You are on page 1of 37

CHEM1031 Study Notes

Assumed Knowledge
Acid - proton donor
Base - proton acceptor
Acidic oxides (non-metals) react with water to make acids or
bases to form salts (CO2). Basic oxides (metals) react with acids to
form salts but do not react with alkaline solutions (CuO, Fe 2O3).
Amphoteric oxides (Al, Zn, Pb, Sn) react with acids or bases to form
salts. Neutral oxides (CO, N2O) dont react.
acid + metal H2 + salt
acid + carbonate CO2 + H2O + salt
Gases
Distinguishing properties of gases:
very compressible
flow rapidly
take shape of and fill a container (liquids only take
shape)
expand and contract with temperature changes (more
so than liquids, solids is near negligible)
infinitely expandable (unlike liquids, solids)
low density
Gas variables:
Pressure (Pa)
=force/area. Due to particles in motion, colliding with
momentum into each other and walls. 1Pa = 1N/m 2 = 1J/m3
(1N = 1kgm/s2)
1atm
=
760mmHg/Torr
Manometer - measures
difference in pressure
= 101325Pa
= 101.325kPa
= 1.01325bar
= 14.7psi

Barometer - measures
atmospheric pressure

Page 1

Volume (m3 - 103L)

Oliver Bogdanovski

number/Amount (mass - kg, moles)


Temperature (always in Kelvin; absolute temperature)
These are dependent upon each other in the three Empirical
Gas Laws:
Boyles Law - V (or P1V1 = P2V2) - as pressure
increases, volume decreases
Charles Law - V T (or = ) - as temperature
increases, volume increases
Avogadros Law (also Gay-Lussacs - found when
gases reacted volumetric ratios were small whole
numbers - a stochiometric ratio) - V n (or = ) - as the
number of moles increases, so does the volume
Combining Boyles and Charles Law:
PV T
or
=
Combining all three forms the Ideal Gas Law: PV nT
or
Constant
mol-1

PV = nRT where

Universal

Gas

= 8.3145 J mol-1 K-1 (SI)


= 0.082057 L atm K-1

Standard Temperature and Pressure (STP): 0 oC (273.15K) and


1 bar (1.00105Pa or 0.98atm). 1 mole of gas at STP is 22.7L. We
can also sub in n=m/M and density ( - rho) = m/V to integrate
other values.
Daltons Law of Partial Pressures - in a mixture of gases,
total pressure is the sum of the pressure each gas would exert if
alone under the same conditions (assuming the gases are
independent and do not react):
PT = Pa + Pb + Pc +
Mole Fraction - for each component A in a mixture, the mole
fraction is (a value between 0 and 1 - not percentage - to express
the percentage of moles of that substance in a mixture):
XA =
Partial Pressure of A
PA = XAPT
Each gas also obeys the Ideal Gas Law independently as if
they took up all the volume, and hence were P T=PA+PB, PAV=nRT
and PBV=nRT. However these conclusions in the 17 th-19th century,
and it wasnt until the 19th-20th century that a theory of atoms began
to form, so these laws all looked at macroscopic ideas, influenced by
what we know to be properties of microscopic atoms.
Kinetic Theory of Gases:
molecule size is negligible compared to distance
between them
molecules move randomly in straight lines in all
directions at various speeds
forces of attraction/repulsion are negligible (because
they are very weak) except in collisions

Page 2

Oliver Bogdanovski

gas particle collisions are perfectly elastic


Ek av absolute temperature
This explains Boyles Law as less space means more frequent
collisions, and hence higher pressure (as collisions result in a force
applied), and Charles Law as increasing temperature, kinetic energy
(molecule speed) increases, so collisions become more frequent
and with greater force.
Kinetic theory states Ek av is only dependent on temperature,
not gas type, and difference gases at the same temperature have
the same average kinetic energy. As k = m2/2, heavier gases will
travel more slowly with the same energy.
It can be found that (dont need to know derivation):
k =
NA is Avogadros number. Remember this is per molecule, so to find
per mole multiply by Avogadros number. Combining this with our
other formula for k:
Rate of Gas Movement:
rms =
Root-mean-square (rms) simply means we have square-rooted the
mean value.
Effusion - escape of molecules through a hole of molecular
dimensions (assuming no collisions between molecules)
Diffusion - mixing of gases until the mixture is homogeneous
Using the above rate and these ideas (in diffusion it could be
two gases reacting and producing a colour located at a particular
point and speeds) we can determine molecular mass.
Grahams Law - The rates of effusion (and diffusion) of two
gases at the same temperature and pressure are inversely
proportional to the square roots of their densities (note time is
inversely proportional to rate):
=
=
=
All gases are actually non-ideal:
all particles do have volume - becomes significant at
high pressure (real volume > ideal volume as ideal
volume hits zero)
they have attractive forces - significant at low
temperatures (real volume < ideal volume as particles
are brought together; gases with low interatomic
dispersion forces like He do not experience this)
particles do interact - negligible at high temperature
(enough energy to keep bonds apart), but significant at
other temperatures (real pressure < ideal gas pressure
as there are less molecules - chemically bonded
together - and hence less collisions)
All known life depends upon the atmosphere, however the
atmosphere doesnt have a definite end, with 99% within 30km,
outer space at ~10,000km.
Atomic Structure
Page 3

Oliver Bogdanovski

Only valence electrons determine chemical properties, and


hence isotopes have nearly identical chemical properties. Light is
electromagnetic radiation (a self-sustaining oscillation of electric
and magnetic fields), and is characterised by its frequency ( - nu;
Hz or sec-1) or wavelength (; m or angstrom==10 -10m), which are
related by c=, with visible light being 3.9-7.0 10-7m, whilst gamma
rays are around 10-12m and long radio waves 104.
Monochromatic Radiation - a selection of one frequency (in
practicality, a narrow band of frequencies) for various scientific
measurements
Polychromatic Radiation - consisting of many frequencies
Light has typical wave-like properties (refraction, diffraction
and interference), however also exhibits the photoelectric effect,
discovered in 1887 by Hertz who found light could eject electrons
from the surface of a metal, and a current could flow to another
electrode in a vacuum. He also found it required a threshold
frequency that was dependent on the type of metal used which was
independent of the intensity, however once above the threshold
frequency the intensity increased current size, and the energy of the
electrons emitted depended on the frequency. In 1905, Einstein
realised light comes in packets or quanta (got Noble Prize in 1921),
where each quantum of energy is proportional to frequency:
E = h
where
h = Plancks constant = 6.62610-34Js
A photon with enough energy could be absorbed and eject
an electron, producing a current, and only one with enough
energy could overcome the attraction of the atom, the
remainder energy being converted to kinetic (E k = h-W).
The energy of a particular orbital can be found by the
Rydberg Equation:
= RH
where
RH = Rydberg constant for H (on data sheet)
n1 = lower shell
n2 = upper shell
For hydrogen the shell is a good indicator of electron energy: E
= -RH/n2
White (polychromatic) light passing through a
gas composed of single atoms gas lines (or specific
frequencies) removed, forming an absorption
spectrum (the release of the photons once electrons
fall is in all directions, and hence much weaker at the
detecting screen; a prism can be used to distinguish
between colours). When heating a gas by electrical
discharge, it produces these series of lines in an
emission spectrum. This is because electrons
occupy discrete energy states that they move up or
down. The spectra vary with the gas used and
pressure (proximity alters energy of shells).

Page 4

Oliver Bogdanovski

Lyman = UV
Balmer = visible light
Paschen = IR
The value of n for each energy level/shell is the principal
quantum number. The ground state is an atoms lowest state,
however it can undergo transitions to higher excited states by
heating or colliding energetically with other bodies. These are
unstable and result in the lowering of energy levels by emission of
photons.
Complete removal of an electron means the electron has been
moved to n=. The energy required to move a valence electron
upwards is called the ionisation energy.
Solutions to the Schrdinger equation have exact analytical
forms for the hydrogen atom:
=E
where
= Hamiltonian (an operator that corresponds to the
total energy of the system
- encompasses nature of proton
and electron particles and their Coulombic attraction [the
electrostatic attraction or repulsion between protons and
electrons])
E = energy of the state (a constant of proportionality)
(psi) = the wavefunction
An eigenstate (or orbital) is an allowed energy (or shell) under
the contraints of the Schrdinger equation (labelled by quantum
numbers - they are the outcomes or results when solving the
equation). These are wave-like states with 3D shape and amplitude
(this form is a direct consequence of the Schrdinger equation. The
electron density (probability of an electron being at a certain point)
is given by the square of the wavefunction (however the Heisenberg
uncertainty principle limits the ability to know both the position and
energy (thus speed) of a quantum particle like an electron):
=2
As the electron could be at any distance from the nucleus (although
further is less probable, the volume with a 90% chance of an
electron being there is called the boundary surface, and this
surface is thought of as the spatial limit of the atomic orbital. There
are four quantum numbers to label each electron:
principal quantum number (n) - 1, 2, 3, ,
azimuthal (angular momentum) quantum number (l) - 0,
1, 2, , (n-1)
magnetic quantum number (ml) - 0, 1, 2, , l (but
counted from negative to positive: -l, -l+1, , 0, , l-1, l)
spin quantum number (ms) -
A set of orbitals with the same n are called a shell (in hydrogen
all orbitals are in the same shell as there is one electron), and a set
of orbitals with the same n and l are part of the same sub-shell,
labelled by letters (called orbital symbols):
if l=0 s orbital
if l=1 its a p orbital
Remember: Spadoof!
if l=2 its a d orbital

Page 5

Oliver Bogdanovski

if l=3 its an f orbital


then g, h, i and so on
This is written as: (n)(orbital symbol)no. of
electrons
e.g. 2s2
s
orbitals
are
spherically
symmetric, however the inner orbitals
often deflect the outer orbitals (by
Coulombic repulsion) resulting in their
electron densities peaking further away.
p orbitals consist of two lobes of
electron density on opposite sides of
the nucleus with a nodal plane (zero
electron density) between them. As
there are three ml values, there are
three types of orbitals: px, py and pz.
d orbitals have either a cloverleaf shape (d xy, dyz, dxz, dx -y ) or
two lobes and a torus (dz ).
2

The Coulombic attraction between the nucleus and electrons


leads to a contraction of shells as you move to the right of the
periodic table, requiring more energy to pull the electron out due to
the increasing nuclear charge. Multi-electron atoms are more
difficult to obtain analytical solutions of through the Schrdinger
equation as the electrons repel each other, however we can still
approximate orbitals that resemble the hydrogen atom. These
repulsions are considered electronic shielding, and do not affect
the electrons of an outer shell equally:
s electrons of an outer shall have at least one smaller lobe
of density closer to the nucleus (inside the region of
shielding electron) and hence are less affected as they are
more often closer in and not further out
the effect is smaller for p orbitals, then d, then f, and so on
Thus the degree of shielding goes s<p<d<f (opposite is
degree of penetration)

Page 6

Oliver Bogdanovski

When determining electron configuration (a particular arrangement


of electrons), orbitals fill up from lowest energy to highest, and this
can be remembered by the filing from the top right (first arrow, then
second, and so on).
The aufbau (building up) principle assembles atoms by
adding one electron for every proton (and usually neutron) into the
lowest energy orbital free. This follows two rules:
Pauli exclusion principle - two electrons cannot have
all the same quantum numbers (and hence a maximum
of two electrons in the same sub-orbital (same m l),
having two different ms)
Hunds rule - in a subshell of orbitals (same l or orbital
symbol, but different ml) electrons distribute one
electron in each orbital alignment (the first spin
quantum number) before they go back and fill it up the
remaining orbital alignments with the other quantum
number
o the only exception is when you can get a half-filled
instead of a nearly filled as that is more stable,
and this only occurs in two transition metals and
the ones below them (Group VI and XI):
Cr (Chromium)
[Ar]4s13d5
NOT
2
4
[Ar]4s 3d
Cu (Copper) [Ar]4s13d10
NOT [Ar]4s23d9
Electron configuration can be written in the arrows in boxes
method (each direction of the arrow shows a different spin, or by
number (which could be replaced by a noble gas configuration
followed by valence electrons):
Oxygen

1s 22s22p4

[He]2p 4

Once you get to the point where 4s is before the 3d due to less
shielding (greater penetration to the nucleus), the 4s is listed before
the 3d.
A different method uses locations on the
periodic table, where the core part of the
configuration is given by the noble gas from the
period above, and the valence part is given by
counting from the left of the table to the element
and filling in shells as you go (staying aware of the
Cr and Cu anomalies).
Anion ground-state electron configuration is simply adds the
number of required electrons to the next available positions,
following the aufbau principle.
Cations are usually done the same way except removing the
last added electrons, with an exception: in neutral atoms, the 4s is
more stable than the 3d and filled up first, but in transition metal
cations, the reverse is true, so all their valence electrons reside in
Page 7

Oliver Bogdanovski

the d orbitals and it is these that are removed whilst the d remain
filled.
Fe
[Ar]4s23d6
Fe+ [Ar]3d7

note how the other s electron


moved to d
Fe2+ [Ar]3d6
Fe3+ [Ar]3d5
In general, the order of electron removal is outermost (n) p,
outermost (n) s, then (n-1) d, and so on.
Paired electrons (two electrons in the same orbital alignment
with different spins) are repelled very weakly from a magnetic field,
and are diamagnetic (dia = going apart; dialysis; basically not
magnetic). Unpaired electrons are attracted more strongly into the
field, and are paramagnetic (para = distinct from, but similar to;
Paralympics). If an atom contains at least one lone electron (one
without an opposite spin) then the paramagnetic effect (attraction)
of just one electron is stronger than the total diamagnetism
(repulsion) of many electrons.
If atoms packing into a solid, then the
atomic radius is considered half that of the
internuclear separation. Going down the
periodic table, extra shells are added and
atoms get bigger, but going across from left
to right, nuclear charge increases, pulling in
outer electrons. Small variations occur due to
new orbitals being used, however these are
not significant due to shielding. Size is often
measured in picometres (one-thousandth of a
nanometre).
Cations are always smaller than their parent atoms as they
form by loss of electrons, causing the 90% density surface to shrink
as there is less shielding, whilst still maintaining the nuclear charge,
resulting in a stronger pull on each electron.
Anions are always larger than their parent atoms as there is
greater shielding (electron-electron repulsions), and they each have
less share of the nuclear charge as there are more electrons,
allowing them to move further away. However they are generally
still smaller than cations from the same period due to the larger
nuclear charge (they have more protons).
The first ionisation energy is the energy required to remove
the least tightly bound electron from a gaseous atom (atom (g)
cation+ (g)+ e- (g)), whilst the second ionisation energy is the
energy required to remove the least tightly bound
electron from the gaseous 1+ ion (cation + (g)
cation2+ (g) +e- (g)). An elements 2nd IE is always
great than the 1st as the nuclear charge per
electron will have increased. Successive IEs
continue to increase, with a stark increase when
the noble gas configuration is forced to ionise.

Page 8

Oliver Bogdanovski

Elements with low 1st IEs usually lose electrons to form cations,
whilst those with high 1st IEs gain electrons to become anions. IEs
decrease down groups (shielding pushes
electrons out, lower energy required), but
increase across a period from left to right
(greater nuclear charge), however whenever a
subshell has reached a filled or half-filled state,
there will be a small jump.
Electron affinity is the opposite of
ionisation energy, being the energy change
when a gaseous atom gains an electron to
form a gaseous anion (atom (g) + e- (g)
anion- (g)). EAs are negative as the system
loses energy (energy is released, exothermic), and hence larger
negative values are favoured (unlike IE), being located at the top
right of the periodic table. Group II elements have the greatest
electron affinity (being large and positive - represented as zero on
the graph as positive values can only be estimated through
calculations, not measured) as this would result in starting a new,
more unstable orbital.
Bonding Theory and States of Matter
When complete transfer of one or more electrons occurs from
one atom to another, it is an ionic bond (although this never
completely happens, and is very weakly covalent). If main electron
density of the bonding electrons moves to a region between the
atoms, it is a covalent bond, and if it is intermediate (electrons
generally closer to one atom than the other), it is a polar covalent
bond (ionic bonds can be thought of as extreme versions of these. A
special case of covalent bonding is metallic bonding, where the
valence electrons are not localised but instead move freely through
the metal body.
Electronegativity is a means to predict which of these a
compound is, where if there difference in the electronegativity of
two atoms is:
>1.7 ionic
0.4-1.7 polar covalent
<0.4 mostly covalent (or simply covalent)
0 non-polar covalent (or simply covalent)
In polar bonds, bond polarity can be shown with a dipole
arrow (pointing from the positive to the negative (sometimes with a
small vertical line at the positive end to make a + shape). As these
are vectors, they can be combined to give an overall molecular
dipole moment.
The attractive forces between molecules (van der Waals
forces) are weak and take little energy to break in melting/boiling
processes, and hence the more ionic a substance the higher its
melting/boiling point. Some binary polar-covalent compounds dont

Page 9

Oliver Bogdanovski

form discrete molecules but networks or lattices with all atoms


linked (e.g. SiO2) and hence have higher melting/boiling points.
Another means of expressing electron configuration is Lewis
electron-dot symbols, where each atom is represented by its
atomic symbol, valence electrons are represented as dots around it,
and a line between two atoms represents a pair of shared electrons.
To draw a Lewis Structure:
1.
Place the least electronegative element in the
middle (except H)
2.
Assemble a bonding network using single bonds
3.
Put 3 lone pairs of electrons on all outer atoms
except H so they each form a Lewis octet (8 valence
electrons)
4.
Assign the remaining valence electrons to the
central atom
5.
Minimise formal charges (formal charge = no. of
valence electrons - number assigned in Lewis structure,
where lone pair=2, bond=1; summing the formal
charges gives the overall charge; if unable to balance
formal charges, negative charges go on most
electronegative,
positive
charges
on
least
electronegative)
Elements in the third period and below can have an
expanded valence shall due to the availability of d orbitals. In
contrast, some elements (B, Be, Al) cannot form Lewis octets in
certain circumstances, and become electron deficient species,
not having a total of 8 electrons (e.g. BeCl 2 or BeCl3). Some
molecules are stable with an odd number of electrons, leaving an
unpaired free radical electron (e.g. NO2).
When completing Lewis structures, there may be more than
one way to minimise formal charges, and this is called resonance,
written as shown in the
No. of
Geometrical
diagram (or in one structure
Name
Pairs
Shape
and stating how many other
2 (AX2)
Linear
structures possible). Note no
electron pair is jumping all
the way around, but instead
3 (AX3)
Trigonal
they jump from the atom to
Planar
being bonded and back. In
reality the bonds do not flip
back and worth, and each NO bond length is somewhere
Tetrahedr
4 (AX4)
between a single and double
al
bond
(approximately
calculated by averaging the
5 (AX5)
lengths in one of the Axial=gree Trigonal
resonance structures).
n
Bipyrami
To find 3D shape, we Equatorial
d
use Valence Shell Electron
=blue

Page 10

6 (AX6)

Octahedr
Oliver Bogdanovski
al

Pair Repulsion (VSEPR), which uses Lewis structures and the


repulsion of bonded (BP) or lone pairs (LP) of electrons. VSEPR
makes no distinction between single, double or triple bonds. Instead
of a bond to a new atom, a blue or green sphere could also
represent a lone pair. The general form of a central atom (A)
surrounded by other atoms (X) is AX n, or if there are lone pairs (E),
AXn-mEm where n is the number of electron pairs (bonded and/or
lone), m is the number of LP.
This geometry is modified due the hierarchy of electron pair
repulsions (from strongest to weakest): LP:LP>LP:BP>BP:BP. This
means two neighbouring pairs will deflect, and depending on what is
surrounding it, may cause the angle to widen (and shorten the
others). Only in repulsion are multiple bonds considered: multiple
bonds are shorter and fatter clouds of electrons, and thus have
stronger repulsion than single bonds.
Because lone pairs can replace outer atoms, and they take the
place furthest from other lone pairs and bonded pairs, the actual
shape of the molecule is different to its geometry (having one or
more outer atoms removed).
Dipol
Compou
Examp
e
Geometry
Shape
nd Class
le
Mome
nt
trigonal
AX2E
bent
SO2
Yes
planar
AX3E
tetrahedral
trigonal pyramidal
NH3
Yes
AX2E2
tetrahedral
bent
H2O
Yes
trigonal
seesaw (remove
AX4E
SF4
Yes
bipyramid
equatorial)
T-shaped
trigonal
AX3E2
(remove two
ClF3
Yes
bipyramid
equatorials)
I3trigonal
AX2E3
linear
(triiodid
No
bipyramid
e)
AX5E
octahedral
square pyramid
ClF5
Yes
AX4E2
octahedral
square planar
XeF4
No
A dipole moment (, measured in coulomb metres: C m overall asymmetric distribution of electron density) only occurs in
atoms where the outer atoms are identical and pull in a net
direction, or have different electronegative charges (being different
atoms) sufficient enough to produce a dipole moment, however this
is not something we particular care about (only the shape part).
Molecules with a dipole align in an electrical field (as opposites
attract). Dipole moments typically span 0-710-30C m, and the
values are directly related (but not exactly proportional) to
electronegativity difference.
In VSEPR, resonance should be treated with an average bond
order (being 1.3 in NO3-), and these are treated as one bond, but

Page 11

Oliver Bogdanovski

when looking at repulsions as an intermediate between a single


bond and double bond. In larger molecules like hydrocarbons, VSEPR
rules apply to each central atom (each carbon in the case of
hydrocarbons), and they should be treated independently.
VSEPR produces quite accurate results, except for transition
metals and some other molecules, however we dont need to know
these.
The angles in VSEPR do not match up those in the
varying orbital alignments (which are often at 90o to each
other) and hence a new model formed by Valence Bond
Theory (VBT) is required. As electrons can be considered
waves, their orbitals have wave-like properties, and orbital
interactions are analogous to the superimposition of
waves. That is, when two atoms bond and overlap
slightly, the electron density is the sum of the two
parts. The new orbital is a composite of the old
orbital, and is called hybridisation. This leads to the
equalisation of energies of the valence orbitals (as
they become one orbital) and allows for the greatest
possible number of unpaired electrons for bonding.
Note how positive (blue) with positive enhances the
final blue (as they are being added; constructive),
whilst when added to a
negative (pink) it is
destructive. Also note
that the number of
orbitals remains the
same
(in this case three ps
and
one
s),
and
the
notation in that the
three
ps become p3.
Overall
in
methane:

Outer atoms (that arent hydrogen) also hybridise before


bonding (e.g. CH2Cl2 - both chlorines hybridise 3s and 3p orbitals to
generate sp3-chlorine).
This has been for four electron pairs,
however in cases like BF3 there are only three
bonds to be made as there are only three free
valence electrons, and so an sp2-orbital is
made (with one electron within each box, and
the composite being an electron from the
fluorine). However, as one p-orbital alignment
is ignored (say pz) it does not conform to the

Page 12

Oliver Bogdanovski

new energy orbital and remains empty at a slightly higher energy


state. This atomic pz orbital remains vacant.
Similarly in BeCl2, as only two orbitals are required, the 2s and
2px hybridise to form sp and leaving a vacant 2p y and 2pz (note how
because hybridisation occurs in the valence shell there is no number
outside of the hybrid orbital).
Sets of
Unused
Hybridisa
Electrons
p-orbitals Geomet
Diagram
tion
(each gold
(blue+pin
ry
lobe)
k)
sp

linear

sp2

trigonal
planar

sp3

tetrahed
ral

Sets of electrons is the number of lone pairs plus the number


of atoms it is bonded to (so multiple bonds count as much as single
bonds). In elements below the 2nd period, d orbitals come necessary
to form sp3d and so on.
To use VBT we:
1. draw Lewis dot structure
2. apply VSEPR to determine shape
3. choose the hybridisation model that fits (sp, sp2, sp3)
4. construct -bonding assembly from partially filled
unhybridised p-orbitals
We can extend the VBT model to account for double or triple
bonds using the unused p-orbitals. For example, in ethene:

Page 13

Oliver Bogdanovski

The electron in the second bond of the double


bond sits in the 2pz orbital, and these two adjacent
orbitals form a -bond, which accounts for the second
bond.
Normal, direct single bonds (called -bond)
have far greater orbital overlap, and hence are
stronger. In addition, the closer the energies between
the bonds the stronger the -bond. -bonds have
poorer orbital overlap and hence weaker bonding, and
they lie around the -bond. In a triple bond, there are two -bonds
(in addition to the -bond) which occur at 90 o to each other. In
resonance structures these are called delocalised -bonds.
In larger molecules such as PCl 5 (trigonal bipyramid) and SF 6
(octahedral) hybridisation can go beyond s- and p-orbitals, and into
d-orbitals. For example, in SF6, the sulfur
must be able to present 6 free places to
bond with, and hence one s orbital, three
p orbitals and two d orbitals are used to
make sp3d2.
Extended -bonding or extended -orbital overlap occurs
when two sets of -bonds are close to each other (as in butadiene,
CH2=CH-CH=CH2, or grapheme, where alternating double bonds
result in extended -bonds), allowing it to conduct electricity due to
the delocalisation of electrons (that is, a larger -cloud of electrons).
However this cannot be explained by VBT, and hence we use
another new model.
Molecular Orbital (MO) Theory is used to account for this,
in addition to the structure of molecules like diborane (in diagram),
and the apparent paramagnetism of some simple molecules like O 2,
when it is predicted they should be diamagnetic.
In MO Theory, every two atomic valence orbitals that combine
form another two orbitals (unlike VBT where only one was formed,
and hence they are different). MO Theory is based upon the overlap
of atomic orbitals. Atomic orbitals (AOs) are mathematical
probability for regions of electron density (determined by squaring
the wavefunction), and as they are waves have positive and
negative amplitude, which simply tell us the side the electron is on.
Being waves, when combining they can be constructive when in
phase (called bonding - when two like signs combine) or
Page 14

Oliver Bogdanovski

destructive when out of phase (antibonding - when two opposite


signs combine; positive and negative). VBT does not account for
antibonding. These can be calculated for H2 by the equations:
bond = ((1s)A + (1s)B)
antibond = ((1s)A - (1s)B)
This can be shown diagrammatically:
Antibonding orbitals (indicated by an asterisk * after the )
always have one more node than their bonding counterparts. In the
bonding of H2, two equal-energy orbitals (1s)
can produce either a bond (which has a
lower energy, and hence is more stable) or
antibond (which has a higher energy and is
less stable). This new electron configuration
is written as 1s . As there are only two
electrons, it is filled from the bottom first,
and as 1s is more stable it is filled first
(labelled as 1s as there are two electrons).
Bond order is also redefined:
Bond Order =
In the case of H2, as there are two
bonding electrons and no antibonding
electrons, bond order = (2-0)/2 = 1. If bond
order is 0, there is no bond (as can be seen
in the case of He2 below, however He2+ does
exist). It cannot be <0 (as bonding electrons
always fill first).
This has all been for elements
in the first period, but it works much
the same in the second period (in
the diagram the p-orbitals have been
generalised, however if it was period
2 you would use 2pz, 2p, *2p, 2py,
2p and *2p). For formalitys sake, pz
is used for horizontal -bonds, whilst
px and py are used for vertical bonds. Note how the number of
nodes is the same before and
afterwards in constructive,
and there is one more in
destructive.
This model can be used
to
explain
the
paramagnetism of oxygen as
using the aufbau principle
(and
abiding
by
Paulis
Exclusion
Principle
and
Hunds Rule) we get lone
electrons that arent paired.
2

Page 15

Oliver Bogdanovski

This is different to VBT as in VBT electrons are involved from


the beginning in the hybridisation of orbitals (whilst now they are
considered two at a time), and hence we never think of an empty
hybridised orbital (whilst MO theory requires unfilled or partially
unfilled antibonding orbitals), and there is no MO anti-bonding
equivalent in VBT.
The bond order in MO is analogous to single, double and triple
bonds in more simple bonding theories, so as bond order increases,
bond energy/enthalpy (the energy required to break the bond) also
increases as there are more bonds to break, however bond length
decreases due to electrostatic attraction:
Bond Order
Bond Energy
Bond Length
There are two critical factors that determine whether an MO
will form and how stable the bonding orbital will be (or how unstable
to antibonding orbital is):
1) degree of overlap (more overlap (>-bonds)=greater
bonding/antibonding character)
2) similarity in energies of contributing orbitals (a 1s
orbital will form a lower energy bonding orbital or even a
higher energy anti-bonding orbital with another
1s rather than going with a 2s)
nd
2 period elements often exist as homonuclear
diatomics (Li2, B2, etc. - all except Be and Ne as antibonding cancels with bonding). However, by this theory, B 2
should be diamagnetic when it shows paramagnetic
properties. This is because the 2s and 2p orbitals are
similar in radii, and hence energies, causing some overlap
(particularly in those before and including N) resulting in
orbital mixing (or s-p hybridisation) and - crossover. Li
and Be are done normally, but B, C and N have a FLIPPED
-bonding and -bonding order, so the -bond is actually
filled up first. This stabilises the 2s (s is just the diamond
of orbitals; 2nd shell, so valence), whilst destabilising the
2p (from the 2pz orbital), or in other words causes them to
move further apart in energy. As the resulting orbitals are
not derived from one
atomic
orbital
from
each atom in the bond,
the labels s and y are
used
to
show
the
predominant type. As
shown in the diagram of
energy levels, mixing
occurs most where the
energies are closest.
It is important to
note that the number of
molecular orbitals is

Page 16

Oliver Bogdanovski

equal to the number of atomic orbitals that made them, the bonding
orbitals are lower than the atomic orbitals from which they were
formed whilst antibonding is higher. It follows both Pauli exclusion
principle and Hunds rule, and they form best when composed of
atomic orbitals of like energies. In heteronuclear diatomics, the
more electronegative atom will have lower atomic orbital energies
(as energies are negative and it requires more energy to pull out an
atom). The bigger the difference in energies (or electronegativity),
the greater the chance of orbital mixing (so NO does not have
orbital mixing, whilst CO does) as the different orbitals (e.g. the 1s
of one and the 2pz of another) are likely to cross. Additionally, if both
atoms are below nitrogen (as in CN -) there will also be orbital
mixing. Another means of deducing this is by considering what
something is isoelectronic (having the same electron configuration)
with. So CO would be the same as N2, and hence has orbital mixing.
Intermolecular Forces
The state of a substance (the boiling/melting point) depends
on the strength and nature of the forces between atoms/molecules
of a substance:
Dispersion forces - a molecule with no
permanent dipole that can become
temporarily polar when collisions induce
unsymmetrical
electron
distributions
(distortions of the electron cloud), occurring
in all molecules; polarisability is the ease of
distortion and determines the strength of the force,
increasing with the number of electrons present (as
more electrons means larger molecule which means less
electronegative); molecular shape also determines
strength of the force, with a higher SA:V ratio (so more
SA) being higher forces as they are more polarisable
Dipole-dipole - two molecules with permanent dipole
moments attract each other by the approach of
oppositely charged ends, which can be caused by
asymmetric structure or peripheral atoms of differing
electronegativity, and we label this with + and - or an
arrow
pointing
from
positivenegative with a
perpendicular line through it at the positive end
Hydrogen bonding - the strongest intermolecular force
is the result of a dipole-dipole interaction between a
strongly polarised H (due to high electronegativity of
O/N/F-H) and an O, N or F (note: requires two of these
high electronegativity atoms, one to create the dipole,
the other to bond with; it is almost 10 stronger than
other intermolecular forces, dominating molecular
properties (evident in binary compounds as boiling point
increases drastically despite smaller molecular size); has
the form: donor-Hacceptor (the donor makes the H+,
Page 17

Oliver Bogdanovski

and the acceptor has a lone pair of electrons for


attraction)

Non-Ideal
Gases
The
ideal gas laws
make two assumptions:
there is no attractive force between atoms/molecules
the volume of the atoms/molecules is negligible
In a compressibility isotherm
(graph), we can plot the gas equation
ratio (which should always be 1 in an
ideal gas) against pressure. At 0
pressure, the ratio is 1, however as it
increases it deviates below 1 as
intermolecular forces play a role,
attracting the molecules and reducing
volume, and hence the ratio. However,
as volume of the total gas decreases
the volume of each particle becomes
significant, stopping the volume from
becoming any smaller until the ratio
begins to increase (as pressure is still
increasing but volume isnt decreasing
as much) so it surpasses 1. In noble gases, there are so few
intermolecular forces (as molecules are monatomic and hence
dispersion forces small), so they dont really dip down. If forces
arent as strong (for example, less available valence electrons in N 2
and CH4 compared to halogens), then the dip down is not as strong.
If the volume of the molecules is larger, then the effects of volume
push up the ratio faster as they come into play with a larger effect
sooner.
To correct for this, we use the van der Waals equation:
The pressure correction term (first bracket) takes into account
the intermolecular forces (note: a = strength of intermolecular
forces; larger a means stronger; added as pressure increases). The
volume correction (second bracket) accounts for the molecular
volume (note: b = volume of gas molecules; larger b means larger
molecules, less free volume; subtracted as free volume decreases).

Page 18

Oliver Bogdanovski

Both melting and boiling point can be used as measures of


intermolecular forces. At the boiling point of a substance, the
intermolecular attractive forces balance with the kinetic energy of
the substance, and if this occurs at P=1.01310 5 Pa (1atm) it is the
normal boiling point. At the melting point, the kinetic energy
energies of the molecules are at a point where solidification and
liquidation are equally possible (also with a normal melting point).
Thus the rates of escape and capture from a phase depends on the
kinetic energy and intermolecular forces, larger intermolecular
forces producing higher points.
Condensed Phases of Matter
Solids have fixed shape, high density and low compressibility
due to the strong bonding between the atoms. The atoms may pack
in a crystalline solid (long range order) like silica/quartz, or
irregularly to give an amorphous solid (disorder) like glass.
Molecular or atomic solids are made of covalently bonded
atoms or free atoms, held together by dispersion forces, dipole
forces and/or hydrogen bonding, and their melting point is
dependent on these forces (although it is generally low). Atomic
solids are held together by just dispersion forces.
Networked/covalent solids have far higher melting points
(like silicon, 1410oC) than molecular solids (like phosphorus, 44 oC,
despite both being sp3-hybridised) as there are many more bonds to
break.. Diamond is an sp3-hybridised 3D network with only covalent
bonding, whilst graphite is an sp2-hybridised 2D network, having
both covalent bonding and dispersion forces (that only exist at the
delocalised electrons that form -bonds between the planes)
between planes (giving it actually a slightly higher melting point of
3652oC vs. 3550oC). They are hard and non-malleable (held in
position).
In metals, there is a sea of mobile valence electrons about a
regular array of metal cations, and the melting point is dependent
upon the number of valence electrons (more means greater forces,
up until after the middle of the d-block, after which electrons begin
filling antibonding orbitals and weaken the bonds) and the cationic
radius of the metal (smaller=closer=stronger). Thus they can have a
wide range of melting points. They are electrical and thermal
conductors, and malleable/ductile as the cations can be arranged
anywhere in the electron cloud.
Ionic solids have electrostatically attraction cations and
anions that drive it to become electrically neutral. Despite lack of
covalency, melting points are quite high, and depends on the charge
of the anions and cations (being greater when the number is bigger)
and the size of the cations and anions (decreasing size leads to a
higher charge density and hence higher melting point; also
important is if they are a similar size it tessellates better, increasing
charge density similarly). They are brittle as any distortion results in
electrostatic repulsion between ions.

Page 19

Oliver Bogdanovski

Liquids are one of the rarest phases, being the narrow phases
where kinetic energy means the substance cannot remain
independent but it higher enough to ensure free movement. They
have a specific volume like solids, but are fluid like gases. They have
three important properties:
Surface Tension - resistance of a liquid to increase its
surface area due to cohesion from the intermolecular
forces between the molecules, with surface molecules
being pulled back into the main body
Capillary Action - adhesive (with cohesion) that leads
water to climb through small-diameter tubes, only
limited by gravity; leads to meniscus
Viscosity - dependent on strength of intermolecular
forces (slows it down) and molecular shape (more
surface area means more places to have forces or to
bump into each other); also dependent upon
temperature, as increasing kinetic
energies allows greater movement by
overcoming intermolecular forces
All substances have a kinetic energy
distribution, so even in a liquid phase some will
continuously vaporise to become gaseous. The
partial pressure of a liquid in a closed container
gradually increases once the system has been
sealed as some gas molecules are released,
however eventually an equilibrium is reached
where the number of molecules that leave the
liquid is the same as the number of gas molecules
that collide with and are recaptured by the liquid. The partial
pressure of the gas above the liquid at equilibrium
is called the vapour pressure of the liquid.
Vapour pressure is approximately proportional to
the a constant from the van der Waals equation.
The temperature at which the vapour
pressure of a substance in an open container
reaches the external pressure is called the boiling
point, and if this occurs at 1atm this is the
normal boiling point.
Phase Changes
Phase
changes
depend
on
intermolecular
and
bonding
(intramolecular) forces. Phase changes
require energy, and hence as heat is
added at particular changes of phase the
energy is not added as temperature but
used as kinetic energy to overcome the
intermolecular forces. Hence temperature
does not increase at these phases.
Page 20

Oliver Bogdanovski

Phase diagrams describe the


most stable phase behaviour of a
substance across T and P. Lines
represent phase boundaries (states
occur at equilibrium). The point of
equilibrium between solid, liquid and
gas is the triple point. At a given
point, at constant pressure adding heat
will not change the temperature of the
substance until it has been completely converted to the new phase
(as seen in the diagram on the previous page. After the critical
temperature (at the critical point; the critical temperature is
determined by the strength of the intermolecular forces), the gas
will not condense into a liquid at any pressure. The line between
solid and liquid is near vertical as it is almost independent of
pressure (however note this would go the other way in water, as ice
takes up more volume and hence increasing pressure (decreasing
volume) would convert from solid to liquid at constant temperature).
Dont forget to learn transitions between phases. To sketch a phase
diagram:
1) Draw P and T axes
2) Place triple point in lower left quadrant
3) Place critical point (if given) in upper right quadrant
4) Plot melting and boiling points at P=1.013105 Pa
5) Draw lines linking the points (starting at (0,0)) to the
extremes (except stopping at the critical point)
If temperature is high enough (past the critical temperature),
no amount of compression can cause liquefaction/condensation. As
temperature increases, the density of the liquid becomes smaller
whilst more molecules are added to the gas, hence increasing its
density. When the density of the liquid and gas is the same, it
becomes a supercritical liquid. This looks like a cross between a
liquid and gas, having the higher levels of diffraction of a liquid
(although not entirely) but no line distinguishing gas from liquid.
Example: supercritical CO2 diffuses rapidly (even through solids) and
transports material quickly due to its low viscosity (plus disposal is
easy), and hence is used for extracting essential oils from plants,
pesticides from meat for analysis, decaffeinating coffee beans and
fat-reduced potato chips.
Properties of Solutions
A solution is a homogenous mixture of two or more pure
substances. In a liquid, the solute dissolves in the solvent (if soluble)
until no more solute can be dissolved, at which point it becomes a
saturated solution. Solutions could also be gas-gas or solid-solid
(alloy) combinations (however well only study liquids). There are
two main thermodynamic factors that determine solubility:
Bond energies
Entropy (disorder/chaos)

Page 21

Oliver Bogdanovski

The process of a solid dissolving in a solvent is scarcely


favoured by entropy. Bond energies are therefore the determining
factor. It consumes energy to separate the solute-solute affinities by
overcoming them (so the solids can slot in), however it also releases
energy as once solute-solvent bonds are formed, their energy
overall is lowered and hence excess energy is produced.
In aqueous solutions, hydrogen bonding is strong, so there is a
substantial energy cost when hydrating a solute, and the
substance will only dissolve if that energy is repaid. However water
also already contains open holes in its network of loose hydrogen
bonds at room temperature, and hence requires very little energy.
Solutes can dissolve if they are ionic (strong ion-dipole bonds) or
ionise in water, form hydrogen bonds (having either H-bond donor or
acceptors, preferably both), or a polar to make strong dipole-dipole
bonds (remember: like dissolves with like).
Non-polar solvents dissolves solute predominantly through
dispersion forces.
The properties of gases can be modelled under the ideal has
law, however in solutions we cannot made this assumption as
density is higher and intermolecular forces are highly influential,
making the influence of temperature and pressure hard to predict.
Instead solution composition and the properties of the parent
species are the best guides for prediction. An ideal solution is one
where all intermolecular forces (solute-solute, solute-solvent,
solvent-solvent) are equivalent, and we can calculate the vapour
pressure of a mixture by assuming the solution is ideal. Raoults Law
states that if the solution is sufficiently dilute, the vapour pressure
of the pure solution multiplied by its mole fraction
is equivalent to the vapour pressure of the solvent
(and hence total solution):
Psolution = (Xsolvent)(Psolvent)
Adding a non-volatile solute to a volatile
solvent increases the boiling point of the solvent
(as boiling point is the temperature at which the
vapour pressure of the substance is equal to the
atmospheric pressure). Using Xsolvent=1-Xsolute, we
can say Psolution=Psolvent-(Xsolute)(Psolvent) or:
P = (Xsolute)(Psolvent)
When two or more components are volatile, the vapour
contains molecules of each species. As the total vapour pressure is
the sum of the partial pressures of each species (which can be
expressed diagrammatically):
Ptotal = XAPA + BBPB

Page 22

Oliver Bogdanovski

If the A:B intermolecular forces are weaker than A:A and B:B,
then the vapour pressure will be higher than expected (as less is
being stored as the solution), resulting in a positive deviation.
Conversely, if they are stronger, more will be stored in solution and
there will be a lower vapour pressure (negative deviation).

The composition of a vapour above a mixture of solvents is


usually different from the solution (as it is richer in more volatile
species), and hence we can use distillation to separate solutions. It
is not practical to do repeated distillations to achieve a complete
separation, so a fractionation column is used, where the high
surface area means there are many mini evaporation-condensation
cycles. This is how crude oil is separated into petroleum, gasoline
and other constituents.
Equilibrium
There are many natural instances of equilibrium,
like the 2NO2
(brown) N2O4 (colourless) in the
atmosphere (dimerisation, the reverse of dissociation),
the binding of O2 to iron within haemoglobin, and the
dissolving of CO2 into oceans. We can plot concentration
as a function of time to observe equilibrium, and they
do not depend on which side of the reaction you begin
with. Once a dynamic equilibrium has been reached, the
concentration of the products and reactants is constant,
however this does not mean the reaction has stopped, but the rate
of the forward reaction is equal to the rate of the reverse reaction.
The equilibrium constant (K) is used to define the ratios of
concentrations at which equilibrium will exist, which in the general
equation aA + bB cC + dD is given by:
is like , except means product of rather than sum of. If you
multiply the stoichiometry by n, you change K by the power of n (so
multiplying by a half means K is square rooted). If you reverse the
chemical equation, it is like multiplying by -1, so you take the
inverse (flip) K. Hence the equilibrium constant must be associated
with a specific stoichiometric chemical equation. K is dimensionless
and has no units. When adding two sequential reactions (like
equilibriums that lead into each other or redox reactions) you

Page 23

Oliver Bogdanovski

MULTIPLY the K values together. A small K means the equilibrium


favours the reactants, whilst a large K means the equilibrium
favours the products. The above formula can also be done in terms
of pressure for a has (as pressure is proportional to concentration as
PV=nRT and we can divide by V to get P=cRT), giving us K p (note
this number will be different to Kc but works similarly). To convert
between them, sub in the alternate value (from P=cRT) into the
equilibrium constant expression (the bit K equals) above, then
simplify, which can be used to show (although you can just do this
first principles):
Kp = Kc (RT)ng
where
ng = total mol(products) - total mol(reactants)
Note that the units of R change with the units used.
In heterogeneous equilibria (where the
reactants are not all in the same phase), if pure
solids or liquids are present they are not included in
the equilibrium constant expressions (as they have
a concentration of 1). Therefore, as the equilibrium
is independent of the amount of solids present,
varying amounts will render the same pressure and
concentration.
The reaction quotient (Q) is calculated in
exactly the same as K, except it uses the current concentrations
rather than the concentrations at equilibrium.
To calculate concentrations from the equilibrium constant use
ICE by writing out the chemical reaction and then underneath write
the:
1) Initial concentrations
2) Change in concentrations (by working out which way
the reaction will go, then adding or subtracting a
multiple of x that reflects the stochiometry of that
substance in the equation)
3) Equilibrium = Initial + Change (gives concentrations at
equilibrium)
4) Using these values as your concentrations, sub into
the equilibrium constant expression, let it equal your K
value, then solve for x (if you have a quadratic, find
which value is impossible - a negative value means the
reaction goes the other way and may be impossible, or
values bigger than the amount of what you are
subtracting from are impossible too)
5) Substitute in x to find final concentrations
To avoid having to solve complex calculations, we can use
small x approximations; that is, letting x=0 when subtracting from
larger numbers (rule of thumb is when x<5% of this larger number).
If Kc is small (e.g. 10-8), the amount of product produced (x) is small
compared to the amount of reactants left, so we can approximate it
to 0 on the denominator. Likewise, if K c is large, then it is likely it will
be small on the top.

Page 24

Oliver Bogdanovski

Le Chateliers Principle states:


If a change is imposed on a system at equilibrium, the position of
the equilibrium will shift in a direction that tends to reduce that
change.
We can induce this in three different ways:
Concentration - changing the concentration of the
reactants or products changes the value of Q; if Q<K,
more products need to be formed and the reaction will
shift to the right, whilst if Q>K, it will shift to the right
(this can also be used to predict which way a reaction
will go)
Pressure - note three effects (works similarly to
concentration)
o adding/removing gas or product (same effect as
doing so in terms of concentration as P c)
o changing the volume of the container (note in
diagram NO2 diminishes twice as fast as N2O4)
o adding an inert gas (NO
EFFECT
as
whilst
it
changes total pressure of
the system, it has no
effect
on
the
partial
pressures of each species;
doesnt increase number
of collisions)
Temperature - determine if reaction is endothermic or
exothermic (H<0 exothermic), then treat heat as
part of the reaction by adding it to the appropriate side
(endothermic is release, so add to products); disturbing
equilibrium by changing temperature is fundamentally
different as it changes the equilibrium constant, related
by the vant Hoff equation which can be integrated (this
can be used to calculate a new K; note in an exothermic
reaction, increasing temperature will make a smaller K):
Adding a catalyst will greatly accelerate a reaction without
being consumed, however they do not appear
in the final overall reaction and therefore
cannot affect the equilibrium position of the
reaction, only the rate of the reaction. They do
this by providing an alternate reaction
path/mechanism with a lower activation
energy. This is important for all modern
chemical products, like the production of
ammonia for fertiliser (Pt catalyst), removing
NO in vehicle exhaust (Pd oxide), and cracking petroleum (zeolite),
and they are important in biology.

Page 25

Oliver Bogdanovski

Acids and Bases


Arrheniuss definition of an acid was that it contains an H
that ionises in water to give an H + ion, whilst a base contains an OH
that ionises in water to give an OH - ion, however this didnt explain
NH3 being basic. Brnsted-Lowrys definition is that an acid is a
proton donor, and thus a base is a proton acceptor, and thus the
products are also conjugate acids and bases themselves (as the

species are in equilibrium).


Water is unique as it is amphiprotic/amphoteric, acting as both
acid and base. At 25oC, the equilibrium constant (also called the
autoprotolysis constant for water due to its self-ionisation) for
pure water is:
Kw = [H3O+][OH-] = 10-14
+
H as a bare proton doesnt really exist, but it is solvated by
surrounding water (all the exact nature of it is still the subject of
research) and forms hydrogen bonds with the oxygen from other
water/hydronium molecules. Equilibrium constant values vary from
10-20 to 1030, and due to its large breadth,
we use the log scale (and apply a negative
to make the small values we commonly
deal with positive). Hence:
pH = -log[H3O+]
(or [H3O+] = 10-pH)
pOH = -log[OH-]
pKa = -logKa
From this we can show that because [H3O+][OH-] = 10-14:
pH
+ pOH = 14
Strong acids and bases dissociate completely in water (use
arrow, not equilibrium), and hence their conjugate species do not
react to any measureable extent with water. Weak acids and bases
react with water but dissociate incompletely as an equilibrium is
established, and hence have the equilibrium constants:
(acidity constant)

Page 26

Oliver Bogdanovski

(basicity constant)
Acids are stronger (easier to remove protons) with a higher K a
but lower pKa, and similarly bases are stronger with a higher K b and
lower pKb. For an acid and its conjugate base (or this could be
written vice versa; base with conjugate acid - same thing):
HA + H2O A- + H3O+ Ka
A- + H2O HA + OHKb
+
2H2O H3O + OH Ka Kb = Kw = 10-14
Therefore:
pKa + pKb = pKw =
14
Many acids (particularly organic ones) have more than one
proton that can react with water, and hence are polyprotic. As the
protons may have different bond strengths, they have different
values for pKa. Those attached to more electronegative atoms (like
oxygen) will have a lower pKa and hence dissociate more (as the Ocan then H-bond with H3O+/H2O, filling in this space and limiting the
ability of an H+ to come back and bond, and thus making it a strong
acid). Even when there are identical protons, the pK a for removing
the first is different to the second (as the molecule is now slightly
negative and has a tighter control on its H +). These are written as
pKa (1), pKa (2), and so on. Whilst in the form HA - (losing a proton in
a diprotic acid), they become amphiprotic (becoming either H 2A or
A2-), and you can determine whether they will go acidic or basic by
looking at the pKa (2) and pKb (14 - pKa (1)), converting to Ka and Kb,
and then the larger equilibrium constant will win out (as it produces
more of those products), determining if it is an acid or base.
Zwitterions are molecules that contain both positive and
negative charges (that do not combine), making it overall neutral
(e.g. the amino acid glycine). This means it has both an acidic and a
basic end.
When looking at the pH of salts, consider the anion and cation
differently. If they form part of a strong acid or base, the anion or
cation (respectively), they will be extremely weak conjugates, and
hence not react with water nor alter pH. Weak acids and bases have
moderate-strong conjugates, which can raise or lower the pH.
However if both ions are weak, then we need to consider their Ks
(NOT pK), and the one with the greater equilibrium constant will
produce more product and hence determine if it is acidic or basic
(that is, if Ka > Kb, it is acidic).
Note that when solving equilibriums with acids and bases, we
can only make a small x approximation for [HA] when the initial [HA]
is 400Ka. Also, if the concentration of the acid or base is less than
10-5, then the autoprotolysis of water must be considered, and
rather than using the initial [H3O+] or [OH-] of water as 0 (as we
have been approximating), we must now use 10-7.
Consider CH3COOH (aq) + H2O (l) H3O+ (aq) + CH3COO- (aq).
If we were the add CH3COONa, be Le Chateliers principle, the
equilibrium will shift left to minimise the change, and hence will

Page 27

Oliver Bogdanovski

move away from disassociation and always tend towards neutral.


This is shift in equilibrium is called the common ion effect.
Rearranging our equilibrium constant expression for K a in
terms of H3O+ and taking the -log of both sides we get the
Henderson-Hasselbalch equation which tells use directly the pH
(useful for common ion effect; uses initial conc.):
pH = pKa + log = pKa + log
The common ion effect also provides the basis of buffers.
Consider:
HA + H2O A- + H3O+
A- + H2O HA + OHAdding a small amount of base (OH -) or acid (H3O+) will result in the
equilibrium shifting in the other direction (however H2O
concentration increasing is irrelevant as it is a liquid and hence its
concentration is always 1), and thus minimising the disturbance of
the acid or base. The pH of a buffer can be calculated using the
Henderson-Hasselbalch equation (or ICE, but takes longer). There
are many important buffers in nature, for example the blood buffer
between carbonic acid and hydrogen carbonate (pKa=6.4), hydrogen
phosphate and dihydrogen phosphate (pKa=7.2), and amino acids
(which act as an acid, pKa7 as various). Blood outside the range
7.2-7.6 is symptomatic, and fatal outside of 6.9-7.9. To get within
these ranges, [base] = 10[acid] (so pH is correct for HendersonHasselbalch equation).
Any mixture of weak acid and conjugate base (or vice versa)
will produce a buffer, however the capacity of the solution to resist a
pH change depends on the relative concentrations. The buffer is
most stable the weak acid and conjugate base are added in equal
concentrations as deviations on either side (adding either one) will
be equal. Thus from the Henderson-Hasselbalch equation, pH=pKa.
Titrations are used to determine the concentration of an
unknown solution by reacting a known volume with a standard
solution (known concentration) of a reactant. Usually the standard
solution is in the burette and titrates a known volume that has been
pipetted into a conical flask. The equivalence point is the point at
which you have the same number of moles of added OH - or H3O+
(depending on what you are
titrating) as the amount of acid or
base (respectively) that you had
initially. In strong acid-strong
base/strong
base-strong
acid
titrations it is 7, in weak acidstrong base titrations it is >7,
whilst in weak base-strong acid
titrations it is <7. Halfway to the
equivalence point, the current
[acid]=[base] (as half of the acid
has
been
reacted
into
its
conjugate base), meaning pH=pKa and hence it is a buffer.

Page 28

Oliver Bogdanovski

Acid-base indicators are weak acids or bases that change


colour with pH, generally occurring at pH = pKa (indicator) 1 (e.g.
phenolphthalein has a pKa of 9.4). They react with water (or the
solution) to gain or lose protons and hence change colour. The pK In
or endpoint should be close to the equivalence point in a titration,
and as little indicator used as possible is better to avoid changing
the pH.
A Lewis acid is an electron-pair
acceptor (inverse of proton donor), whilst a
Lewis base is an electron-pair donor (inverse
of proton acceptor), however the strengths of
these acids and bases arent as readily
quantified as BrnstedLowry counterparts.
Thermochemistry
Thermochemistry is the study of energy change in a chemical
reaction, which helps us understand the direction of chemical
change (including how to approach equilibrium) and how much
energy it takes to drive a chemical reaction. Note 1 calorie = 4.184J
(the energy required to heat 1g of H 2O by 1oC), whilst 1
Calorie=1000 calories. Heat is the amount of energy transferred
between two objects, whilst absolute/thermodynamic temperature is
what can actually be measured (in K). For example, heating two
differently sized bodies of water with the same amount of energy
will result in a higher temperature in the smaller one. Similarly,
different materials will yield different temperatures for the same
energy.
System - the reaction (or thing) we are interested in (open
systems can pass energy/mass across boundaries into surroundings,
whilst closed systems can pass energy but not mass, whilst isolated
systems cannot pass either mass or energy)
Surroundings - everything else (we only worry about this if
affected by the system, like in thermal contact)
Universe - system + surroundings
Boundary - the place where the energy (e.g. heat) flows
across
There are four thermodynamic state functions (only depend
on current state, not affected by how it cam to be or will be, hence
the change in each is only dependent on final-initial, not the
pathway taken; as opposed to path functions that
depend on the pathway taken, like heat and work, as
some method may require more of one or the other):
Internal Energy (U) - sum of nuclear,
electronic,
vibrational,
rotational,
translational and interaction energies of
all the individual particles in a sample of matter;
absolute U cannot be measured, only U (in the cases
shown due to bond energies)

Page 29

Oliver Bogdanovski

Enthalpy (H) - related to the heat


absorbed or produced in a chemical
system,
determined
by
measured
temperature change under constant
pressure
Entropy (S) - measure of the number of ways energy is
distributed throughout a chemical system (related to
enthalpy)
Gibbs free energy (G) - relates enthalpy and entropy

The First Law of Thermodynamics states:


U = q + w
where
q = heat absorbed by the system (J)
w = work done by the system (J)
Work is the energy required to move something against a
force. This energy can be electrical, light, spring energy or piston
energy (the main focus of thermochemistry). In pistons, if a
compressed gas is placed under a piston, it will expand to
atmospheric pressure and move the piston up. Hence work is
dependent upon the change in volume of the pistons and the
opposing pressure, given by: w = -pV (dont think we need to use
this).
This comes from the conservation of energy, as energy cannot
be lost or gained, simply transferred, and hence is used in the form
of heat or movement.
Heat can be determined from temperature by:
q = cT
where
c = heat capacity (dependent upon both the type and
amount of substance present). For pure substances:
q = mcT
q = nCT
where c = specific heat capacity (J/K/g)
C = molar heat capacity (J/K/mol)
Rearranging U = q - pV gives us the enthalpy or heat of
reaction:
H (=q) = U (or E) + PV
Many chemical reactions occur under constant pressure (not
constant volume) like laboratory experiments in open containers,
biological reactions in living systems, atmospheric reactions and
combustion reactions that arent in closed systems, so measuring
the heat change would give H, not U.
Calorimetry can a bomb calorimeter to
measure U (usually for combustion reactions) by
having a thermally insulated constant volume from
the rest of the universe and a known
heat capacity (and hence we can predict
the
effect
of
the
surroundings).
Alternatively, to measure H by having
constant pressure, we can use a coffeePage 30

Oliver Bogdanovski

cup calorimeter that is also thermally insulted, and usually used for
liquids in heat of dissolution, heat capacity of solids and aqueous
reactions.
We can have enthalpy of vaporisation
(Hvap>0 as you have to overcome
intermolecular forces) is the energy
required to go from liquid to gas (in J/mol),
enthalpy of combustion (Hc<0 as you are
releasing energy by forming new bonds) is
burning in oxygen, and enthalpy of
atomisation (Hatom>0 as you have to put
in energy to break the bonds) is splitting a
substance into each individual atom (not even into H 2, but 2H).
Atomisation enthalpies can be found by summing all the individual
bond enthalpies. In any chemical reaction (assuming states remain
the same), such as combustion, we can approximate the enthalpy
by finding the atomisation enthalpy and then adding the negative
values of the products bond enthalpies (basically final-initial but
using an alternate pathway). This is Hesss Law: if you add up
chemical equations to form a new overall equation, then the overall
enthalpy is the sum of the individual enthalpies. However in this
case it is only approximate as bond energies depend on the entire
molecule, not just the two atoms involved, and the values on tables
are mere averages (for example we would predict CHClBr 2 wouldnt
decompose in the stratosphere as bond energies appear higher than
the energy of the light there, but in actuality the bond energies are
lower). So instead of using bond enthalpies (which are hard to
measure as isolated atoms in the gas are difficult to measure
experimentally), we use atoms in the state they are most commonly
found in, or their standard state, occurring at 1 bar (105 Pa), 298K
and 1M, and this is denoted by a o (e.g. Ho). Now we use
enthalpy/heat of formation (fHo). For an element in its standard
state (e.g. O2 (g), Fe (s)), fHo = 0. In this, we look at the energy
required to make the products and subtract the energy required to
make the reactants:
rH = fH (products) - fH (reactants)
We can use this to predict the properties of different
substances, for example the thermite reaction (Fe 2O3 + 2Al) is highly
exothermic so we know the safety precautions to use. The thermite
reaction is used to weld railroad tracks (when there is no electricity),
and a variant is used as an igniter for rocket fuel.
A spontaneous reaction means once it has started it will
continue on its own, whilst a non-spontaneous reaction must
constantly have energy applied for it to run, or otherwise it will stop.
H is not the only criterion for spontaneity (despite most exothermic
reactions (H<0) being spontaneous, and the more exothermic the
more vigorous), as endothermic reactions like that between barium
hydroxide and ammonium nitrate can also be spontaneous. The
other criterion is entropy.

Page 31

Oliver Bogdanovski

If we look at H2O (s) H2O (l), H=6.02kJ/mol (hence


endothermic, add heat to left). At T>273K, it will shift to the right,
whilst below this it will shift left. This is because energy (heat) is
flowing from the surroundings to the system (or vice versa), and this
energy flow is key to understanding spontaneous change. Entropy is
the tendency for energy to spread out as far as possible (so having a
hot object next to a cold one will result in energy moving from the
hot to the cold until they are equal - a result of random probability,
like diffusion). The energy can spread out in two main ways: the
molecules themselves move further apart, or the energy is spread
across more molecules. Therefore entropy is greater in:
gas > solid
particles spread further
solution > solid + liquidenergy localised in solid spreads
gas + liquid > solution energy spreads even further in gas
C2H6 > CH4
more bonds to spread energy across
3mol > 2mol
entropy amount (spreads across
more molecules)
20K>10K
more kinetic energy=collisions, spreads
energy
Note that phase (position
entropy)
tends
to
dominate
molecular
complexity, as the molecules can
spread further apart. Knowing which
is
greater, we can deduce whether the
S of a reaction will be positive or
negative (by doing final-initial).
Standard entropies (S0) are
the entropy change from T=0K (where S=0) to T=298K (25oC).
We can find entropy by:
S =
The propensity for energy to spread out is known in the 2nd
Law of Themodynamics:
Suniverse = Ssystem + Ssurroundings
For any spontaneous process, Suniverse > 0. Using data from tables
(having fH0 in kJ/mol and S0 in J/K/mol - REMEMBER UNITS), we can
then determine the difference between reactants and products, and
hence if a reaction will be spontaneous at 298K. Using the above
value of S for surroundings, and using q=H, it can be shown that
for a spontaneous process:
G = Hsystem - TSsystem < 0
This is called Gibbs Free Energy, a measure
of the spontaneity of a process and the useful
energy available from it. For ALL chemical
reactions, a graph of G against the mole
fraction of a reactant/product will have a
minimum, and that minimum occurs at
equilibrium. As G<0 for any spontaneous
reaction, this makes sense as G will only
decrease. If the system does not form an

Page 32

Oliver Bogdanovski

equilibrium, it will just be a straight line from the reactant to product


(and depending on which direction it goes it could be diagonally up,
down, or horizontal like water going between solid and liquid at 0oC.
As we can see in this case of the equilibrium, when one side has a
higher value of G0 than the other, it raises the graph and the
equilibrium shifts towards the reactants (as more energy is required
to shift towards the raised end). If less raised, the equilibrium will be
more centred, and always below both sides. The relationship
between Go and the equilibrium constant (K) is:
Go = -RT ln(K)
where
R = universal gas constant (J/mol/K - make sure G uses
same units)
When rGo>0, K<1, and the reactants are favoured, shown in
the graphs.
Galvanic Cells
The concept of oxidation from reactions with oxygen, in which
the metal was oxidised to form an ionic compound (although
originally thought to be equal sharing of electrons, however we now
know electron density is higher on the oxygen). Oxidation is loss,
reduction is gain (of electrons). Note that the reducing agent
reduces the other species and it itself is oxidised (and vice versa). In
oxidation, the oxidation number decreases, whilst in reduction the
oxidation number decreases (oxidation numbers are also used for
covalent substances but are not ionic charge, only used for
convenience). Oxidation numbers are used for naming compounds,
deducing properties and identifying redox reactions. Free elements
are neutral, and therefore have oxidation numbers of 0, as do
neutral molecules, but the parts within those molecules and ions are
equivalent to the ionic charge (or charge it would be if ionic),
including the sign. Then follow the rules in this order:
1. F is always -1
2. Group 1 is always +1, Group 2 always +2
3. O is usually -2 (except in peroxides where it is -1)
4. Halogens are usually -1
5. H is -1 with metals and +1 with non-metals
To balance redox reactions we:
1. Determine half-reactions
2. Balance atoms and charges
a. Balance atoms other than O and H
b. Balance O by adding H2O
c. Balance H by adding H+
d. Balance charge by adding e3. Multiply each half-reaction by an integer for the same
number of e4. Add half-reactions (including states), cancel excess,
check balanced
a. If basic, add OH- to cancel H+ (cancel any
additional H2O)

Page 33

Oliver Bogdanovski

The activity series shows the order of


most likely to be oxidised (or more correctly,
the strongest reducing agents). The higher
species on the reactivity series determines
the direction of the spontaneous reaction. In
redox reactions, electrons are transferred
from one element to another, and we can
harness these electrons by separating the
half-reactions in a galvanic cell. As
electrons
are
gained
(reduction)
through the circuitry in the cathode,
cations in the solution bump into these
electrons and become part of the solid.
The opposite occurs at the anode. The
shorthand nomenclature for standard
cell notation is (note it shows the
direction of reactions/flow of electrons;
the middle salt bridge ):
Using a table of cell potentials (E 0 (V or J/C - the number of
joules transported by 1 amp in 1 second)), we can add the cell
potentials together to find the overall potential of the reaction by
determining their direction, balancing electrons and summing. A
spontaneous reaction always has E0>0. The tables
provided are for reduction potentials (not oxidation),
so everything will be backwards and upside down
from HSC. Note that unlike the equilibrium constant,
this does not depend on the stoichiometry, and
hence multiplying by a specific number does not
change the cell potential.
When the electrodes themselves are part of
the chemical reaction, they are called active
electrodes. We can also use inert materials like
graphite (often used for halogen gases) or platinum,
called inactive electrodes. They conduct electrons,
but do not partake in the reaction. Generally their
reactions involve a gas or another ion. Electrodes are
always placed on outside of the shorthand notation,
regardless if active or inactive.
Initially a standard hydrogen electrode was used to make
measurements off, however this is difficult to replicate accurately
(as concentration may vary with the gas), so it was initially replace
with a normal and then saturated calomel electrode, however this
contained mercury, so now we use a silver/silver chloride standard
electrode, where E0=0.22V. Using a number line, and knowing which
reaction is more positive, we can deduce the other electrodes
standard potential.
As concentration affects cell potential, we use a standard
concentration in E0, being 1M. Varying the concentration will involve
Le Chateliers principle, and if it shifts to the right the cell potential

Page 34

Oliver Bogdanovski

will increase, whilst shifting to the left it will decrease. This is


quantified in the Nernst Equation:
Ecell = E0 - ln(Q)
where
Ecell = the maximum potential a cell can generate (V)
E0 = cell potential if c=1M
R = 8.314 J/K/mol (universal gas constant - note units)
T = temperature (K)
n = number of electrons transferred per mole of reagent
F = Faraday constant = 96485C/mol (on data sheet) = charge
of 1 mol(e-)
Q = reaction quotient (current ratio of [products]/[reactants]
Alternatively, at standard conditions (T=25oC):
Ecell = E0 - log10(Q)
If Q>1 (more product than reactants), E cell<E0, and the
converse is true. The values are proportional to a log scale, so
plotting Q=10x along the bottom will produce a decreasing line.
As the reaction approaches equilibrium, the cell potential is
constantly lowering (as less reaction is happening), until it stops,
and Ecell=0. Therefore (this makes sense as a larger K favours
products, so a larger E0 is produced):
E0 = ln(Kc)
Using the relationship G = -RT lnK:
G = -nFE0
Using the relationship between E0 and concentration from the
Nernst equation, we can build concentration cells, where the halfreaction in each cell is identical but they consist of different
concentrations. We can also use this to determine the concentration
of a cell by measuring it against a cell of known concentration
(usually 1M) and measuring the cell potential. This is how pH meters
work (comparing 2H+ (aq, 1M) + 2e- H2 (g) or more commonly a
silver/silver chloride reference to an unknown solution and
measuring current can be used to calculate concentration and
hence pH). A similar thing can be done with other ions for ion
selective electrodes (to measure the concentration of particular
ions). Concentration cells are also used in nerve signalling, ion
pumps across cell membranes, and energy production and storage
in cells.
Forcing the electrons in the
opposite direction (that is, against the
spontaneous
reaction)
is
called
electrolysis, forming an electrolytic
cell. As oxidation always occurs at the
anode, the anode is now the other
electrode which is becoming positive
(as it is now losing electrons). In a
galvanic cell, the anode was already
negative, and electrons flowed from it
to the cathode.

Page 35

Oliver Bogdanovski

Faradays First Law of Electrolysis states that the mass of


a substance (m; in grams) liberated at an electrode during
electrolysis is proportional to the quantity of charge (q; in Coulombs)
passing through the electrolyte: m q
Faradays Second Law of Electrolysis states the number
of Coulombs needed to liberate one mole of different products
occurs in whole number ratios: q = It (I in amps, t in seconds). We
can divide this q (C) by Faradays constant (C/mol) to find the
number of moles of electrons, which can be used to deduce the
number of moles of a substance being reduced or oxidised.
There are three main classes of batteries:
Primary Batteries - non-reversible (cant charge)
o Dry Cell - carbon cathode coated in MnO2 (with
graphite powder for conductivity), a paste of NH4Cl
and ZnCl2 as an electrolyte, and Zn as active
anode; ammonium gradually turns to ammonia
which evaporates away
o Alkaline Battery - uses alkaline environment
(KOH electrolyte) to prevent build up of gas and
preserves the zinc electrode
Secondary Batteries - reversible and rechargeable
o Lead-Acid - alternating Pb and PbO2 plates that
provide large surface area to H 2SO4 electrolyte
(delivers large currents)
o Nickel-Cadmium - Cd anode, NiO(OH) anode, OHelectrolyte, same cell potential as primary
batteries so can be used interchangeably
o Lithium - light with high cell potential (meaning
little is needed) with Li anode and MNO2 cathode
Fuel Cells - fuels/chemicals can constantly be passed
into the battery
o reactants are burnt (overall like combustion),
however the two half reactions (anode: H2, CH4,
etc.; cathode: O2 + 4H+ + 4e- H2O), with
graphite electrodes surrounded
by platinum catalysts
The oxidation of a metal to its oxide is often
spontaneous (as they are generally exothermic).
The most economically important corrosion is that
of steel (more specifically, the iron inside) as it
results in loss of structural strength (rust holes in
cars, or structural supports like bridges).
Oxidation occurs through an active anode (the Fe) to form a pit, and
yielding e- which travels through the metal, at which point Fe
located elsewhere acts as an inactive cathode to reduce O 2 to OH-. It
consists of two sets of reactions:
Anode
2Fe 2Fe2+ + 4eE0=+0.44V

Page 36

Oliver Bogdanovski

Cathode O2 + 4H+ + 4e- H2O


E0=+1.23V
Overall
2Fe (s) + O2 (g) + 4H+ (aq) 2Fe2+ (aq) + 2H2O (l)
E0=+1.67V
Anode
2Fe2+ 2Fe3+ + 2e0
E =-0.77V
Cathode O2 + 4H+ + 4e- H2O
E0=+1.23V
Overall
2Fe2+ (aq) + O2 (g) + 2H+ (aq) 2Fe3+ (aq) +
0
H2O (l)
E =+0.46V
Iron(III) then forms a very insoluble oxide which is deposited at the
edge.
2He3+ (aq) + (3+n) H2O (l) Fe2O3 nH2O (s) + 6H+ (aq)
OVERALL 2Fe (s) + 3/2O2 (g) + nH2O (l) Fe2O3nH2O
As H+ is cancelled out of the reaction, it must be a catalyst.
Features include: rusting doesnt occur in dry air (no water=no salt
bridge), rusting does not occur in oxygen-free water like ocean
depths (no oxygen=no oxidant), iron rusts more quickly in acidic
environments (as H+ is a catalyst, takes a route with a lower energy
requirement), and iron rusts more quickly at the seaside (more
salt=more conductivity for salt bridge). To protect against rusting we
can use anodic inhibition - greasing (oiling the surface) or allowing
a thin metal oxide film to form (often done with sodium chromate to
form Cr2O3), preventing oxygen from reaching
the metal.
Electro-refining
is
the
principle
method of obtaining Cu of a high purity, as
pure Cu is less easily oxidised, and hence
less easily reduced metals remain in solution,
and noble metals (those less reactive than
Cu) are not oxidised and fall to the bottom as
mud.

Page 37

Oliver Bogdanovski

You might also like