molecule, up and down groups will all be reversed. In
both examples below, the OH group switches from up to down after being flipped.

Comment: Do not worry about whether the rotation is 30 or 120 or that the flip is
horizontal or vertical. The point of this exercise is to be able to manipulate the
molecules so that you can convince yourself that all the molecules are the same
molecules.

Introduction:
When deciding if two groups are stereoisomers or the same, you need to answer the
question whether the two molecules are mirror images. One way to do this is to draw
the mirror image of the first molecule and see if it is the same as the second.

How to:
When deciding if two groups are stereoisomers or the same, you need to answer the
question whether the two molecules are mirror images. One way to do this is to draw
the mirror image of the first molecule and see if it is the same as the second.
Mirror below: All the atoms will stay in the same place but any stereocenter (up or
down group) will be flipped.
Mirror mirror to one side: The atoms will be flipped. Groups on the left will be on
the right (and visa versa). However, all the stereocenters will stay the same.
Comment: In the above examples, the mirror image is different from the original
molecule. This means that each molecule is chiral. The relationship between the
original molecule and its mirror image are enantiomers.

Introduction:
Chemists like to categorize the similarities between two molecules just as you would
for the relationship between two people. The level of similarity between two
molecules can help predict their similarity in properties and chemical reactivity.

Two molecules that are very similar are called isomers. The problem is that there are
many different kinds of isomers. These are listed below in order of increasing
similarity. Note that there are two kinds of stereoisomers: diastereomers and
enantiomers.

How to:
Below is a flow chart to help you categorize the relationship between two molecules.
The possible answers are: a) not isomers, b) two different depictions of the same
molecule, c) constitutional isomers, d) diastereomers, and e) enantiomers.
relationship between two molecules
not isomers
constitutional isomer
stereoisomer (diastereomer)
stereoisomer (enantiomer)
the same*

more similar

analogy to human relationships


not related
distant cousins
siblings (brother or sister)
siblings (identical twins)
two pictures of same person
*Conformational isomers or rotamers are the same molecules that have their been folded
or twisted into different conformations. A human analogy would be two pictures of the
same person with their hands up and hands down. They are the same people but the
three dimensional

Example 1:

Same molecular formula?


Same connectivity?
Are they the same?
Yes
Yes

Yes
Comment: These two are the same molecules. If you pick up compound 1 and flip it
over (around the horizontal-axis), you get compound 2. Notice that when you flip
over the molecule the groups pointing up will point down and visa versa.
Example 2:
Same molecular formula?
Same connectivity?
Are they the same?
Are they mirror images?
Yes
Yes
No
No
Comment: These molecules are diastereomers. They are not mirror images. We can
try to test this by drawing the mirror image of compound 1 and checking to see if it is
the same as compound 2 (which it is not).

Example 3:
Same molecular formula?
Same connectivity?
Are they the same?
Are they mirror images?
Yes
Yes

No
Yes
Comment: These molecules are enantiomers. The hard question is whether these
molecules are the same. It looks like you could flip over compound 1 and it would
become compound 2. However, if you did that, then the CH3 and the OH groups
would switch places and be at the wrong positions. We can also tell these are mirror
images if we place the mirror behind compound 1. This would flip all of the chiral
centers (without moving their positions) and this mirror image of compound 1 is the
same as compound 2. Therefore, 1 and 2 are mirror images of each other and are
different, which is the definition of enantiomers.

Example 4:
Same molecular formula?
Same connectivity?
Are they the same?
Are they mirror images?
Yes
Yes
No
No
Comment: These molecules are diastereomers. The hard question is: "are they the
same?" If you flip over compound 1, then the Br's that point up will point down just
like in compound 2. However, remember that the OH will also change
stereochemistry from down to up, which is different from compound 2.

Example 5:
Same molecular formula?
Same connectivity?
Are they the same?
Comment: The molecules are the same. This problem is difficult because the two
cyclohexanes (compound 1 and 2) are depicted from different perspectives. In
compound 1, we are looking from the top of the cyclohexane and in compound 2 we
are looking from the side the cyclohexane.) The hard question is again: "are they the
same." The difficulty is in deciding in compound 2 if the methyl groups are both up,
both down, or one up and one down. Remember that each carbon in a ring will have
two substituents (one up and one down). We can see that in compound 2, one methyl
group is up and the other is down.
Yes
Yes
Yes

Example 6:
Same molecular formula?
Same connectivity?
Are they the same?
Comment: The molecules are the same. Again the difficult question is: "Are they the
same?" If you flip over compound 1, then it will be the same as compound 2.

Yes
Yes
Yes

Example 7:
Same molecular formula?
Same connectivity?
Comment: The molecules are constitutional isomers. Like example 5, the problem
is translating the two different perspectives of the cyclohexane ring. It is helpful to
redraw the chair depiction (compound 1) in the top-view depiction (like compound
2). When we do this, we can clearly see that the connectivity of these isomers is
different and therefore, these compounds are constitutional isomers. (Hint: it is
helpful to number the carbons when doing this translation.)
Yes
No

Note: It is hard to tell if the OH and CH3 groups on carbons 2 and 4 are up or down.
However, it is easier to see that the H's on these carbons are both down. Therefore,
we know that the other substituent (the OH and CH3) must be up.

Comment: The molecules are enantiomers. These molecules are shown as Fischer
projections. The nice thing about Fischer projections is that it makes comparing
stereochemistry easier. You can just compare chirality centers simply by looking at
which groups are on the left and right sides. For example in compound 1, the top
chirality center has the OH group on the right; whereas, in compound 2, the top
chirality center has the OH on the left. So, the top chiral centers of the two
molecules are opposites, as are the bottom chirality centers. Therefore, these
molecules are enantiomers.
Example 8:
Same molecular formula?
Same connectivity?
Are they the same?
Are they mirror images?
Yes
Yes
No
Yes

KEN SHIMIZU

(SHIMIZU@SC.EDU)

Introduction:
In the IUPAC naming system, each chiral center is assigned as either R or S.
Remember that chiral centers (also known as stereogenic centers) are tetrahedral
atoms with four different attached groups. Usually, chiral centers are SP3 hybridized
carbons.* The four groups can be attached in two different ways, leading to
enantiomeric forms (R and S).

How to:

There are two steps to assigning a chiral center.


1) Start with the first atom of each group that is directly attached to the chiral
tetrahedral center. The atoms with higher atomic number will have highest priority.
For example, the ranking for the four groups around the chiral center of the molecule
CHBrClF would be:

Hint: If hydrogen is one of the groups attached to a chiral center, then it will be the
lowest priority group.

2) If two groups have the same first atom, then compare the second atom from the
chiral center. If there are multiple second atoms, then compare them in order of
atomic number. Stop, when there is a difference. For example:

Although -CBr3 has more atoms that are higher in atomic number, it is the lower priority
group. The first atom in -CBr3 is a carbon, which has a lower atomic number (6) than -F
(9).
The first atom of the two groups are the same (carbons). So, we compare the "second"
atoms. However, in each case there are three "second" atoms so we have to proceed in
order of atomic number (high to low). The first pair of "second" atoms are the same (Br
and Br) so we proceed to the next highest pair (Cl and F), which are different. Chlorine
has a higher atomic number and so the group on the left will have the higher priority.
HINT: It is important to stop as soon as you find a difference. Dont get fooled by
seeing higher atomic number atoms on the substituent that are further from the chiral
center, like in the top example.

3) If all the first and second atoms from the chiral center are the same, then proceed
to the next furthest atom (in order of atomic number) until you find a difference.

4) Treat double and triple bonds as if they are a series of single bonds to the same
atom. This will involve the creation of singly bonded dummy atoms (highlighted in
blue) of the same type as involved in the double or triple bonds. For example:
Note: The first carbon is singly bonded to hydrogen and doubly bonded to carbon 2. For
the purposes of ranking priorities, we would consider the first carbon as being singly
bonded to hydrogen, singly bonded to carbon 2, and singly bonded to another 'dummy'
carbon.
The same rules can be applied to carbon two, it is singly bonded to a H and Br and doubly
bonded to carbon one. Therefore, we would consider carbon two as being singly bonded
to H, Br, and carbon one. In addition, it also forms a single bond to a another 'dummy'
carbon.
(STEP 2) Rules for assigning R and S.

1) After assigning priorities to the four different groups attached to the chiral center,
make sure that the lowest priority group (4) is pointing away from you. All three
examples below have the lowest priority (4) group point to the back.
Hint: One trick that I use is to use my right hand. I point my thumb in the direction of
the lowest priority group (4), and then curl my fingers (like in a grabbing motion) in the
direction of the groups 1, 2, and 3. If my fingers follows the direction of 1, 2, and 3, then
the chiral center is R. This is easy to remember (right = R). If my right hand does not
follow the direction of 1, 2, and 3, then the chiral center is S.
2) Connect the groups in the order: 1, 2, and 3. If the direction is clockwise, then the
chiral center is R. If the direction is counter-clockwise, then the chiral center is S.

3) Sometimes the lowest priority group is not facing away from you. In these cases,
you have to reorient the groups around the chiral center to put the 4 group in the
back. There are two methods you could use.
Method A: Switch any three groups. What you are doing here is holding the
tetrahedral atom by one group and spinning the other three (like an umbrella).
Method B: Switch any two groups, which will flip the chiral center. This method is
easier to implement but it is a bit confusing because it will invert the chiral center.
So, if you use this method for assigning the chiral center, you need to switch your
final assignment.
STEP 1: The four different groups attached to the chiral center atom are ranked from
highest priority (1) to lowest priority (4).
STEP 2: Chiral center is reoriented so that the lowest priority group is placed in the
back and the remaining groups are connected in order of priority. If these groups (1,
2, and 3) are in a clockwise order then the chiral center is R. If the groups (1, 2, and
3) are in a counterclockwise order then the chiral center is S.
Rules for each of these steps are described in more detail below.
(STEP 1) Rules for ranking the priority of groups attached to the chiral center

Isomers:Definitions
You are already familiar with the concept
of isomers: different compounds which have
the same molecular formula. In this chapter
we learn to make distinctions between various
kinds of isomers, especially the more subtle
kind of isomers which we call stereoisomers.

Constitutional Isomers: Isomers which


differ in "connectivity". The latter term
means that the difference is in the sequence
in which atoms are attached to one another.
Examples of isomers pairs which are
consitutional isomers are (1)butane and
methylpropane,i.e., isobutane, which are
different in that butane has a sequence of
four carbon atoms in a row, but isobutane
has a three carbon chain with a branch
(2)dimethyl ether and ethanol--the former
has a C-O-C chain, while the latter has a CC-O chain (3) 1-pentene and cyclopentane--

the former has an acylic chain of 5 carbons,


while the latter has a 5-membered ring.

Stereoisomers: Isomers which have the


same connectivity. Thus all isomers are
either constitutional or stereoisomers.
Stereoisomerism is a more subtle kind of
isomerism in which the isomers differ only
in their spatial arrangement, not in their
connectivity. Cis- and Trans-1,4dimethylcyclohexane are a good example of
a pair of stereoisomers.
Stereoisomers

We have just seen that there are two major types


of isomer, but now it is necessary to further
notice that their are two sub-types of
stereoisomers:

Enantiomers: Stereoisomers which are


mirror images

Diastereoisomers: Stereoisomers which


are not mirror images

The examples of cis- and trans-1,4dimethylcyclohexane are of the latter type, that
is , they are diastereoisomers. Cis- and transisomers in general are diastereoisomers. They
have the same connectivity but are not mirror
images of each other. Enantiomers are mirror
image isomers. This is the very most subtle way
in which two chemical compounds can differ:In
an overal sense, then , there are three types of
isomers: (1)constitutional isomers
(2)diastereoisomers and (3)enantiomers in
order of increasing subtlety of
difference. Since we have previously
considered constitutional isomerism, and since
the difference between diastereoisomers and
enantiomers rests upon the concept of mirror
image isomerism, we must now consider this
latter phenomenon in greater detail.
Mirror Image Isomerism

To be isomers, molecules must not be


identical. The test for "identicality" is one
ofsuperimposability. In a sample of butane, all
of the molecules are identical because they can
be superimposed upon one another in some
conformation. The same is true of ethanol or
propanol or 1-butanol, but in the case of 2butanol there are two isomeric forms which can
not be superimposed. They do not differ in
connectivity, obviously, or they wouldn't both be
called by the same name (2-butanol). They also
don't have a cis or trans prefix, to indicate that
they are diastereoisomers. They have a very
specific, unique relationship to one another, the
same relationship which exists between an
object and its mirror image. A key aspect of this
difference, as we all know, is that a mirror acts
to interchange left and right hands.
CHIRALITY

A molecule or object which is not identical


to(i.e., non-superimposable upon) its mirror
image molecule or object is said to

be chiral. This means it resembles a human


hand in that the left and right hands are not
superimposabile but can be readily
distinguished (at least by some of us). By
the same token, a molecule or any object is
said to be achiral if it is identical to
(superimposable upon) its mirror image
molecule or object. Many molecules are
achiral, but many are chiral, especially
complex molecules such as are found in
biological systems.How can we anticipate
when a molecule is chiral and therefore has
an isomer (an enantiomer) or when it is
achiral and has no enantiomer?

Consider 2-butanol, which is an example of


a chiral molecule. The illustration below
(hopefully) shows that the mirror image of
one 2-butanol isomer is nonsuperimposable upon the original molecule.
Your can verify this by making models, but
you can also visualize trying to superimpose
the two by sliding one structure over

(mentally) on top of the other.We can, for


example, slide B over to A and superimpose
the OH, the central C, and its attached H of
the B molecule over the corresponding
gorups of the A molecule, but the ethyl
group on B sits over the methyl group of A,
and the methyl group on B superimposes
upon the ethyl group of A. The two
molecules have all the same kinds of bonds
and are extremely similar, but are mirror
image isomers. We will learn how to name
the two different enantiomers shortly.

Although 2-butanol is a chiral molecule and


therefore has two enantiomers, the very
similar molecule2-propanol is achiral and
does not exist as an enantiomeric pair. In the
illustration, you can see that B slides over
onto A with all corresponding groups
superimposing perfectly. Many simple
molecules are of this kind. How can we
predict whether a molecule is chiral or
achiral?

One of the simple ways is to use the concept


of a stereogenic center. If a molecule has a
single stereogenic center it will necessarily
be chiral. The most common kind of
stereogenic center is a carbon (or other
atom) which has four different atoms or
groups directly attached to it. You can see
that the central carbon of 2-butanol (the one
marked by an asterisk) is a stereogenic
center, having H,OH,methyl, and ethyl
groups attached. Since it has just a single
stereogenic center , it must be chiral. On
the other hand, 2-propanol has no
stereogenic center and is achiral. The
corresponding carbon atom of 2-propanol
has an OH,H, and two methyl groups
attached. Of course, no methyl carbon atom
or methylene carbon can be chiral since
these groups automatically have at least two
identical groups (H's) attached. We will see
a little later what happens when we have
more than one stereogenic center.

The second method, especially useful when


there is more than one stereogenic center, is
the use ofsymmetry elements.If the
molecule or object has either a plane of
symmetry or a center of symmetry it is
achiral. The examples shown below refer
to cis- and trans-1,2-dimethylcyclobutane,
The former of which is achiral and the latter
chiral. They both have two stereogenic
centers, viz., the ring carbons which have
the methyl and hydrogen groups attached,
but one molecule is chiral and the other
achiral. This emphasizes the point that a
molecule or object is guaranteed to be chiral
only if it has a single stereogenic center. If it
has more than one stereogenic center, it may
be either chiral or achiral. Note that in
the cis isomer, the two methyls are on the
same side of the ring and are equidistant
from the plane of symmtery which runs
through the center of the ring perpendicular
to the ring. In thetrans isomer, the methyls
are on opposite sides of the ring, so that

where there is a methyl group on the right


there is a H on the left.

What is the relationship between


the cis and trans isomers of 1,2dimethylcyclobutane??? They
arediastereoisomers, having the same
connectivity but obviously not being mirror
images of each other. To sum up, there are
three isomers of 2,3-dimethylcyclobutane, a
single cis isomer, and two enantiomeric
trans isomers.
The plane of symmetry is relatively easy to
find and is the most common one to look
for, but one other element of symmetry also
guarantees an achiral molecule, and that is
the center of symmetry. This is a point in
the molecule for which any line drawn
through the point will encounter identical
components of the object at equal distances
from the center of symmetry.In the case
illustrated, 2,3-dimethylbutane (the so-called
meso isomer), the center of symmetry is at

the center point of the C2-C3 carbon-carbon


bond. One of the dotted lines shown
connects the equivalent bromines on of the
two carbons,another connects equivalent
methyl groups, and a third connects
equivalent hydrogens (not shown).The meso
isomer is just one of the three stereoisomers
of this system. Again, there is one
enantiomeric pair plus this meso isomer,
which is achiral. A center of symmetry will
be encountered in any molecule which has
two equivalent chiral centers (i.e., both
carbons have the same set of four distinct
substituents) and in a conformation of such a
molecule in which all identical groups are
anti to one another. The two carbons of this
molecule both have H,methyl,bromine, and
1-bromoethyl substituents.

Please note that the stereogenic center


need not be carbon. It can be a quaternary
nitrogen atom ( the nitrogen of an

ammonium salt, if there are four different


groups attached to the nitrogen.

Symmetry Elements Which Guarantee


Achirality

R,S Nomenclature
NAMING ENANTIOMERS
Since two enantiomers are different
compounds, we will need to have
nomenclature which distinguishes them from

each other. The convention which is used is


called the (R,S) system because one
enantiomer is assinged as the R enantiomer
and the other as the S enantiomer. What are
the rules which govern which is which??

Priorities are assigned to each of the four


different groups attached to a given
stereogenic center (one through four, one
being the group of highest priority). (It
should be understood that each stereogenic
center has to be treated separately.)
Orient the molecule so that the group of
priority four (lowest priority) points away
from the observer.
Draw a circular arrow from the group of first
priority to the group of second priority.
If this circular motion is clockwise, the
enantiomer is the R enantiomer. If it is
counterclockwise, it is the S enantiomer.

HOW TO ASSIGN GROUP PRIORITIES

There is also a set of conventions (rules) which


govern the setting of group priorities, which is a
part of the R,S system of nomenclature.

Priority is based upon atomic number,


i.e., H has the lowest priority, O over C, F
over O, and so on. Priority assignment is
based upon the four atoms directly
attached to the stereogenic center. For
example, in 2-butanol, the example we
considered previously, the four atoms are
H,O, and two C's. Oxygen gets the first
priority, and H the fourth. But the methyl
and ethyl groups both are attached through
carbon , so there is initially a tie for the
second and third priorities.
In this kind of tie situation, priority
assignments proceed outward to the next
atoms, which we will call the beta atoms.
(The directly attached atoms are the alpha
atoms). For the methyl group, the alpha
atom is carbon and the beta atoms are three
H's, while for the ethyl group the alpha atom

is also carbon and the beta atoms are two H's


and 1 carbon. This beta C of the ethyl group
wins the priority competition because there
is no beta atom on the methyl group which
has an atomic number greater than 1 (all
three beta atoms are H). In general, the
competition contines from alpha to beta to
gamma to delta atoms until a tie-breaker is
found.

Some additional conventions are necessary


for handling multiple bonds and aromatic
bonds, and these are a little tricky to learn.
As an example, take the vinyl group. Each
carbon of this double bond is considered to
have two bonds to carbon, because of the
double bond. In the case of a carbonyl
group, the carbon is considered to be bonded
to two oxygens, and the oxygen is
considered to be bonded to two carbons. For
this reason, a vinyl group has priority over
an isopropyl group, as shown in the
illustration.

Two Stereogenic Centers


Non-Equivalent Stereogenic Centers

When a molecule has two stereogenic


centers, each of them can be designated as R
or S. Thus there are four possible
stereoisomers. If we designate one
stereocenter as "a" and the other as "b" just

for labelling purposes, the four


stereoisomers can be designated as
RaRb,RaSb,SaRb, and SaSb These designations
correspond to the cirucumstance theat
stereocenter "a" can have the R or S
configuration ,and stereocenter "b" can have
either configuration.

In general, if there are n such stereogenic


centers , there will be a maximum of
2n stereoisomers. For example, with three
stereogenic centers, there are eight possible
stereoisomers. The maximum of 2noccurs
when there are all non-equivalent
stereocenters. Stereogenic centers are
equivalent when all four substituents
attached to the center are identical. For
example, in 2,3-dibromobutane, both
stereogenic carbons have a H, a Br, a
methyl, and a 1-bromoethyl substituent. The
maximum of four stereoisomers is not
observed here, as we saw before. In fact
there are three stereoisomers, including one

achiral stereoisomer. This is because the


2R,3S molecule is identical to the 2S,3R
molecule, since carbons 2 and 3 are
equivalent.

On the other hand, 2,3-dibromopentane has


two non-equivalent stereogenic centers and
there are four stereoisomers, consisting of
two pairs of enantiomers. It should be noted
that the relationship between one
enantiomeric pair and the other pair of
enantiomers is that they are
diastereoisomers..

TWO EQUIVALENT STEREOGENIC


CENTERS

As noted above, when both stereogenic


centers are equivalent, the number of
stereoisomers is less than the maximum of
2n, but in fact is n + 1. In the case of two
stereogenic centers (n = 2), there are 3
stereoisomers, as we saw for 2,3dibromobutane. There is, first of all , a pair
of enantiomeers: these are the (2R,3R) and
(2S,3S) isomers. Note that the mirror image
of 2R,3R is 2S,3S ( i.e., the mirror image
inverts the configuration at each
stereocenter).
There is also an achiral stereoisomer. A
molecule which has stereocenters but is
achiral is called ameso compound. We saw
in an earlier diagram that this molecule has a
point of symmetry in its most stable
conformation.

It should be noted carefully that the meso


isomer is a diastereoisomer of the two
enantiomers.
COMPARATIVE PROPERTIES OF
ENANTIOMERS AND
DIASTEREOISOMERS
DIASTEREOISOMERS

Diastereoisomers are not mirror image


isomers. They are essentially like any other
pair of isomers (e.g., constitutional isomers)
in that they have distinct chemical and
physical properties. Since they have the
same functional groups, however, they are
usually rather similar to one another in their
reactions and properties.
Two diastereoisomers can usually be
separated from one another by , e.g.,
recrystallization, since they have different
solubilities.

Although their chemical


properties(reactions) are similar, the two
diastereoisomers will typically react at
different rates.
ENANTIOMERS

Since two enantiomers are mirror images of


each other, they are not distinguished by any
physical or chemical means which cannot
distinguish mirror images, i.e., which are not
themselves chiral (handed, meaning can
distinguish left from right).
Therefore 2 enantiomers have exactly the
same energy, solubility in typical achiral
solvents, boiling and melting points, NMR
and IR spectra, etc.
Their chemical properties, including both
the qualitative reactions and the quantitative
rates of reaction are identical when reacting
with achiral chemical species.

In general, then, both chemical and


physical properties of 2 enantiomers are
exactly identical twoard achiral
agents,chemical or physical. ,li>It is
important to realize, however, that when 2
enantiome4s react with a pure single
enantiomer of another chiral compound, the
rates of reaction of the 2 enantiomers will be
different (more later).
Also, one physical property which can
distinguish them is "optical activity" (see
below).
OPTICAL ACTIVITY

Since enantiomers are "handed" or "chiral",


they can be distinguished by other agents
which are chiral. Thus, we can easily tell, in
using our right hand to shake hands with
another person, whether that person is using
his left or right hand. There is a better "fit"

of the two right hands than there is of right


hand to left hand.

Chemically this occurs, as noted above,


when enantiomers react with another chiral
compound. Both the original enantiomer and
its reactant distinguish left from right , so
then one of the original enantiomers will
find a better energetic fit with the chiral
compound than will the other.
One physical property which distinguishes
2 enantiomers is "optical activity". This term
refers to the property of chiral compounds
(exclusively) of rotating the plane of planepolarized light to the right (clockwise) or to
the left (counterclockwise).
The two enantiomers have exactly the
same ability to rotate this plane,
quantitatively, but they rotate it in
opposite senses. Thus, if one enantiomer
rotates the plane by 10.5 degrees clockwise
(considered a positive rotation), the other

rotates it by -10.5 degrees (i.e., in the


counterclockwise direction).

Since the exact amount of the rotation of the


plane by a given enantiomer depends upon
how much of that enentiomer the light
encounters as it passes through the solution,
the measured rotation is divided by the
concentration of the enantiomer and by the
path length of the polarimeter cell to give a
true measure of the inherent ability of the
enantiomer to rotate the plane of polarized
light. This number is called the specific
rotation. Note that in deriving the specific
rotation, the concentration is taken in grams
per mL, and the path length in decimeters.
The magnitude of the rotation also depends
upon the wave length of the plane polarized
light, so the a single wave length is usually
used, i.e., the sodium D line (529 nm),the
line responsible for the yellow color of
sodium-vapor lamps.

A positive (clockwise) rotation is sometimes


called dextrorotation and a ngetaive
rotation is sometimes called levorotation
RACEMIC MIXTURES

A racemic mixture is a 50:50 mixture of


the 2 enantiomers of a chiral compound.
Because the two enantiomers have equal and
opposite specific rotations, a racemic
mixture has a specific rotation of zero,
i.e., it is optically inactive
In nature, most naturally occurring
compounds occur as a single enantiomer,
not as racemic mixtures. The importance of
racemic mixtures is that ordinary
laboratory synthesis which generate a
stereogenic center produce a racemic
mixture. For example,if 1-butene is
converted to 2-butanol by the addition of
water catalyzed by acid, a stereogenic center
is created in a molecule where none

previously existed. Since both enantiomers


have equal energy, and since there is nothing
in the catalyst or solvent or reactant that is
chiral, both enantiomers are formed in equal
amounts(for a mechanistic explanation, see
later).

Whereas racemic mixtures are not


particularly desirable, they are not
problematic in many labaoratory organic
syntheses. However, in the manufacture of
drugs, usually only a single enantiomer is
effective, so that it is desirable to synthesize
only a single enaniomer. Nevertheless,
racemic drugs are often used anyway
because the other enaniomer is harmless,
and racemic mixtrues are easier(read,
cheaper) to synthesize.
OPTICAL PURITY

If the specific rotation of a pure single


enantiomer is known, it is easy to determine
the purity of a sample containing both

enantiomers in unequal amounts. The


%OPTICAL PURITY = specific rotation of
the sample/specific rotation of the pure
enantiomer. This particular measure of
optical purity is often called
ENANTIOMERIC EXCESS( or ee) because
it gives %R - %S. A small problem
(admittedly very small, mathematically)
arises in converted the ee (enantiomeric
excess) into a specific composition given in
terms of %R and %S. One simple way of
doing this is as follows: If the enantiomeric
excess of the R enantiomer is, for example,
80%, this means that there is 80% of the R
enantiomer plus 20% of the racemic mixture
(not 20%S). Since the racemic mixture is
10%R and 10%S, the composition of the
mixture is 90% R and 10%S. Remember:
ee represents not the % of one of the
enantiomers, but the difference between
the % of one pure enaniomer and the %
of racemic mixture).

SEPARATION OF ENANTIOMERS

The separation of 2 enantiomers present


in a racemic mixture or any mixture of
enantiomers, is called resolution .
Enantiomers are not readily separated by
conventional means, such as
recrystallization or fractional distillation,
since they have the same solubilities, m.p.'s,
b.p.'s, etc. So, special means are required for
"resolution" of two enantiomers.
One common strategy for resolution is often
to take advantage of the circumstance that,
while enatiomers have the same solubilities
and cannot be readily separated by simple
recrystallization, diastereoisomers have
different solubilites. The two enantiomers
present in a racemic mixtrue can be
reacted with a pure enantiomer of a chiral
compound (called a resolving agent)
which we have on hand (many occur in
pure form in nature). This will form a

compound with two chiral centers, and


will give rise to 2 different
diastereoisomers which can be separated
from each other. Following this separation
the chiral resolving agent rcan be removed
by through some chemical reaction to give
the two separate enantiomers. The chiral
resolving agentcan also be recovered for reuse.

As an example, consider the generalized


case shown in the illustration below.

KINETIC RESOLUTION USING


ENZYMES

Enzymes are proteins which have many


chiral centers and which occur in nature as a
single enantiomer (out of all the myriads of
possible stereoisomers).
The rates of reaction of two enantiomers
with a single enantiomer of any chiral
substance are different. We can see that the
products will be diastereoisomeric, and so of
different energies, and the rates of formation
of these products will in general be different.
In other words, a "handed" molecule can
distinguish chemically between 2 mirror
image isomers. Enzymes are particularly
effective in making this distinction, so that a
racemic mixture can often be easily resolved
by reaction with some simple substance in
the presence of the chiral enzyme as
catalyst. The enantiomer whiich reacts faster
will be converted to a new compound
having an entirely different functional
group, while the enantiomer which reacts
more slowly will remain unreacted. The

separation of the two compounds is then


quite easy.

As an example, if the compound which is


the racemic mixture has an alochol function,
it can be converted to an acetate ester by
reaction with acetic acid in the presence of a
suitable esterifying enzyme. The separation
of the ester of one enantiomer from the
alcohol of the other is then very easy. Then
ester can then be hydrolyzed to the alcohol,
if desired, by either simple chemical means
or by enzyme catalyzed reaction.

RETURN TO THE TOP OF THIS PAGE


ON TO THE NEXT CHAPTER:ALKENES I

BACK TO THE PREVIOUS CHAPTER


BACK TO THE BAULD HOME PAGE

in the molecule - not necessarily on an atom - through which all other atoms can be reflected 180 degrees
into another, identical, atom. (In more accurate symmetry terms, an inversion through a center is
equivalent to rotating groups by 180 degrees and then reflecting the groups through a plane
perpendicular to the rotation axis.) This type of symmetry is rare in organic molecules, and is more
common in inorganic molecules. The inversion center is represented by the blue circle in the above
example. The same molecule is shown three-dimensionally below. The inversion center is in the center of
the middle carbon-carbon bond. This molecule is not chiral because of itsinversion center.

Вам также может понравиться