1. Kinetics of biological reaction. Influence of different
factors on the chemic reaction rate.
3. Catalysis and catalysts.
Influence of inorganic catalysts and enzymes on chemical reaction rate.
Chemical kinetics
Chemical kinetics, also known as reaction kinetics,
is the study of rates of chemical processes. Chemical kinetics includes
investigations of how different experimental conditions can influence the speed
of a chemical reaction and yield information about the reaction's mechanism and
transition states, as well as the construction of mathematical models that can
describe the characteristics of a chemical reaction. In 1864, Peter Waage and Cato Guldberg pioneered
the development of chemical kinetics by formulating the law of mass action,
which states that the speed of a chemical reaction is proportional to the
quantity of the reacting substances.
Chemical
kinetics deals with the experimental determination of reaction rates from which
rate laws and rate constants are derived. Relatively simple rate laws exist for
zero-order reactions (for which reaction rates are independent of
concentration), first-order reactions, and second-order reactions, and can be
derived for others. In consecutive reactions, the rate-determining step often
determines the kinetics. In consecutive first-order reactions, a steady state
approximation can simplify the rate law. The activation energy for a reaction
is experimentally determined through the Arrhenius
equation and the Eyring equation. The main factors
that influence the reaction rate include: the physical state of the reactants,
the concentrations of the reactants, the temperature at which the reaction
occurs, and whether or not any catalysts are present in the reaction.
Factors affecting reaction rate
Nature of the reactants
Depending upon what substances are reacting, the
reaction rate varies. Acid/base reactions, the formation of salts, and ion
exchange are fast reactions. When covalent bond formation takes place between
the molecules and when large molecules are formed, the reactions tend to be
very slow. Nature and strength of bonds in reactant molecules greatly influence
the rate of its transformation into products.
Physical state
The physical state (solid, liquid, or gas) of a
reactant is also an important factor of the rate of change. When reactants are
in the same phase, as in aqueous solution, thermal motion brings them into
contact. However, when they are in different phases, the reaction is limited to
the interface between the reactants. Reaction can occur only at their area of
contact; in the case of a liquid and a gas, at the surface of the liquid.
Vigorous shaking and stirring may be needed to bring the reaction to
completion. This means that the more finely divided a solid or liquid reactant
the greater its surface area per unit volume and the more contact it makes with
the other reactant, thus the faster the reaction. To make an analogy, for
example, when one starts a fire, one uses wood chips and small branches — one
does not start with large logs right away. In organic chemistry, on water
reactions are the exception to the rule that homogeneous reactions take place
faster than heterogeneous reactions.
Concentration
The reactions are due to collisions of reactant
species. The frequency with which the molecules or ions collide depends upon
their concentrations. The more crowded the molecules are, the more likely they
are to collide and react with one another. Thus, an increase in the
concentrations of the reactants will result in the corresponding increase in
the reaction rate, while a decrease in the concentrations will have a reverse
effect. For example, combustion that occurs in air (21% oxygen) will occur more
rapidly in pure oxygen.
Temperature
Temperature usually has a major effect on the
rate of a chemical reaction. Molecules at a higher temperature have more
thermal energy. Although collision frequency is greater at higher temperatures,
this alone contributes only a very small proportion to the increase in rate of
reaction. Much more important is the fact that the proportion of reactant
molecules with sufficient energy to react (energy greater than activation
energy: E > Ea) is significantly higher and is explained in detail by the
Maxwell–Boltzmann distribution of molecular energies.
The 'rule of thumb' that the rate of chemical
reactions doubles for every
A reaction's kinetics can also be studied with a
temperature jump approach. This involves using a sharp rise in temperature and
observing the relaxation time of the return to equilibrium. A particularly
useful form of temperature jump apparatus is a shock tube, which can rapidly
jump a gases temperature by more than 1000 degrees.
Catalysts
A catalyst is a substance that accelerates the
rate of a chemical reaction but remains chemically unchanged afterwards. The
catalyst increases rate reaction by providing a different reaction mechanism to
occur with a lower activation energy. In autocatalysis
a reaction product is itself a catalyst for that reaction leading to positive
feedback. Proteins that act as catalysts in biochemical reactions are called
enzymes. Michaelis–Menten
kinetics describe the rate of enzyme mediated
reactions. A catalyst does not affect the position of the equilibria,
as the catalyst speeds up the backward and forward reactions equally.
In certain organic molecules, specific substituents can have an influence on reaction rate in neighbouring group participation.
Agitating or mixing a solution will also
accelerate the rate of a chemical reaction, as this gives the particles greater
kinetic energy, increasing the number of collisions between reactants and,
therefore, the possibility of successful collisions.
Pressure
Increasing the pressure in a gaseous reaction
will increase the number of collisions between reactants, increasing the rate
of reaction. This is because the activity of a gas is directly proportional to
the partial pressure of the gas. This is similar to the effect of increasing
the concentration of a solution.
In addition to this straightforward mass-action
effect, the rate coefficients themselves can change due to pressure. The rate
coefficients and products of many high-temperature gas-phase reactions change
if an inert gas is added to the mixture; variations on this effect are called
fall-off and chemical activation. These phenomena are due to exothermic or
endothermic reactions occurring faster than heat transfer, causing the reacting
molecules to have non-thermal (non-Boltzmann) energy
distributions. Increasing the pressure increases the heat transfer rate between
the reacting molecules and the rest of the system, reducing this effect.
Condensed-phase rate coefficients can also be
affected by (very high) pressure; this is a completely different effect than
fall-off or chemical-activation. It is often studied using diamond anvils.
A reaction's kinetics can also be studied with a
pressure jump approach. This involves making fast changes in pressure and
observing the relaxation time of the return to equilibrium.
Equilibrium
While chemical kinetics is concerned with the
rate of a chemical reaction, thermodynamics determines the extent to which
reactions occur. In a reversible reaction, chemical equilibrium is reached when
the rates of the forward and reverse reactions are equal and the concentrations
of the reactants and products no longer change. This is demonstrated by, for
example, the Haber–Bosch process for combining
nitrogen and hydrogen to produce ammonia. Chemical clock reactions such as the Belousov–Zhabotinsky reaction
demonstrate that component concentrations can oscillate for a long time before
finally attaining the equilibrium.
Free energy
In general terms, the free energy change
(ΔG) of a reaction determines whether a chemical change will take place,
but kinetics describes how fast the reaction is. A reaction can be very
exothermic and have a very positive entropy change but will not happen in
practice if the reaction is too slow. If a reactant can produce two different
products, the thermodynamically most stable one will in general form, except in
special circumstances when the reaction is said to be under kinetic reaction
control. The Curtin–Hammett principle applies when determining the product
ratio for two reactants interconverting rapidly, each
going to a different product. It is possible to make predictions about reaction
rate constants for a reaction from free-energy relationships.
The kinetic isotope effect is the difference in
the rate of a chemical reaction when an atom in one of the reactants is
replaced by one of its isotopes.
Chemical kinetics provides information on
residence time and heat transfer in a chemical reactor in chemical engineering
and the molar mass distribution in polymer chemistry.
Applications
The mathematical models that describe chemical
reaction kinetics provide chemists and chemical engineers with tools to better
understand and describe chemical processes such as food decomposition,
microorganism growth, stratospheric ozone decomposition, and the complex
chemistry of biological systems. These models can also be used in the design or
modification of chemical reactors to optimize product yield, more efficiently
separate products, and eliminate environmentally harmful by-products. When
performing catalytic cracking of heavy hydrocarbons into gasoline and light
gas, for example, kinetic models can be used to find the temperature and
pressure at which the highest yield of heavy hydrocarbons into gasoline will
occur. Kinetics is also a basic aspect of chemistry.
Concentration Changes
Chemical kinetics is the study of the speed with
which a chemical reaction occurs and the factors that affect this speed. This
information is especially useful for determining how a reaction occurs.
What is meant by the speed of a reaction? The
speed of a reaction is the rate at which the concentrations of reactants and
products change.
Consider the following hypothetical example. The
letters A, B, and C represent chemical species (in this context, the letters do
not represent elements). Suppose the following imaginary reaction occurs:
A + 2 B →
The simulation below illustrates how this
reaction can be studied. The apparatus at the left is called a stopped-flow
apparatus. Each syringe contains a solution filled with a different reactant (A
or B). When the two solutions are forced out of the syringes, they are quickly
mixed in a mixing block and the reaction starts. The reacting solution passes
through the tube at the bottom. An analytical technique such as spectrophotometry is used to measure the concentrations of
the species in the reaction mixture (which is in the tube at the bottom) and
how those concentrations change with time.
In this example, the syringe at the left contains
a solution of species A, which has a yellow color. The syringe at the right
contains a solution of species B, which has a light blue color. The product C
has a red color.
Chemical Kinetics
Chemical kinetics is the study and discussion of
chemical reactions with respect to reaction rates, effect of various variables,
re-arrangement of atoms, formation of intermediates etc. There are many topics
to be discussed, and each of these topics is a tool for the study of chemical
reactions. By the way, the study of motion is called kinetics, from Greek
kinesis, meaning movement.
At the
macroscopic level, we are interested in amounts reacted, formed, and the rates
of their formation. At the molecular or microscopic level, the following
considerations must also be made in the discusion of
chemical reaction mechanism.
Molecules or atoms of reactants must collide with
each other in chemical reactions.
The molecules must have sufficient energy
(discussed in terms of activation energy) to initiate the reaction.
In some cases, the orientation of the molecules
during the collision must also be considered.
Reaction Rates
Chemical reaction rates are the rates of change
in concentrations or amounts of either reactants or products. For changes in
amounts, the units can be one of mol/s, g/s, lb/s, kg/day etc. For changes in
concentrations, the units can be one of mol/(L s),
g/(L s), %/s etc.
With respect to reaction rates, we may deal with
average rates, instantaneous rates, or initial rates depending on the
experimental conditions.
Thermodynamics and kinetics are two factors that
affect reaction rates. The study of energy gained or released in chemical
reactions is called thermodynamics, and such energy data are called
thermodynamic data. However, thermodynamic data have no direct correlation with
reaction rates, for which the kinetic factor is perhaps more important. For
example, at room temperature (a wide range of temperatures), thermodynamic data
indicates that diamond shall convert to graphite, but in reality, the
conversion rate is so slow that most people think that diamond is forever.
Factors Influence Reaction
Rates
Many
factors influence rates of chemical reactions, and these are summarized below.
Much more extensive discussion will be given in other pages.
Nature of Reactants
Acid-base
reactions, formation of salts, and exchange of ions are fast reactions. Reactions
in which large molecules are formed or break apart are usually slow. Reactions
breaking strong covalent bonds are also slow.
Temperature
Usually, the higher the temperature, the faster the reaction.
The temperature effect is discussed in terms of activation energy.
Concentration Effect
The
dependences of reaction rates on concentrations are called rate laws. Rate laws
are expressions of rates in terms of concentrations of reactants. Keep in mind
that rate laws can be in differential forms or integrated forms. They are
called differential rate laws and integrated rate laws. The following is a
brief summary of topics regarding rate laws.
rate
laws: differential and integrated rate laws.
Integrated rate laws: First Order Reactions
Second
Order Reactions
Rate laws
apply to homogeneous reactions in which all reactants and products are in one
phase (solution).
Heterogeneous reactions: reactants are present in
more than one phase
For
heterogeneous reactions, the rates are affected by surface areas.
Catalysts: substances used to facilitate
reactions
By the
nature of the term, catalysts play important roles in chemical reactions.
Reaction Mechanisms
The
detailed explanation at the molecular level how a reaction proceeds is called
reaction mechanism. The explanation is given in some elementary steps. Devising
reaction mechanisms requires a broad understanding of properties of reactants
and products, and this is a skill for matured chemists. However, first year
chemistry students are often given a mechanism, and be asked to derive the rate
law from the proposed mechanism. The steady-state approximations
is a technique for deriving a rate law from the proposed mechanism.
Rate of Reaction
The rate
of a reaction is the speed at which a reaction happens. If a reaction has a low
rate, that means the molecules combine at a slower speed than a reaction with a
high rate. Some reactions take hundreds, maybe even thousands, of years while
others can happen in less than one second. The rate of reaction depends on the
type of molecules that are combining. If you want to think of a very slow
reaction, think about how long it took dinosaur bones to become fossils through
breakdown. You can thank chemical processes in bacteria for most of those
dinosaur bones in the museum.
There is another big idea for rates of reaction
called collision theory. The collision theory says that as more collisions in a
system occur, there will be more combinations of molecules bouncing into each
other. If there are a higher number of collisions in a system, more
combinations of molecules can occur. The reaction will go faster and the rate
of that reaction will be higher. Even though they are both liquids, think about
how slowly molecules move in honey when compared to your soda. There are a
lower number of collisions in the honey.
Reactions
happen - no matter what. Chemicals are always combining or breaking down. The
reactions happen over and over, but not always at the same speed. A few things
affect the overall speed of the reaction and the number of collisions that can
occur.
Concentration:
If there is more of a substance in a system, there is a greater chance that
molecules will collide and speed up the rate of the reaction. If there is less
of something, there will be fewer collisions and the reaction will probably
happen at a slower speed. Sometimes when you are in a chemistry lab, you will
add one solution to another. When you want the rate of reaction to be slower,
you will add only a few drops at a time instead of the entire beaker.
Temperature:
When you raise the temperature of a system, the molecules bounce around a lot
more because they have more energy. When they bounce around more, they are more
likely to collide. That fact means they are also more likely to combine. When
you lower the temperature, the molecules are slower and collide less. That
temperature drop lowers the rate of the reaction. Back to the chemistry lab!
Sometimes you will mix solutions in ice so that the temperature of the system stays
cold and the rate of reaction is slower.
Pressure:
Pressure affects the rate of reaction, especially when
you look at gases. When you increase the pressure, the molecules have less
space in which they can move. That greater density of molecules increases the
number of collisions. When you decrease the pressure, molecules don't hit each
other as often. The lower pressure decreases the rate of reaction.
Rates of Reaction
What factors influence the rate of a chemical
reaction?
1.
Temperature
2.
Catalysts
3.
Concentrations of reactants
3.
Surface area of a solid reactant
4.
Pressure of gaseous reactants or products
If you are planning an investigation, I suggest
that you investigate the effects of temperature or the effects of the
concentration of the reactants because these will allow you to choose a
suitable range of values for the controlled or independent variable. The
dependent variable will be the rate of the reaction. Keep all the other
variables fixed.
To make a prediction for your investigation you
will have to ask yourself the question: What will happen to the rate of the
reaction when I increase the temperature? or What will
happen to the rate of the reaction if I increase the concentration of one of
the reactants? The answer to that question is your prediction. The next thing
to do is to explain your prediction. You will have to answer the question: Why
will the reaction go faster if I increase the temperature? or
Why will the reaction go faster if I increase the concentration of one of the
reactions? The answer to this question is your explanation, and to get the
highest possible marks, you will have to provide a full scientific explanation.
Once you have written your hypothesis (prediction
with explanation) you will decide how to do the experiments, i.e. write the
proposed method.
How does
temperature affect the rate of a chemical reaction?
When two chemicals react, their molecules have to
collide with each other with sufficient energy for the reaction to take place.
This is collision theory. The two molecules will only react if they have enough
energy. By heating the mixture, you will raise the energy levels of the
molecules involved in the reaction. Increasing temperature means the molecules
move faster. This is kinetic theory. If your reaction is between atoms rather
than molecules you just substitute “atom” for “molecule” in your explanation.
How do
catalysts affect the rate of a reaction?
Catalysts speed up chemical reactions. Only very
minute quantities of the catalyst are required to produce a dramatic change in
the rate of the reaction. This is really because the reaction proceeds by a
different pathway when the catalyst is present. Adding extra catalyst will make
absolutely no difference. There is a whole page on this site devoted to catalysts.
How does
concentration affect the rate of a reaction?
Increasing the concentration of the reactants
will increase the frequency of collisions between the two reactants. So this is
collision theory again. You also need to discuss kinetic theory in an experiment
where you vary the concentration. Although you keep the temperature constant,
kinetic theory is relevant. This is because the molecules in the reaction
mixture have a range of energy levels. When collisions occur, they do not
always result in a reaction. If the two colliding molecules have sufficient
energy they will react.
If reaction is between a substance in solution
and a solid, you just vary the concentration of the solution. The experiment is
straightforward. If the reaction is between two solutions, you have a slight
problem. Do you vary the concentration of one of the reactants or vary the
concentration of both? You might find that the rate of reaction is limited by
the concentration of the weaker solution, and increasing the concentration of
the other makes no difference. What you need to do is fix the concentration of
one of the reactants to excess. Now you can increase the concentration of the
other solution to produce an increase in the rate of the reaction.
How does
surface area affect a chemical reaction?
If one of the reactants is a solid, the surface
area of the solid will affect how fast the reaction goes. This is because the
two types of molecule can only bump into each other at the liquid solid
interface, i.e. on the surface of the solid. So the larger the surface area of
the solid, the faster the reaction will be.
Smaller particles have a bigger surface area than
larger particle for the same mass of solid. There is a simple way to visualize
this. Take a loaf of bread and cut it into slices. Each time you cut a new
slice, you get an extra surface onto which you can spread butter and jam. The
thinner you cut the slices, the more slices you get and so the more butter and
jam you can put on them. This is “Bread and Butter Theory”. You should have
come across the idea in your biology lessons. By chewing your food you increase
the surface area so that digestion can go faster.
What affect
does pressure have on the reaction between two gasses?
You should already know that the atoms or molecules
in a gas are very spread out. For the two chemicals to react,
there must be collisions between their molecules. By increasing the pressure,
you squeeze the molecules together so you will increase the frequency of
collisions between them. This is collision theory again.
In a diesel engine, compressing the gaseous
mixture of air and diesel also increases the temperature enough to produce
combustion. Increasing pressure also results in raising the temperature. It is
not enough in a petrol engine to produce combustion, so petrol engines need a
spark plug. When the petrol air mixture has been compressed, a spark from the
plug ignites the mixture. In both cases the reaction (combustion) is very fast.
This is because once the reaction has started, heat is produced and this will
make it go even faster.
The Rate Law
The rate law for a chemical reaction links the
reaction rate with concentrations or pressures of reactants and constant
parameters.
Rate
Laws for Various Reactions
A variety of reaction orders are observed, and they cannot be easily correlated with the stoichiometry of the reaction.
Rate Law
Rate = K[A]x[B]y
Rate = Reaction Rate Reaction
Order
The sum of x
and y.
The reaction of bromine and formic acid is
first order in bromine, zeroth order in formic acid,
and first order overall.
K = rate constant
x/y = determined
numbers
A/B = concentrations
How fast is Fast?
The Mathematics of Change
Consider a reaction of Red molecules
(A) to make Blue molecules (B), i.e. A -> B. If we were able to see the
reaction on a molecular scale, the reaction of each individual molecule of
occurs very rapidly, but the overall color of the vessel changes more slowly.
Snapshots of the reaction in progress might look like this:
The number of reactants and products in the
reaction vessel changes with time, with the relative number of reactant
molecules destroyed and number of products formed per reaction event determined
by the reaction stoichiometry. Each reactant molecule
is identical to every other one, but they all don't react at the same instant.
At each point in time, the probability of reaction per unit time is the same
for each molecule in the sample, and that probability influences the overall
reaction rate. But that isn't the only thing that determines the overall
reaction rate. The total number of reactions at any instant is the probability
of reaction per unit time multiplied by the number of reactants remaining in
the vessel. Thus the reaction proceeds quickly at first, when there are lots of
reactants around, and appears to slow as the reactants are consumed. A simulation of a
similar reaction involving reactive collisions between molecules can be run on
you browser. A plot of the time dependence of the number of molecules of each
type looks 'smooth' when there are lots and lots of molecules in the sample, so
individual reaction events get 'averaged' out. The concentrations of the
reactants and products change in time like this:
The
Rate Law
The equation that describes the dependence of the
reaction rate on the concentrations of the species in the reactor is called a
rate law. The rate law for a given reaction is determined from the reaction
mechanism. Several important kinds of simple rate
laws are worth noting
First Order Rate Law
The simplist reaction
mechanism is that of unimolecular decomposition (or isomerisation). In such a process, a single reactant
undergoes a transformation at a constant probability per unit time. Such a mechanism leasds to a first-order
reaction rate law. Examples of reactions such as these are radioactive
decay, bacterial growth, and compound interest. Let's assume the reaction has a
simple stoichimetry:
A → B
A First-Order Rate Law is called such because the
rate of product formation ( or reactant depletion ) is
proportional to the first power of the number of available reactants (or
reactant concentration):
rate
=k[A]
where
[A] represents the concentration (number density) of species A in the sample.
Second Order
Rate Law
If two molecules undergo a bimolecular reaction
such as a reaction that involves a collisional
encounter to produce products, and has a stoichiometry
like this:
A + A → B + C
Zeroth Order Rate Law
If a reaction is catalysed
by a surface and has enough (excess) reactant, the rate of the reaction depends
on the area of the catalyst, not on how much reactant is present. This is an
unusual circumstance outside of the realm of catalyzed reactions and is
described by a Zeroth Order rate law:
rate
=k
THE EFFECT OF
CONCENTRATION ON REACTION RATES
This page describes and explains the way that
changing the concentration of a solution affects the rate of a reaction. Be
aware that this is an introductory page only. If you are interested in orders
of reaction, you will find separate pages dealing with these. You can access
these via the rates of reaction menu (link at the bottom of the page).
For many reactions involving liquids or gases,
increasing the concentration of the reactants increases the rate of reaction.
In a few cases, increasing the concentration of one of the reactants may have
little noticeable effect of the rate. These cases are discussed and explained
further down this page.
Don't assume that if you double the concentration
of one of the reactants that you will double the rate of the reaction. It may
happen like that, but the relationship may well be more complicated.
The
Temperature Dependence of Reaction Rates
Chemical Activation
Consider the reaction
H2 + Cl2 ->
2HCl
On a molecular level, bonds must be broken (H-H
and Cl-Cl) before the reaction can proceed too far
into products. This means that as the reactant molecules come together, the
collision must have enough energy to initiate the bond breakage for the
reaction to occur. Not all collisions will have this amount of energy. The collisions that do not have sufficient energy to react end up
as elastic scattering events.
Only collisions with enough energy react to form
products. The energy of the system changes as the reactants approach each
other. The critical amount of energy to make the reaction proceed is called the
Activation Energy.
The Reaction Coordinate is the 'distance' along
the path of the reaction, and is plotted along the horizontal axis. The energy
of interaction of the reactive system is plotted vertically, and is called the
Chemical potential, or just potential energy. You fight gravitational potential
energy when you try to roll a boulder over a mountain.
A chemical potential of interaction usually looks
like something like the graph above, which is similar to the 'pushing a boulder
over a hill' graph above. The Graph above is drawn for the isomerization
of an isonitrile that we discussed before. The
barrier to the isomerization keeps the unstable CH3NC
from reacting away quickly at low temperature, even though energy is released
upon the net reaction.
Catalyst and Catalysis
A catalyst increases the rate of a particular
reaction without itself being used up. A catalyst can be added to a reaction
and then be recovered and reused after the reaction occurs. The process or
action by which a catalyst increases the reaction rate is called catalysis. The
study of reaction rates and how they change when manipulated experimentally is
called kinetics.
The term catalysis was proposed in 1835 by the
Swedish chemist Jöns Berzelius
(1779-1848). The term comes from the Greek words kata
meaning down and lyein meaning loosen. Berzelius explained that by the term catalysis he meant
"the property of exerting on other bodies an action which is very
different from chemical affinity. By means of this action, they produce
decomposition in bodies, and form new compounds into the composition of which
they do not enter."
Most chemical reactions occur as a series of
steps. This series of steps is called a pathway or mechanism. Each individual
step is called an elementary step. The slowest elementary step in a pathway
determines the reaction rate. The reaction rate is the rate at which reactants
disappear and products appear in a chemical reaction, or, more specifically,
the change in concentration of reactants and products in a certain amount of
time.
While going through a reaction pathway, reactants
enter a transitional state where they are no longer reactants, but are not yet
products. During this transitional state they form what is called an activated
complex. The activated complex is short-lived and has partial bonding
characteristics of both reactants and products. The energy required to reach
this transitional state and form the activated complex in a reaction is called
the activation energy. In order for a reaction to occur, the activation energy
must be reached. A catalyst increases the rate of reaction by lowering the
activation energy required for the reaction to take place. The catalyst forms
an activated complex with a lower energy than the complex formed without
catalysis. This provides the reactants a new pathway which requires less
energy. Although the catalyst lowers the activation energy required, it does
not affect reaction equilibrium or thermodynamics. The catalyst does not appear
in the overall chemical equation for a pathway because the mechanism involves
an elementary step in which the catalyst is consumed and another in which it is
regenerated.
Catalysts exist for all types of chemical
reactions. A specific catalyst can be classified into one of two main groups;
homogeneous and heterogeneous. A catalyst that is in the same phase as the
reactants and products involved in a reaction pathway is called a homogeneous
catalyst. When a catalyst exists in a different phase than that of the
reactants, it is called a heterogeneous catalyst. For example, nickel is a
catalyst in the hydrogenation of vegetable oils. Nickel is a solid, while the
oil is a liquid, therefore nickel is a heterogeneous
catalyst. An advantage of using heterogeneous catalysts is their ease of
separation from the reactants and products involved in a pathway.Metals
are often used as heterogeneous catalysts because many reactants adsorb to the
metal surface, increasing the concentration of the reactants and therefore the
rate of the reaction. Ionic interactions between metals and other molecules can
be used to orient the reactants involved so that they react better with each
other, or to stabilize charged reaction transition states. Metals also can
increase the rate of oxidation-reduction reactions through changes in the metal
ion's oxidation state.
Another group of catalysts are called enzymes.
Enzymes are catalysts that are found in biological systems. The role of
catalysts in living systems was first recognized in 1833. French chemists Anselme Payen (1795-1871) and
Jean François Persoz isolated a material from
malt that accelerated the conversion of starch to sugar. Payen
called the substance diatase. A half century later
German physiologist Willy Kühne suggested the
name enzyme for biological catalysts.
Enzymes are proteins and therefore have a highly
folded three-dimensional configuration. This configuration makes an enzyme
particularly specific for a certain reaction or type of reaction. Synthetic
catalysts, on the other hand, are not nearly as specific. They will catalyze
similar reactions that involve a wide variety of reactants. Enzymes, in
general, will lose activity more easily than synthetic catalysts. Very slight
disturbances in the protein structure of enzymes will change the
three-dimensional configuration of the molecule and, as a result, its
reactivity. Enzymes tend to be more active, i.e., they catalyze reactions
faster, than synthetic catalysts at ambient temperatures. Catalytic activity
for a reaction is expressed as the turnover number. This is simply the number
of reactant molecules changed to product per catalyst site in a given unit of
time. When temperature is increased, synthetic catalysts can become just as
active as enzymes. With an increase in temperature, many enzymes will become
inactive because of changes to the protein structure.
There are endless reactions that can undergo
catalysis. One example is the decomposition of hydrogen peroxide (H2O2).
Without catalysis, hydrogen peroxide decomposes slowly over time to form water
and oxygen gas. A 30% solution of hydrogen peroxide at room temperature will
decompose at a rate of 0.5% per year. The activation energy for this reaction
is 75 kJ/mol. This activation energy can be lowered to 58 kJ/mol with the addition of iodide ions (I-). These
ions form an intermediate, HIO-, which reacts with the hydrogen peroxide to
regenerate the iodide ions. When the enzyme catalase
is added to the hydrogen peroxide solution, the activation energy is lowered even further to 4 kJ/mol. The catalase
is also regenerated in the reaction and can be separated from the solution for
reuse. This example shows how a catalyst can lower the activation energy of a
reaction without itself being used up in the reaction pathway.
Another example of catalysis is the catalytic
converter of an automobile. Exhaust from the automobile can contain carbon
monoxide and nitrogen oxides, which are poisonous gases. Before the exhaust can
leave the exhaust system these toxins must be removed. The catalytic converter
mixes these gases with air and then passes them over a catalyst made of rhodium
and platinum metals. This catalyst accelerates the reaction of carbon monoxide
with oxygen and converts it to carbon dioxide, which is not toxic. The catalyst
also increases the rate of reactions for which the nitrogen oxides are broken
down into their elements.
A well-known example of catalysis is the
destruction of the ozone layer. Ozone (O3) in the upper atmosphere serves as a
shield for the harmful ultraviolet rays from the Sun. Ozone is formed when an
oxygen molecule (O2) is split into two oxygen atoms (O) by the radiation from the
Sun. The free oxygen atoms then attach to oxygen molecules to form ozone. When
another free oxygen atom reacts with the ozone molecule, two oxygen molecules
are formed. This is the natural destruction of ozone. Under normal
circumstances, the rate of destruction of ozone is the same as the rate of
ozone formation, so no net ozone depletion occurs. When chlorine (Cl) atoms are present in the atmosphere, they act as
catalysts for the destruction of ozone. Chlorine atoms in the atmosphere come
from compounds containing chlorofluorocarbons, or CFCs. CFCs are compounds
containing chlorine, fluorine, and carbon. CFCs are very stable and can drift
into the upper atmosphere without first being broken down. Once in the upper
atmosphere, the energy from the Sun causes the chlorine to be released. The
chlorine atom reacts with ozone to form chlorine monoxide (ClO)
and an oxygen molecule. The chlorine monoxide then reacts with another oxygen
atom to form an oxygen molecule and the regenerated chlorine atom. With the help
of the chlorine catalyst, the degeneration of ozone occurs at a faster rate
than its formation, which has caused a net depletion of ozone in the
atmosphere.
The previous examples illustrate some of the many
practical applications of catalysis. Almost all of the chemicals produced by
the chemical industry are made using catalysis. Catalytic processes used in the
chemical industry decrease production costs as well as create products with
higher purity and less environmental hazards. A wide variety of products are
made using catalytic processes. Catalysis is used in industrial chemistry,
pharmaceutical chemistry, and agricultural chemistry, as well as in the
specialty chemical industry. Useful chemicals such as sulfuric acid,
penicillin, and fructose are made more efficiently using catalytic processes.
Research and development efforts in the chemical industry are significantly
more productive with the use of catalysis in fields such as fuel refining,
petrochemical manufacturing, and environmental management.
The majority of manufacturing processes in use
today by the chemical industry employ catalytic reactions. These reactions are
highly efficient, but research is continuing to increase the efficiency even
more. The focus of this research is on separation and regeneration of the
catalysts in order to decrease costs of production while increasing the purity
of the product. The field of catalysis research is rapidly growing and will
continue to do so as new catalysts and catalytic processes are discovered.
Factors affection the
reaction rate.
The fate of any
particular reaction depends upon the following factors:
1.
Nature of the reactants. Consider the following two reactions
These reactions appear to be
similar but the first is fast while the second is slow. This is because
different amounts of energies are required for breaking of different bonds and
different amounts of energies are released in the formation of different bonds.
2. Concentration of the reactants. Greater are
the concentrations of the reactants, faster is the
reaction, as the concentrations of the reactants decrease, the rate of reaction
also decreases.
3. Temperature. The rate of reaction increases with
increase of temperature. In most of the cases, the rate of reaction becomes
nearly double for 10Ê
rise of temperature. In êîòå ñìåÿ, reactions
do not take place at room temperature but take place at higher temperature.
4. Presence of Catalyst. À catalyst
generally increases the speed of à reaction
without itself being consumed in the reaction. In case of reversible reactions,
à
catalyst helps to attain the equilibrium quickly without disturbing the state
of equilibrium.
5. Surface area of the reactants. For à reaction
involving à
solid reactant or catalyst, the smaller is the particle size i.å.,
greater is the surface area, the fast r is the reaction.
6. Presence of light. Some
reactions do not take place in the dark but take place in the presence of light
e.g.,
Í2 +
Ñ12
= 2ÍÑ1.
Such reactions are called “photochemical reactions”
Rate laws and rate constants. It is often
found that the rate of reaction is proportional to the concentrations of the
reactants raised to à power. For example, 1ñ may be found that the rate is
proportional to the concentrations of two reactants À and Â, and that:
where each concentration is raised
to the first power. The coefficient k is called the rate constant for the reaction
or velocity constant. The rate constant is independent of the concentrations
but depends on the temperature. An experimentally determined equation of this
kind is called the
rate law of the reaction. More formally, à rate law is an equation
that expresses the rate of reaction as à function of the concentrations of all
the species present in the overall chemical equation for the reaction.
If
all concentrations are take as unity, [A] = [B] = 1 mole/liter, then rate = k.
Hence
rate constant may be defined as the rate of the reaction when the concentration
of each reactants is take as unity. That is why the
rate constant is also called specific reaction
rate.
Characteristics of rate constant.
Some important characteristics of the rate constant are as follows:
1.
Rate constant is a measure of
the rate of reaction. Greater is the value of the rate constant, factors is the
reaction.
2.
Each reaction has a definite
value of the rate constant at particular temperature.
3.
The value of the rate constant
for the same reaction changes with temperature.
4.
The value of the rate constant
of a reaction does not depend upon the concentration of the reactants.
5.
The units of the rate constant
depend upon the order of reaction.
À practical application of à rate law is that, once we know it and the
value of the rate constant, we can predict the rate of reaction from the
composition of the mixture. Moreover, as we shall see later, by knowing the
rate law we can go on to predict the composition of the reaction mixture at à
later stage of the reaction. The theoretical usefulness of a rate law is that
it is à guide to the mechanism of the reaction, for any proposed mechanism must
be consistent with the observed rate law.
Order of reaction
The sum of the concentration terms on which the rate of à reaction
actually depends as observed experimentally is called the order of the
reaction. For example, in the above case, order of reaction = à + p. Thus the
orders a reaction may also be defined as the sum of the exponents (powers) to
which the concentration terms in the rate law equation are raised to express
the observed rate of the reaction.
The power to which the concentration of à species is raised in à rate
law is the order of the reaction
with respect to that species. À reaction with the rate law is first-order in À
and first-order in Â. The overall order of à reaction is the sum of the
orders of all the components. The rate law is therefore second-order overall.
Some reactions obey à zero-order rate law, and therefore have à rate
that is independent of the concentration of the reactant (so long as some is
present). Thus, the catalytic decomposition of phosphine
(ÐÍ3) on hot tungsten at high pressures has the rate law: u =
k
The PH3 decomposes at à constant rate until it has almost
entirely disappeared. Only heterogeneous reactions can have rate laws that are zero-order
overall.
These remarks point to three problems. First, we must see how to
identify the rate law and obtain the rate constant from the experimental data.
We shall concentrate on this aspect in this chapter. Second, we must see how to
construct reaction mechanisms that are consistent with the rate law. k2 +
k3[Â]0
It is most important to distinguish molecularity
from order: Reaction
order is an empirical quantity, and obtained from thy experimental rate law.
The molecularity refers to an elementary reaction
proposed as an individual step in à mechanism.
In contrast to reactions in general, the rate law of an elementary
reaction can be written down from its chemical equation. Thus, the rate law of
à unimolecular elementary reaction is first-order in
the reactant;
À ® Products : d[À]/dt = - k [À]
À unimolecular reaction is first-order because
the number of À molecules that decay in à short interval is proportional to the
number available to decay. (Ten times as many decay in the some interval when
there are initially 1000 À molecules than when there are only 100 present).
Therefore, the rate of decomposition of À is proportional to its molar
concentration.
An elementary bimolecular reaction has à second-order rate law:
À bimolecular reaction is second-order because its rate is proportional
to the rate at which the reactant species meet, which is proportional their
concentrations. Therefore, if we believe (or simply postulate) that a reaction
is à single-step, bimolecular process, then we can write down the rate law (and
then go on to test it). Bimolecular elementary reactions are believed to
account for many homogeneous reactions, such as the dimerizations
of alkenes and dienes and reactions such as:
CH3I(alc) + ÑÍ3ÑÍ2Î-
(alc) ®ÑÍ3ÎÑÍ3ÑÍ2(àlñ) + I-(à1ñ);
(where 'alc' signifies alcohol
solution). The mechanism of the last reaction is believed to be the single
elementary step: CH3I + ÑÍ3ÑÍ2Î- ®ÑÍ3ÎÑÍ3ÑÍ2
+ I-
u =
k[CH3I] [CH3CH2O-]
The interpretation of à rate law is full of pitfalls, partly because à
second-order rate law, for instance, can also result from à complex reaction
scheme. We shall see below how to string simple steps together into à mechanism
and how to arrive at the corresponding rate law. For the present we emphasize
that if the reaction is an elementary bimolecular process, then it has
second-order kinetics, but if the kinetics are second-order, then the reaction
might be complex. The postulated mechanism can be explored only by detailed
detective work on the system, and by investigating whether side products or
intermediates appear during the course of the reaction. Detailed analysis of
this kind was one of the ways, for example, in which the reaction H2(g) + I2(g) ®
2HI(g) was shown to proceed by à complex reaction after many years during which
it had been accepted on good, but insufficiently meticulous evidence, that it
was à fine example of à simple bimolecular reaction in which atoms exchanged
partners during à collision.
It is found that the rates of most reactions increase as the temperature
is raised. Many reactions fall somewhere in the range spanned by the hydrolysis
of methyl ethanoate (where the rate constant at 350Ñ
is 1.82 times that at 250Ñ) and hydrolysis of sucrose (where the
factor is 4.13).
The Arrhenius parameters. An empirical
observation is that many reactions have rate constants that follow the Arrhenius equation.
That
is, for many reactions it is found that à plot of ln k against
1/Ò
gives à
straight line.
The
factor À is called the pre-exponential
factor or the frequency factor; Åa
is called the activation
energy. Collectively, the two quantities are called the Arrhenius parameters
of the reaction. This equation is sometimes written in an alternative form that
combines the two parameters:
The quantity D+G is
called the activation
Gibbs energy. In this form, the expression for the rate constant
strongly resembles the formula for the equilibrium constant in terms of the
standard reaction Gibbs energy.
For the present chapter we shall regard the Arrhenius
parameters as purely empirical quantities that enable us to discuss the
variation of rate constants with temperature. There we shall see that the
activation energy is the minimum energy that reactants must have in order to
from products. For example, in à gas-phase reaction there are numerous
collisions each second, but only à tiny proportion of them are sufficiently
energetic to lead to reaction. The fraction of collisions with a kinetic energy
in excess of an energy Åa is given by the Boltzmann distribution as å-Ea/RT. Hence, the
exponential factor can be interpreted as the fraction of collisions that have
enough energy to lead to reaction.
The analogous interpretation of the pre-exponential factor is that it is
a measure of the rate at which collisions occur irrespective of their energy.
Hence the product of À and the exponential factor gives the rate of successful
collisions.
The temperature dependence of some reactions is not Arrhenius-like.
This
definition reduces to the earlier one (as the slope of an Arrhenius
plot) for à temperature-independent activation energy. Thus, by using d(l/Ò) = - dT/Ò2 we can rearrange equation:
CATALYSIS.
It is found that the rates of many reactions are increased by the
presence of à catalyst, à substance that increases the rate of à reaction without being consumed by it.
Although at first thought this may seem impossible, it can indeed occur,
because à
catalyst is à
substance that is used in one step in the mechanism for à reaction and
is regenerated in à
subsequent step. À
catalyst acts by making available à new reaction
mechanism with à
lower activation energy.
Figure.1.shows
the uncatalyzed path of à reaction
contrasted with its catalyzed path. (Each potential-energy maximum corresponds
to the formation of an activated complex.) Note that ÜÍ for the
reaction is independent of the reaction mechanism, and depends only upon the
identity of the reactants and products. However, the activation energy for the
catalyzed path is less than that for the uncatalyzed
path. Thus, at any given temperature more reactant molecules possess the
activation energy for the catalyzed reaction than for the uncatalyzed
one. The catalyzed mechanism thus predominates. À catalyst does not eliminate
à
reaction mechanism; rather, it offers à new, faster one. Mîre molecules, often almost all of
them, will follow the new (catalyzed) pathway the products, instead of the old.
If the activation energy of à
reaction is high, at normal temperatures only à small proportion of molecular
encounters result in reaction. À catalyst lowers the activation energy of the
reaction by providing an alternative path that avoids the slow,
rate-determining step of the uncatalysed reaction,
and results in à higher reaction rate at the same temperature. Catalysts can be
very effective; for instance, the activation energy for the decomposition of
hydrogen peroxide in solution is 76 kJ/mol, and the
reaction is slow at room temperature. When à little iodide is added, the
activation energy falls to 57 kJ/mol, and the rate increases by à factor of
2000. Enzymes, which are biological catalysts, are very specific and can have à
dramatic effect on the reactions they control. The activation energy for the
acid hydrolysis of sucrose is 107kJ/mol, but the enzyme saccharase
reduces it to 36 kJ/mol, corresponding to an acceleration of
the reaction by à factor of 100 at blood temperature (310 Ê).
À homogeneous
catalyst is à catalyst that is in the same phase as the
reaction mixture (å.g. an acid added to an aqueous
solution).
À heterogeneous
catalyst is in à di6erent phase (å.g.
à solid catalyst for à gas-phase reaction).
Homogeneous catalysis.
In homogeneous catalysis, the
catalyst and the reactants are present in the same phase. Consider the
elementary process
À +
 ®
products (slow)
Assume
that this process has à
high activation energy. If we now add catalyst C the reaction mixture, à new,
two-step mechanism is possible, in which rate-determining step (step 1, below)
has à lower
activation energy:
Step 1: À + Ñ ® ÀÑ (fast)
Step 2: ÀÑ + Â ® products + Ñ (faster)
Here,
both activation energies are low, and each reaction is faster than original, uncatalyzed reaction. Notice that the overall net equation
is changed, and that while catalyst Ñ is used up in step 1, it is regenerated step 2.
The rate law for the uncatalyzed reaction is: rate = k[A][B]
and
for the catalyzed reaction, rate = k'[À][C]
An
example of homogeneous catalysis is found in the oxidation of sulfur dioxide to
sulfur trioxide by oxygen, using nitrogen oxide, NO, as à catalyst.
The net equation for the
reaction is
2SO2
(g) + O2 (g) ®2 SO3
(g)
The uncatalyzed reaction is very slow,
either because it is termolecular (unlikely) or
because one step in its reaction mechanism has à very high activation energy.
Addition of nitrogen oxide, NO, to the mixture greatly speeds the reaction by
making the following mechanism available:
Step 1: O2 (g) + 2NO(g) ® 2NO2 (g)
Step 2: [NO 2 (g) +
SO2 (g) ®
NO (g) + SO3(g)] õ 2
The sum of these gives the original
net equation, and because the activation energy for each step is fairly low,
the reaction proceeds more rapidly than via the uncatalyzed
path.
Sîmå idea of the
mode of action of homogeneous catalysts can be obtained by examining the
kinetics of the bromide-catalysed decomposition of
hydrogen peroxide:
2Í2Î2(aq) ®2Í2Î2(aq) + Î2(g)
Heterogeneous catalysis.
À heterogeneous catalyst is one which provides à surface on which molecules can readily combine.
The process of heterogeneous catalysis begins with the adsorption of à molecule on
the surface of the catalyst. There are two general types of adsorption: the
relatively weak physical, or van-der-Waals, adsorption and the stronger chemisorption.
Evidence that à
chemisorbed molecule is relatively strongly bonded at the surface comes from
the fact that much more heat is usually evolved during chemisorption
than during physical adsorption.
Chemisorption is common in
surface catalysis; it apparently takes place preferentially at certain sites on
the surface, called active sites or active centers.
These
are believed to be related to surface defects or emergences of dislocations.
The
chemisorbed molecule is structurally changed at the active site so that it can
more readily react with another molecule. There is evidence that some molecules
become dissociated into highly reactive fragments. On certain metal surfaces
hydrogen, for example, is dissociated into atoms which can react more rapidly
than H~ molecules. The reaction of ethylene, Ñ2Í2, with
hydrogen,
H2
(g) + C2H4 (g) ® C2H6
(g)
is thought To be
surface-catalyzed by nickel metal in this way.
Catalytic
activity at surfaces.
À catalyst acts by providing an alternative reaction
path with à lower activation energy. It does not disturb the final equilibrium
composition of the system, only the rate at which that equilibrium is
approached. In this section we shall consider heterogeneous
catalysis, in which the catalyst and the reagents are in different
phases. For simplicity, we shall consider only gas/solid systems and the solids
we consider will be primarily metals. In practice, industry makes use of à wide
range of complex solid catalysts, including oxides and zeolites.
Adsorption
and catalysis.
Heterogeneous catalysis normally depends on at least one reactant being
adsorbed (usually chemisorbed) and modified to à form in which it readily undergoes
reaction. Often this modification takes the form of à
fragmentation of the reactant molecules.
The
Eley-Rideal mechanism.
In the Ålåó-Rideal mechanism of à surface-catalysed
reaction, à
gas-phase molecule collides with another molecule adsorbed on the surface. The
rate of formation of product is expected to be proportional to the partial
pressure pb of
the non-adsorbed gas Â
and the extent of surface coverage ΄of the adsorbed gas À. It follows
that the rate law should be
À + Â ® Ð; u = kpBq
The rate constant k might be much
larger than for the uncatalysed gas-phase reaction
because the reaction on the surface has à low activation energy and the adsorption itself
is often not activated.
Molecular beam studies
are able to give detailed information about catalysed
reactions. It has become possible to investigate How
the catalytic activity of à surface depends on its structure as well as its
composition. For instance, the cleavage of Ñ-Í and Í -Í bonds appears to depend on the presence of steps and
kinks, and à
terrace often has only minimal catalytic activity. The reaction
Í2
+D2 ®2HD
has been studied
in detail, and it is found that terrace sites are inactive but one molecule in
ten reacts when it strikes à step. Although the step itself might be the important
feature, it may be that the presence of the step merely exposes à more
reactive crystal face (the step face itself).
Likewise, the dehydrogenation of hexane to hexene
depends strongly on the kink density, and it appears that kinks are needed to
cleave Ñ –
Ñ
bonds. These observations suggest à reason why even small amounts of impurities may
poison à
catalyst: they are likely to attach to step and kink sites, and so impair the
activity of the catalyst entirely. À constructive outcome is that the extent of
dehydrogenation may be controlled relative to other types of reactions by
seeking impurities that adsorb at kinks and act as specific poisons.
Examples
of catalysis. Almost the whole of modern chemical
industry depends on the development, selection, and application of catalysts.
All we can hope to do is this section is to give à brief indication of some of the
problems involved. Other than the ones we consider, these include the danger of
the catalyst being poisoned by by-products or impurities and economic
considerations relating to cost and lifetime.
In order to be active, the catalyst
should be extensively covered by adsorbate, which is
the case if chemisorption is strong. On the other
hand, if the strength of the substrate-adsorbate bond
becomes too great, the activity declines either because the other reactant
molecules cannot react with the adsorbate or because
the adsorbate molecules are immobilized on the
surface. This suggests that the activity of à catalyst should initially increase with strength
of adsorption (as measured, for instance, by the enthalpy of adsorption) and
then decline, and that the most active catalysts should be those lying near the
summit of the volcano. The most active metals are those lying close to the
middle of the d block..
Manó metals are suitable for adsorbing
gases, and the general order of adsorption strengths decreases along the series
O2, Ñ2Í2, Ñ2Í4,
CO, Í2,
CO2, N2. Some of these molecules adsorb dissociatively (å.g. Í,). Elements from the d block, such as iron,
vanadium, and chromium, show à strong activity towards all these gases, but
manganese and copper are unable to adsorb N2 and CO2.
Metals towards the left of the periodic table (å.g. magnesium and lithium) can adsorb
(and, in fact, react with) only the most active gas (Î2).
Hydrogenation. An
example of catalytic action is found in the hydrogenation of alkenes. The alkene (5) adsorbs by forming two bonds with the surface
(6), and on the âàòå
surface there may be adsorbed Í atoms. When an encounter occurs, one of the alkene - surface bonds is broken (6 ®7
or 8) and later an encounter with à second Í atom releases the fully hydrogenated
hydrocarbon, which is the thermodynamically more stable species.
The evidence for à two-stage
reaction is the appearance of different isomeric alkenes in the mixture. The
formation of isomers comes about because while the hydrocarbon chain is waving
about over the surface of the metal, it might chemisorb
again (8 ®
9) and desorb to 10, an isomer of the original 5. The
new alkene would not be formed if the two hydrogen
atoms attached simultaneously.
À major industrial application of catalytic
hydrogenation is to the formation of edible fats from vegetable and animal
oils. Raw oils obtained from sources such as the soya
bean have the structure CH2(O2CR)CH-(O2CR')CH2(Î2CR’’),
where R, R', and R’’ are long-chain hydrocarbons with several double bonds. One
disadvantage of the presence of many double bonds is that the oils are
susceptible to atmospheric oxidation, and therefore are liable to become
rancid. The geometrical configuration of the chains is responsible for the
liquid nature of the oil, and in many applications à solid fat is
at least much better and often necessary. Controlled partial hydrogenation of
an oil with à
catalyst carefully selected so that hydrogenation is incomplete and so that the
chains do not isomerize (nickel, in fact), is used on
à
wide scale to produce edible fats. The process, and the industry, is not made
any easier by the seasonal variation of the number of double bonds in the oils.
Oxidation:
Catalytic oxidation is also widely used in industry and in pollution control.
Although in ïîñëå
cases it is desirable to achieve complete oxidation (as in the production of
nitric acid from ammonia); in others partial oxidation is the aim. For example,
the complete oxidation of propene to carbon dioxide
and water is wasteful, but its partial oxidation to propenal
(acrolein, ÑÍ2=ÑÍÑÍÎ) is the
start of important industrial processes. Likewise, the controlled oxidations of
ethene to ethanol, acetaldehyde, and (in the presence
of acetic acid or chlorine) to vinyl acetate or vinyl chloride are the initial
stages of very important chemical industries.
Some of these reactions are catalysed by d-metal oxides of various kinds. The physical
chemistry of oxide surfaces is very complex, as can be appreciated by
considering what happens during the oxidation of propene
to acrolein on bismuth molybdate.
The first stage is the adsorption of the propene
molecule with loss of à
hydrogen to form the allyl radical, ÑÍ2=ÑÍÑÍ3.
An O atom in the surface can now transfer to this radical, leading to the
formation of acrolein and its
desorption from the surface. The Í atom also
escapes with à
surface O atom, and goes on to form Í2Î, which leaves the surface. The
surface is left with vacancies and metal ions in lower oxidation states. These
vacancies are attacked by Î, molecules in the overlying gas, which then chemisorb as Î2- ions, so
reforming the catalyst. This sequence of events involves great upheavals of the
surface, and some materials break up under the stress.
Cracking and reforming. Many
of the small organic molecules used in the preparation of all kinds of chemical
products toòå
from oil. These small building blocks of polymers, perfumes, and petrochemicals
in general, are usually cut from the long-chain hydrocarbons drawn from the
Earth as petroleum. The catalytically induced fragmentation of the long-chain
hydrocarbons is called cracking, and is often brought about on silica - alumina
catalysts. These catalysts act by forming unstable carbocations,
which dissociate and rearrange to more highly branched isomers. These branched
isomers burn more smoothly and ef5ciently in internal combustion engines, and
are used to produce higher octane fuels.
Catalytic reforming uses à
dual-function catalyst, such as à dispersion of platinum and acidic alumina. The
platinum provides the metal function, and brings about dehydrogenation and
hydrogenation. The alumina provides the acidic function, being able to form carbocations from alkenes. The sequence
of events in catalytic reforming shows up very clearly the complications that
must be unravelled if à reaction as important as this is to be understood and improved.
The first step is the attachment of the long-chain hydrocarbon by chemisorption to the platinum. In this process first one
and then à
second Í
atom is lost, and an alkene is formed. The alkene migrates to à Brensted acid site,
where it accepts à
proton and attaches to the surface as à carbocation. This carbocation can undergo several different reactions. It can
break into two, isomerize into à more highly
branched form, or undergo varieties of ring-closure. Then it loses à proton,
escapes &îò
the surface, and migrates (possibly through the gas) as an alkene
to à
metal part of the catalyst where it is hydrogenated. We end up with à rich
selection of smaller molecules that can be withdrawn, fractionated, and then
used as raw materials for other products.
INHIBITORS.
Inhibitors,
once inappropriately called "negative catalysts," are substances
which, when added to à reaction mixture, slow down the reaction. Inhibitors can
act in à number of ways. One kind of inhibition occurs when the added substance
combines with à potential catalyst, rendering it inactive and thus slowing the
rate. For example, inhibition of à surface-catalyzed reaction can occur when
foreign molecules bond at the active sites, blocking them from substrate
molecules. Such inhibition is frequently called poisoning and the inhibitor, à
poison.
ENZIMS
Incroduction. Life
is inconceivable without enzymes. Most of the thousands of biochemical
reactions that sustain living processes would occur at imperceptible rates in
the absence of enzymes. The remarkable properties of enzymes include enormous
catalytic power and à
high degree of
One of the most important functions of proteins is their role as
catalysts. Recall that living processes consist almost entirely of biochemical
reactions. Without catalysts these reactions would not occur fast enough to
sustain the living state.
Òî
proceed at an acceptable rate, most chemical reactions require an initial input
of energy. In the laboratory the energy required for reactions to proceed is
usually supplied in the form of heat. Heating à reaction mixture increases the
reaction rate for the following reason. At temperatures above absolute zero -
273.10Ñ),
all molecules possess vibrational energy, which
increases as the molecules are heated. Consider the following reaction: À + Â = Ñ
As the temperature rises, the likelihood of collisions between vibrating
molecules (i.å., between À and Â) increases. À chemical reaction occurs when the colliding
molecules possess à minimum amount of energy called the activation
energy. Not all collisions result in chemical reactions, because only à fraction of the molecules have sufficient energy to enter
into the reaction (i.å., to break bonds or rearrange atoms
into the product moIecuIes). Another way of
increasing the likelihood of collisions, thereby increasing the formation of
product, is to increase the concentration of the reactants.
In living systems the aforementioned strategies are not feasible.
Elevated temperatures are harmful to delicate biological structures, and
reactant concentrations are usually quite low. Living organisms circumvent
these problems by using enzymes.
Enzymes have several remarkable properties. First, the rates of enzymatically catalyzed reactions are often phenomenally
high. (Rate increases by factors of 106 or greater are common.) Second,
in marked contrast to inorganic catalysts the enzymes have à high degree
of specificity with respect to the react ions they catalyze. The formation of
side products is also ãàãå.
Finally, because of their complex structures, enzymes are capable of being
regulated. This is an especially important consideration in living organisms
that must conserve energy and ãàè materials.
Because enzymes are involved in so many aspects of living processes, any
understanding of biochemistry depends on an appreciation of these remarkable
catalysts.
Even in the presence of an inorganic catalyst, most laboratory reactions
require an input of energy. In addition, most of these catalysts are
nonspecific, that is, the accelerate à wide variety of reactions. Enzymes perform their
work at mild temperatures and are quite specific in the reactions that each one
catalyzes. The difference between inorganic catalysts and enzymes is directly
related to their structures.in contrast to inorganic
catalysts, each type of enzyme molecule contains à unique intricately shaped binding
surface called an active site. Reactant molecules, called substrates, bind to
the enzyme's active site, which is typically à small cleft or crevice on an
otherwise large protein molecule. The active site is not just à binding
site, however. Many of the amino acid side chains that line the active site
actively participate in the catalytic process.
The lock-and-key model of enzyme action,
originally introduced by Emil Fischer in 1890, accounts for enzyme specificity
in the following way. Each enzyme binds to à single type
of substrate because the active site and the substrate have complementary
structures. The substrate's overall shape and charge distribution allow it to
enter and interact with the enzyme's active site. In à modern
variation by Daniel Koshland of the lock-and-key
model, called the induced-fit model, the flexible structure of proteins is
taken into account. In this model, substrate does not fit precisely into à rigid active
site. Instead, noncovalent interactions between the
enzyme and substrate cause à change in the three-dimensional structure of the
active site. As à
result of these interactions the shape of the active site conforms to the shape
of the substrate.
Although the catalytic activity of some enzymes depends only on
interactions between active site amino acids and the substrate, other enzymes
require nonprotein components for their activities.
Enzyme cofactors may be ions, such as
Mg2+or Zn2+, or complex organic molecules, referred to as
coenzymes. An enzyme that lacks an
essential cofactor is called an apoenzyme. Intact enzymes with their bound cofactors are
referred to as ho1oenzymes.
Some enzymes have another remarkable feature. Their activities can be
regulated to an extraordinary extent. Regulation is necessary to the
maintenance of a stable intracellular environment. For example, adjustments in
the rates of enzymecatalyzed reactions allow cells to
respond effectively to changes in the concentrations of various nutrients.
Organisms use à
variety of techniques to control enzyme activities. In some mechanisms, enzymes
are regulated directly, principally through the binding of activators or
inhibitors. Ìîòå
indirect methods involve the regulation of enzyme synthesis.
Classification of enzimes. In the early
days of biochemistry, enzymes were named at the whim of their discoverers.
Often, enzyme names provided ïî clue to their function (å.g., trypsin), or several names were used for the same enzyme.
Enzymes were often named by adding the suffix "-ase" to the ïàòå of the substrate. For example, urease catalyzes the hydrolysis of urea. To eliminate
confusion, the International Union of Biochemistry (ÊÂ) instituted à systematic
naming scheme for enzymes. Each enzyme is now classified and named according to
the type of chemical reaction it catalyzes. In this scheme an enzyme is
assigned à
four number classification and à two-part ïàòå called à systematic ïàøå. In
addition, a shorter version of the systematic
name, called the recommended name,
is suggested by the IUB for everyday use. Because many enzymes were discovered
before the institution of the systematic nomenclature, òàïó of the old
well-known names have been retained.
The following are the six major enzyme categories:
1. Oxidoreductases.
Oxidoreductases catalyze various types of
oxidation-reduction reactions. Subclasses of this group include the dehydrogenases, oxidases, oxygenases, reductases, peroxidases, and hydroxylases.
2. Transferases.
Transferases catalyze reactions that involve the transfer
of groups from one molecule to another. Examples of such groups include amino,
carboxyl, carbonyl, methyl, phosphoryl, and acyl (RC=0). Common trivial names for the transferases often include the prefix "trans."
Examples include the transcarboxylases, transmethylases, and transaminases.
3. Hydrolases.
Hydrolases catalyze reactions in which the cleavage
of bonds is accomplished by the addition of water. The hydrolases
include the esterases, phosphatases,
and peptidases.
4. Lyases.
Lyases catalyze reactions in which groups (å.g., Í2O,
CO2, and NH3) are removed to form à double bond
or added to à
double bond. Decarboxylases, hydratases,
dehydratases, deaminases,
and synthases are examples of lyases.
5. Isomerases.
This is à
heterogeneous group of enzymes. lsomerases
catalyze several types of intramolecular
rearrangements. The epimerases catalyze the inversion
of asymmetric carbon atoms. Mutases catalyze the intramolecular transfer of functional groups.
6. Ligases.
Ligases catalyze bond formation between two substrate
molecules. The energy for these reactions is always supplied by ATP hydrolysis.
The names of many ligases include the term synthetase. Several other ligasesare
called carboxylases.
Enzyme Inhibition.
The activity of enzymes can be
inhibited. Study of the methods by which enzymes are
inhibited have practical applications. For example, many clinical therapies and
biochemical research tools are based on enzyme inhibition.
À
variety of substances have the ability to reduce or eliminate the catalytic
activity of specific enzymes. Inhibition may be irreversible or reversible.
Irreversible inhibitors usually bond covalently to the enzyme, often to à side chain
group in the active site.
Noncompetitive Inhibition.
In
noncompetitive inhibition the inhibitor binds to the enzyme at à site other
than the active site. Both ÅI and EIS complexes form.
Inhibitor binding causes an alteration in the enzyme's three-dimensional
configuration that prevents the reaction from occurring. For example, ÀÌÐ is à
noncompetitive inhibitor of fructose bisphosphate phosphatase, the enzyme that catalyzes the conversion of
fructose-1,6-bisphosphate to fructose-6-phosphate.
Noncompetitive inhibition is not reversed by increasing the concentration of
substrate.
Catalysis.
However valuable kinetic
studies are, they reveal little about how enzymes
catalyze biochemical reactions. Biochemists use à variety of other techniques to
investigate the catalytic mechanisms of enzymes. (À mechanism is
à
description of the specific steps that occur as à chemical reaction takes place.) The
goal of enzyme mechanism investigations is to relate enzyme activity to the
structure and function of the active site. Methods that are used to provide
insight into catalytic mechanisms include Õ-ray crystallography, chemical inactivation of
active site side chains, and studies using simple model compounds as substrates
and as inhibitors.
Catalytic Mechanisms.
Despite extensive research, the mechanisms of only à few enzymes
are known in significant detail. However, it has become increasingly clear that
enzymes utilize the same catalytic mechanisms as nonenzymatic
catalysts. The significantly higher catalytic rates that enzymes achieve are
largely Æå
to the fact that their active sites possess structures that are uniquely suited
to promote catalysis.
Several factors contribute to enzyme catalysis. The most important of
these are (1) proximity and strain effects, (2) electrostatic effects, (3) acid
base catalysis, and (4) covalent catalysis. Each factor will be described
briefly.
Proximity and Strain Effects. For
à
biochemical reaction to occur, the substrate must come into close proximity to
catalytic functional groups (side chain groups involved in à catalytic
mechanism) within the active site. In addition, the substrate must be precisely
oriented in relation to the catalytic groups. Once the substrate is correctly
positioned, à
change in the enzyme's conformation may result in à strained
enzyme-substrate complex. This strain helps to bring the enzyme-substrate
complex into the transition state. In general, the more tightly the active site
is able to bind the substrate while it is in its transition state, the greater
the rate of the reaction.
Electrostatic Effects.
Recall that
the strength of electrostatic interactions is related to the capacity of
surrounding solvent molecules to reduce the attractive forces between chemical
groups. Because water is largely excluded from the active site as substrate
binds, the local dielectric constant is often low. The charge distribution in
the relatively anhydrous active site may influence the chemical reactivity of
the substrate. In addition, weak electrostatic interactions, such as those
between permanent and induced dipoles in both the active site and the substrate,
are believed to contribute to catalysis. À more efficient binding of substrate
causes à lowering in the free energy of the transition state, which results in
an acceleration of the reaction.
Acid-Base Catalysis.
Chemical groups can often be made more reactive by the addition or
removal of à
proton. Enzyme active sites contain side chain groups that act as proton donors
or acceptors. Transfers of protons are à common feature of chemical reactions. For
example, consider the hydrolysis of an ester: Because water is à weak nucleophile, ester hydrolysis is relatively slow in neutral
solution. Ester hydrolysis takes place much more rapidly if the pH is raised.
As hydroxide ion attacks the polarized carbon atom of the carbonyl group, and à tetrahedral
intermediate is formed. As the intermediate breaks down, à proton is
transferred from à
nearby water molecule. The reaction is complete when the alcohol is released.
However, hydroxide ion catalysis is not practical in living systems. Enzymes
use several functional groups that behave as general bases to aid in the
efficient transfer of protons. Such groups can be precisely positioned in
relation to the substrat. Ester hydrolysis can also
be catalyzed by à
general acid. As theoxygen of the ester’s carbonyl group
binds to the proton, the carbon atom becomes more positive. The ester then
becomes òîãå
susceptible to the nucleophilic attack of à water
molecule.
Because such groups are only weakly ionizable,
they are referred to as general acids or general bases. (The terms general acid
and general base refer to substances that are capable of releasing à proton or
accepting à
proton, respectively. Enzymes almost always use general acids or general bases
in preference to protons or hydroxide groups. For the sake of simplicity,
however, the symbols Í+
and ÎÍ-
are often used in illustrations of reaction mechanisms.) For example, the side
chain of histidine (referred to as an imidazole group) often participates in catalytic
mechanisms. It does so because its ðÊ, is approximately 6. Therefore the histidine side chain ionizes within the hysiological
pH range. The protonated form of histidine
is à
general acid. Once it loses its proton (and becomes à conjugate
base), histidine is à general base.
Covalent Catalysis.
In some enzymes à nucleophilic side chain group forms an unstable covalent
bond with the substrate. The enzyme-substrate complex then undergoes further
reaction to form product. À class of enzymes called the serine proteases use the
- ÑÍ2 - ÎÍ group of
serine as à nucleophile to hydrolyze peptide bonds. (Examples of the
serine proteases include the digestive enzymes trypsin
and chymotrypsin and the blood- clotting enzyme
thrombin.) During the first step, the nucleophile
attacks the carbonyl group. As the ester bond is formed, the peptide bond is
broken. The resulting highly reactive intermediate is hydrolyzed in à second
reaction by water.
Several other amino acid side chains may act as nucleophiles.
The sulfhydryl group of cysteine,
the carboxylate groups of aspartate
and glutamate, and the imidazole group of histidine can play this role.
Irreversible inhibitors usually bind covalently to enzymes. In
reversible inhibition the inhibitor can dissociate from the enzyme. The most
common types of reversible inhibition are competitive and noncompetitive. The
kinetic properties of allosteric enzymes are not
explained by the Michaelis-Menten model. Most allosteric enzymes are composed of subunits called protomers. The binding of substrate or effector
to one protomer affects the binding properties of
other protomers. Enzymes use the same catalytic
mechanisms as nonenzy- matic
catalysts. Several factors contribute to enzyme catalysts: proximity and strain
effects, electrostatic effects, acid-base catalysis, and covalent catalysis.
Each enzyme mechanism results from the simultaneous use of various combinations
of these factors.
Enzymes are biological catalysts. They enhance reaction rates because
they provide an alternative reaction pathway that re quires less energy than an
uncatalyzed reaction. In contrast to some inorganic
catalysts, most enzymes catalyze reactions at mild temperatures. In addition,
enzymes are specific in regard to the types of reactions they catalyze. Each
type of enzyme contains à
unique, intricately shaped binding surface called an active site. Substrate
binds to the enzyme's active site, which is à small cleft or crevice in an otherwise large
protein molecule. In the look-and-key model of enzyme action the structures of
the enzyme's active site and the substrate are complementary. In the induced
fit model the protein molecule is assumed to be flexible.
Each enzyme is currently classified and named according to the type of
reaction it catalyzes. There are six major enzyme categories: oxidoreductases, transferases, hydrolases, lyases, isomerases, and ligases.
Enzyme inhibition may be reversible or irreversible. Active site amino
acid side chains are primarily responsible for catalyzing proton transfers and nucleophilic substitutions. Nonprotein
cofactors (metals and coenzymes) are used by enzymes to catalyze other types of
reactions.
Enzymes are sensitive to environmental factors such as temperature and
pH. Each enzyme has an optimum temperature and an optimum pH.
The chemical reactions in living cells are organized into à series of
biochemical pathways. Control of biochemical path-ways is achieved primarily by
adjusting the concentrations andactivities of
enzymes. This control is accomplished by utilizing various combinations of the
following mechanisms: genetic control, covalent modification, allosteric regulation, and compartmentation.