Sei sulla pagina 1di 27

“MODELS GOVERNING CHEMICAL KINETICS”

NAME : SHASHI PAUL


REG.NO. : 11006142
ROLL.NO : RP8003-B-15
COURSE : MSc.MICROBIOLOGY
COURSE CODE : 2403
SUBJECT CODE : BTY602

Submitted to:
Dr. Avadh Kumar.
ACKNOWLEGMENT

It has been a great challenge but a plenty of learning


and opportunities to gain huge knowledge on the way preparing
this term paper. I would not succeed without my teacher
Dr.Avadh kumar , who seemed to be with me always; and
prepared to give me feedback and guidelines whenever I needed
it.
Thank You
Sir!
I also would like to thank all my friends.
I hope you will find my working as interesting and
knowledge earning. And it will be useful for others wanting to
learn about apparel industry and retailers’ policies and
strategies about planning.

SHASHI PAUL................
CONTENTS

• ABSTRACT
• CHEMICAL KINETICS
• FACTORS AFFECTING REACTION RATE
• REACTION MECHANISM
• CHEMICAL KINETICS – REACTION RATE
• INSTANTANEOUS RATES OF REACTION
• RATE LAWS & RATE CONSTANTS
• COLLISION THEORY MODEL OF CHEMICAL REACTIONS
• KINETICS
• MODELING CHEMICAL KINETICS WITH STELLA
• A KINETIC MODEL OF CHxCL4-x/CH4 COMBUSTION
• EXGAS
• THE LEADS METHANE OXIDATION MECHANISM
ABSTRACT
The differential equations governing the propagation in time of the sensitivity
matrix for a mathematical model given by a system of ordinary differential
equations are derived. These equations are used to perform a statistical sensitivity
analysis of models for chemical reactors. The behavior of the sensitivities at
equilibrium is analyzed. It is shown that the sensitivity equations for linear kinetics
may be solved using an analytic representation. The numerical solution of these
equations is discussed, and illustrative examples are presented. The lognormal
distribution is presented as being representative of errors in rate constants.

CHEMICAL KINETICS INTRODUCTION

Reaction rate tends to increase with concentration - a phenomenon explained by


collision theory.

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
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 influences the rate of its transformation into
products. The reactions which involve lesser bond rearrangement proceed faster
than the reactions which involve larger bond rearrangement.

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
only occur 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 doesn't 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
Concentration plays a very important role in reactions, because according to the
collision theory of chemical reactions, molecules must collide in order to react
together. As the concentration of the reactants increases, the frequency of the
molecules colliding increases, striking each other more frequently by being in
closer contact at any given point in time. Think of two reactants being in a closed
container. All the molecules contained within are colliding constantly. By
increasing the amount of one or more of the reactants it causes these collisions to
happen more often, increasing the reaction rate

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 10 °C
temperature rise is a common misconception. This may have been generalized
from the special case of biological systems, where the Q10 (temperature coefficient)
is often between 1.5 and 2.5.

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 rate of an
equilibrium process.

CATALYSTS
Generic potential energy diagram showing the effect of a catalyst in an
hypothetical endothermic chemical reaction. The presence of the catalyst opens a
different reaction pathway (shown in red) with a lower activation energy. The final
result and the overall thermodynamics are the same.
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.
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 generally 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.

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.
REACTION MECHANISM

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.

CHEMICAL KINETICS – REACTION RATES


Chemical kinetics is the branch of chemistry which addresses the question: "how
fast do reactions go?" Chemistry can be thought of, at the simplest level, as the
science that concerns itself with making new substances from other substances.
Or, one could say, chemistry is taking molecules apart and putting the atoms and
fragments back together to form new molecules. (OK, so once in a while one uses
atoms or gets atoms, but that doesn't change the argument.) All of this is to say
that chemical reactions are the core of chemistry.

Here are some examples. Consider the reaction,


2 H2(g) + O2(g) → 2 H2O(l).
We can calculate ΔrGo for this reaction from tables of free energies of formation
(actually this one is just twice the free energy of formation of liquid water). We
find that ΔrGo for this reaction is very large and negative, which means that the
reaction wants to go very strongly. A more scientific way to say this would be to
say that the equilibrium constant for this reaction is very very large.

However, we can mix hydrogen gas and oxygen gas together in a bulb or other
container, even in their correct stoichiometric proportions, and they will stay there
for centuries, perhaps even forever, without reacting. (If we drop in a catalyst -
say a tiny piece of platinum - or introduce a spark, or even illuminate the mixture
with sufficiently high frequency uv light, or compress and heat the mixture, the
mixture will explode.) The problem is not that the reactants do not want to form
the products, they do, but they cannot find a "pathway" to get from reactants to
products.
REACTION RATES
Consider the reaction,

2 NO(g) + O2(g) → 2 NO2(g).


We can specify the rate of this reaction by telling the rate of change of the partial
pressures of one the gases. However, it is convenient to convert these pressures
into concentrations, so we will write our rates and rate equations in terms of
concentrations, where square brackets, [ ], mean concentration in mol/L.

We might try to write the rate variously as,

or as

but these are not the same because each molecule of O2 gives two molecules of
NO2. To arrive at an unambiguous definition of reaction rate we define the
"reaction velocity," v, as

(1)
This is unambiguous. The negative sign tells us that that species is being consumed
and the fractions take care of the stoichiometry. Any one of the three derivatives
can be used to define the rate of the reaction.

For a general reaction,

aA + bB → cC + dD, (2)
the reaction velocity can be written in a number of different but equivalent ways,

(3)
As in our previous example, the negative signs account for material that is being
consumed in the reaction and the positive signs account for material that is being
formed in the reaction. The stoichiometry is preserved by dividing the rate of
change of concentration of each substance by its stoichiometric coefficient.
Instantaneous Rates of Reaction and the Rate Law for a Reaction
The rate of the reaction between phenolphthalein and the OH- ion isn't constant; it
changes with time. Like most reactions, the rate of this reaction gradually
decreases as the reactants are consumed. This means that the rate of reaction
changes while it is being measured.

To minimize the error this introduces into our measurements, it seems advisable to
measure the rate of reaction over periods of time that are short compared with the
time it takes for the reaction to occur. We might try, for example, to measure the
infinitesimally small change in concentration d(X) that occurs over an
infinitesimally short period of time dt. The ratio of these quantities is known as
the instantaneous rate of reaction.

The instantaneous rate of reaction at any moment in time can be calculated from a
graph of the concentration of the reactant (or product) versus time. The graph
below shows how the rate of reaction for the decomposition of phenolphthalein
can be calculated from a graph of concentration versus time. The rate of reaction
at any moment in time is equal to the slope of a tangent drawn to this curve at that
moment.

The instantaneous rate of reaction can be measured at any time between the
moment at which the reactants are mixed and the reaction reaches equilibrium.
Extrapolating these data back to the instant at which the reagents are mixed gives
the initial instantaneous rate of reaction.

Rate Laws and Rate Constants


An interesting result is obtained when the instantaneous rate of reaction is
calculated at various points along the curve in the graph in the previous section.
The rate of reaction at every point on this curve is directly proportional to the
concentration of phenolphthalein at that moment in time.

Rate = k(phenolphthalein)

Because this equation is an experimental law that describes the rate of the
reaction, it is called the rate law for the reaction. The proportionality constant, k,
is known as the rate constant.

Different Ways of Expressing the Rate of Reaction


There is usually more than one way to measure the rate of a reaction. We can
study the decomposition of hydrogen iodide, for example, by measuring the rate at
which either H2 or I2 is formed in the following reaction or the rate at which HI is
consumed.

2 HI(g) H2(g) + I2(g)

Experimentally we find that the rate at which I2 is formed is proportional to the


square of the HI concentration at any moment in time.

What would happen if we studied the rate at which H2 is formed? The balanced
equation suggests that H2 and I2 must be formed at exactly the same rate.

What would happen, however, if we studied the rate at which HI is consumed in


this reaction? Because HI is consumed, the change in its concentration must be a
negative number. By convention, the rate of a reaction is always reported as a
positive number. We therefore have to change the sign before reporting the rate of
reaction for a reactant that is consumed in the reaction.

The negative sign does two things. Mathematically, it converts a negative change
in the concentration of HI into a positive rate. Physically, it reminds us that the
concentration of the reactant decreases with time.

What is the relationship between the rate of reaction obtained by monitoring the
formation of H2 or I2 and the rate obtained by watching HI disappear? The
stoichiometry of the reaction says that two HI molecules are consumed for every
molecule of H2 or I2 produced. This means that the rate of decomposition of HI is
twice as fast as the rate at which H2 and I2 are formed. We can translate this
relationship into a mathematical equation as follows.

As a result, the rate constant obtained from studying the rate at which H2 and I2
are formed in this reaction (k) is not the same as the rate constant obtained by
monitoring the rate at which HI is consumed (k')

The Rate Law Versus the Stoichiometry of a Reaction


In the 1930s, Sir Christopher Ingold and coworkers at the University of London
studied the kinetics of substitution reactions such as the following.

CH3Br(aq) + OH-(aq) CH3OH(aq) + Br-(aq)

They found that the rate of this reaction is proportional to the concentrations of
both reactants.

Rate = k(CH3Br)(OH-)

When they ran a similar reaction on a slightly different starting material, they got
similar products.
(CH3)3CBr(aq) + OH-(aq) (CH3)3COH(aq) + Br-(aq)

But now the rate of reaction was proportional to the concentration of only one of
the reactants.

Rate = k((CH3)3CBr)

These results illustrate an important point: The rate law for a reaction cannot be
predicted from the stoichiometry of the reaction; it must be determined
experimentally. Sometimes, the rate law is consistent with what we expect from the
stoichiometry of the reaction.

2 HI(g) H2(g) + I2(g) Rate = k(HI)2

Often, however, it is not.

2 N2O5(g) 4 NO2(g) + O2(g) Rate = k(N2O5)

Collision Theory Model of Chemical Reactions


The collision theory model of chemical reactions can be used to explain the
observed rate laws for both one-step and multi-step reactions. This model assumes
that the rate of any step in a reaction depends on the frequency of collisions
between the particles involved in that step.

The figure below provides a basis for understanding the implications of the
collision theory model for simple, one-step reactions, such as the following.

ClNO2(g) + NO(g) NO2(g) + ClNO(g)

The kinetic molecular theory assumes that the number of collisions per second in a
gas depends on the number of particles per liter. The rate at which NO2 and ClNO
are formed in this reaction should therefore be directly proportional to the
concentrations of both ClNO2 and NO.

Rate = k(ClNO2)(NO)

The collision theory model suggests that the rate of any step in a reaction is
proportional to the concentrations of the reagents consumed in that step. The rate
law for a one-step reaction should therefore agree with the stoichiometry of the
reaction.

The following reaction, for example, occurs in a single step.

CH3Br(aq) + OH-(aq) CH3OH(aq) + Br-(aq)

When these molecules collide in the proper orientation, a pair of nonbonding


electrons on the OH- ion can be donated to the carbon atom at the center of the
CH3Br molecule, as shown in the figure below.

When this happens, a carbon-oxygen bond forms at the same time that the carbon-
bromine bond is broken. The net result of this reaction is the substitution of an OH-
ion for a Br - ion. Because the reaction occurs in a single step, which involves
collisions between the two reactants, the rate of this reaction is proportional to the
concentration of both reactants.

Rate = k(CH3Br)(OH-)

Not all reactions occur in a single step. The following reaction occurs in three
steps, as shown in the figure below.

(CH3)3CBr(aq) + OH-(aq) (CH3)3COH(aq) + Br-(aq)


In the first step, the (CH3)3CBr molecule dissociates into a pair of ions.

First step

The positively charged (CH3)3C+ ion then reacts with water in a second step.

Second step

The product of this reaction then loses a proton to either the OH- ion or water in
the final step.

Third step

The second and third steps in this reaction are very much faster than first.
(CH3)3CBr (CH3)3C+ + Br- Slow step
(CH3)3C+ + H2O (CH3)3COH2+ Fast step
(CH3)3COH2+ + OH- (CH3)3COH + H3O Fast step

The overall rate of reaction is therefore more or less equal to the rate of the first
step. The first step is therefore called the rate-limiting step in this reaction
because it literally limits the rate at which the products of the reaction can be
formed. Because only one reagent is involved in the rate-limiting step, the overall
rate of reaction is proportional to the concentration of only this reagent.

Rate = k((CH3)3CBr)

The rate law for this reaction therefore differs from what we would predict from
the stoichiometry of the reaction. Although the reaction consumes both (CH3)3CBr
and OH-, the rate of the reaction is only proportional to the concentration of
(CH3)3CBr.

The rate laws for chemical reactions can be explained by the following general
rules.

• The rate of any step in a reaction is directly proportional to the


concentrations of the reagents consumed in that step.
• The overall rate law for a reaction is determined by the
sequence of steps, or the mechanism, by which the reactants are
converted into the products of the reaction.
• The overall rate law for a reaction is dominated by the rate law
for the slowest step in the reaction.
Kinetics

Kinetics is the area of chemistry concerned with reaction rates. The rate can be
expressed as:
rate = change in substance/time for change to occur (usually in M/s)

There are several factors that determine the rate of a specific reaction and those
are expressed in the "collision theory" that states that for molecules to react, they
must:

1. collide
2. have the right energy
3. have the right geometry

To increase the rate, you must make the above more likely to occur. This is
possible by changing other factors such as:

• increasing the surface area (of solids)-this allows for more collisions and
gives more molecules the right geometry
• increasing the temperature-this gives more molecules the right energy (also
called the activation energy, Ea)

• increasing the concentration (of gases and solutions)-this allows for more
collisions and more correct geometry
• using a catalyst-helps molecules achieve the correct geometry by providing
a different way to react
A catalyst is a substance added to a reaction that comes out of the reaction
unchanged. As mentioned earlier, catalysts help lower the activation energy as
shown in the following graph. They do this by changing the reaction mechanism.

Modeling Chemical Kinetics with Stella


The models below were contributed by Shawn Sendlinger, Associate Professor of
Chemistry at North Carolina Central University. The basics of each of the models
is described below. The Stella models are in a zip file that can be downloaded by
clicking on the link below. You will need Winzip for Windows or Stuffit Expander
for the Macintosh to unpack the file.

Chemical Kinetics using STELLA

A simple chemical equation can be described as:

A→B

The rate at which this reactions proceeds depends on a number of things. These
factors include k (the rate constant), and Ao (the initial amount of compound A). In
differential form, the rate equations are:
These equations can also be written as integrated rate equations:

The STELLA software program can readily be used to build a model of this
reaction. The stock icon is used to represent the chemical species “A” and “B”. A
flow icon represents the conversion of stock A into stock B. A converter icon is
used to hold a value for the rate constant k. Finally, the connector icon is used to
show the mathematical dependence shown in the above equations.

The STELLA model (SimpleAtoB.STM) would look like:

<!--[if !supportEmptyParas]--> <!--[endif]-->

Notice that the flow out of stock “A” represents a decrease in the stock, so the
negative sign from the above equation is taken into account. Similarly, the flow
into stock “B” is automatically considered an increase (a positive sign in the
above equation). To give good results that reflect those determined experimentally,
the best integration method available (Runge-Kutta 4) should be used in
conjunction with a small “dt” value (0.01 or smaller). Try setting the initial
amount of “A” to 1000, “B” to zero, and a small value for the rate constant k (1 x
10-3). A graph of A will show the expected exponential decrease, while the graph of
B will show the expected exponential increase.

The above model can be easily modified in a number of ways. These are
enumerated below.

Modification #1: Equilibrium(A ↔ B)

In addition to the forward reaction where “A” turns into “B”, now the reverse is
also true: “B” can turn back into “A”. Now we have forward (kf) and reverse (kr)
rate constants. The differential form
In addition to the forward reaction where “A” turns into “B”, now the reverse is
also true: “B” can turn back into “A”. Now we have forward (kf) and reverse (kr)
rate constants. The differential form of the rate equations look like:

The above STELLA model now includes a second flow in the opposite direction, as
well as a second converter for the additional rate constant. This model is shown
below (AequilB):

To check if this model is working properly, the sum of “A” and “B” at any time
should be a constant.

Modification #2: Consecutive Reactions (A → B → C)

It is quite easy to modify the first model to include an additional reaction step by
adding another stock and another flow. A second converter for the new rate
constant is also required. The equations for the consecutive reaction model look
like (A to B to C):

>

The model is shown here:


Again, note that the direction of the flows automatically takes care of the signs
from the differential equations shown above.

If needed, each of the reactions steps in this model could be made reversible.
Additional chemical species could also be added (“D”, “E”, “F”, etc.).

Modification #3: Parallel Reactions from One Reactant

Sometimes, a given reactant might form two different products, as shown below:

A → BandA → C

In differential form, the rate equations are:

The STELLA model will now have two flows originating from the “A” stock, and
each flow will have its respective rate constant

The STELLA model will now have two flows originating from the “A” stock, and
each flow will have its respective rate constant:

Modification #4: Parallel Reactions to Form One Product


This situation is opposite to that of Modification #3 shown above. In this case,
reactants “A” and “B” combine together to form one product “C”:

A→C←B

The rate equations are:

The STELLA model is shown below: The STELLA model is shown below:

Be careful to notice the directions of the flows in the above model.

Summary:

STELLA is a convenient and easily understood tool for constructing and studying
chemical kinetics. All of the simple models here can have “real” chemical names
substituted for “A”, “B”, etc., and experimentally determined rate constants can
be entered. The resulting models accurately predict the concentration of the
various chemical species at any time of interest. In their “general” form, the
models can be experimented with to gain an understanding of the important
factors that govern kinetics.
A KINETIC MODEL OF CHxCl4-x/CH4 COMBUSTION
Laminar Flame Speeds and Oxidation Kinetics of Tetrachloromethane

ABSTRACT
The laminar flame speeds of tetrachloromethane (CCl4) with methane in air at
room temperature and atmospheric pressure were experimentally determined
using the counterflow twin-flame technique, varying both the amount of CCl4 in the
fuel and the equivalence ratio of the unburned mixture. Comparison between the
experimental results and the previous data of CH3Cl-, CH2Cl2-, and CHCl3-CH4-
air flames demonstrates the dominant influence of the atomic Cl-to-H ratio on the
propagation rate of laminar flames with chlorinated methane addition. A detailed
kinetic model previously employed for CH3Cl, CH2Cl2, and CHCl3 combustion was
expanded to include additional pathways pertinent to tetrachloromethane
combustion. Numerical simulation shows that the model predicts the laminar flame
speeds reasonably well. Carbon flux and sensitivity analyses indicate that the
oxidation kinetics of CH4 flames doped with CCl4 are essentially the same as those
doped with other chloromethanes.

1.
EXGAS : Automatic generation of detailed and reduced mechanism
The development of validated and reliable kinetic models to represent the
oxidation and the combustion of organic compounds is of particular interest. For
instance, in the case of spark ignited engines, this type of model could help the
research works aiming at formulating gasolines which present optimal octane
number properties and which lead to minimal pollutants formation. In the purpose
of modelling the combustion of an organic compound or, what is the actual case in
most practical applications, of a mixture of organic compounds, in the large
temperature field which can be observed in an engine, it is necessary to consider
several thousands of reactions. Since automatic procedure would be a convenient
and rigorous way to write such huge mechanisms, EXGAS was designed for a
computer aided construction of mechanisms.

CO-C2 Reactions
DataBase
Detailed Primary
Mechanism
Lumped Primary
Mechanism
Lumped Secondary
Mechanism

Corresponding kinetic data


are calculated by KINGAS
or estimated by
correlations
Global of scheme EXGAS Description
THE LEEDS METHANE OXIDATION MECHANISM
The common feature of all published mechanisms is that although they are up-to-
date at the time of compilation, they might become out-of-date at the time of
publication and they surely become more-and-more out-of-date as the time passes.
Here we are trying to maintain a methane oxidation mechanism that is always up-
to-date. We are inviting everyone to criticise this mechanism (we prefer private e-
mails to public criticism). This mechanism will be continously updated according
to the suggestions from the chemical kinetic/combustion community and whenever
new information about reactions becomes available.

In order to help checking the information in the mechanism, each reaction is


annotated with a classification of the rate data (which spans from A (well know
data) to U (entirely uncertain data)), an estimate of the uncertainty, the
temperature interval of recommendation, and a reference to the origin of data.

A manuscript has been published in Int.J.Chem.Kinet. that contains a description


of the mechanism and its testing and a comparison of selected parameters with the
GRI, Warnatz, and Konnov mechanisms.

The methane mechanism has since been extended to include nitrogen and sulphur
containing species.

The most recent version of the Leeds Methane Oxidation Mechanism is version 1.5
Earlier versions can be found in the Archive.

REFERENCE
• Preparing for the Chemistry AP Exam. Upper Saddle River, New Jersey:
Pearson Education, 2004. 131–134. ISBN 0-536-73157-8
• K.J. Laidler, Chemical Kinetics (3rd Ed.), Benjamin Cummings, San
Francisco, 1987.
• Model Files to Download: kinetics
• http://en.wikipedia.org/wiki/Chemical_kinetics"
• Alberty, R. A., Advances in Enzymol., 17, I (1956).
• Koshland, D. E., Jr., in McElroy, W. D., and Glass, B., The mechanism of
enzyme action, Baltimore, 608 (1954).
• Chance, B., in Friess, S. L., and Weissberger, A., Investigations of rates and
mechanisms of reactions, New York, 627 (1953).
• Kistiakowsky, G. B., and Lumry, It., J. Am. Chem. Sot., ‘71, 2006 (1949).
• Maier, V. P., Tapper, A. L., and Volman, D. H., J. Am. Chem. Sac., 77, 1278
(1955).
• Malmstrom, B. G., Biochim. et biophys. acta, 18, 285 (1955).
• Gutfreund, H., Discussions Faraday Sot., 20, 167 (1955).
• Roughton, F. J. W., and Chance, B., in Friess, S. L., and Weissberger, A.,
Investigations of rates and mechanisms of reactions, New York, 669 (1953).
• Amis, E. S., Kinetics of chemical change in solution, New York (1949).
• Hammett, L. P., Physical organic chemistry, New York (1940).
• Malmstrom, B. G., Arch. Biochem. and Biophys., 46, 345 (1953).
• Malmstriim, B. G., The mechanism of metal-ion activation of enzymes;
studies on enolase, Uppsala (1956).
• Malmstriim, B. G., Arch. Biochem. and Biophys., 70, 58 (1957).

Potrebbero piacerti anche