Introductory Maths For Chemists: Chemistry Maths 1 J. E. Parker
Introductory Maths For Chemists: Chemistry Maths 1 J. E. Parker
Introductory Maths For Chemists: Chemistry Maths 1 J. E. Parker
for Chemists
Chemistry Maths 1
J. E. Parker
J.E. PARKER
INTRODUCTORY MATHS
FOR CHEMISTS
CHEMISTRY MATHS 1
CONTENTS
Acknowledgements 6
www.simcorp.com
9 References 134
ACKNOWLEDGEMENTS
I was pleased to respond to bookboon.com to write a textbook (which is split into 3
more manageable books, Introductory Maths for Chemists, Intermediate Maths for Chemists,
and Advanced Maths for Chemists which should be studied in sequence) that would help
chemistry students survive, and even enjoy, the maths required for a chemistry degree. I
developed and presented tutorials on maths to our first year chemistry students over several
years at Heriot-Watt University, Edinburgh, Scotland. These tutorials have formed the basis
for these workbooks. I would like to thank the staff of Heriot-Watt University for their
help; and thank the students who “suffered” these tutorials, I hope they helped them with
their degrees and later careers. Most of all I would like to thank my wife Jennifer for her
encouragement and help over many years.
I shall be delighted to hear from readers who have comments and suggestions to make,
please email me. So that I can respond in the most helpful manner I will need your full
name, your University, the name of your degree and which level and year of the degree
you are studying. I hope you find these textbooks helpful and I wish you good luck with
your studies.
J.E.Parker@hw.ac.uk
http://johnericparker.wordpress.com/
The author, John Parker, has 39 years experience of teaching chemists, biologists, pharmacists,
chemical and other engineers at Heriot-Watt University Edinburgh. The author’s research
interests from 1966 onwards are physical and analytical chemistry, especially chemical kinetics,
spectroscopy, statistical mechanics, quantum mechanics, computational chemistry and the
use of mass spectrometric techniques to study single-collision kinetics. The author’s teaching
was mostly in physical and analytical chemistry for years 1 to 4 of BSc and MChem and
PhD students using combinations of lectures, tutorials and lab classes. During this time
John Parker developed the content and taught maths to first year chemists for several years.
The author believes that maths must be put into a chemistry context for the students to
grasp its significance, usefulness, and its application in science and engineering.
1 WEEK 1: CHEMISTRY
AND ALGEBRA 1
1.1 INTRODUCTION
The three books; Introductory Maths for Chemists, (Parker 2018), Intermediate Maths
for Chemists (Parker 2012) and Advanced Maths for Chemists (Parker 2013) are tutorial
workbooks intended for first year undergraduates taking a degree in chemistry, chemical
engineering, chemical physics, molecular biology, biochemistry, or biology. From now on I
will be using the term “chemistry” as a short-hand to cover chemical engineering, chemical
physics, molecular biology, biochemistry, or biology as well as chemistry itself. The texts
may also be very useful for final year school or college students prior to them starting an
undergraduate degree and also to their chemistry teachers and lecturers. The text is published
as three books in order to reduce file size and make handling on a laptop or tablet computer
much easier. Introductory Maths for Chemists roughly covers the first 8 weeks of semester 1;
Intermediate Maths for Chemists the remainder of semester 1 and the beginning of semester
2; and Advanced Maths for Chemists the rest of semester 2. They each have chapters named
as Week 1, Week 2, etc. This is purely to help you self-pace your work on a weekly basis.
Go through the questions and work out the solution yourself on paper then check your
solution. Full solutions are given to show you the method of solving the problem. Initially the
solutions give most of the steps but as you progress through the workbook the explanations
become less detailed. When you do finally cover the chemistry involved in the examples
during your chemistry degree you won’t be blinded or scared by the maths as by then you
will be happy playing around with equations and graphs.
One final and important comment, a common mistake of many students is thinking you
need to memorize all the equations you come across in any area of the subject. This is
impossible and I know that I, or any other member of staff, can’t remember them all. There
are a very small number of equations that become familiar simply by usage and which you
remember without really trying. All the rest comes from being able to apply your maths to
this small number of familiar equations or to the equations supplied in an exam or from
a textbook and this enables you to get to your target.
Maths is a convenient and fast shorthand language that summarizes the details of a particular
topic of chemistry. It is the language of chemistry, it is also the underlying language of
all the sciences, engineering, economics, and many other subjects. So we won’t be able to
become really fluent in chemistry, other science subjects, or engineering until we understand
the “shorthand” language of maths.
At the beginning of your university chemistry degree you may find that many of the
chemistry examples used in this workbook have not yet been covered in your chemistry
course. Don’t worry we are trying to understand the maths at the moment more than the
chemistry, and the chemistry details will come later as you progress in your degree. Just
treat these examples as maths examples, which is what they are, and solve the maths. The
chemistry will add meaning to the maths, which otherwise can be a bit abstract.
We start with reviewing chemistry and maths, this is spread over Weeks 1 and 2 and covers
the underlying skills that all chemists should be good at but which needs some practice.
Some of this material you might have covered before but stick in there as it is vital to the
rest of your degree course.
In your maths lessons in school or college the variables used were probably x and y and
angles θ or α as these are the general symbols used by mathematicians. But in science and
engineering all the variables we use are physical quantities, such as mass, length, time, force,
work, enthalpy, entropy, and so on. These physical quantities usually have a conventional
symbol agreed by usage of the international community of scientists, for example, IUPAC
the International Union of Pure and Applied Chemistry and IUPAP for physics. These
symbols are used in the maths equations describing the phenomenon of interest. A few
examples of the symbols used for physical quantities are m for mass, c for the velocity of
light and E for energy. These come together in the equation that everyone has met, E =
mc2. In maths this is equivalent to y = ax2 which could apply to many situations, however,
E = mc2 only applies to the specific process of converting mass into energy. So this book
will get you accustomed to using maths in the real world of manipulating equations made
up of physical quantities written in the accepted scientific way.
#
- #
0C*#6
Definition: a physical quantity
//%/.
+
A physical quantity may be a variable or a constant they consist of two parts; a pure number
(which in scientific notation may includes a “multiplier” of ten raised to a power) and units.
The pure number and the units are inseparable. An example would be c = 2.9979×108 m s–1
where c is the symbol for the velocity of light. In order to clearly distinguish the three parts
of the above equation, the symbol for physical quantity is written in italics (sloping) font
and the number and the symbols for the units are in roman (upright) font. Don’t worry
too much about hand written material but for typed material it should be typed correctly.
Some units are named after people such as Sir Isaac Newton or Lord Kelvin. When we
refer to the person then their name has an initial capital letter, Newton or Kelvin, but
the unit name is all lower case, newton or kelvin, except for the unit of “degrees Celsius”
for historical reasons. The unit symbol is an initial capital of one or sometimes two letters
(N or K). The use of initial capital for units has a few exception for historical reasons e.g.
the second (s), the kilogram (kg), metre (m) and centimetre (cm). When we substitute a
physical quantity for its symbol in an equation we must substitute both the number and
the units into the equation, they are inseparable, they go together like the two sides of a
piece of paper.
Straight line graphs are important in science. Linear graphs are particularly useful as being
strong evidence that the maths equation really does model the chemistry. Linear graphs are
much easier to use in this way than curved graphs. They also allow you to spot any “rogue”
or outlying experimental data points which can be checked back in the lab. The equation
for a linear graph is below.
Where x and y are the variables, m is the gradient of the line (how quickly y increases, or
decreases, as x is increased), and c is the intercept on the y-axis (the “offset” that y starts
with when x = 0).
!
The vertical axis in Fig. 1.1 is the y-axis and the horizontal axis is the x-axis. Whatever
names are used for the variables in a particular situation, we should always say “plotting y
against x” or “plotting y versus x” where the variable names are used instead of y and x so
the axes will always have an unambiguous meaning.
Graph axes, table headings, and the table data are the physical quantity divided
by the units and any numerical multiplier, they are pure small numbers.
�e Graduate Programme
I joined MITAS because for Engineers and Geoscientists
I wanted real responsibili� www.discovermitas.com
Maersk.com/Mitas
I joined MITAS because fo
I wanted real responsibili�
Month 16
I was a construction
supervisor ina co
I was
the North Sea su
advising and the
Real work he
helping foremen ad
International
al opportunities
Internationa
�ree wo
work
or placements ssolve problems
Real work he
helpin
International
Internationaal opportunities
�ree wo
work
or placements ssolv
Download free eBooks at bookboon.com
Click on the ad to read more
11
Click on the ad to read more
INTRODUCTORY MATHS FOR CHEMISTS:
CHEMISTRY MATHS 1 Week 1: Chemistry and Algebra 1
Where possible is it best to use a spreadsheet to plot the graph and find the equation of the
line, both linear and non-linear least squares curve fitting are discussed in Advanced Maths
for Chemists (Parker 2013). But you also need to be able to carry out the process by hand
on graph paper, perhaps in an exam where computers, phones, and graphic calculators are
normally not allowed except for some “open-book” type of exams, so practice with some of
these tutorial questions! Once you have used your judgement to draw the best straight line
e.g. using a transparent ruler then the equation of the best straight line is found as follows.
Using the line itself not the data points (otherwise what was the point of drawing a best
straight line) pick two positions on the line near the left hand side and near the right hand
side of Fig. 1.1 (x1, y1) = (2, 5) and (x2, y2) = (8, 14).
The two points on the line give two simultaneous equation are (1) and (2) for y = mx + c.
Subtraction of (1) from (2) gives the gradient (3) m = 1.5. Addition of (1) and (2) gives
(4) into which we substitute m = 1.5 and find the intercept (5) c = 2. The equation of the
best straight line (6) is y = 1.5x + 2.
Quite often the data is such that the graph has a very large intercept, not like Fig. 1.1,
then it is necessary to plot the graph with the intersection of the axes not at the origin (0,
0) but at a more convenient point. Also the scale of the axes must be adjusted so the data
points occupy a reasonably large area (about half an A4 page in portrait mode is a useful
size for a lab notebook). Then the only way to obtain gradients and intercepts is by finding
the equation of the line as in the above method unless you are using a spreadsheet.
Graphs should be scaled so the data points fill most of the available space,the axes
may not intersect at (0, 0) and the origin might be a long distance away from your data
points. Do not extrapolate but find the equation of the best-fit straight line you draw.
In chemistry often the equation that models the chemical behaviour is not a straight line
function. Then maths is necessary to rearrange the equation to obtain a linear function. A
linear plot as supporting evidence that the model (a mathematical representation or theory
of a phenomenon) agrees with the experimental data (the facts). This is the first test of a
scientific theory, the second test is that the it predicts further experiments to test the theory
or model.
!$
!%
Fig. 1.2 is a typical plot of data from a rearranged Arrhenius equation, which we will
discuss later in the book. The rearranged Arrhenius equation is a plot of the natural log
of the rate constant against the reciprocal of temperature. Notice how the axes are labelled
as small pure numbers. The axes scales fit all the data and the intersection of the axes is a
long way from the origin (0, 0).
Perhaps a little confusing at first sight, is the fact that the x-axis in Fig. 1.2 is 1/T and it is
multiplied and not divided by a numerical factor and units. As some people have problems
with this type of reciprocal labelling I have shown the steps from (7) to (12) below, using
T = 300 K as an example for our x-axis labelling.
www.skoda-career.com
7%8 % !.. L 7+8 7/8 ...!!!! L
% !.. L %
%
7.8 !!!!. ! L 78 !
!!!! 78 % .! L !!!!
% . L
The axis are clearer plotted as small numbers hence the division by the multiplier (11).
The x-axis is better for later processing of results when written as (1/T)(103 K) = 3.333 as
in (12) rather than (11).
7!8
# & //+ !
. L % +% 7&8 //+. ! L
"
7,8 //+. ! L 7
8 " //+.! L+!& ?L #
'
!
7%8 " +/%. ?# 7+8 " +/% ?#
The best-fit straight line equation of Fig. 1.2 is (13). In Section 3.1.1 on logs they are of
pure numbers so in ln(k/s−1) the brackets cannot be separated out. Although the x-axis is
written as (1/T)(103 K) so that we can plot a small pure number however the gradient m
is a physical quantity with a multiplier and units (14). For the Arrhenius equation we will
discover later on that the gradient equals −Ea /R where R is the gas constant and Ea is the
activation energy (15). So cancelling out the negatives and taking the gas constant over to
the right (16) we obtain the activation energy Ea (17) and (18) in the conventional units
of kJ mol−1.
From the equation of the best-fit straight line for Fig. 1.2, the intercept c = 28.7 in (19).
From the Arrhenius equation the intercept is c = ln(A/s−1). Don’t worry if at the moment
you are unsure of logs and antilogs as we will cover all of that area in Week 3 of this book.
In (20) we take the antilog of the intercept. Only then may we take the units over to the
right hand side to give the pre-exponential factor (21).
Introductory Maths for Chemists may be used with any maths textbook, however, the
students and I found the textbook (Stroud and Booth 2013) very useful. Despite its name
of Engineering Mathematics Stroud and Booth’s book covers all the maths needed by all the
sciences and engineering subjects. Introductory Maths for Chemists gives chemical examples
of the maths concepts. If you want to look up any first year chemistry then any general
chemistry textbook is useful but the textbook (Blackman, Bottle, Schmid, Mocerino and
Wille 2015) is excellent. For later on in your course the following textbooks have many
examples of the interplay between chemistry and maths mainly in physical chemistry,
Atkins, de Paula and Keeler 2017 and Atkins and de Paula 2016, and Levine 2009. For
drawing and manipulating curves I use Graph 4.4.2 (Johansen 2013) which is an easy way
of finding the areas under a curve between two limits, tangents, normals, and first and
second differentials of a curve. I also use LibreOffice Writer, Calc spreadsheet, Draw, and
Math for maths equations (LibreOffice 2018). For visualizing molecules I use Jmol (Jmol
2018). These three software packages are free.
Chemists use spectroscopic analysis all the time in the lab to determine what molecules are
present in a sample and what are their concentrations. Spectroscopy is also used as a tool
to understand the structure and reactions of molecules. Fig. 1.3, light consists of a sine
wave of electric field (E) and a sine wave of magnetic field (B) moving in phase together at
right angles to one another and also at right angles to the direction of travel at the velocity
of light (c), hence the alternative name for light of electromagnetic radiation. In order to
be able to use spectroscopy in chemistry we must understand light along with the maths
equations which summarize the properties of light.
The wavelength λ (italic Greek “lambda”) is the distance between equivalent points on
neighbouring waves, the units are sub-multiples of a metre. The correct symbol for frequency
is υ (italic Greek “nu”) it is the number of complete waves which pass a fixed point in a
second, the units are “per second” or s−1 sometimes called hertz (symbol Hz). The velocity
of light c = 2.9979×108 m s−1 is related to its wavelength and its frequency by the equation
below. Calculate the wavelength λ of the light which is absorbed by a molecule if the light
has a frequency of υ = 2.78×1014 s−1.
Wavelength and frequency of light
Chemists routinely need to know and alter the concentrations of solutions in the lab. The
concentration of a solute is c mol L–1 and is given in terms of the number of moles of the
solute, n mol, and the volume of the solution V measured in litres, symbol L. The IUPAC
unit of volume is the metre cubed m3 which is 1000 L, however, chemists often use the
careers.slb.com/recentgraduates
named sub-multiple unit litre as it is more convenient than the cubic metre and is identical
to the decimetre cubed dm3.
Definition: concentration in mol L−1
+
3 5 !
At a given temperature this reaction eventually comes to equilibrium. The double half-
arrows or “dynamic equilibrium arrows” indicate that the reaction is at equilibrium, both
the forward and the reverse reaction are still occurring simultaneously and the rates of the
forward and reverse reactions have become equal to one another. Once the reaction has
reached equilibrium the concentrations of the reactants and products are then constant and
are related to one another by the reaction’s “standard concentration equilibrium constant”
Koc which is characteristic of the reaction at a given temperature. The concentration for
example cA may also be written as the molecule’s symbol inside square brackets [A] so cA
= [A] mol L−1.
3 3 #' 5 5 #'
Molar concentration
#' #'
At school or college you may have been told that equilibrium constants Koc are dimensionless
pure numbers, here is the reason. In the equation for the standard equilibrium constant Koc
each concentration is divided by its own standard state concentration. The two alternative
symbols for the standard state is either a superscript “o” or a superscript “o” (pronounced
“standard”). For solutes, their standard state concentrations are co = 1 mol L−1. As the
concentration is divided by the standard concentration the units and the numerical value
of “1” will cancel out to give the pure number “dimensionless concentrations” e.g. [A]. So
Koc the equilibrium constant must also be dimensionless. In chemical equilibria each of the
concentrations is raised to the same power as its stoichiometric coefficient.
!
, Standard concentration equilibrium constant, a pure number
3 5
1) Calculate the value of the equilibrium constant Koc for the above equilibrium in a
solution where all four of the concentrations are equal to 0.1 mol L−1.
2) Calculate the new concentration of [C] if we alter the concentrations to [A] = 0.2
mol L−1, [B] = 0.15 mol L−1 and [D] = 0.25 mol L−1.
- - % %
The ΔrCpo shows how sensitive the change in reaction enthalpy is to a change in temperature
and ΔrCpo is called the change of standard heat capacity at constant pressure for the chemical
reaction. Over a small temperature change ΔrCpo may be treated as constant. This last point
is exemplified by Fig. 1.4 the graph of Cpo for N2 against temperature where the tangent
at 25 oC shows that the heat capacity varies and can only be treated as a constant over a
small range of temperatures.
! + &
%
%
!%
Each of the reactants and products molecules of a reaction have a Cpo which will vary
differently with T so the range of linearity for the reaction ΔrCpo would be expected to be
even smaller than that of a single molecule such as N2. With the above provision that the
linear relation between enthalpy of reaction and temperature is only valid over about a 60
o
C range at the most, rearrange the above equation to get T1 as the subject of the equation.
The reaction at high temperature between H2 and Br2 molecules in the gas phase gives
HBr gas by a complex (multi-step) chain reaction which will be covered in your chemical
kinetics lectures.
Studying the chemical kinetics of this reaction chemists found the rate of reaction (v) is given
by the following rate law. The symbol v for rate of reaction (italic “vee” as in velocity) should
not be confused with the Greek italic letter ν (“nu”), the context should remove any confusion.
& 5 !
( 5 ( 5 ( Stoichiometric equation and rate law
5 & / 5
Where k and k' are parameters which depend upon the temperature of the reaction. Note
that the stoichiometric coefficients are not related to the powers of the concentrations except
for single-step “elementary reactions”. The concentration of e.g. H2 is [H2] mol L−1 but for
clarity reaction rate equations are usually written without the units being explicitly shown.
Rearrange this equation for plotting as a straight line graph of 1/v against [HBr] by taking
reciprocals of both sides and simplifying the right hand side of the new equation to get
the final form which is ready for a linear plot. What are the gradients and intercepts of
this linear plot?
The van der Waals equation models the behaviour of many non-ideal gases. Let us write
the ideal gas law pV = nRT in the form p = nRT/V. The ideal gas law assumes that the
molecules have zero radii and do not attract or repel one another, the ideal gas law is clearly
just a first approximation.
Let us look at the molecular repulsion. If the molecules have a van der Waals radius r due
to electron-electron repulsion at short intermolecular distances, then around a molecules is
an excluded volume, a sphere of radius 2r if all the molecules are identical. The volume V
of the container of the gas molecules is reduced by nb where the empirical van der Waals
parameter b reflects the “excluded volume” of a given species of molecule and n is the
number of moles of gas. The ideal gas law is then modified to p = nRT/(V−nb).
We also know that at long range molecules attract one another by van der Waals forces
which are made up of contributions from London dispersion forces, permanent-dipole
induced-dipole forces, and permanent-dipole permanent-dipole forces. These van der Waals
attractive forces reduce the pressure of the gas in two ways, firstly by reducing the frequency
'%
+
+
+
+ '% Two versions of the van der Waals equation
An equation such as the van der Waals needs to be seen as a graph in order to be meaningful.
Carbon dioxide has parameters a = 3.64 L2 bar mol−2 and b = 0.04267 L mol−1. Using the
same units the gas constant is R = 8.3145×10−2 L bar K−1 mol−1, use a spreadsheet to plot
the pressure p from 30 to 90 bar against the molar volume Vm from 0.07 to 0.3 L mol−1 for
the following temperatures of 273.15, 298.15, 304.19, and 313.15 K. Comment upon the
shapes of these curves, you will need to use a textbook to help formulate your comments.
When a volatile liquid or solid is placed into an evacuated container then some of the liquid
evaporates or some of the solid sublimes into the gas phase. When the liquid or solid and
its vapour have come to equilibrium at a temperature T the gas phase pressure is called
the vapour pressure p.
Vapour pressure increases with increasing temperature. The enthalpy of vaporization ΔvapHo
is the amount of heat required to vaporize a mole of liquid in its standard state. For a solid
ΔvapHo is replaced by ΔsubHo the enthalpy of sublimation. The Clausius-Clapeyron equation
approximately models the change in vapour pressure with temperature for solid-gas or
liquid-gas equilibria in which the vapour is not close to its critical temperature.
#
-
* Clausius-Clapeyron equation
% '%
The two approximations are firstly that the molar volume of a vapour is much larger than
the molar volume of the liquid or solid so that the liquid or solid molar volumes can then
be ignored. Secondly, that the vapour can be treated as an ideal gas except near its critical
temperature. From the Clausius-Clapeyron equation if we assume that ΔvapHo is constant
over a small temperature range then the equation may be integrated to give the equation
below where T1 and T2 are low and high temperatures having vapour pressures p1 and p2.
Rearrange this equation to get T2 as the subject of the equation.
* - % %
#
'
% %
Jump to Solution 7 (see page 31)
KG 0 # KG# # 0
The rate of reaction of a second-order chemical reaction between two different reactants A
and B depends on the concentrations of both the molecules A and B.
3 5 *
The rate of reaction is summarized by the rate law which includes the following term
where a0 and b0 are the initial concentrations of molecules A and B, and x is the change in
concentration of both reactants. Simplify this expression using partial fractions.
. .
The shorthand notation which is on the left of the above equation is called the “binomial
bracket”, there is no dividing line and it does not mean n divided by k. The binomial
bracket is pronounced as “n objects, choose k of them”. The exclamation mark such as in
n! is pronounced as “en factorial”.
There is only one arrangement of zero objects 0! = 1 this is called the empty product by
mathematicians. The symbol “…” means “the equation continues in the same way” it is
called an “ellipsis” and is the maths equivalent to “etc” in text. The binomial theorem is
used in chemistry for: (1) calculating the intensities of NMR (nuclear magnetic resonance)
lines due to spin-spin splitting, (2) in mass spectrometry for calculating the ion abundances
(intensities) for molecules containing elements with several isotopes, (3) to model the
biochemical transition between the helix and coiled structures of proteins, and (4) in statistical
mechanics to calculate the probabilities of a given distribution of molecules over a set of
quantum states. You will meet all of these techniques later on in your chemistry degree.
The binomial theorem expands the equation (x + y)n into a polynomial of terms axbyc where
the powers b and c are positive integers and n = b + c. The coefficients “a” in the terms
axbyc are known as the binomial coefficients. We already know (x + y)1 = x + y also that (x
+ y)2 = x2 + 2xy + y2 and you should be able to work out for yourself by multiplying the
last two equations together that (x + y)3 = x³ + 3x²y + 3y²x + y³. The binomial coefficients
of the terms in these three equations are respectively (1, 1) then (1, 2, 1) and (1, 3, 3,
1) if we carried on to (x + y)n then these binomial coefficients can be arranged in Pascal’s
triangle where each entry is the sum of the two above it.
Pascal’s triangle of binomial coefficients
! !
&
&
, . . ,
The natural abundance of the stable carbon isotopes are about 13C ≈ 1% and 12C ≈ 99%.
Calculate the isotopic abundances arising from just the three carbon atoms in propane (ignoring
the H-atom isotopes) by expanding the binomial formula (1 + 99)3 into the first 4 terms.
These 4 terms correspond to the relative amounts of the following four isotopic molecules
13
C3H8, 13C212CH8, 13C12C2H8, and 12C3H8 these isotopic variants are called isotopologues.
78
78
7!8
//%/. +
7&8
7,8 7
8 .%+&.
%+.&
Rearranging the equation (1) so that the unknown quantity λ is the subject of the equation
is most clearly done by dividing both sides of the equation by the frequency ν to give (2)
Cancelling top and bottom within any single term, υ (in red) cancels to give (3) and then
rearranging (3) to make λ the subject (4). We say that λ is the subject of the equation (4)
when it is on the left of the equal sign and it is equal to the rest of the equation. To make
this physically meaningful we must substitute the physical quantities for the symbols (5)
both numbers and units and then cancel the units within any single term. In this case it is
s–1 that cancels (in red) to give (6) the wavelength with the units being determined by the
correct use of physical quantities. The wavelength is in the near infrared part of the spectrum.
The number of moles of the solute remains constant when the concentration is altered by
diluting the solution with more solvent, n is constant as V and c vary. We rearrange the
equation so that the unknown but constant quantity n becomes the subject of the equation
and then substituting in the original physical quantities.
+ . #' ..., ' ,.& #
+
It is usually less confusing to carry out calculations in the base unit, here the litre (L), rather
than in multiples or sub-multiples such as the mL. The volume of the diluted solution is
15 mL = 0.015 L. The diluted concentration may now be calculated.
! .! .
,
! !,
. .
, ..
3 5
. .
, , , ..
3 5 . .
! , 3 5
78! , 78 ,
3 5
!
7!8
, 3 5 !
3 5
!
!
78 , 78 , 3 5 7!8
3 5! , 3 5 ! ....,
7&8
,
3 5
7,8 7
8 ! %.&
!
! ...., ., !
7&8 7,8 7
8 %.&
7%8 +/
. 7+8 .,
#'
7/8 +/
. #'
7%8 +/
. 7+8 #' 7/8 +/
. #'
The equilibrium constant equation (1) must be rearranged so that [C] becomes the subject.
First we multiply both sides of equation (1) by [A][B]2 and then cancelling out top and
bottom within any single term (not shown) to give (2). Divide both sides of (2) by [D]2
and cancel top and bottom within any single term gives (3). Write [C]3 as the subject in
(4). In (5) substitute in the numerical value for Koc and those for [A], [B], and [D]. We
have in (6) the numerical value for [C]3. Calculate in (7) the cube root of [C]3 to find
[C]. Remember (8) that [C] is the symbol for the pure number of the concentration. The
concentration (9) is thus cC = 8.96×10−2 mol L−1.
Increasing the concentrations of [A], [B], and [D] has decreased the concentrations of [C] in
order to maintain the equilibrium constant Koc at its constant value for the given temperature.
-
-
78
-
-
% % 78 % %
-
-
-
-
7!8 % % 7&8 % %
(1) The lower temperature T1 is inside the bracket. Remove the bracket (2) by diving both
sides by ΔCpo. Add T1 to both sides (3). Subtract (ΔH2o−ΔH1o)/ΔCpo from both sides (4) to
make T1 the subject of the equation.
7!8
5
& / 5
7&8
&/
5
!
& 5 & 5 ! & 5 & 5 !
(1) This is the rate law for the reaction and in order to obtain a linear plot we use algebra
as follows. (2) We take reciprocals on both sides of the equation. (3) The right hand term
is split into two terms with the same denominator. The first of these two terms may then
be simplified by cancelling out [Br2] top and bottom. (4) This is our linear graph with y =
1/v and x = [HBr] as in Fig 1.10.
!7)87*8!
7)*8!& 6
Table 1.1: the first few terms of the van der Waals spreadsheet.
www.foss.dk
I am plotting more temperatures than asked for in the question as this is easy when using
a spreadsheet. Table 1.1 shows the first few entries for the spreadsheet for drawing the van
der Waals plots for CO2 Fig. 1.11. The molar volumes in column A are typed in manually
for a few of them and then selected and copied downwards by dragging the bottom-right
handle of the selected cells in column A from 0.07 to 0.3 and then release the mouse
button. The formulae for calculating the pressures are based upon the van der Waals
equation with pressure as the subject p = nRT/(V − nb) − an2/V2. Remember the headings
and all the cells of the spreadsheet are dimensionless pure numbers, e.g. the cell B2 formula
is =(0.083145*273.15)/(A2−0.04267)−(3.64/A2^2) which contains the two van der Waals
parameters for CO2 a/L2 bar mol−2 = 3.64 and b/L mol−1 = 0.04267. The formulae in row
2 in typed in manually with temperatures in kelvin not degrees Celsius, and then selected
and copied downwards by dragging the bottom-right handle to complete the spreadsheet.
2
2
2
!0
2
2
&!6&
2
At temperatures below the critical temperature Tc = 304.19 K (31.04 oC) in green in Fig.
1.11 the van der Waals equation is giving unrealistic loops upwards then downwards called
van der Waals loops. This is an artefact in the p-V plot due to the van der Waals equation
only involves two terms in Vm and Vm2. Normally any given loop is replaced by a horizontal
line so that the area of the downward loop cancels the area of the upward loop. This artefact
is removed by using the virial equation p = RT(1/Vm + B/Vm2 + C/Vm3 + …) which is at
least a cubic equation in Vm with empirically derived virial coefficients B, C, etc. which are
functions of temperature.
The van der Waals loops get smaller until they disappear as CO2 approaches the critical
temperature Tc of 31.04 oC. At the “critical point” CO2 has a critical pressure of pc =
73.825 bar, a critical molar volume of Vm,c = 0.092 L mol−1 and a critical temperature of
Tc = 304.19 K (Acree & Chickos 2017). The critical point is defined mathematically as a
point of inflection that is where the first and second derivatives of the p-V curve are zero,
below. A point of inflection is where a curve changes from a positive gradient to a negative
gradient or the other way round. The curly Greek “delta” means the partial derivative with
the variable temperature held constant. Don’t worry we will soon cover differentiation in
this book.
+ %
.
+ %
. Point of inflection in a curve of p = f(V m)
Equations (1) is the van der Waals equation in term of the critical pressure pc the critical
temperature Tc and the critical molar volume Vmc. The first partial derivative set equal to
zero and then rearranged is (2). The second partial derivatives set equal to zero and then
rearranged is (3). These three simultaneous equations (1), (2), and (3) can be solved for the
three unknowns pc Vmc Tc the algebra for solving them is as follows. Equation (2) is divided
by (3) to give (4) which rearranges to (5) with Vmc = 3b. Substitute critical molar volume
from (5) Vmc = 3b into equation (2) to give (6) which rearranges to (7) Tc = 8a/27Rb.
Substituting the critical molar volume from (5) Vmc = 3b and the critical temperature from
(7) into the van der Waals equation (1) gives (8) which rearranges to (9) for the critical
pressure pc = a/27b2.
The algebra of simultaneous equations has allowed us to find the critical values for the
pressure, molar volume, and temperature from the van der Waals constants. Conversely,
when experimental values for pc and Tc are measured for a gas we can calculate the two
van der Waals parameters “a” and “b” by solving the two simultaneous equations (7) and
(9), you can prove this for yourself.
% ' % '%
The van der Waals parameters from experimental pc and Tc
& +
Above the critical point the CO2 is neither a gas nor a liquid but it is a “supercritical fluid”.
Supercritical fluids have chemical and physical properties which are intermediate between
those of a liquid and a gas, they are denser materials than what we would normally expect
for a gas which gives them the ability to dissolve large nonvolatile molecules. CO2 is an
industrially important supercritical solvent which is easily recycled after use (by just lowering
the pressure below pc) and it is an environmentally friendly industrial technique as the CO2
is a waste product of industry which is captured and reused and replaces organic solvents
which may have environmental problems. Supercritical fluids are also used in supercritical
chromatography for very efficient separation and analysis of nonvolatile mixtures.
This question is not as easy as it might first appear. The method suggested is one of several
solutions but with this type of question it is easy to make a mistake, so a step by step
approach is best.
* - % %
* -
* -
78 #
' %% 78 #
'%
'%
* -
* - ' #
7!8 # 7&8
'% '% % %
* -
- '% # %
* -
7,8 * 7
8 %
% %
* -
* - ' % #
Stating with the variation of vapour pressure with temperature equation (1) then multiply
out the brackets and cancel were possible gives (2). Rearrange (2) so that the term containing
T2 becomes the subject in (3). Divide by ΔvapHo/R throughout in (3) gives (4). Before we
can take the reciprocal of we need to bring the right hand side of (4) to the common
denominator of T1ΔvapHo as in (5). Finally we can now take reciprocals on both sides of
equation (5) to give T2 as the subject in (6).
www.job.oticon.dk
In order to simplify the expression using partial fractions we use two undetermined
coefficients A and B whose values we shall determine later on. In the partial fraction, note
that the symbol “≡” means “identical to” and that an identity is true for all values of x. On
the other hand, for an equation the left hand side is only equal to the right hand side for
certain values of x which are called the solutions or roots of the equation.
(
Definition: partial fraction
. . . .
Multiply both sides of the identity above by (ao − x)(bo − x) to give equation (1) and cancel
out where possible gives (2). The identity (2) is true for all values of x and if we set x =
b0 as in (3) we find the unknown constant B in (4). Alternatively if we set x = a0 in the
identity (2) we have (5) which gives us the unknown constant A in (6). Note that the two
unknown constants in (4) and (6) are A = −B.
Returning to our original partial fraction identity let us substitute for A and B for the
partial fractions gives (7). Then in (8) we use brackets to clean up the right hand side of
the solution of the partial fraction.
(
7%8
. . . . . . . . . .
7+8
. . . . . .
Return to Question 8 (see page 23)
The carbon atom isotopic abundances in propane are found by expanding the binomial
bracket below for the four different isotopologues with “3-C-atoms, choose 0, or 1, or 2,
or 3” to be 12C.
78 //.!
!// !// !. //!
! !
78 // !
7!8 // !//
!//
/% /&.! //
/%.//
!
7!8 // /% /&.! /%.//
Expand the binomial bracket (1) using the binomial theorem. Cancelling out terms top
and bottom gives (2) with the expected 1, 3, 3, 1 binomial coefficients for having 0, 1, 2,
3 12C atoms, that is the fourth row of Pascal’s triangle. Multiplying by the 99 terms (the
approximate percentage of 12C) we obtain (3) the isotopologues abundancies.
Mathematically we have finished, but in chemistry the mass spectrometric abundances are
normally expressed as a percentage of the largest abundance to give the following predicted
mass spectrum. Notice how rarely we would find in our mass spectrometer a propane
molecule with three 13C atoms (about one in a million propane molecules). On the other
hand, just over 3% of the propane molecules contain one 13C atom.
! !
! + + ! + ! +
/% /&.! /%.//
..... Y ..! Y !.!Y .. Y
Return to Question 9 (see page 23)
2 WEEK 2: CHEMISTRY
AND ALGEBRA 2
Chemical spectroscopy and analytical chemistry use light of various types, not just visible
light, but also other regions of the electromagnetic spectrum. In quantum mechanics one
view of light is that it is made up of a stream of photons which are particles travelling at
the speed of light (c). The energy of a single photon, E, is the product of Planck’s constant
h = 6.6261×10−34 J s and the frequency of the light. The accepted symbol for frequency is
ν (italic Greek “nu”). Don’t confuse frequency, italic Greek “nu” (the Greek “n”) with the
italic roman “vee”, you can rely on the context to remove any confusion.
"
Photon energy
The photon energy equation combines the particle view of light on the left hand side (the
energy of a photon) with the wave view of light on the right hand side of the equation (the
frequency of a wave), i.e. this equation embodies wave-particle duality! Photons also have
spin, as below.
Figure 2.1: photons with right-handed (+1) and left-handed (−1) polarization.
In section 1.2.1 we saw that quantum mechanics also views light as a wave, we have an
equation that relates the frequency ν and the wavelength λ of light to the velocity of light
c = 2.9979×108 m s−1.
Wavelength and frequency of light
Combine these two equations and calculate the energy of a single photon of wavelength
λ = 427 nm (the symbol nm stands for a unit called a nanometre or 10–9 m).
35
This stoichiometric (the overall) reaction may have a mechanism consisting of several
elementary steps. The experimental rate law for this reaction is found to depend upon the
square of the concentration of the reactant A, the cube of the concentration of the reactant
B, and is inversely proportional to the concentration of the product C. The rate of reaction
v (italic roman “vee” as in velocity) is generally given in terms of the concentrations of the
reacting species (indicted by square brackets) and the rate constant k. Purely for clarity in
chemical kinetics it is conventional to indicates the concentration by just the pure number
[A] and not explicitly show the mol L−1 units terms.
3 5!
&
Calculate the rate of reaction if the concentrations of A, B, and C are all doubled by finding
the ratio of the final rate of reaction, vfinal to the initial rate of reaction vintial.
An ionic crystal MX (M+ being the metal cation and X− the non-metal anion) is shown in
Fig. 2.2, a small section of a KCl crystal.
The ionic crystal has a certain amount of intrinsic strength that holds the solid crystal
structure together. This is called the lattice enthalpy ΔlatHo and is given by the equation of
the Born-Haber cycle for KCl.
# -
0 -
- _
- 7#4#8
-
( - KCl Born-Haber cycle
These enthalpy changes will be discussed early in your chemistry degree. The enthalpy changes
are for: (1) electron gain of a chlorine atom in the gas phase, (2) ionization of a gaseous
potassium atom, (3) bond dissociation of Cl2 gas, (4) sublimation of solid potassium, (5)
formation of KCl, and (6) the lattice energy of KCl.
Given the following data for potassium chloride crystals, rearrange the Born-Haber equation
and calculate the electron gain enthalpy of chlorine.
0 -
-
- 7#4#8
-
( -
&!% ?# +/ ?# && ?# &+ ?# %% ?#
The double full-arrows indicate that the reactions may not necessarily have come to
equilibrium but the reaction is reversible and can be started from either A or B molecules.
Writing A for cyclopropane and B for propene for clarity, the Lindemann mechanism for
the isomerization follows first-order kinetics in the concentration of A.
3 5
The Lindemann mechanism for first-order reaction mechanisms involves the collision between
A and another molecule either A or B, let’s call it M in general. This produces a highly
vibrationally excited molecule A symbolized as A* which has enough energy to isomerize to
B but it is not the transition state of the reaction. This is because the excitation energy of
A* is dispersed over the vibrational modes of A*, 21 modes of vibration for cyclopropane.
The energized molecule A* may do one of two things. If A* collides before it can isomerize
it may lose enough vibrational energy that it can no longer isomerize, i.e. it has returned to
a molecule A. Alternatively, if the energized molecule A* doesn’t collide quickly, it may have
enough time to isomerize to B by the vibrational energy accumulating into the transition
state reaction coordinate.
3 3Z 3Z 5 Lindemann mechanism for unimolecular reactions
During the algebraic manipulation of the three-step Lindemann mechanism the following
kinetic equation is obtained.
The square brackets indicate concentrations of the molecules A and A*. Rearrange the
equation to get [A*] the concentration of the energized reactant as the subject. The rate
of formation of the product B is from the mechanism v = k3[A*], hence find the rate of
reaction by substituting in your expression for [A*].
,
G
G ! G
The standard reaction enthalpy ΔrHo is equal to the sum of the standard enthalpies of
formation ΔfHo of the products minus the sum of the standard enthalpies of formation of
the reactants. Each enthalpy of formation being multiplied by the stoichiometric coefficient
ni for that molecule. Enthalpy of formation is covered during your first year of your degree.
-
0 - *
0 - Definition: standard reaction enthalpy
The standard enthalpy of combustion of benzene is shown below, rearrange the equation
so that the enthalpy of formation of benzene is the subject.
-
0 - G !
0 - G
0 -
,
0 - G
Fig. 2.5 is the visible Balmer spectrum that is observed for H-atoms when an electric
discharge is passed through H2 gas.
4 Definition: wavenumber
This e-book
is made with S E TA S I G N
SetaPDF
www.setasign.com
Download free eBooks at bookboon.com
Click on the ad to read more
40
Click on the ad to read more
INTRODUCTORY MATHS FOR CHEMISTS:
CHEMISTRY MATHS 1 Week 2: Chemistry and Algebra 2
& 4&
*
=(
*
*
Figure 2.7: Bohr model of light emission as jumps between quantum states.
The quantized energy levels are called atomic orbitals (AOs) and for the H-atom are
characterized by the principal quantum numbers, n = 1, 2, 3, … ∞ where the infinite energy
level corresponds to ionization of the H-atom. The Bohr model for light emission is the
transition from is n2 to n1 and for light absorption from n1 to n2 where the energy of n2 >
n1. In Fig. 2.5 the Balmer visible spectrum are transitions down to n1 = 2.
4 '
Rydberg equation for H-atom electronic spectra
The Rydberg equation accurately summarizes the UV-visible electronic spectrum of H-atoms.
It gives the wavenumber of the light emitted or absorbed where the Rydberg constant is
RH = 109677 cm−1. Rearrange the Rydberg equation to obtain n2 as the subject.
( ( (
The above chemical reaction when carried out at a temperature of 425 oC has not only
the H2 but also the I2 and HI all in the gas phase. The chemical symbol of the “two half-
arrows” indicates that the reaction is at a dynamic equilibrium, the reaction is going in both
directions simultaneously with equal rates of reaction. This reaction has an equilibrium
constant Kco = 55.64 at 425 oC (see section 1.2.3 for a discussion about dimensionless
equilibrium constants). If we originally placed 2 mol L−1 of H2 and 2 mol L−1 of I2 in a
flask at 425 oC with no HI present, calculate the equilibrium concentrations of H2 I2 and
HI when equilibrium is achieved at 425 oC.
, ,,
&
The way to tackle this type of question is to assume that x moles of H2 and I2 are consumed
in going from the original conditions to equilibrium and 2x moles of HI are formed when
equilibrium is achieved. The concentrations at equilibrium are then shown below.
,
,,
&
The rather difficult chemistry problem comes down to solving the maths for x in this
equation. Solve this equation in the most efficient manner which may not necessarily be
by the use of the quadratic formula. Then using the chemically sensible value of x from
your solution, calculate the equilibrium concentrations for the three species of molecule.
When a molecule forms a crystalline solid we can use a beam of monochromatic X-rays to
determine the positions of equivalent atoms within the crystal i.e. the same atom within
the identical molecules making up the crystal. X-rays are very short wavelength light with
wavelengths of a few ångstroms (Å = 10−10 m) which are of the order of the spacing between
equivalent atoms in crystals.
The crystal consists of equally spaced planes of equivalent atoms, each of which may be part
of a large molecule. Most of the X-rays pass through the crystal without any interaction.
Some X-rays interact with the planes of equivalent atoms but will have different path-
lengths depending upon which plane of atoms “reflect” them. These different path-length
X-rays will interfere, mainly in a destructive manner. A small number of photons will have
a path-length difference of 2d sin(θ) equal to an integral multiple n of the wavelength λ
of the X-rays and they constructively interfere and give a diffraction pattern. Clearly the
path-length difference depends upon the angle of incidence θ of the X-rays to the planes of
atoms the crystal and the distance between the planes of atoms d (called the lattice spacing).
Where n = 1, 2, 3,… is the order of the diffraction and is a dimensionless number. Calculate
the lattice spacing d when using radiation of wavelength λ = 0.154 nm for n = 1 (called a
first-order reflection) at a scattering angle of θ = 11°. Specify the units for your calculated
d value.
Long polymer molecules tend to coil up rather than being stretched out straight, Fig. 2.10 is
an example of a short polythene (polyethylene) molecule. Averaging a large number of such
randomly coiled molecules the average end-to-end distance <r> of the polymer molecules is
given blow. The angled brackets in <r> is the normal way of indicating the mean or average
of the quantity inside the brackets.
7 Average end-to-end distance <r> of an ideal random coil polymer
The polymer consists of N chemical bonds making up the backbone of the polymer chain,
with a monomer length of l, and θ is the C-C-C atom bond angle. Calculate the average
end-to-end distance <r> for a polymer with θ = 109.5° (the tetrahedral angle), l = 154 pm,
and N = 5×103. Specify the units for your calculated <r> value.
Figure 2.11: a polyene conjugated molecule H (light grey) and C (dark grey).
A conjugated molecule is one with alternating single and double carbon-carbon bonds in its
backbone. Conjugated molecules are important parts of large biological molecules for many
different functions including your ability to detect light within your eyes’ retina and for
plant photosynthesis. Conjugated molecules are important industrial dyes. A simple example
of a conjugated molecule is a “linear” polyene, Fig.2.11. The quantum mechanical equation
The Wake
the only emission we want to leave behind
.QYURGGF'PIKPGU/GFKWOURGGF'PIKPGU6WTDQEJCTIGTU2TQRGNNGTU2TQRWNUKQP2CEMCIGU2TKOG5GTX
6JGFGUKIPQHGEQHTKGPFN[OCTKPGRQYGTCPFRTQRWNUKQPUQNWVKQPUKUETWEKCNHQT/#0&KGUGN6WTDQ
2QYGTEQORGVGPEKGUCTGQHHGTGFYKVJVJGYQTNFoUNCTIGUVGPIKPGRTQITCOOGsJCXKPIQWVRWVUURCPPKPI
HTQOVQM9RGTGPIKPG)GVWRHTQPV
(KPFQWVOQTGCVYYYOCPFKGUGNVWTDQEQO
for the π-electrons in a simple conjugated polyene molecule has a wavefunction ψ(x) that
depends upon the position x along the length L of the molecule. The ideas of wavefunctions
and quantum mechanics for such systems are covered early on in your chemistry degree.
The general equation that applies to all wave motion is given below (Parker 2015, section
1.7.2 and Parker 2016, section 1.7).
Where A, B, and k are constants whose values need to be determined for the polyene
molecule and the sine and cosine functions are in radians not degrees. Find the constant k
and any constraints that A or B may have using the boundary conditions that at the two
ends of the polyene molecule the wavefunctions must be zero in order to have the electron
confined within the polyene, ψ(0) = 0 and ψ(L) = 0.
The water molecule has an O−H bond distances of 0.9687×10−10 m and the distance between
the two hydrogen atoms is 1.5391×10−10 m. Calculate the H−O−H bond angle.
! GG G ! GG ! G
Figure 2.13: acetic acid, water, acetate ion, and hydronium ion.
Acetic acid (ethanoic acid) is a weak acid which partially dissociates in water to the acetate
ion and hydronium ion. The standard equilibrium constant is Kao (“a” is for acid). See
section 1.2.3 for a discussion about dimensionless equilibrium constants.
GG ! G
,
! GG G
The solvent water is in a large excess and to a very good approximation its concentration
is constant. We can incorporate [H2O] with the standard equilibrium constant Kao to give
the “acidity constant” Ka = Kao [H2O].
GG ! G
, Acetic acid acidity constant Ka
! GG
Acetic acid’s acidity constant is Ka = 1.7539×10−5 at 298.15 K. Calculate the hydronium ion
concentration [H3O+] for a 0.15 mol L−1 solution of acetic acid? We can solve this chemical
problem by saying the fraction of acid dissociated at equilibrium is x. For each molecule of
acid that dissociates it gives an equal number of acetate ions and hydronium ions.
%,!/.,
.,
This chemical equilibrium problem has been reduced to solving the quadratic equation
written in the standard form ax2 + bx + c =0 using the quadratic formula.
&
The quadratic formula
Ammonia is an important industrial chemical as a starting material for many other compounds
including fertilizer, nitric acid, urea, and ammonium nitrate. Ammonia is synthesized using
the following equilibrium reaction where the reactant and products are in the gas-phase.
Its makes sense to define the standard pressure equilibrium constant Kpo in terms of the
partial pressures of the gases at equilibrium, which is Kpo = 1.64×10−4 at 400 oC.
K !6
, Ammonia standard pressure equilibrium constant
K 6 6 !
Kpo is given in terms of the partial pressures of the reactants and products in pascals or bars
where each partial pressure is divided by the standard pressure of po = 1×105 Pa or po = 1
bar and Kpo is dimensionless. Each of the terms is raised to the power of its stoichiometric
coefficient. As the standard pressure base unit in pascals is not unity po = 1×105 Pa and
appears in each of the po terms it affects the value for Kop and we cannot make a simplification
similar to the one we made for Kco in terms of concentration where co = 1 mol L−1 (see
section 1.2.3). For any given equilibrium reaction each Kpo in terms of partial pressures in
pascals must be written in full with the po values included. This suggests that it is better
to use the standard pressure po = 1 bar (which is unity) rather than 1×105 Pa and that all
the pressures are also quoted in bar we can then cancel out the units “bar” and greatly
simplify the arithmetic!
For our reaction we initially have partial pressures of pN2 = 1 bar and pH2 = 3 bar of H2 which
are mixed, there being no NH3 present and the mixture is allowed to come to equilibrium
at 400 oC. If x is the fraction of a mole of N2 which is lost due to the reaction then the
equilibrium constant becomes as below.
K (
! (
K ! ( , !
&. &
!!
Note that “each time the reaction occurs” we form two molecules of NH3 hence the term
2x. Likewise, each time the reaction occurs we lose 3 molecules of H2 thus the term (3–3x)
and remember that each terms is raised to the power of its stoichiometric coefficient.
We can simplify the equation by taking out any common factors then calculate the fraction
by solving for x using the quadratic formula. Using this value for x calculate the partial
pressures of all the reactants and products.
"
The common variable in these two equations is the frequency υ. The energy of a single
photon is found by: rearranging c = λυ to get υ as the subject υ = c/λ, substituting this
expression for υ into E = hυ gives E = hc/λ.
(1) Substitute the physical quantities into E = hc/λ note that in the calculation we have
replaced nanometres by the base units metres in order that the units cancel. (2) The energy
of a single photon is incredibly small, the power of −19 is a minute fraction of a joule, and
one joule is about the energy required to lift a medium-size tomato (100 g) 1 m vertically
from the surface of the Earth. (3) In practice we normally only meet photons in enormous
numbers (a mole of photons) which can give chemically significant amounts of total energy
of hundreds of kJ of light energy. A mole of photons is called an einstein (note the lower
case “e” for the named unit, the 1921 Nobel Prize winner Einstein is an upper case “E”).
(4) An einstein of 427 nm photons is 280 kJ which for comparison is of the order of
magnitude of the Cl-Cl bond dissociation energy of 242 kJ mol−1.
3 5! 3 5! & 3 + 5!
78 # & 78 0# & 7!8 0# &
!
3 5 0#
7&8 0#
& 7,8 0#
# 7
8
#
(1) the original rate law. (2) The final rate law with the powers applying to the new doubled
concentrations. (3) and (4) Take the numeric factors outside the brackets. (5) Replace the
bracketed terms with the initial rate law. (6) Doubling all the concentrations has made this
particular reaction go 16 times faster.
78
# -
0 -
- _
- 7#4#8
-
( -
78
( -
# -
0 -
- _
- 7#4#8
-
7!8
( - %% &!% +/ _ && &+ ?#
7&8
( - %% &!% +/ _&& &+ ?#
7,8
( - !&/ ?#
Technical training on
WHAT you need, WHEN you need it
At IDC Technologies we can tailor our technical and engineering
training workshops to suit your needs. We have extensive OIL & GAS
experience in training technical and engineering staff and ENGINEERING
have trained people in organisations such as General
ELECTRONICS
Motors, Shell, Siemens, BHP and Honeywell to name a few.
Our onsite training is cost effective, convenient and completely AUTOMATION &
customisable to the technical and engineering areas you want PROCESS CONTROL
covered. Our workshops are all comprehensive hands-on learning
experiences with ample time given to practical sessions and MECHANICAL
demonstrations. We communicate well to ensure that workshop content ENGINEERING
and timing match the knowledge, skills, and abilities of the participants.
INDUSTRIAL
We run onsite training all year round and hold the workshops on DATA COMMS
your premises or a venue of your choice for your convenience.
ELECTRICAL
For a no obligation proposal, contact us today POWER
at training@idc-online.com or visit our website
for more information: www.idc-online.com/onsite/
(1) The lattice enthalpy equation is rearranged to give (2) with the electron gain enthalpy as
the subject of the equation. (3) Substitute the physical quantities for the symbols, note the
difference between the mathematical operations of addition or subtraction as shown by plus
or minus signs with a space before the number, and the positive or negative values of the
physical quantities themselves shown by plus or minus without a space before the number.
(4) Multiply out the brackets. (5) The value for the electron gain enthalpy ΔegHo = −349 kJ
mol−1 means the reaction Cl(g) + e−(g) → Cl−(g) is exothermic, that is the isolated chloride
ion is more stable than the isolated chlorine atom and electron, all in the gas phase. Also
there is in the literature the term “electron affinity” Eea which is the energy released in the
above reaction. The word “released” means that for the exothermic reaction is written as
Eea = +349 kJ mol−1.
Return to Question 3 (see page 37)
Starting with the original kinetics equation (1) we move the term which does not involve [A*]
to the right hand side by subtracting k1[A]2 from both sides, (2). Multiply (2) throughout
by −1 to make all three terms positive, (3). Take [A*] outside a bracket in (4) for the two
terms on the left hand side. (5) Divide left and right by (k2[A] + k3). The rate of reaction
in (6) is k3[A*], so (7) is the exact rate of reaction but can we approximate it?
Looking at the left hand side of (3) the [A*] « [A] so k3[A*] « k2[A*][A] which cancelling out
[A*] in this inequality gives k3 « k2[A]. We can now approximate the exact rate of reaction
(7) by ignoring the negligible term k3 in the denominator we get the approximation for the
rate in (8). For this approximate rate (8) we may now cancel out the [A] in the denominator
with the square term to obtain a good approximation for the rate of isomerization.
&!&
3 Lindemann rate of “unimolecular” isomerization reactions
&
had a minus sign) and so the denominator consists of two separate terms, not one single
term. You can show this is true, look at the numeric example below.
/
! ! / ! !
' ' ' ' ' '
78 4 ' 78 4 7!8 4 7&8 4
7,8
'
' 4
7
8
' 4
'
7%8
'
' 4
7+8
'
' 4
Rearranging the Rydberg equation (1) to obtain n2 as the subject of the equation is a bit
more difficult than it may appear at first sight. Expand out the bracket (2) on the right.
(3) Take RH/n22 over to the left and then move nu-bar over to the right, gives (4). (5) Bring
the right hand side to a common denominator of n12 this is a crucial step as otherwise you
cannot later take reciprocals. (6) Divide left and right by RH. (7) Take reciprocals on both
sides and (8) take the square root of both sides. In the last step of square rooting, the n1
in the numerator of (7) can be square rooted and taken outside the square root bracket.
On the other hand n1 in the denominator of (7) cannot be square rooted because of the
minus sign, the denominator is made up of two terms which must both be treated equally.
78 , ,,
& 78 ,,
&
7!8 %&,/ 7&8 %&,/
(1) Note that the [HI] is all squared [2x]2 it is not 2x2. The equilibrium constant equation
is solved most easily by (2) writing the denominator as a square term and then in (3) take
the square root of both sides and note the ±7.4592. (4) Move [2−x] to the right.
0
7,8
%&,/
7
8 %&,/
7%8 &/+& %&,/ 7+8 &/+& %&,/
7/8 /&,/ &/+& 7.8 ,&,/ &/+&
78 ,%% 78 %!%
The equations (5), (7), (9), and (11) carry out the algebra for the +7.4592 square root. The
equations (6), (8), (10), and (12) process the −7.4592 square root. The concentrations of
H2 and I2 which equal [2−x] must be positive as you cannot have negative concentrations,
so the chemically sensible root is x = 1.5571. The equilibrium concentrations are calculated
by substituting x = 1.5571 as below.
www.simcorp.com
.,&
78 78 7!8
.,&
7&8 7,8 .&.!
./
The Bragg equation (1) is rearranged so that d is the subject (2) and then substitute in the
values of the variables (3). (4) Note that we do not need to change the units to the base
unit of metres, they may be left as nanometres (nm) but whichever unit of length is used
it will need to have been inserted in the equation in order to know the units of the lattice
spacing (5).
./,
78 7 78 ,.! ,& *
./,
+ &
7!8 !%&. * 7&8 ,&.+. *
The average end-to-end distance equation (1) has the physical quantities substituted (2) for
the variable symbols. It was not necessary to convert to metres, but it was necessary to have
the units of the length in the equation. Clearly when we take the square root of (3) it is
only the positive distance in (4) that is chemically sensible. The average end-to-end distance
of the polymer <r> = 1.5408×104 pm which is about 2% of the stretched out “linear”
length of polymer 77×104 pm. Macromolecules both natural (e.g. DNA, RNA, proteins, and
carbohydrates) and synthetic polymers will be covered in detail in your chemistry lectures.
!!
!
The wavefunctions for the π-electrons (Fig. 2.15) have to be zero at both ends of the polyene
molecule so that the wavefunctions are “standing waves” which persists through time, these
are called the “boundary conditions” in quantum mechanics. Another way of looking at the
boundary conditions is that in order for the π-electrons to be confined within the polyene
the wavefunctions must be zero at the two ends of the polyene.
�e Graduate Programme
I joined MITAS because for Engineers and Geoscientists
I wanted real responsibili� www.discovermitas.com
Maersk.com/Mitas
I joined MITAS because fo
I wanted real responsibili�
Month 16
I was a construction
supervisor ina co
I was
the North Sea su
advising and the
Real work he
helping foremen ad
International
al opportunities
Internationa
�ree wo
work
or placements ssolve problems
Real work he
helpin
International
Internationaal opportunities
�ree wo
work
or placements ssolv
Download free eBooks at bookboon.com
Click on the ad to read more
54
Click on the ad to read more
INTRODUCTORY MATHS FOR CHEMISTS:
CHEMISTRY MATHS 1 Week 2: Chemistry and Algebra 2
The general wave equation (1) needs to fit with the two boundary conditions. (2) The left
hand boundary condition x = 0 and ψ(0) = 0 but in radians cos(0) = 1 and sin(0) = 0. So
the sine function agrees with the quantum mechanics but the cosine does not agree. Hence
in order to remove the cosine term from the general wave equation we must set A = 0 to
give (3). (4) The right hand boundary condition of the polyene molecule x = L and ψ(L)
= 0 and so in radians sin(kL) must be equal to zero. This means kL (in radians) must be
equal an integral number n = 1, 2, 3,… of π so kL = nπx. The integer is a quantum number
and in general quantization arises because of the fulfilling of boundary conditions such as
the ones in this question. (5) Rearranging kL = nπx gives the value of the constant k =
nπ/L. (6) Finally we have the (incomplete) wavefunction for the particles, the π-electrons,
in a one-dimensional box, the polyene conjugated molecule. The only constant that is yet
to be found is B the normalization constant. The value of the constant B is calculated later
in section 8.2.2.
Return to Question 10 (see page 44)
The water molecule forms an equilateral triangular structure. We can then bisect the H–O–H
angle to make two right angled triangles. The O−H distance is 0.9687 Å and the half the
H−H distance is 0.76955 Å with θ the half angle for H–O–H.
@ @
@
.%
/,, 3
78 .%/&& 78 .%/&& ,
. .,
7!8 G
./
+% 3
(1) The distance multiple of 10–10 m is called an ångstrom with the symbol Å, it is a
common and convenient unit used by chemists as it is of the order of bond distances and
atomic radii. (2) As θ is only half of the water bond angle, the calculated bond angle (3)
is 105.2o close to but not equal to the tetrahedral angle 109.47o.
78 %,!/., 78 %,!/.,.,
.,
7!8
!./.
%,!/., 7&8 %,!/.,
!./.
.
(1) The fraction x of acid lost in going to the equilibrium is given by this quadratic equation.
(2) Rearrange the equation by moving (0.15 − x) to the left hand side. (3) Multiply out
the brackets. (4) Write the quadratic in the standard form.
&
7,8
%,!/., %,!/., &
!./.
7
8
%,!/. !.%
. .,&.
, . ,
7%8
(5) The quadratic formula where the symbols have their conventional meanings. (6) Substitute
our quadratic equation coefficients. (7) Note the change of sign to the second term within
the square root and also the very large difference in magnitude between b2 and 4ac terms.
(8) I will temporarily be keeping some extra significant figures in order to show the effect of
this relative difference in magnitudes between b2 and 4ac. (9) Once we have taken the square
root of the second term in the numerator then we only need to retain a sensible number of
4 or 5 significant figures. So which value of x applies in this question? Remember x is the
fraction of acetic acid that has dissociated. A fraction is a positive number of less than one
(unity). Another way of looking at this is that you cannot have a negative concentration,
thus the chemically correct result is (10) that is, the concentration of hydronium ions in
the acetic acid dissociation.
Although not part of the question, this concentration of H3O+ has a pH = −log(1.6133×10−3)
= 2.792 which is acidic, as expected.
& &
&
78 !
&. 78 ! !
&. 7!8 &
.%..!
!! !
7&8
.%.. !
7,8
!!%.
7
8
The solution to this problem is long but the solution is not difficult and working with
pressures in bar rather than pascals simplifies the arithmetic. The equilibrium equation (1)
may be solved for x by (2) taking out the numeral factors. (3) The numerical factors are
taken over to the right and evaluated. (4) Square root both sides of the equation to give (5).
(6) In order not to get lost it helps to keep the quadratic equation looking simple, let g =
3.3272×10−2 and take the denominator of (5) to the right and side and expand out (1−x)2.
www.skoda-career.com
Multiply out the bracket of (6) and rearrange to (7) and then into the standard quadratic
form (8). The quadratic formula (9) has our variables substituted to give (10). (11) Square
the term in the square root bracket and cancel out 4g2. (12) Substitute for g = 3.3272×10−2
and calculate the two roots (13) of the quadratic equation. As the mole fraction must be
positive and less than unity then x = 0.031 is the chemically sensible result. This gives the
three equilibrium partial pressures.
K ..!
./
/
!! !!..! /.%
K
!
..! ..
3 WEEK 3: CHEMISTRY,
LOGARITHMS AND
EXPONENTIALS
This material should be completed in 1 week, and I thought it would be perhaps convenient
for you to have a summary about logs, exponentials, and acid-base equilibrium.
We are familiar with expressing a number as a power (also called an exponent) of a base
number. These numbers may also be written as logarithms, the log is the power required
to produce a given number.
(1) 102 = 100 where “2” is the power and “10” is the base number and writing it as a log
is equation (2). A second example (3) 53 = 125 with “3” the power and “5” is the base
number and (4) is written as a log. A third example (5) is 109 = 1000000000 which is a
billion with “9” the power and “10” the base number and (6) its log equivalent. A final
example (7) is that of the byte a sequence of eight bits of information 28 = 256 which is
the standard for computer architectures and (8) its log. Logarithms (logs for short) are the
inverse process to exponentiation.
The rules of logs cover (1) the multiplication of the two numbers, (2) the division of the
two numbers, and (3) the power of the number.
78 #(
#( #( Rules of logs
careers.slb.com/recentgraduates
The additional rules of logs are just special cases of the general rules above, (1) the number
is the same as the base, (2) the number is the same as the base raised to a power, (3) the log
is an exponent of the base number, and (4) when the number and the base are exchanged.
Logs to base 10 are used in science and engineering to express either very large numbers
or very small numbers and are normally written as “log” without a 10 subscript, they are
pronounced as “logs” or “common logs”. Large numbers have positive powers e.g. 1023 with
log(1×1023) = 23, and very small numbers have negative powers e.g. 10−34 with log(1×10−34)
= −34. An example is pH which range from 0 to 14, from a high concentration of protons
pH = 0 or [H+] = 100 = 1 mol L−1 to the very low concentration of pH =14 or [H+] =
1×10−14 mol L−1.
Logs to base “e” are normally written as “ln” rather than loge and they are pronounced as
“natural logs” or “Napierian logs” after John Napier (1550–1617) a Scot from Edinburgh
who invented logs and popularized the modern usage of the decimal point in writing
numbers. Natural logs are important for their maths properties which model many natural
phenomena, hence they appear often in science and engineering. The base number of
natural logs is Euler’s number “e” which comes from the infinite series below and is named
after the Swiss mathematician Leonard Euler (pronounced as “oiler” in English). Euler’s
number is easy to remember to 16 significant figures (if you like that sort of thing), it is
2.7 followed by 1828 twice then 45, 90, 45 the angles of a right-angled isosceles triangle
e = 2.718281828459045… after which it is really boring. Below are two alternative ways
of defining Euler’s number (truncated at small values of n).
#
,
&++
,/!
%.&
%
%+
j`
Euler’s number
M . ! & ,
% + / .
%+
.
The n = 0 term of the first definition uses the fact that anything raised to the power 0 equals
one. The second definition has the symbol “!” which means the factorial of the number
(see section 1.2.9). Note that e is the symbol for a numerical constant it is not a variable
and e is similar to π (also a symbol for a numerical constant) or any other number such
as 10 is written in upright roman script not in italic script. In science and engineering we
quite often find equations with ex or e−x in them. The negative power is just a special case
of ex in which the sign is positive or negative for even and odd power of x, respectively.
! &
! & Euler’s number raised to the powers x and −x
! &
! &
The equation y = ex may also be written as y = exp(x) with “exp” all lower case. This exp
notation is used for greater clarity particularly in printed material. For example the Arrhenius
equation k = A e−Ea/RT can also be written as k = A exp(−Ea /RT). However, be careful in
using some calculators with a key marked EXP, all capitals, which means “10 raised to an
exponent”, where you type in the power (the exponent).
An important and common process is to change the base of the logs. (1) Shows the general
formula for changing from base number (b) to base number (a). (2) The specific change
between natural logs and base-10 logs and in (3) this change is written in the conventional way.
The equation 122x = 35.4 is an example of an indicial equation, the variable is present as
an index (or exponent or power). To solve (1) for x, we take logs on both sides (2), from
the rules of logs we can take the power down (3), and evaluate the logs in (4). The solution
of the indicial equation is (5).
78 !,& 78 #( #(!,& 7!8 #( #(!,&
7&8 .%/+ ,&/.. 7,8 .%%
+
The concentrations of protons H+ (present as hydronium H3O+ ions) and hydroxyl ions
OH− are very small in aqueous solution and involve many negative powers of ten in mol L−1.
Humans are normally happier dealing with positive numbers that are in the range of 0-100.
The “p” part of the variables pH and pOH refers to the power (index or exponent) of these
small concentrations which are made more manageable by using logarithms. Taking the
negative of the log converts the negative number into a positive number. The symbols pH
and pOH are by convention written with upright roman script not as italics.
* #( .
#' *G #(.
G
#'
Definitions: pH and pOH
Strictly speaking pH and pOH should be defined in terms of the “thermodynamic activity
a” of H+ and OH− but for most purposes concentrations are of adequate precision.
! G G
78 G G ! G G 78 ,
G
The self-ionization or autoprotolysis of water (1) has an equilibrium constant Kc (2) but
as the concentration of the solvent water is to a very good approximation constant, we
may combine the water concentration with the equilibrium constant [H2O]2×Kc = Kw the
autoprotolysis (or self-ionization) constant of water.
G 3
7!8 3 G ! G 3 7&8 , ! ! G 3
7!8 3 G ! G 3 7&8 , 3
3
3
7,8 * , #( . , 7
8 * , * #( 3
7,8 * , #( . , 7
8 * , * #(
3
3
For an acid (3) we can define an acid dissociation constant (4) Ka again incorporating the
approximately constant solvent concentration [H2O] into Ka. The pKa of the acid dissociation
constant is defined (5) in a familiar way. The pKa is related (6) to the pH from equation
(4). Rearranging the last equation (6) we obtain the Henderson-Hasselbalch equation which
applies to weak acids and buffer solutions.
3
* * , #( Henderson-Hasselbalch equation
3
The molar concentration of OH− ions in a certain solution is 1.04×10−5 mol L−1. Calculate
the pOH and then the pH of the solution.
Death occurs if the pH of human blood plasma changes by more than about ±0.4 from its
normal value of about pH = 7.4 the “safe” range is from neutral to slightly alkaline. What
is the corresponding range of molar concentrations of hydronium ions for which human
life can be sustained?
K& G K ! ! G
The concentration of the reactant, say A, in a first-order chemical reaction decreases with
time as it is consumed to form products and the concentration of [A] varies with time
as follows.
Where [A]0 is the initial concentration of the reactant at time t = 0, [A] is the concentration
of the reactant at a later time t, and k is the rate constant for this particular reaction taking
place at constant temperature. Sketch the shape of the graph of [A]/[A]0 plotted as the
y-axis against kt plotted as the x-axis (k is constant at a given temperature) and label the
two axes. You do not need to calculate anything. As a hint Fig. 3.3 is a plot of the negative
exponential curve y = exp(−x). The negative exponential has an intercept at (0, 1) and has
two asymptotes shown in Fig. 3.3 where the arrows mean “in the limit”.
,4`
,
,
4`
`B ,4
For a first-order rate law reaction write the natural log form of the rate law with ln([A]/
[A]0) as the subject.
For a first-order rate law reaction sketch three possible graphs that give straight line plots for
the rate law. Label the axes and give the meanings of the gradient and intercept. Comment
on which is the best of these three graphs.
If a first-order reaction has a rate constant, k = 1.83×10−4 s−1, calculate how long it would
take for the concentration of reactant [A] to decrease from 3.25×10−3 mol L−1 to 2.18×10−3
mol L−1.
So far we have looked at the rate law which shows the variation of the rate of reaction v
with concentration and time, under constant temperature conditions. However, the rate
constant k of a chemical reaction increases with the absolute temperature T, this is modelled
by the Arrhenius equation.
Write the Arrhenius equation in the form of natural logs. Sketch the straight line graph
with the axes labelled for the Arrhenius equation. Give the meanings of the gradient and
intercept in terms of the Arrhenius equation.
On your graph for Question 9, sketch the results for a linear plot for two different chemical
reactions, one with a large activation energy and the other with a small activation energy,
clearly label the two lines. Assume that both chemical reactions have the same value for
the pre-exponential factor A.
The activation energy for the decomposition of hydrogen peroxide on a platinum catalyst
is Ea = 48.9 kJ mol-1 and R = 8.3145 J K−1 mol−1.
G G G
The effect of increasing the temperate is shown by calculating the ratio of the rate constant
k2 at 30oC to the rate constant k1 at 20oC.
The following redox chemical reaction (redox is short for reduction-oxidation) can be made
into an electrochemical cell to derive electrical energy, Fig. 3.5. The cell is called the Daniell
cell and has a voltage difference between the zinc and copper metal electrodes which is
given by the Nernst equation. This voltage difference drives electrons through an external
circuit. The zinc sulfate and copper sulfate solutions are separated by a porous barrier which
allows the ions to pass through.
. .
D1.
2#
2#D1
ED1
2#
E E
The standard electrode potentials Eo are with solutions of 1 mol L−1 and 298.15 K. The
cell potential E varies with the temperature of the cell and the concentrations and is given
by the Nernst equation.
'% l
" " # Nernst equation for the Daniell cell
=
Where n is the number of electrons transferred in the stoichiometric reaction (in this case
n = 2), R is the gas constant, T the temperature in kelvins, and F is the Faraday constant
(one mole of elementary charge).
Calculate the concentration of zinc ions [Zn2+] for E = 1.21 V, [Cu2+] = 0.1 mol L−1, T =
298.15 K and Eo = 1.10 V. Make use of the conversion 1 joule = 1 coulomb volt (1 J =
1 C V).
The standard Gibbs energy of reaction ΔrG° is the maximum non-expansion work that a
chemical reaction can supply at constant temperature and pressure, Kco is the equilibrium
constant. Calculate ΔrG° with appropriate error limits, for a reaction which has an equilibrium
constant Kc° = 1.8×10−5 at T = 298.15 K and R = 8.3145 J K–1 mol−1 is the gas constant.
The decay of 238U has a half-life of 4.4683×109 years which is close to the age of the earth
4.54 ± 0.05 billion years (4.54 × 109 years ± 1%). As the name implies the half-life is the
time for n/n0 to equal 0.5. Uranium-238 decays predominantly to thorium-234 by an alpha
particle decay mechanism, where an alpha particle (α) is a fast moving helium-4 nucleus
4
He2+. Calculate the rate constant for this uranium-238 decay process.
!+
: !&
An alkaline solution.
From the definition of a logarithm we have [H3O+] = 10−7.8. The easiest way of solving this
on your calculator is to use “shift-log” which is 10x, then type −7.8, then equals. Take care
to use “shift-log” which equals 10x not “shift-ln” which equals ex. Human blood covers only
a small range of hydronium ion concentrations as shown below.
78 * - %+ 78 %+ #(. ! G - 7!8 ! G - .%+ ,+.+ #'
7&8 * %. 7,8 %. #(. ! G 7
8 ! G .%. ...% #'
3 ..% #'
* * , #( . * /, #( . * /+
3 .., #'
This alkaline buffer has a pH of 9.28. When using your calculator make sure that you
use brackets, type log(0.027/0.025) = 0.03342 and don’t type log 0.027/0.025 which may
“incorrectly” give −62.745.
3
3 3. -*& -*&
3.
Remember that the rate constant k is constant for a given reaction at a fixed temperature
and plotting [A]/[A]0 versus kt is a negative exponential (Fig. 3.6). The ratio [A]/[A]0 is a
dimensionless variable as is kt for a first-order reaction. The reaction starts at time equals
zero kt = 0 and there are no negative times for the reaction, the intercept at kt = 0 is [A]/
[A]0 = 1 that is [A] = [A]0. Note that using this non-linear plot it is not easy to deduce
the rate constant k, we will return to this later on. Fig. 3.6 shows that with an exponential
decrease, that increasing numbers of half-lives the concentration of the reactant decreases
by ½ of the previous value each time. As the time tends towards infinity the concentration
asymptotically approaches zero.
F
7"8!7"8
!
F
! F
! F
!
Starting with the first order kinetics equation (1), take natural logs (2) on both sides, expand
out the right hand side of (2) into the sum of the log of each term (3). The natural log of
the exponential in (3) is just the exponent –kt, (4). Move ln[A]0 to the left hand side (5)
and convert the difference of two logs into the log of the ratio (6).
Any of the following three graphs are correct and each of them are linear graphs allowing
the rate constant to be easily found. Firstly, from section 3.3.5 we have the rate law in terms
of a linear equation (Fig. 3.7 left). Secondly, from the rules of logs, section 3.1.2, we can
invert the log on the left hand side of the above equation and the right hand side changes
www.foss.dk
sign (Fig. 3.7 centre). Thirdly, an the intermediate stage section 3.3.5 equation (4) allows
a third and equally valid straight line graph to be plotted (Fig. 3.7 right).
7"8!7"8
7"8!7"8
7"8
7"8
This plot on the right has the advantage that you don’t need to measure the value of ln[A]0
which you would for the other two plots and which may be difficult to experimentally
accurately measure. Also with the plot on the right we are not plotting a log of a ratio
which may show only small variations in [A]/[A0] (left) or very large variations in [A0]/[A]
(middle) for small changes in [A].
We write the rate law as a log of a ratio and note that logs are dimensionless as they are a
ratio of similar physical quantities, so the units cancel and we are taking the natural log of
a pure number (as we must do).
3 +.! #'
# & # !
+!. &
3. !,. #'
.!//!! +!.& + !
!$
!%
Figure 3.8: plot of k against T for first-order reaction data (smoothed data).
Fig. 3.8 shows the curved plot of the rate constant against the temperature which for clarity
is with smoothed data (no random errors). This plot visually emphasizes the exponential
nature of the rate constant’s dependence upon temperature. Otherwise this plot is not very
useful as at high temperatures a small change in T gives drastic changes in the rate constant
k, but at low temperatures a large change in T gives a negligible change in k. It would
be much better if we can rearrange the Arrhenius equation to enable us to have a linear
(straight line) graph as in section 3.3.9.
78 & -*
"
'%
78 # & # -*
"
'%
7!8 # & # # -*
"
'%
7&8 # & #
"
'%
We can go from the exponential form of the Arrhenius equation (1) to the log form by
taking natural logs on both sides (2). The log on the right of (2) is a product of two terms
which can be separated as the sum of two log terms (3) the second term is now the natural-
log of an exponential, which is just equal to the exponent (4). Let us write the log form
of the Arrhenius equation (4) with the 1/T term written explicitly, this helps us see that
the log form of the Arrhenius equation is in the form of a straight line equation written
as y = c + mx.
"
Log form of the Arrhenius equation
# & #
' %
Plotting ln k as the y-axis and 1/T as the x-axis is called an “Arrhenius plot” the intercept
c = ln A and the gradient m = −Ea/R.
!$
! %
Fig. 3.9 uses the same smoothed data as Fig. 3.8. Notice a best straight line would end a
long way from the y-axis intercept (the axes cross at x = 4.0 in this example). So drawing an
extension to the best straight line is not the way forward as it will lead to very large errors.
The equation of the best line was obtained using a spreadsheet is shown below.
#
&
//!!
.!.. L
!
%
The graphs axes are plotted as pure numbers (5). Logs are pure numbers, as in ln(k/s−1)
and we cannot separate the units inside the log term. However, we must separate the units
and multiplier of 1/T as the units (kelvin) and the multiplier are an integral part of the
gradient m in (6). The gradient of the Arrhenius equation in log form is m = −Ea/R and in
(7) we calculate the activation energy by moving R over to the right. The activation energy
in (8) is quoted in the conventional units of kJ mol−1.
The intercept occurs when x = 0 which is long way outside our plotted area, from the equation
of the best-fit straight line c = ln(A/s−1) = 29.9336 in (9). We antilog both sides in (10)
only then are we allowed to take the units over to the right in (11) as it is no longer a log.
Fig. 3.10 shows that the reaction with the larger activation energy Ea = 50 J mol−1 is more
affected by changes in temperature (steeper gradient) than the one with a smaller activation
energy Ea = 20 J mol−1. In the limiting case of zero activation energy the rate constant
would be independent of temperature! Both plots in Fig. 3.10 have the same intercept c
= ln(k/s−1) = ln(A/s−1) = 29.9336 which as we have seen gives A = exp(29.9336) s−1 and A
= 1.0×1013 s−1.
3+&
!$
3+&
!%
Figure 3.10: plot of ln k versus 1/T for high and low activation energies.
& " & &+/. ?#
7&8 # 7,8 #
& ' % % &
+!&, ?L #
!.!, L /!, L
& & &
7
8 # .
+ 7%8 -*.
+ 7+8 /!+!
& & &
www.job.oticon.dk
The increase in temperature from T1 to T2 for the Arrhenius equation is most easily handled
using the log forms of the Arrhenius equation, (1) and (2). Subtracting the log equations
and write the subtraction of the two log terms as the log of the ratio of the terms (3) and
on subtraction the two ln(A) terms cancels out. The common factor −Ea/R is taken out of
the bracketed term in (4). Substitute the physical quantities in (5) for the variables gives
(6). Take the antilog of (6) in (7) to obtain the ratio of rate constants in (8).
Note that the activation energy is quoted in kJ mol−1 but must be converted to the base unit
of J mol−1 in order to cancel out the units, and also that the temperature T must always be
in kelvin. For this particular reaction its rate constant has almost doubled and the reaction
is going almost twice as fast for a 10oC rise in temperature from 20oC to 30oC. This is only
true for reactions with activation energies of about 50 kJ mol−1.
'% l l " "=
78 " " # 78 #
=
'%
l . S /
&+, ?S # l
7!8 #
7&8 # +,
%
. #' +!&, ?L # /+, L . #'
7,8 l /..& . #' 7
8 l /.., #'
The Nernst equation for the Zn2+/Cu2+ redox reaction (1) is rearranged in (2) so that the
log term containing [Zn2+] is the subject. Substitute for the variables in (3) gives (4). Take
antilogs of (4) and then you are able to multiply by 0.1 mol L−1 on the right hand side of
(5). The unknown zinc ion concentration (6) has been measured electrochemically.
Electrochemistry has given us power sources such as batteries and fuel cells. Electrochemical
analysis is used in measuring the concentrations of metal ions in drinking water, rivers, lakes,
off-shore water, blood, drinks, food products, and in environmental pollution monitoring.
Electrochemistry is used in the industrial production of many important metals e.g. aluminium
and magnesium, used in aircraft fuselages, laptop and camera cases, and thin-shell buildings
in modern architectural designs.
78
> ? '% # , 78
> ? +!&, ?L # /+, L # +.,
7!8
> ? %.+!. & ?# 7&8
> ? %.+.& ?#
The equilibrium constant Ko is only quoted to two significant figures and a variation in Ko
of ± 0.1×10−5 gives ΔrGo with a precision of ± 14 J mol−1. Results are normally quoted in
multiples of 103 or 10–3 and with the appropriate multiple symbol. The precision and the
value are quoted in brackets with both of them having the same units and multiple as in (4).
78 . -*& 78 -*& 7!8 ., -*&
.
7&8 # ., & 7,8 .
/!, & &&
+!./ 7
8 & ,,!..
Clearly the natural decay rate constant of uranium-238 is is very small (k ≈ 10−10 year−1),
indeed during the lifetime of the earth only about half of the uranium-238 originally
present at its formation has been converted to thorium-234 which is itself unstable and
decays further by fission reactions.
4 WEEK 4: EXPERIMENTAL
DATA ANALYSIS
As there is a lot of graph plotting in analysing experimental data, it would be useful to
revise section 1.1.4 before commencing.
Light of a selected frequency ν is directed at a sodium metal target inside a vacuum system.
Electrons are ejected from the sodium photocathode and a variable “stopping” voltage is
applied to turn back all but those electrons with the highest kinetic energy. This is the
maximum kinetic energy of the ejected electrons EKE. The experiment consists of altering the
frequency of the light and measuring the maximum kinetic energy of the ejected electrons,
the photoelectrons.
# '
n" '
G
H
The UV photons must have a minimum frequency for them to eject any photoelectrons
from the photocathode, the UV photon hν is annihilated when it is absorbed. The minimum
frequency corresponds to a minimum energy hv of the photons, this minimum energy is
called the “work function” of the metal Φ (upper case italic Greek phi) the work function
is the solid-state equivalent to the ionization energy of a gas phase atom. Different metals,
for example caesium, sodium, or potassium have different work functions.
" L2 Photoelectric effect
Plot the graph of EKE against the frequency ν either using a spreadsheet or manually (good
practice for exams). Does the data form a straight line plot, i.e. is the maximum energy of the
ejected electrons proportional to the frequency of the light hitting the sodium metal target?
Determine the values of the gradient of the line which is Planck’s constant h. Determine
the work function Φ from the x-axis intercept (not the y-axis intercept).
The volume of one mole of CO2 gas was measured Vm/L mol−1 as the pressure p/atm
(atmospheres) was varied at a constant temperature of 313 K. The former unit of pressure was
the atmosphere 1 atm = 1.01325×105 Pa the standard pressure is 1 bar = 1×105 Pa (N m−2).
Either using a spreadsheet or manually plot: (1) p/atm versus Vm/L mol−1, (2) log(p/atm)
against log(Vm/L mol−1) and (3) p/atm versus (1/Vm)(L mol−1) and comment upon each of
the plots.
The Arrhenius equation was introduced in last week’s questions. Firstly, using the following
experimental data for the rate constants and temperatures of a kinetically first-order reaction
draw a graph of ln(k/s−1) versus (1/T )(103 K). Secondly, determine the values of A and Ea
for this reaction. Thirdly, use the equation of the line from the graph to calculate the rate
constants at 283 K (do not extrapolate the graph).
The rate of a chemical reaction depends upon the rate constant and the concentrations of
the reagents and products each raised to a power, the powers are called the order of reaction
with respect to that compound. For example a particular chemical reaction involving H2
has an experimental rate of the reaction is below.
&
The italic roman “vee” v (from velocity of reaction) is the rate of reaction with units of mol
L−1 s−1 whatever the order of reaction. The k is the rate constant whose units are determined
by the order of reaction (x is the order with respect to H2). Don’t confuse rate of reaction
with rate constant. The following data for the rate of reaction is given in terms of seconds
and pressure in the non-SI unit of torr, pressure is proportional to concentration at a constant
temperature. A torr is a named unit for Evangelista Torricelli who invented the mercury
in glass barometer (symbol Torr) is a convenient unit of pressure when using a mercury
manometer with a ruler calibrated in millimetres as in the undergraduate chemistry lab,
750 Torr = 1 bar to three significant figures.
. &. !&.
. .,
,
&.. !
&+ /
,,
Firstly, plot a linear graph of the experimental data to determine x the order with respect
to [H2]. Secondly, find the rate constant for the reaction at the temperature of the H2 gas
and thirdly, calculate the rate of reaction for a pressure of 100 Torr.
Fig. 4.4 left, has a liquid added to a cylinder which is closed by a piston. The liquid and
the apparatus are thermostated at a constant temperature. Fig. 4.4 right, some of the
liquid has evaporated at equilibrium and pushes the piston out with a pressure p called
the vapour pressure of the liquid at that temperature. The pressure of the vapour above a
volatile liquid or volatile solid varies with temperature approximately as in the Clausius-
Clapeyron equation, see section 1.2.7. The Clausius-Clapeyron equation may be integrated
assuming that the enthalpy of vaporization is constant over a small temperature range (C
is the constant of integration).
* -
# Integrated Clausius-Clapeyron equation
'%
%=! +
.
!$
Fig. 4.6, there is a linear dependence of the maximum kinetic energy of the ejected electron
with the frequency of the light with a threshold frequency for the photoelectric effect.
Although it is better to use a spreadsheet when possible, I am using this solution to explain
the manual process (exam practice). To find the equation of the best-fit straight line y =
mx + c by hand, we need to pick two points on the line and use their x-y coordinates, see
section 1.1.4. Do not use the data points themselves, or else you are wasting your time
drawing a graph!
78
... & .!&,./ ? 78 /...& /
./ ?
7!8 .!&,./ ?
... & 7&8 /
./ ? /... &
7,8 /,./ ? !. & 7
8
,.!.!& ?
/ / /
7%8
&. ? /%,&,. ? 7+8 !,,%. ?
The two points on the line are (1) and (2) which gives two simultaneous equations (3)
and (4) respectively. the third decimal point is imprecise. Subtracting the two simultaneous
equations (3) and (4). Subtracting these two simultaneous equations gives (5) then the
gradient (6) of the best straight line. Adding the two simultaneous equations (7) and also
substituting in the student’s experimental value for the gradient leads to the intercept on
the y-axis (8).
Planck’s constant from the gradient (6) is h = 6.503×10−34 J s whilst the accepted value for
Planck’s constant is h = 6.626×10−34 J s the student measured value for Planck’s constant
has an accuracy of about 2%, so the gradient is reasonably accurate for an undergraduate
teaching experiment with only four data points.
We have the equation of the line (9). The intercept on the x-axis (10) is v0 the value of
x when y = EKE =0 and is calculated in (11). The work function Φ is hv0 (12) and using
the student’s value for Planck’s constant is Φ = 3.557×10−19 J. The accepted value for the
work function of sodium metal is Φ = 3.78×10 −19 J, so the student’s work function has an
accuracy of about 6% and needs some improvement to the apparatus and the experimental
method with more data points over a wider frequency range.
! &
! &
: : :
&! 6&
&! 6&
!&6&
The plot of pressure against molar volume (Fig. 4.7, left) is a curve which the spreadsheet
gives a best fit with a power function of p = 25.66×V −1 this suggests a function of the
form p ∝ 1/V similar to the ideal gas law. However, the p-V plot is not very useful at high
pressures as the plot is nearly vertical and at low pressures it is nearly horizontal.
The data covers a wide range of values which suggest a log-log plot might be useful (Fig. 4.7,
centre) this allows the data to be viewed more clearly and it also allows any “rogue” points
in the experimental results to be noted for remeasuring and perhaps also for interpolating
between data points. However, the log-log plot does not reflect any physical model for the
p-V data.
As suggested by the power function p = 25.66×V −1 that we noted earlier, we have plotted
the pressure against the reciprocal of the volume (Fig. 4.7, right). This linear plot means
that p×V equals a constant which agrees with the ideal gas law pV = nRT. Volumes are in
litres and the pressures are in atmospheres and the gas constant in these units is equal to
R = 0.0820573 L atm K−1 mol−1 and at T = 313 K the gradient nRT should equal 25.7 L
atm. The gradient of our plot is 25.2 L atm the difference is due to the slight scatter in the
data. This experimental data fits the model of the ideal gas law over this range of p-V-T.
Revise section 1.1.4 where we saw why the reciprocal 1/T is tabulated and plotted as (1/T)
(103 K). The rate constant has to be converted to a natural log. Logs are only for pure
numbers not physical quantities. For example, k/(10−3 s−1) = 5.26 must be converted to
ln(k/s−1) = ln(5.26×10−3) = −5.25 note that any multiple must be incorporated before taking
a log (or before taking an antilog) because the multiple is a part of the pure number.
In Fig. 4.8 the Arrhenius plot is a straight line. The best-fit straight line is y = –9.978x +
28.70. The Arrhenius equation has a gradient m = –Ea /R where R is the gas constant and
Ea is the activation energy. So from the gradient we obtain the activation energy Ea we must
also include the units and their multiples of the axes in the calculation, but as logs have to
be pure numbers, so the units s−1 cannot be separated in the ln(k/s−1) term.
The intercept in Fig. 4.8 is c = 28.70 which for the Arrhenius equation is equal to c = ln(A).
The intercept is on the y-axis and so the intercept will have the same units and multiples
as the y-axis. Many first-order reactions have pre-exponential factors around 1012-1013 s−1.
To calculate the rate constant at T = 283 K then (1/T) = (3.534×10−3 K−1) and so x = (3.534×10−3
K−1), the equation for the best fit straight line is y = –9.978x + 28.70. Substitute x = (3.534×10−3
This e-book
is made with S E TA S I G N
SetaPDF
www.setasign.com
Download free eBooks at bookboon.com
Click on the ad to read more
88
Click on the ad to read more
INTRODUCTORY MATHS FOR CHEMISTS:
CHEMISTRY MATHS 1 Week 4: Experimental Data Analysis
K−1) into the equation for the best straight line in (7) which is evaluated in (8). Take antilogs of
both sides in (9). We may now move the units across to the right hand side (10).
7%8 # & //%+.! L!,!&.! L +%. 7+8 #&
,
!
7/8 & -*
,
&!. 7.8 & &!.!
78 & 78 #( #( & 7!8 #( #(& #(
#( #(& #(
From the rate law (1) v = k[H2]x take logs (2) on both sides of the equation. (3) Expand out
the log of the product of two terms as the sum of two log terms. Bring down the exponent
x of the log gives us a linear equation y = c + mx shown above. Don’t confuse the order of
reaction x and the general variable as in x-axis.
. &. !&.
. .,
,
&.. !&..
.. .,.
,.
#(
!%% &
/ ,+,
++ %+!
&.. !
&+ /
,,
#(
. ,. !/ / /
In the log of the rate of reaction we must take any multiplier over to the right hand side
before taking logs as the multiplier is part of the pure number.
The equation of the best straight line is y = 1.0005x − 3.9772. The order of reaction is the
gradient = 1.0005 which is clearly first-order in H2 within the scatter of the limited data.
The rate constant at the temperature of the measurements is equal to the intercept c =
−3.9772. The y-axis is a log of a pure number so for the first-order rate constants c = log(k/
s−1) = −3.9772. Taking antilogs of both sides k/s−1 = 10−3.9772 = 1.0534×10−4 and moving the
units across k = 1.0534×10−4 s−1.
78 #(
...,#( !.%%
78 #(
..., !.%% /%
7!8
./%
.,
!.
7&8
.,
!.
To calculate the rate of reaction for a p = 100 Torr we start with the equation of the best
straight line (1). In (2) substitute log(p/Torr) = log(100) = 2 and calculate the log(rate).
Take antilogs in (3) and then (4) move the units to the right hand side.
In order to have a linear graph we covert the temperatures from degrees Celsius to kelvin
and take the reciprocal (1/T)(103 K) and the pressures in Torr to pascals then take the
natural-log ln(p/Pa). Fig. 4.10 has a best-fit straight line of y = –4.0464x + 23.4428 but
the axes cross a long way from the origin.
!H
!%
# " # "
78 !
&.&
& 78 &.&
&. ! L
% . L %
-
7!8 * &.&
&.! L 7&8
* - &.&
&. ! L+!&, ?# L
'
7,8
* - !!
&&. & ?# 7
8
* - !!
& ?#
The enthalpy of vaporization is calculated from the gradient (1) of the best straight line. The
multiplier and units for the reciprocal temperature are moved across in (2) but the log term
cannot be separated as logs have to be a pure number. The Clausius-Clapeyron equation
has a gradient (3) of m = −ΔvapH/R and in (4) negative signs cancel and the gas constant is
moved over to the right. The student’s measured value for the enthalpy of vaporization of
hexane (6) is higher by about 8% than the accepted literature value of 31.1 kJ mol−1 (Acree
& Chickos, 2017) and possible improvements to the experimental method such as many
more data points and also possible improvements to the apparatus should be considered.
The Wake
the only emission we want to leave behind
.QYURGGF'PIKPGU/GFKWOURGGF'PIKPGU6WTDQEJCTIGTU2TQRGNNGTU2TQRWNUKQP2CEMCIGU2TKOG5GTX
6JGFGUKIPQHGEQHTKGPFN[OCTKPGRQYGTCPFRTQRWNUKQPUQNWVKQPUKUETWEKCNHQT/#0&KGUGN6WTDQ
2QYGTEQORGVGPEKGUCTGQHHGTGFYKVJVJGYQTNFoUNCTIGUVGPIKPGRTQITCOOGsJCXKPIQWVRWVUURCPPKPI
HTQOVQM9RGTGPIKPG)GVWRHTQPV
(KPFQWVOQTGCVYYYOCPFKGUGNVWTDQEQO
L
A function y = f(x) has tan(θ) = Δy/Δx between the points 1 and 2. As point 2 approaches
point 1 the tan(θ) increases as Δy and Δx change as shown by the blue lines. We reach
in the limit the differential (or derivative) dx/dy at point x1 which may also be written as
f ′(x1) and tan(θ) approaches the slope of the tangent to the function at x1 the black line.
o # Definition: differentiation of a function at x1
j .
The function used in Fig. 5.1 is y = 0.5 + 5x − 2x2 so dy/dx = 5 − 4x and the derivative
at x1 = 0.2 is equal to dy/dx = 5 − 4×0.2 and the derivative of our function at x1 = 0.2 is
f ′(0.2) = 4.2.
Fig. 5.2 shows from the left the two π-bonds at right angles to one another and the σ-bond
which constitute the N2 triple bond. Let us start by considering a simple maths situation
that perhaps familiar from school or college. Consider a N2 molecule at 298 K moving
with a typical velocity v = 515 m s–1. The velocity at any point in time is the derivative of
the distance-time curve (d-t graph). Note the difference between the differential operators
d in roman upright text and the variable symbol in italic d for distance, operators operate
on variables such as d or t.
Velocity is the differential of distance with time
!&
!
&$
$
Calculate the time for the nitrogen molecule to travel a distance of 10 km in a straight line
with no collisions (corresponds to an extremely low pressure as in the earth’s outer atmosphere).
Acceleration is the differential of the velocity with time
!&$
!
&$
!$
The N2 molecule collides with another molecule in a collision lasting 100 ps (picoseconds
where 1 ps = 10–12 s). As a result of the collision the velocity of our N2 molecule increases
from 515 to 630 m s–1. What is the acceleration of the molecule if it is still travelling in
the same direction?
The unified atomic mass unit (symbol u) also called a dalton (symbol Da) is u = 1.6605×10−27
kg to five significant figures, note it is in kilograms not grams and is in roman type not
italic. The most common isotopic form of nitrogen is 14N2 which has a mass of 2×14.0031
u = 28.0062 u. The units of force comes from mass times acceleration and are kg m s−2
it is a named unit called the newton with a lower case “n” (symbol N) whereas Sir Isaac’s
surname is Newton. Calculate the force in newtons which was required to bring about the
acceleration of section 5.1.2.
The N2 molecule now travelling at 630 m s–1 has a collision with the walls of its containing
vessel which again lasts for 100 ps. As a result of the wall collision the N2 molecule is now
travels at 630 m s–1 in the opposite direction, assume the wall of the container has “infinite” mass
compared with the N2 molecule. What has been the N2 molecule’s acceleration in this collision?
At normal pressures of about 1 bar and ambient temperatures of around 298 K the time
between a gas phase molecule colliding is approximately 10−10-10−9 s or 0.1-1 ns (10−9 s
is a nanosecond, symbol ns). Each of these collisions alters the speed and direction of the
molecules. For a gas consisting of identical molecules the fraction f(c) of them with a speed
c (ignoring the direction) is given by the Maxwell-Boltzmann distribution of speeds.
Where m is the molecular mass (the mass of one molecule in kg), kB is Boltzmann’s constant
(the gas constant per molecule) kB = R/NA = 1.38065×10–23 J K–1. The subscript “B” is to
remove any confusion with the symbol for rate constant or force constant. T the temperature
of the gas in kelvin.
-*(
For simplicity and in order to curve sketch let’s arbitrarily assume the constants
A = B = 1. Firstly, sketch the curves of c2 then exp(−c2) and finally the product c2×exp(−c2)
against c using a spreadsheet (curve sketching a graph means show the basic shape of it).
Secondly, differentiate the equation with respect to c and find the maximum of the graph
which is called the most probable speed cmp of the molecules.
In a chemical reaction the concentration of a reagent [A] is found to decrease with time t
as it is consumed to form products. A student measured the data below in the teaching lab.
3
Definition: rate of reaction for a single reactant species A.
The minus sign is to convert the negative slope (it is a decrease in reactant with time) into
a positive rate of reaction. Firstly, draw a graph of [A] against time, either by hand or using
a spreadsheet. Secondly, draw tangents to the graph and find values for the rate of reaction
at t = 0, 2.5×10–3, 5.0×10–3, and 7.5×10–3 s. What conclusion can you draw from these
rates of reaction as a function of time?
Technical training on
WHAT you need, WHEN you need it
At IDC Technologies we can tailor our technical and engineering
training workshops to suit your needs. We have extensive OIL & GAS
experience in training technical and engineering staff and ENGINEERING
have trained people in organisations such as General
ELECTRONICS
Motors, Shell, Siemens, BHP and Honeywell to name a few.
Our onsite training is cost effective, convenient and completely AUTOMATION &
customisable to the technical and engineering areas you want PROCESS CONTROL
covered. Our workshops are all comprehensive hands-on learning
experiences with ample time given to practical sessions and MECHANICAL
demonstrations. We communicate well to ensure that workshop content ENGINEERING
and timing match the knowledge, skills, and abilities of the participants.
INDUSTRIAL
We run onsite training all year round and hold the workshops on DATA COMMS
your premises or a venue of your choice for your convenience.
ELECTRICAL
For a no obligation proposal, contact us today POWER
at training@idc-online.com or visit our website
for more information: www.idc-online.com/onsite/
. &
78 78 7!8
7&8 /&
,,
For uniform motion in a straight line the differential dd/dt is constant with time, that is the
velocity is constant so v = d/t as shown in (1) above which is rearranged in (2) to make time
the subject. Substitute the physical quantities for the variables (3), it is advisable to carry
out calculations in the base units (m not km) so that the units may cancel out correctly.
(4) Notice how in the absence of any collision a molecules can travel a long distance very
quickly, 10 km in 19 s. The absence of collisions corresponds to very low pressures as in
the earth’s atmosphere above about 700 km (exosphere) where most satellites are in orbit.
Of course at ambient pressures the molecules are rapidly colliding with one another and
changing their directions as well as their speeds but nevertheless the average speed of N2 at
room temperature is still around 515 m s–1.
The acceleration in this 100 ps collision of molecules is calculated below. Note when
calculating a change in a variable by convention it is defined as final minus initial value
of the variable. Molecules are capable of very rapid acceleration and deceleration by means
of collisions.
As expected it only takes a minute force to accelerate the N2 molecule because of its small
mass. How large is a newton of force? An average-sized apple (about 100 g) in the earth’s
gravitational field exerts about one newton of force, which we measure as the apple’s weight.
A person of 80.7 kg weight exerts a gravitational force of 791 N.
The final velocity is in the opposite direction to the initial velocity, note the difference
between the operation of subtraction and a negative quantity. Collisions with the vessel walls
as well as molecule-molecule collisions cause molecules to undergo very rapid acceleration
or deceleration.
Cells A2 to A6 were manually typed into the spreadsheet. A2 to A6 were then selected with
the mouse and the bottom-right handle of the selection was dragged down the A column
to dynamically expand the selection down to c = 3.2. Then the cells B2 = A2^2, C2 =
EXP(-B2), and D2 = B2×C2 where manually typed. Then B2, C2, and D2 were selected
with the mouse and the bottom-right handle of the selection was dragged down the D
column to dynamically expand the selection down to the row with c = 3.2.
Fig. 5.5 shows the shapes of the two constituent curves in blue and green and the product
of these two terms is the shape of the Maxwell-Boltzmann speed distribution of molecules
in red. Remember for simplicity and in order to curve sketch we have arbitrarily assumed
the constants A = B = 1. The Maxwell-Boltzmann speed distribution is the product of two
terms in c that is f(c) = Ac2exp(−Bc2). The c2 term (green) increase from zero upwards with
increasing c. The negative exponential term (blue) exp(−Bc2) decrease from 1.0 downwards
asymptotically towards zero with increasing c. So the f(c) = Ac2exp(−Bc2) being the product
of these two terms in red will start from zero at c = 0, then drop back to zero at large values
of c, and at intermediate values of c the fraction will pass through a maximum.
,
(
,
To find the most probable speed, the Maxwell-Boltzmann equation must be differentiated,
and df(c)/dc and then set equal to zero. As the Maxwell-Boltzmann equation is the product
of two terms involving c we must use the “product rule for differentiation”.
www.simcorp.com
The product rule for differentiation
Let’s substitute u = Ac2 and v = exp(–Bc2) so that f (c) = uv and use the product rule for the
differentiation of a product of two functions.
-*( ( -*(
The dv/dc term has come from using the general form for the differential of an exponential
term using the chain rule of differentiation.
Chain rule for differential of a function y =f (n) where n = f (x)
The chain rule applied to y = en where n = f (x)
(1) Differentiate the Maxwell-Boltzmann speed distribution using the product rule for
differentiation and the chain rule for derivative of an exponential. (2) We take the common
factor exp(−Bc2) outside the bracket and at the maximum the derivative must be equal to zero.
78 (-* ( -* (
78 ( -*( .
7!8 * ( * * .
*
&5 %
Definition: most probable speed of gas molecule
7"8!&
6
! $
Fig. 5.6 shows the reagent concentration against time graph and the gradients to the curve.
The spreadsheet gradients from Graph 4.4.2 (Johansen 2013) which are negative, however,
rates of chemical reactions are defined as always positive. The gradient has units and a
multiplier for the time, for example, at zero time the spreadsheet gradient is d[A]/mol L−1/
dt/(10−3 s−1) = −0.25133, the rate of disappearance of A is −d[A]/dt = 0.25133×103 mol L−1
s−1. The rate of reaction at zero time is v = 251.33 mol L−1 s−1 which is tabulated above as
a pure number as v/mol L−1 s−1 = 251.33.
!&
6$
!$
Fig. 5.7 plots the rate of reaction against time. The rate of reaction v decreases with time
because the concentration of reactant [A] is decreasing as it is consumed in forming
the product.
The non-bonding interaction between two neutral atoms or two neutral molecules (called
the intermolecular interaction) is given by the Lennard-Jones equation.
Lennard-Jones potential
+ & ! !
Where V is the potential energy of the interacting molecules, r is the distance between the
two molecules’ centres of mass, ε (italic Greek epsilon) is the depth of the potential well
and σ is the value of r for V = 0. The 6th-power term describes the long-range attraction
between the molecules due to the van der Waals or dispersion force plus any dipole-dipole
force and any dipole-induced dipole force, it lowers the potential energy (the term is
negative). The 12th-power term represents the short-range repulsion between the molecules
(the term is positive) from the overlap of their electron orbitals when the distance between
the molecules becomes small, it rapidly increases the potential energy. Fig. 6.1 shows the
MOPAC2016 (PM7) calculated structure for the minimum of the potential energy well
for the interaction between two N2 molecules.
In Fig. 6.1 the surfaces are the van der Waals “size” of the molecules, in this relative position
there is a balance of the repulsive and the attractive forces between the molecules. Use a
spreadsheet to plot the following three graphs: firstly, the attractive long range potential
Vat = −4ε(σ/r)6, secondly, the short range repulsive potential Vre = 4ε(σ/r)12, and thirdly,
the combined Lennard-Jones potential VLJ for the interaction between two N2 molecules.
The N2 parameters are ε = 763.6 J mol–1 and σ = 3.919×10–10 m (3.919 Å) and plot for
intermolecular distances from 3.8 Å to 10 Å in steps of 0.01 Å.
+ & ! !
For a given molecular interaction the variables in the Lennard-Jones equation are V and
r. Calculate the derivatives dV/dr then differentiate the first derivative with respect to r to
find the second derivative d2V/dr2 for the intermolecular potential between N2 molecules.
Use the solutions of Question 2 to find the separation at which the Lennard-Jones potential
V is at a minimum, that is the distance between two N2 molecules at which the potential
energy is a minimum.
Firstly, show that d(pH)/d[H+], the small change in pH arising from a small change in [H+]
is proportional to the ratio d[H+]/[H+]. Secondly, a solution originally with [H+] = 0.02
mol L–1 had an additional concentration of d[H+] = 5×10–4 mol L–1 added to it, calculate
the change in pH, d(pH).
The rate constant k for a chemical reaction is related to the temperature T by the Arrhenius
equation. Where the pre-exponential factor A and activation energy Ea are constants for a
given chemical reaction, R is the gas constant. A particular biochemical reaction is second-
order, first let’s find the units for the second-order rate constant. The units for any rate of
reaction v are concentration per time, mol L−1 s−1 and all concentrations are mol L−1. The
dimensional analysis below gives the units of k as L mol−1 s−1 for a second-order reaction.
78
&
P3QP5Q 78
&
P3QP5Q
�e Graduate Programme
I joined MITAS because for Engineers and Geoscientists
I wanted real responsibili� www.discovermitas.com
Maersk.com/Mitas
I joined MITAS because fo
I wanted real responsibili�
Month 16
I was a construction
supervisor ina co
I was
the North Sea su
advising and the
Real work he
helping foremen ad
International
al opportunities
Internationa
�ree wo
work
or placements ssolve problems
Real work he
helpin
International
Internationaal opportunities
�ree wo
work
or placements ssolv
Download free eBooks at bookboon.com
Click on the ad to read more
104
Click on the ad to read more
INTRODUCTORY MATHS FOR CHEMISTS:
CHEMISTRY MATHS 1 Week 6: Chemistry and Differentiation 2
Consecutive reactions are where two or more chemical reactions follow on from one another.
Consecutive reactions are very common in all branches of chemistry and biochemistry. The
simplest type is where both reactions are kinetically first-order.
5
3
The molecule A is the reactant, C the product, and B is called an intermediate. If k1 and k2
are first-order rate constants and assuming that we start at time zero with pure A and no
B or C present then this mechanism gives the concentration of the intermediate at a later
time t as follows. Where [A]0 is the initial concentration of A. Find the time tmax at which
[B] is a maximum [B]max.
&
5 3.
& &
-*& -*&
For a first-order reaction the reactant concentration varies with time as [A] = [A]0 exp(−k1t)
at a constant temperature.
1) Construct a spreadsheet of the following data, [B], t, and d[B]/dt (from section
6.2.6). The time should run from zero to one second in intervals of 0.1 s.
2) Assuming that [A]0 = 1 mol L–1, k1 = 7.0 s–1 and k2 = 5.0 s–1 draw a graph of [A],
[B], and [C] versus time.
3) Draw a graph of d[B]/dt versus t and find tmax on your graph.
4) Find the value of [B]max From your graph.
For two consecutive first-order chemical reactions A → B → C consider each of the three
terms in the equation below for the intermediate concentration [B] and show that if k1 is
much smaller than k2 (k1 « k2) then [B] will be a small concentration and also d[B]/dt will be
relatively constant in absolute terms compared with the reactant and product concentrations.
&
5 3.
& &
-*& -*&
&
5 3.
& &
-*& -*&
By consideration of each of the terms of the equation find approximate expressions for
[B] and d[B]/dt when k1 is very much larger than k2 (k1 » k2), that is, the reverse of the
situation in section 6.1.8.
+ & ! !
!@ 6+!+&
!+&
!+&
:
:
:
:
The spreadsheet (Table 6.1) has the intermolecular distance varying from 3.8 Å to 10 Å
in steps of 0.01 Å. The distances A1 to A5 are manually typed into the spreadsheet. Cells
A1 to A5 are selected with the mouse and the right-bottom handle in dragged down the A
column until 10 Å is reached and the mouse button is released. The formulae B2 to D2 are
typed manually as B2 = 4*763.6*(3.919/A2)^12 − (3.919/A2)^6, C2 = −4*763.6*(3.919/
Fig. 6.2 shows the repulsive part of the potential in black, the attractive part of the potential
in blue, the combined Lennard-Jones potential is the sum of these two terms is shown in
red. Note that both σ = 3.919 and the intermolecular distance r are in angstroms hence
the units cancel, the potential energy is in J mol−1. The two parameters give the position
where the potential energy is zero σ = 3.919 Å, and from the spreadsheet the potential
energy well depth is ε = 763.6 J mol–1 at an intermolecular distance of rmin = 4.40 Å to a
precision of 0.01 Å.
! + &
6+
@
&
www.skoda-career.com
78 + & ! ! 78 + & ! !
7!8
+
& ! !
!
% 7&8
+
&
!
!
!
%
7,8
+
& ,
! &
& !
+
7
8
+
& ,
! & !
&
+
Starting with the Lennard-Jones potential (1) write the denominators as numerators (2),
the first derivative gives (3) is converted in (4) to the r−13 and r−7 terms as denominators.
A similar approach is used to calculate the second derivative. From the first derivative of
the Lennard-Jones potential (3), the second derivative is (5) and is rewritten with the r−14
and r−8 as denominators in (6).
!
78
+
&
!
!
! % .
78
!
!
!
%
7!8
%
!
!
7&8
!
7,8 !//
7
8 &!/+/ d
# *
78 * 78
!.
!.
,.& #'
7!8 * 7&8 * ..
!.
!.
.. #'
In order to differentiate the log10 equation, pH = −log[H+] we must first convert it to natural
logs, (see section 3.1.4) using the conversion 2.3026 = ln(x)/log(x) which gives (1). In (2)
we can now differentiate the pH with respect to hydrogen ion concentration. We separate
the variables of the derivative in (3) and take d[H+] over to the right. Substitute of the
values for the variables in (4). As expected increasing the hydrogen ion concentration leads
to a decrease in pH.
In the Arrhenius equation (1) we know the activation energy and we must first calculate the
pre-exponential factor A. Substitute the values for the variables in their base units in (2) and
cancel units where possible to obtain (3) which gives us the pre-exponential factor in (4).
To differentiate the Arrhenius equation let x = 1/T as in (5). The differential of an exponential
(6) is the exponential term multiplied by any constant factor within the exponential term.
In (7) substitute x with 1/T. We use the chain rule of differentiation (8) to convert dk/dx
to dk/dT as x = T−1 then dx/dT = −T −2 = −1/T 2 to give us (9). We clean-up the equation
(9) to obtain dk/dT in (10).
78 & %
'%
"
-*
"
'%
78 & L
.,!.. '# ,.... ?#
+!&, ?L # !.. L
-*
,.... ?#
+!&, ?# L !.. L
7!8 & !!
& '#
We separate the variables of the derivative in (10) and take dT to the right hand side
to get (11). In (12) we now substitute the values for A and Ea along with dT = 1 K.
Equation (13) shows the change in the rate constant due to the change in temperature.
As the temperature of 300 K is increased to 301 K the rate constant has increased from
20 L mol−1 s−1 to 21.3364 L mol−1 s−1 which is about a 6% increase. Those chemical and
biochemical reactions which have activation energies around 50 kJ mol−1 are sensitive to
very small ambient temperature changes.
&
78 5 3.
& &
-*& -*&
5 &
78
3.
& &
& -*& & -*&
7!8 & -*& - & -*& - 7&8 # & -*& - # & -*& -
7,8 # & & - # & & - 7
8 #& #& & - & -
& &
7%8 # - & & 7+8 - #
& & & &
%
7/8 - # 7.8 - .
+&
% , ,
Take the natural logs of both sides of (3) gives (4). From (4) we use the rule of logs that
the log of a product is equal to the sum of two logs, and also ln[exp(x)] = x gives (5). (6)
We rearrange (5) so that the log terms are on the left in (6) and the non-log terms are on
the right. The subtraction of two log terms in (6) equals the log of the ratio of the terms
(7). Rearrange (7) to (8), the time at which the intermediate is a maximum concentration
depends only on the values of the two first-order rate constants, and is independent of the
initial concentration of A. Substituting the values for the rate constants in (9) gives the
time at which the concentration of the intermediate is a maximum (10) as 0.16824 s (the
“excess” significant figures are to lessen rounding errors for section 6.2.7).
It is always worth drawing graphs to be able to visualize how the algebraic equations (which
may be a bit “abstract”) are modelling the chemical reactions. Also using a spreadsheet to plot
the curve allows us to use non-linear curve fitting (Parker 2013, p. 33) to find the maximum
of a function or where a function crosses an axis. To calculate [C] the concentration of the
product we use the conservation of matter, matter is not destroyed or created in a chemical
reaction, with an initial concentration [A]0 = 1 mol L−1 = [A] + [B] + [C]. Section 6.2.6
has the equation for d[B]/dt and substituting in the values for the variables gives d[B]/dt
in Table 6.2.
!$
7"8!&
6
7*8!&
6
728!&
6
7*8!
!&
6$ : : : : : : : : :
7"8
7*8
728
728
7"8
7*8
!$
Fig. 6.3, the reactant concentration [A] is an exponential decrease with time and initially
the intermediate concentration [B] increases. As the [B] increases the rate of forming the
product [C] increases rapidly but then starts to flatten off as the [B] concentration decreases
as it is being consumed and converted to [C].
Fig. 6.4 plots [B] (on a different scale to Fig. 6.3) and d[B]/dt against time. When d[B]/dt
equals zero is the time tmax = 0.17 s at the maximum concentration of [B]max = 0.43 these
values agree with Section 6.2.6 (the non-linear curve fitted values are tmax = 0.16824 s and
[B]max = 0.4312 mol L−1).
7*8!
7*8
7*8!
7*8& ,
& , 7*8
:
&
78 5 3. -*& -*&
& &
%.
78 5- #' -*%. .
+& -*,. .
+&
,. %.
7!8 5- !, #'.!.+. .&! .&! #'
careers.slb.com/recentgraduates
Calculating [B]max from the equation for the reaction intermediate (1) we substitute
tmax = 0.16824 s along with [A]0 = 1 mol L–1, k1 = 7.0 s–1, and k2 = 5.0 s–1 gives (2) and
(3) and [B]max = 0.4312 mol L–1.
,
,
4`
`B ,4
The maths involve negative exponentials and it is worth graphing the function y = exp(−x)
again to show its characteristics. The negative exponential has two asymptotes. An asymptote
is a straight line that is the limiting value of a curve, it can be considered as the tangent at
infinity. The first asymptote is as x tends to plus infinity the negative exponential tends to
zero. Maths symbol for “tends to the limiting value” is an arrow. The second asymptote is
as x tends to minus infinity then the negative exponential tends to plus infinity. A negative
exponential crosses the y-axis at y = 1, the y-intercept has Cartesian coordinates of (0,1).
If k1 is much smaller than k2 (k1«k2) then the following approximations may be made. As
k2 is dominantly large then exp(−k2t) ≈ 0 and as k1 is small and positive then exp(−k1t) ≈
exp(0) ≈ 1. Finally (k2 − k1) ≈ k2.
3. & 3. &
78 5 -*& -*& 78 5
& & &
5 3. & 5 3. &
7!8
& -*& & -*&
7&8
& & &
Inserting these three approximations exp(−k2t) ≈ 0, exp(−k1t) ≈ 1, and (k2−k1) ≈ k2 into the
intermediate concentration (1) gives us (2) the approximate expression for [B] when k1«k2.
As k1/k2 is a small number the intermediate concentration [B] is small compared to [A]0.
We use the same three approximations exp(−k2t) ≈ 0, exp(−k1t) ≈ 1, and (k2−k1) ≈ k2 and
substitute them into the rate of formation of the intermediate (3) gives us (4) the approximate
expression for d[B]/dt when k1 « k2. As k1/k2 is a small number then k12/k2 is a smaller
number and thus the rate d[B]/dt is very small, [B] is relatively constant compared with
the initial concentration of [A]0. Thus the intermediate concentration [B] is both small and
relatively constant compared with [A]0 under the approximations where k1 « k2 and this is
called the “steady state approximation” in chemical kinetics. The steady state approximation
corresponds to where there is a relatively slow reaction k1 forming the intermediate followed
by a fast reaction k2 of the intermediate to form the product, Fig. 6.6.
728
7"8
7*8
728
7"8
7*8
In Fig. 6.7 the concentrations are plotted against k1t which is a dimensionless variable
for first-order reactions. The steady state approximation does not assume the reaction
intermediate concentration to be constant nor that its time derivative is zero. The steady
state approximation assumes that the concentration of the intermediate is near-constant over
the middle part of the reaction time frame after the initial “induction period” (where [B]
increases from zero to just past its maximum value). The concentration of the intermediate
is so low that even a large relative variation in its concentration is small if compared on an
absolute quantitative basis against those of the reactant and product concentrations.
When k1 » k2 (the opposite conditions from section 6.2.8) then the following approximations
may be made from Fig. 6.5 for the properties of negative exponential functions. The condition
k1 » k2 means exp(–k1t) ≈ 0 and as k2 is small and positive then exp(–k2t) ≈ exp(0) ≈ 1 and
(k2–k1) ≈ –k1.
3. &
78 5 -*& -*& 78 5 3 .
& &
5 3. & 5
7!8 & -*& & -*& 7&8 3 . &
& &
Inserting these three approximations exp(–k1t) ≈ 0, exp(–k2t) ≈ 1, and (k2–k1) ≈ –k1 into
the intermediate concentration (1) gives us the approximate expression (2) for [B] ≈ [A]0
when k1 » k2. The concentration of the intermediate is of the same order of magnitude as
the initial concentration of the reactant.
We use the same approximations exp(–k1t) ≈ 0, exp(–k2t) ≈ 1, and (k2–k1) ≈ –k1 and
substitute into the rate of forming the intermediate d[B]/dt (3) which gives (4). From (4),
after the induction period when [B] is a maximum, for the rest of the reaction time the
rate of loss of the intermediate to form the product is approximately −d[B]/dt ≈ [A]0k2. But
from (2) this is the same as −d[B]/dt ≈ [B]k2 that is the rate of formation of the product
is approximately the rate of the slow second step B→C of the consecutive reactions and
is approximately independent of the rate of the fast first step A→B. This second step is a
reaction “bottle neck” and is called the “rate determining step”, Fig. 6.8.
Figure 6.9: [A], [B], and [C] versus k1t for k1 = 10k2.
Fig. 6.9 the relatively large value of k1 causes the [A] reactant concentration to decrease
rapidly and a large concentration of the intermediate [B] to be is quickly formed. The
product concentration [C] does not start to appreciably increase until the intermediate
concentration [B] has passed its maximum by which time the reactant concentration [A]
is negligible. That is the formation of the product C is approximately determined by the
large intermediate concentration [B] with a B→C rate determining step.
7 WEEK 7: CHEMISTRY
AND INTEGRATION 1
This introduction to integration is spread over both weeks 7 and 8. The tutorials questions
and solutions will show some of the more important chemical applications of the maths
technique of integration. One very common maths technique used in chemistry is the
“separation of variables of a derivative” method for integrating a differential equation.
The decay of a radioactive element A to a product P, A→P, has a rate constant k. Radioactive
decay is a first-order chemical reaction with [A] and t being the variables of the equation.
3
& 3 Rate of reaction for radioactive decay
The rate of reaction is equal to the product of the rate constant k and the concentration [A]
of the reactant at that particular time. The negative sign is to give a positive rate constant for
the reaction which involves the loss of the reactant. In order to calculate the concentration
of A remaining at time t we need to integrate this equation from t = 0 at which [A] = [A]0
to a general time t = t where [A] = [A]t. Carry out this definite integration for this first-
order chemical radioactive decay reaction.
We have looked at chemical reactions with first-order reaction rate laws, another common
type is a second-order rate law. This may occur in two forms depending upon whether there
are one or two species of reactant molecule.
12 3 1 2
P3Q P3Q
78 & 3 78
& P3QP5Q Definition: second-order rate laws
NO2 (brown gas) is involved in important chemical reactions in polluted atmospheres and
also in the industrial manufacture of nitric acid and hence nitrates. The rate of loss of NO2
is expressed by the second-order rate law below.
KG
& KG
The fact that the reaction is second-order and the stoichiometric coefficient of NO2 is also
2 is purely a coincidence and does not imply anything about the mechanism of reaction. To
calculate the concentration of NO2 at any time during this reaction we need to integrate this
INTRODUCTORY MATHS
d [NOFOR CHEMISTS:
2] 2
CHEMISTRY MATHS
− 1 = k [NO 2 ] Week 7: Chemistry and Integration 1
dt
equation from
The fact that tthe
=0 whereis[NO
reaction ] = [NOand
second-order
2
] to a general time t = t when [NO ] = [NO2]t. a
2 0 the stoichiometric coefficient of NO 2 is2 also 2 is purely
Carry out this definite
coincidence and doesintegration
not imply for this second-order
anything chemical ofreaction.
about the mechanism reaction. To calculate the
concentration of NO2 at any time during this reaction we need to integrate this equation from t = 0
where [NO2] = [NO2]0 to a general time t = t when [NO2] =Jump
[NO2]to
t. Carry out this definite integration
Solution 2 (see page 122)
for this second-order chemical reaction.
7.1.3 QUESTION
7.1.3 3: EXPANDING
QUESTION GAS AND
3: EXPANDING THERMODYNAMIC
GAS WORK
AND THERMODYNAMIC WORK
In chemistry it is very common to have gases involved in reactions, they may be formed,
In chemistry it is very common to have gases involved in reactions, they may be formed, or react, or
or react, or be compressed, or expand. A simple way of understanding such processes is to
be compressed, or expand. A simple way of understanding such processes is to begin with a
begin with a “mechanical” picture of a gas trapped inside a cylinder by a movable piston.
“mechanical” picture of a gas trapped inside a cylinder by a movable piston.
, pex V
In thermodynamics
In thermodynamicsthe thematerial
materialweweareareinvestigating
investigating (gas
(gas in
in this
this case)
case) is
is called the “system”
called the “system” and
and everything
everythingother
other
thanthan the system
the system is the
is called called the “surroundings”.
“surroundings”. The systemTheplussystem plus the is
the surroundings
surroundings is called the
called the “universe”. “universe”.
If the If the external
external pressure pressure ofis the
of the surroundings surroundings
pex and is “gas
the gas in our pex and
piston”
the is
gasalways
in ourkept
“gasatpiston”
the same is pressure of the
always kept at external
the same surroundings
pressure ofbythesupplying
externalheat as the volume
surroundings
increases it is called a “reversible expansion”. If the initial volume of the system is Vi the pressure-
by supplying heat as the volume increases it is called a “reversible expansion”. If the initial
volume work (p-V) done on the gas (w) in expanding the gas to a final volume Vf is given by the
volume of the system is Vi the pressure-volume work (p-V) done on the gas (w) in expanding
integral below and Fig. 7.3 (left).
the gas to a final volume Vf is given by the integral below and Fig. 7.3 (left).
Vf
w = −∫ p ex d V Work doneononthethe
Work done system
system by aby a reversible
reversible expansion
expansion
V
i
p , p p , p
ex ex
V V
Vi Vf Vf Vi
Figure 7.3: reversible expansion of a gas (left),
and reversible compression (right).
In Fig. 7.3 (left) starting at pex and Vi we heat the gas to keep the pressure constant and
increase the volume to Vf the gas follows the red line and the work done on the gas is equal
86
to minus the area under the p-V curve, called the line integral of the p-V plot. The sign
convention is that work done on the gas by expansion the gas is negative, the system (gas)
has done work in pushing the surroundings backwards out of the way. The work done on
the system (gas) in compression Fig. 7.3 (right) is plus the area under the line, the system
In Fig. 7.3 (left) starting at pex and Vi we heat the gas to keep the pressure constant and increase the
follows the red line and has decreased its volume by the surroundings pushing the gas into
volume to Vf the gas follows the red line and the work done on the gas is equal to minus the area under
a p-V
the smaller volume
curve, andline
called the theintegral
gas losing heat
of the p-Vin order
plot. Thetosign
maintain constant
convention is thatpressure.
work done on the
gas by expansion the gas is negative, the system (gas) has done work in pushing the surroundings
Assumingoutthe
backwards gas way.
of the is anThe
ideal gas,done
work calculate the work
on the system (gas)done when the Fig.
in compression gas 7.3
expands
(right)from
is plusan
initial
the area volume
under the to a the
Vi line, finalsystem
volume Vf by the
follows integrating
red line the
and p-V
has work equation.
decreased its volume by the
surroundings pushing the gas into a smaller volume and the gas losing heat in order to maintain
constant pressure. Jump to Solution 3 (see page 123)
Assuming the gas is an ideal gas, calculate the work done when the gas expands from an initial
volume Vi to a final volume Vf by integrating the p-V work equation.
7.1.4 QUESTION 4: IDEAL GAS EXPANSION
Jump to Solution 3 (see page 89)
We have one mole of ideal gas at an initial volume of 4.90×103 cm3 which expands to a
final volume
7.1.4 of 2.45×10
QUESTION
4
cm3 at aGAS
4: IDEAL constant temperature T = 298 K in a thermostated bath.
EXPANSION
Using your solution from section 7.2.3 calculate the work done on the gas in this expansion.
We have one mole of ideal gas at an initial volume of 4.90×103 cm3 which expands to a final volume
Jump bath.
of 2.45×104 cm3 at a constant temperature T = 298 K in a thermostated to Solution 4 (see
Using your pagefrom
solution 123)
section 7.2.3 calculate the work done on the gas in this expansion.
7.2.1 SOLUTION
7.2.1 SOLUTION1:1:
RADIOACTIVE DECAY
RADIOACTIVE DECAY
Where
Where C isCthe
is constant
the constant of integration.
of integration. Using
Using this thisI solution
solution am going Itoam goingthe
introduce to “separation
introduce of
the
“separation
the variables ofof the variables
a derivative” of afor
method derivative”
integratingmethod for integrating a function.
a function.
3 3
78 d [A] & 3 78 d [ A] &
(1) = −k [A] (2) 3 = −k d t
dt
3 [A]
[A]t
∫"
3 3
d [A] & " 7&8 # 3 [A]
t
7!8 t & t .
(3) 3 = −k ∫. d t (4) [ ln[ A] + C ] [A]
3
.= −k [ t ] 0
3 [A] 0
0
[A] .
0
(1) Treat the derivative d[A]/dt as two separate variables d[A] and dt each of which follow
(1) Treat the derivative d[A]/dt as two separate variables d[A] and dt each of which follow the normal
the normal rules of algebra. Importantly you cannot separate “d” an operator, from “t” the
rules of algebra. Importantly you cannot separate “d” an operator, from “t” the operand (a variable
operand (a variable which is operated on), as the operator “d” has to operate on its operand.
which is operated on), as the operator “d” has to operate on its operand. Similarly with any other
Similarly withcombination
operator-variable any other operator-variable combination
such as d[A]. Rearrange such
equation (1) as d[A].
to bring Rearrange
everything equation
that involves
(1) to bring everything that involves [A] to the left hand side and everything that involves
t or are constants or a minus sign to the right hand side of equation (2). Integrate both
sides of the equation (3) separately, using the two sets of appropriate integration limits,
87
be careful to check these limits! Take any constants and the negative sign outside of the
integral for clarity, note the rate constant k is independent of time and is constant. Both the
left and right hand sides of the equation are standard integrals. The left hand side making
use of the standard integral of a reciprocal. Evaluating the definite integral (4) will cancel
out the constant of integration. This gives us four equivalent ways of writing the definite
integral, which one we choose to use to solve a particular problem will depend upon the
circumstances.
3
#3 # 3. & # &
3.
Integrated first-order rate law
3
-*& 3 3. -*&
3.
Return to Question 1 (see page 117)
We have integrated first-order rate laws above but what about second-order chemical kinetic
rate laws?
3
3 3
7!8 " 3 3 & "
78 & 3 78 &
3 3 .
.
3 3
3
7&8 & . 7,8 & 7
8 &
3 3 3 3 3.
. .
To make the solution a general one, in (1) let’s replace NO2 by A for the integration. We
use the method of “separation of variables” for the integration of this second order rate law.
Treat “d[A]” and “dt” as separate variables and bring everything that involves [A] to the left
hand side and everything that involves t or is a constant or a minus sign to the right hand
side of (2). Now in (3) integrate both sides separately with the appropriate limits. An easy
and foolproof way of integrating the left hand side of the equation is to write it as [A]−2
and then integrate as normal. You can use this “trick” to integrate many equations which
at first sight might look a bit difficult. In (5) write [A]−1 as 1/[A]. Evaluate the definite
integral (6) and the constant of integration cancels. Multiply (6) by minus throughout to
obtain the integrated second-order rate law with one reactant species as shown below.
& Integrated second-order rate law with one reactant
3 3.
For the specific example of the nitrogen dioxide decomposition this corresponds to the
following.
&
KG KG .
+0
'%
78 + '% 78 7!8 " - +
+ +
+ +
0 0
The ideal gas law (1) gives us the pressure of the gas in (2) which we substitute into the
pressure-volume work equation (3) to obtain (4). Take the constants outside the integral
(5) and integrate using the standard integral of a reciprocal ∫dx/x = ln x to (6) and evaluate
the integral between its two limits (7). Finally write the difference of two log terms as the
log of the ratio of the two terms.
+0
' % # p-V work done on the system in an ideal gas expansion
+
+0 &,. & !
78 '% # 78 #+!&, ?L # /+ L# ! !
+ &/..
7!8 !//. ! ? 7&8 !// ?
Isothermal just means the change takes place at constant temperature. The work done
on the ideal gas is the area under the curved arrow between the two limits of volume in
equation (1). Substituting the values for physical quantities in (2). Quote the work done
on the gas (4) in the conventional multiples of kilojoules −3.99 kJ. The gas is doing work
on the surroundings in the isothermal reversible expansion process as the gas is “pushing
back” the surroundings.
8 WEEK 8: CHEMISTRY
AND INTEGRATION 2
NASA has fitted high quality experimental data for the standard heat capacity at constant
pressure Cpo (also called the isobaric heat capacity), the standard enthalpy Ho, and the
standard entropy So for more than 2000 solid, liquid and gaseous chemical species over an
extremely wide temperature range to three polynomial equations called the NASA polynomials
(McBride et al, 2002). The polynomial equations are shown below in a dimensionless form
where R is the gas constant, and b1 and b2 are constants of integration. The polynomials
each have seven-terms with temperature raised to the powers running from −2 to +4 and
they are designed for easy computer manipulation of the data.
www.foss.dk
% % ! & % , %
% ! % % &
'
- # % % % %! % & NASA polynomial equations
% ! & ,
%
'% % ! & , %
! &
F % % % %
% ! # % & % ,
%
' ! &
The polynomial equation for enthalpy is obtained by integrating Cpo with respect to T
and for entropy by integrating Cpo/T with respect to T. These integrations arise from the
thermodynamic definitions below.
%
-
%
- " % Constant pressure heat capacity and enthalpy
%
%
"
-
F
F % Constant pressure heat capacity and entropy
% %
%
The coefficients a1 to a7 for N2 between 200 K and 1000 K are tabulated by NASA to 10
significant figures for computer manipulation. I have rounded the data to four significant
figures for use with a calculator.
! & ,
%
& ! , /
.. !++.
.+! +,!. !+,. /
. ,..
1) Integrate the heat capacity polynomial equation between 273 K and 373 K (0 oC
to 100 oC). This integrated equation is equal to the enthalpy change ΔHo(373-273
K) of nitrogen.
2) Substitute in the values of the coefficients for N2 into your integrated equation to
find heat capacity change between 273 K and 373 K, (the area under the curve)
which is the enthalpy change ΔHo (373-273 K) of N2 for this constant-pressure
temperature change.
3) Using a spreadsheet plot the graph of Cpo for N2 against T at 20 K intervals between
200 and 1000 K. Calculate the enthalpy change for N2 on heating from 273 to
373 K.
4) Using a spreadsheet plot the graph of Cpo/T against T at 20 K intervals between 200
and 1000 K. Calculate the entropy change for N2 on heating from 273 to 373 K.
A polyene molecule is a linear hydrocarbon chain of alternating single and double bonded
C-atoms. Polyene molecules are important components or moieties of many biological
molecules, e.g. β-carotene and lycopene.
Figure 8.2: C18 linear polyene molecule (left) and β-carotene (right).
Because of the alternation of the single and double bonds, the π-electrons are mobile and
may move along the whole length of the polyene. We may treat the polyene molecule as a
“one-dimensional” box of length L which confines the mobile π-electrons. This interesting
and important quantum mechanics concept will be discussed in your chemistry degree.
Fig. 8.3 shows some of the different wavefunctions for the allowed solutions for the mobile
π-electrons as the quantum number n increases so does the energy of the π-electrons in
that level.
!!
!
"
.
( :
Where the sine is in radians not degrees hence the need for π, x is the distance along the
molecule, n is a constant (a quantum number) which may take any of the values n = 1,
2, 3, ∙∙∙ and B is a constant to be evaluated. Use the standard integral below (where C is a
constant of integration) to integrate the probability equation from x = 0 to x = L and thus
find B in terms of L the length of the molecule. Substitute for B to obtain the complete
equation for a particle in a one-dimensional box.
The radial distribution function P(r) is the probability density that the electron is in a thin
spherical shell which is a distance r from the nucleus and thickness dr. The H-atom 1s
electron radial distribution function is shown below.
3
&
!
. Radial distribution function for a 1s electron of a H-atom
-*
.
www.job.oticon.dk
Where a0 is a constant called the Bohr radius (a0 = 0.5292 Å or a0 = 0.5292×10–10 m). Fig.
8.4 plots the hydrogen atom 1s electron radial distribution function. The maximum in Fig.
8.4 is at the Bohr radius.
!&
Calculate the probability of finding the electron at a radius of less than 2a0, that is integrate
the 1s H-atom radial distribution function from r = 0 to r = 2a0. Note that 4/a03 is a
constant and may be taken outside the integral.
.
"
&
"# -*
!. .
.
We need to integrate the product of two function of r. We will integrate ∫uv dx using
the “integration by parts tabular method” with u = x2 and v = exp(bx) (x = r and b = −2/
a0) and we set up the table below with each row the differential of the term above or the
integral of the term above.
3 5
"
-*
-* Integration by parts, tabular method
-*
. -*
!
We pair the first entry of column A with the second entry of column B; then the second
entry of column A with the third entry of column B, etc. alternating the signs starting with
a positive sign and carry on until the product of the pairing is zero.
The integral from 273 K to 373 K is shown below where ΔHo(373 K − 273 K) is written
as ΔHo for brevity.
!%!L
78
- ' " %
%
! & % ,%
% % % %
! &
%!L
!%!L
78
- ' % # % ! % & % , % ! !
% & & % % , , %!L
Notice that the constant of integration in (2) has cancelled out in the definite integral (3).
Each of the terms in the definite integral (3) has units of kelvin which are not shown for
clarity. We now substitute the values of the NASA polynomial coefficients a1 to a7 for N2.
$ ! %
+ % &
! + % &
=
:=
=
:=:
=:
: =:
=:
Table 8.1 is the top-left corner of the spreadsheet with A2 to A8 the polynomial coefficients
for N2 typed manually. The temperatures B2 to B10 were typed in manually. Cells B2
to B10 were selected with the mouse, and the right-bottom handle was pulled down to
dynamically copy the selection to B42 to give the temperature range 200-1000 K in steps
of 20 K and the selection handle was released. The formulae C3 to E3 were manually
typed, C2 is Cpo/R = (A$2*B2^-2)+(A$3*B2^-1)+A$4+(A$5*B2^1)+(A$6*B2^2)+(A$7*B
2^3)+(A$8*B2^4), D2 is Cpo/(J K−1 mol−1) = 8.314*C2, and E2 is (Cpo/T)/(J K−2 mol−1) =
D2/B2. These three formulae are selected and the right-bottom handle of the selection is
dragged down to dynamically copy down to E42 and then the selection handle is released.
!+%&
!%
Fig. 8.5 is the spreadsheet plot of the heat capacity Cpo against temperature. Integration of
Fig. 8.5 from 273 to 373 K, (the area under the curve) obtained by using a spreadsheet
is ΔHo(373−273 K) = 2.914 kJ mol−1 which agrees to within three figures with the direct
integration. Calculating an area under a curve is best accomplished using Graph 4.4.2
(Johansen, 2013).
!!+%&
In Fig. 8.6 the heat capacity divided by the temperature Cpo/T is plotted against the
temperature T. The area under the curve, the integral, between any two temperatures is
the entropy change between these temperatures ∆So(T2−T1), and between 273 and 373 K
it is ∆So(373−273) = 9.0903 J K−1 mol−1.
The thermodynamic quantity most “easily” experimentally measured is the heat capacity
which is the amount of heat (usually supplied as electrical energy) needed to increase the
temperature by one degree Celsius or one kelvin. Once Cpo is measured using an adiabatic
constant-pressure calorimeter at different temperatures we can graphically find the enthalpy
and the entropy of the system, as above.
( (
: :
. .
78 ( ( .
& &
78 (
:
(
:
(
:
Let us write a = nπ/L and then the integral of the wavefunction squared is B2∫sin2(ax)dx
for a particle in a box. In (1) we use the standard integral B2∫sin2(ax)dx = B2[x/2 − (1/4a)
sin(2ax) + C] between the limits of x equals 0 and L. This definite integral is equal to
unity, the total probability for the mobile π-electrons trapped in the polyene. The term
sin(2nπ) in (1) is in radians not degrees and as n is an integer, sin(2nπ) = 0, (prove this by
using your calculator to find sin(π) in radians) so the lower limit is zero also the second
term of the upper limit is zero. The constant B is called the normalization constant as it
makes the total probability of our π-electron being located in the polyene equal to unity,
we have normalized the probability. We now have the full equation for the wavefunctions
of a π-electron in a given quantum level n of a polyene molecule as shown in Fig. 8.3.
:
: Wavefunctions for a particle in a one-dimensional box
.
& !
78 *# ! -* . . !.
. . .
!
& ! ! . !.
78 *# -*& . . -*. . .
!
. & &
7!8 *#
&
!.
-*&
! !.
&
&
!.
!.
&
! -*& .%
/
We are calculating the integral of the radial distribution function between r equals 0 and
2a0 using the integration by parts by the tabular method explained in the section 8.1.3.
Substitute in (2) the two limits for our integral and the integration constant cancels out.
(3) The probability of the electron being between 2a0 and 0 from the nucleus for the 1s
atomic orbital of a H-atom is 0.7619. This is the same as saying there is a 76.19% chance
that the 1s electron is between the nucleus and a distance of 2a0. See Fig. 8.4 for the radial
distribution function which has an asymptote at a radius of infinity and Fig. 8.7 for the
3-dimensional 1s atomic orbital drawn to enclose 90% of the total probability.
Postscript
This workbook is the first of three books which continue with Intermediate Maths for
Chemists (Parker 2012) and Advanced Maths for Chemists (Parker 2013). The three books
together cover a typical first year maths course for your chemistry or related science or
engineering degree. Also the three books are very useful for future reference in later years of
your degree(s). Go to my web page for links to download the other two maths books and
also books on quantum mechanics and spectroscopy.
http://johnericparker.wordpress.com/
9 REFERENCES
Acree, WE & Chickos JS 2017, Phase Transition Enthalpy Measurements of Organic and
Organometallic Compounds in NIST Chemistry WebBook, NIST Standard Reference
Database Number 69, Eds. Linstrom, PJ & Mallard WG, National Institute of Standards
and Technology, Gaithersburg MD, 20899, doi:10.18434/T4D303, (retrieved July 4, 2017).
Atkins, P, de Paula, J & Keeler, J 2017, Atkins’ Physical Chemistry, 11th edn. Oxford University
Press, Oxford.
Atkins, P & de Paula, J 2016, Elements of Physical Chemistry, 7th edn. Oxford University
Press, Oxford.
Blackman, A Bottle, SE Schmid, S Mocerino, M & Wille, U 2015, Chemistry, 3nd edn.
John Wiley, Australia.
Jmol, 2018, an open-source Java viewer for chemical structures in 3D. http://www.jmol.org/
McBride, BJ, Zehe, MJ, & Gordon, S, 2002, NASA Glenn Coefficients for Calculating
Thermodynamic Properties of Individual Species, NASA/TP—2002-211556, Cleveland, Ohio.
Parker, JE 2018, Introductory Maths for Chemists, 3rd edn. BookBooN.com, Copenhagen.
Stroud, KA & Booth, DJ 2013, Engineering Mathematics, 7th edn. Palgrave MacMillan,
Basingstoke.
10 LIST OF FORMULAE
The formulae in order of appearance.
#
- #
Definition: a physical quantity
//%/.+
Wavelength and frequency of light
+ Definition: concentration in mol L
−1
3 3 #' 5 5 #'
Molar concentration
#' #'
& 5 !
( 5 ( 5 ( Stoichiometric equation and
5 & / 5
rate law
'%
+
+
+ + '% Two versions of the van der Waals
equation
#
-
* Clausius-Clapeyron equation
% '%
Pascal’s triangle of binomial coefficients
! !
&
&
, . . ,
+ %
.
+ %
. Point of inflection in a curve of p = f (V m)
% ' % '%
The van der Waals parameters from experimental pc and Tc
& +
(
Definition: partial fraction
. . . .
"
Photon energy
# -
0 -
- _
- 7#4#8
-
( - KCl Born-Haber cycle
3 3Z 3Z 5 Lindemann mechanism for unimolecular reactions
-
0 - *
0 - Definition: standard reaction
enthalpy
4 Definition: wavenumber
4 '
Rydberg equation for H-atom electronic spectra
11 !1 The Bragg equation for X-ray diffraction
7 Average end-to-end distance <r> of an ideal random coil polymer
GG ! G
, Acetic acid acidity constant Ka
! GG
K !6
, Ammonia standard pressure equilibrium constant
K 6 6 !
&!&
3 Lindemann rate of “unimolecular” isomerization reactions
&
78 #(
#( #( Rules of logs
#
,
&++
,/!
%.&
%
%+
j`
Euler’s number
M . ! & ,
% + / .
%+
.
! &
! &
Euler’s number raised to the powers x and −x
! &
! &
* #( .
#' *G #(.
G
#'
Definitions: pH and pOH
3
* * , #( Henderson-Hasselbalch equation
3
" L2 Photoelectric effect
* -
# Integrated Clausius-Clapeyron equation
'%
o # Definition: differentiation of a function at x1
j .
Velocity is the differential of distance with time
Acceleration is the differential of the velocity with time
3
Definition: rate of reaction for a single reactant species A.
The product rule for differentiation
Chain rule for differential of a function y =f (n) where n = f (x)
The chain rule applied to y = en where n = f (x)
*
&5 %
Definition: most probable speed of gas molecule
Lennard-Jones potential
+ & ! !
3
& 3 Rate of reaction for radioactive decay
12 3 1 2
P3Q P3Q
78
& 3 78
& P3QP5Q Definition: second-order rate laws
+0
& Integrated second-order rate law with one reactant
3 3.
+0
' % # p-V work done on the system in an ideal gas expansion
+
% % ! & % , %
% ! % % &
'
- # % % % %! % &
% ! & ,
% NASA polynomial equations
'% % ! & , %
! &
F % % % %
% ! # % & % ,
%
' ! &
%
-
%
- " % Constant pressure heat capacity and enthalpy
%
%
"
-
F
F % Constant pressure heat capacity and entropy
% %
%
3
&
!
.
-*
. Radial distribution function for a 1s electron of a H-atom
3 5
"
-*
-* Integration by parts, tabular method
-*
. -*
!
:
: Wavefunctions for a particle in a one-dimensional box