23.03.2013 Views

I. Rates and Rate Laws Definition of Reaction Rate: So far we have ...

I. Rates and Rate Laws Definition of Reaction Rate: So far we have ...

I. Rates and Rate Laws Definition of Reaction Rate: So far we have ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

I. <strong><strong>Rate</strong>s</strong> <strong>and</strong> <strong>Rate</strong> <strong>Laws</strong><br />

<strong>Definition</strong> <strong>of</strong> <strong>Reaction</strong> <strong>Rate</strong>:<br />

<strong>So</strong> <strong>far</strong> <strong>we</strong> <strong>have</strong> talked about the thermodynamics <strong>of</strong> process <strong>and</strong> reactions, as <strong>we</strong>ll as their<br />

equilibrium. We now want to consider the timescales with which chemical <strong>and</strong> physical<br />

processes occur. The study <strong>of</strong> the rates <strong>of</strong> reactions is called kinetics.<br />

The first thing <strong>we</strong> <strong>have</strong> to do is define a rate, or velocity, <strong>of</strong> reaction. This can be expressed<br />

as the time derivative <strong>of</strong> the concentration for a particular component. In this formalism, the<br />

velocity <strong>of</strong> the reaction is related to a change in concentration with time:<br />

dc<br />

v= = rate<br />

dt<br />

Note that for a particular reaction, the concentrations <strong>of</strong> each component are coupled, so <strong>we</strong><br />

can express the rate in terms <strong>of</strong> the time rate <strong>of</strong> change for any reactant or product. In other<br />

words, if a reactant transforms into product, the concentrations at any point in the reaction<br />

are related by their stoichiometric coefficients. As a result, for the reaction:<br />

aA + bB → cD + dD<br />

The rate can be expressed in terms <strong>of</strong> any component:<br />

−1 dA [ ] −1<br />

dB [ ] 1 dC [ ] 1 dD [ ]<br />

rate = = = =<br />

a dt b dt c dt d dt<br />

Note that the negative derivative for the reactants corresponds to their disappearance <strong>and</strong> the<br />

positive derivative for the products corresponds to their appearance.<br />

<strong>Rate</strong> <strong>Laws</strong>:<br />

We can also express the rate in terms <strong>of</strong> the concentrations <strong>of</strong> the stoichiometric reactants<br />

<strong>and</strong> products (i.e. those that appear in the net reaction.) In this way, the reaction rate can be<br />

thought <strong>of</strong> in terms <strong>of</strong> collisions – if there is a higher concentration <strong>of</strong> a particular<br />

component, then it is more likely to react, but there is some proportionality that takes into<br />

account physical parameters <strong>of</strong> the system. We will discuss this later. For now, let’s<br />

represent the reaction rate in terms <strong>of</strong> a rate law.<br />

Example – consider the reaction:<br />

The rate law for this reaction is:<br />

A+ B→ P<br />

m n q<br />

rate =<br />

k[ A] [ B] [ P]<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 1 <strong>of</strong> 33


Where k is the rate constant <strong>and</strong> m , n <strong>and</strong> q are reaction orders with respect to a<br />

particular component. For example, if m = 2 <strong>and</strong> n = 1,<br />

<strong>we</strong> say that the reaction is second<br />

order in A <strong>and</strong> first order in B. <strong>Reaction</strong> orders are typically integers (i.e. 0, 1, 2, …), but<br />

they could be fractions.<br />

Since the rate <strong>of</strong> reaction has units <strong>of</strong> concentration/time (M/s), the units for k are:<br />

1 1<br />

k[<br />

] x 1<br />

s M −<br />

⎛ ⎞⎛ ⎞<br />

= ⎜ ⎟⎜ ⎟<br />

⎝ ⎠⎝ ⎠<br />

where x is the total reaction order, or the sum <strong>of</strong> the individual reaction orders.<br />

II. Temperature Dependence <strong>of</strong> <strong>Rate</strong> Constants<br />

Preliminaries – Arrhenius equation:<br />

We would now like to consider the temperature dependence <strong>of</strong> the rate constants in our<br />

reactions. Qualitatively, <strong>we</strong> expect that the reaction will occur with a faster rate if <strong>we</strong><br />

increase temperature, i.e.: there will be more collisions, thus a faster reaction. It has been<br />

found empirically that the rate constants for reactions follow the Arrhenius Equation:<br />

Ea/ RT<br />

k Ae −<br />

=<br />

where A is called the pre-exponential factor (<strong>we</strong>akly T-dependent), Ea is called the<br />

activation energy <strong>and</strong> R = 8.314 J/mol K is the universal gas constant.<br />

The Arrhenius equation can be rearranged into:<br />

Ea<br />

ln k = ln A− RT<br />

The rate constants at two different temperatures are related by the expression:<br />

1 1<br />

ln<br />

E k ⎛ ⎞ ⎛ ⎞<br />

=− −<br />

2<br />

a<br />

⎜ ⎟ ⎜ ⎟<br />

⎝ k1 ⎠ R ⎝T2 T1⎠<br />

<strong>and</strong> if you <strong>have</strong> the rate constants at many temperatures, you can plot ln k vs. 1/T to obtain<br />

a linear relationship where the slope is equal to − E / R <strong>and</strong> the intercept is equal to ln A .<br />

The activation energy is typically on the order <strong>of</strong> 50 – 200 kJ/mol (<strong>we</strong>aker than a chemical<br />

bond) <strong>and</strong> the pre-exponential factor is on the order <strong>of</strong> 10 10 – 10 14 s -1 , <strong>and</strong> is related to the<br />

collision cross section. We will discuss this later.<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 2 <strong>of</strong> 33<br />

a


Energetics <strong>of</strong> activation – changing reaction rate with temperature:<br />

We first want to underst<strong>and</strong> what the activation energy is <strong>and</strong> how <strong>we</strong> can affect the rate <strong>of</strong> a<br />

reaction. We can express energetic changes throughout a reaction on a potential energy (or a<br />

reaction coordinate) diagram<br />

The reaction rate is dependent on the height <strong>of</strong> the activation barrier. By increasing the<br />

temperature <strong>of</strong> the reaction, the thermal energy <strong>of</strong> the system increases <strong>and</strong> it becomes more<br />

probable that a particular collision can overcome the activation barrier, resulting in a faster<br />

reaction, consistent with empirical observations. Be careful, though, because the equilibrium<br />

conditions depend on the thermodynamic changes, independent <strong>of</strong> the activation barrier!<br />

Catalysis – changing reaction rate by changing Ea:<br />

Having defined the activation energy, <strong>we</strong> can consider ways to manipulate the reaction rate<br />

by decreasing the activation barrier. One common way to speed up a reaction is to add a<br />

catalyst. Energetically, this serves to decrease the activation barrier, making it more<br />

probable that a particular collision can overcome the lo<strong>we</strong>red barrier. Mechanistically, a<br />

catalyst is a substance that is consumed in an early step in the mechanism <strong>and</strong> produced in a<br />

later step. A catalyst may appear in the rate law even though it is not necessarily a<br />

stoichiometric product or reactant, <strong>and</strong> it has been observed that reaction rates do typically<br />

depend on the catalyst concentration.<br />

One example is the Michaelis-Menten mechanism <strong>of</strong> enzyme catalysis. The mechanism is<br />

as follows:<br />

E + S←⎯⎯ ⎯⎯→ ES<br />

k1<br />

k−1<br />

k2<br />

ES ⎯⎯→ E+ P<br />

Notice in this mechanism that the net reaction is S → P,<br />

where the enzyme E is the catalyst<br />

(consumed initially, then formed) <strong>and</strong> ES is the intermediate. Physically, ES is a complex<br />

bet<strong>we</strong>en enzyme <strong>and</strong> substrate.<br />

The rate law can be determined using the SSA (since <strong>we</strong> <strong>have</strong> no information about relative<br />

rates):<br />

kk 1 2 rate = [ E][ S]<br />

k + k<br />

−1<br />

2<br />

This rate law is first order in catalyst concentration. The total enzyme (catalyst)<br />

concentration occurs either as bound or unbound, <strong>and</strong> the total amount is conserved, so<br />

[ E] = [ E] +<br />

[ ES]<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 3 <strong>of</strong> 33<br />

0


This relation can be used in the SSA step to obtain a more useful form:<br />

where V k2[ E]<br />

0<br />

= <strong>and</strong> ( )<br />

−1<br />

2 1<br />

VS [ ]<br />

rate =<br />

K + [ S]<br />

K = k + k / k is the Michaelis Constant.<br />

m<br />

Accounting for the pre-exponential factor – collision theory:<br />

We now want to turn our attention to the physical meaning <strong>of</strong> the pre-exponential factor A .<br />

This can be described by the kinetic theory for collisions. Consider the simple bimolecular<br />

reaction in the gas phase:<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 4 <strong>of</strong> 33<br />

m<br />

k<br />

A+ B⎯⎯→ P<br />

This reaction rate constant is determined by two parameters: A <strong>and</strong> Ea. In order for a<br />

reaction to occur, A <strong>and</strong> B molecules <strong>have</strong> to collide with enough energy to overcome the<br />

activation barrier. We expect that the number <strong>of</strong> collisions per unit volume per unit time<br />

1/2<br />

⎛8kT B ⎞<br />

(collision density) is related to the RMS velocity <strong>of</strong> the gas particles urms<br />

= ⎜ ⎟ where<br />

⎝ π m ⎠<br />

−23<br />

kB= 1.38× 10 J / K is Boltzmann’s constant <strong>and</strong> m is the mass <strong>of</strong> the particle, the number<br />

<strong>of</strong> particles N i <strong>and</strong> the collision cross-section σ . The collision cross basically says that if<br />

two particles are within some critical area, there will be a collision.<br />

If the two particles are within ( ) 2<br />

σ π R R<br />

= A + B , there will be a collision.<br />

Now, the collision density Z AB bet<strong>we</strong>en two unlike particles is:<br />

1/2<br />

⎛8kT B ⎞ 2<br />

AB σ<br />

av<br />

Z = ⎜ ⎟ N [ A][ B]<br />

⎝ πμ ⎠<br />

where N av is Avogadro’s number <strong>and</strong> μ is the reduced mass <strong>and</strong><br />

1 1 1<br />

= + .<br />

μ m m<br />

1 2<br />

Now, the rate <strong>of</strong> change in the molar concentration <strong>of</strong> [A] is equal to the collision density<br />

multiplied by the fraction <strong>of</strong> collisions that occur with a kinetic energy in excess <strong>of</strong> some<br />

threshold value Ea:<br />

dA [ ] ZAB<br />

− = ×<br />

f<br />

dt N<br />

av


where<br />

Now,<br />

f e −<br />

Ea/ RT<br />

= is given by the Boltzmann distribution <strong>of</strong> particles with energy Ea.<br />

dA [ ] ⎛8kT B ⎞<br />

− = σ ⎜ ⎟<br />

dt ⎝ πμ ⎠<br />

1/2<br />

−Ea/<br />

RT<br />

Nave A B<br />

[ ][ ]<br />

Comparing this with the rate law rate = k[ A][ B]<br />

, <strong>we</strong> find that:<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 5 <strong>of</strong> 33<br />

1/2<br />

⎛8kT B ⎞<br />

a<br />

k = σ ⎜ ⎟ Nave ⎝ πμ ⎠<br />

−E<br />

/ RT<br />

Comparing this with the Arrhenius form <strong>of</strong> the rate constant <strong>we</strong> find that:<br />

1/2<br />

⎛8kT B ⎞<br />

A= σ ⎜ ⎟ N<br />

⎝ πμ ⎠<br />

as long as the exponential temperature dependence dominates the <strong>we</strong>ak square root<br />

dependence <strong>of</strong> the pre-exponential factor.<br />

This form is slightly different than that used in LMS. They express the rate in terms <strong>of</strong> the<br />

numbers <strong>of</strong> particles, while this representation uses molar concentrations. They also define<br />

dAB = ( RA+ RB) = σ / π , <strong>and</strong> group the π term into the contribution from u rms .<br />

Kinetic isotope effect:<br />

From collision theory, <strong>we</strong> can see that the rate constant depends on a quantity called the<br />

reduced mass μ . If you consider the relative rate constants bet<strong>we</strong>en different isotopes, you<br />

expect qualitatively that there should be a difference. For example, if you are running a<br />

reaction where you break a C-H bond, you may expect a different rate if you broke a C-D<br />

bond instead, because <strong>of</strong> the differences in reduced masses <strong>and</strong> the relative location <strong>of</strong> the<br />

zero-point energies. This difference is rate constants results in the kinetic isotope effect,<br />

which states that different isotopes react with different rates. This difference is most<br />

pronounced when replacing H with D (or T), <strong>and</strong> is quantified by a quantity called the<br />

isotopic ratio kH/ k D . Table 9.2 in LMS lists some values for isotopic ratios for H as <strong>we</strong>ll as<br />

other atoms. For heavier elements, the kinetic isotope effect is much less pronounced.<br />

Thermodynamic description <strong>of</strong> kinetics? Eyring transition state theory:<br />

Another treatment used to describe the temperature dependence <strong>of</strong> reaction rates is called<br />

activated complex theory, or transition state theory. Consider the simple bimolecular<br />

reaction:<br />

av


k<br />

A+ B⎯⎯→ P<br />

This reaction can be represented as a two-step process:<br />

†<br />

K †<br />

† k<br />

A+ B AB ⎯⎯→ P<br />

†<br />

where the species AB is called an activated complex, or transition state. Note that the<br />

transition state is not the same thing as an intermediate!<br />

Our goal now is to try to relate kinetic parameters to thermodynamic ones.<br />

The first step in the process forms a fast equilibrium, so<br />

considered to be irreversible.<br />

<strong>So</strong>lving the rate law:<br />

† †<br />

rate = k [ AB ]<br />

†<br />

† [ AB ]<br />

= <strong>and</strong> the second step is<br />

[ A][ B]<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 6 <strong>of</strong> 33<br />

K<br />

† †<br />

rate = k K [ A][ B] = k [ A][ B]<br />

† †<br />

Now <strong>we</strong> only consider kobs = k K . We know that the Gibbs energy bet<strong>we</strong>en the reactants<br />

†<br />

<strong>and</strong> the transition state is related to K :<br />

Eyring theory states that the rate constant<br />

obs<br />

Δ G =− RTln K<br />

† †<br />

Planck’s constant. Now, the net rate constant is:<br />

But since Δ G =ΔH −TΔ S,<br />

If <strong>we</strong> compare this to the Arrhenius equation:<br />

†<br />

B<br />

k is equal to<br />

kT<br />

h<br />

†<br />

B −ΔG<br />

/ RT<br />

kobs = e<br />

kT<br />

, where<br />

h<br />

−34<br />

h 6.626 10 J s<br />

= × i is<br />

† † kT B<br />

−Δ ⎡ H −TΔS ⎤/<br />

RT<br />

⎣ ⎦<br />

kobs = e<br />

h<br />

kT † †<br />

B ΔS / R −ΔH<br />

/ RT<br />

kobs = ⎡e ⎤⎡e ⎤<br />

h ⎣ ⎦⎣ ⎦ [1]<br />

Arrhenius: A plot <strong>of</strong> ln k vs. 1/T yields a slope <strong>of</strong> − E / R<br />

†<br />

TS Theory: A plot <strong>of</strong> ln k vs. 1/T yields a slope <strong>of</strong> −( Δ +<br />

)<br />

a<br />

H RT / R


By comparison:<br />

E = Δ H + RT ΔH<br />

a<br />

† †<br />

since Ea~ 100 kJ / mol <strong>and</strong> RT ~2.5 kJ / mol . Plugging this into equation [1] <strong>and</strong><br />

comparing with the Arrhenius form, <strong>we</strong> find that:<br />

†<br />

⎡ ΔS<br />

/ R⎤<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 7 <strong>of</strong> 33<br />

kT B<br />

A= e<br />

h ⎣ ⎦<br />

Thus from information about the reaction rates at different temperatures, <strong>we</strong> can apply TS<br />

theory to obtain information about the energetics <strong>of</strong> the transition state in the reaction!<br />

III. Integrated <strong>Rate</strong> <strong>Laws</strong><br />

<strong>Rate</strong> <strong>Laws</strong>:<br />

We can also express the rate in terms <strong>of</strong> the concentrations <strong>of</strong> the stoichiometric reactants<br />

<strong>and</strong> products (i.e. those that appear in the net reaction.) In this way, the reaction rate can be<br />

thought <strong>of</strong> in terms <strong>of</strong> collisions – if there is a higher concentration <strong>of</strong> a particular<br />

component, then it is more likely to react, but there is some proportionality that takes into<br />

account physical parameters <strong>of</strong> the system. We will discuss this later. For now, let’s<br />

represent the reaction rate in terms <strong>of</strong> a rate law.<br />

Example – consider the reaction:<br />

The rate law for this reaction is:<br />

A+ B→ P<br />

m n q<br />

rate = k[ A] [ B] [ P]<br />

Where k is the rate constant <strong>and</strong> m , n <strong>and</strong> q are reaction orders with respect to a<br />

particular component. For example, if m = 2 <strong>and</strong> n = 1,<br />

<strong>we</strong> say that the reaction is second<br />

order in A <strong>and</strong> first order in B. <strong>Reaction</strong> orders are typically integers (i.e. 0, 1, 2, …), but<br />

they could be fractions.<br />

Since the rate <strong>of</strong> reaction has units <strong>of</strong> concentration/time (M/s), the units for k are:<br />

1 1<br />

k[<br />

] x 1<br />

s M −<br />

⎛ ⎞⎛ ⎞<br />

= ⎜ ⎟⎜ ⎟<br />

⎝ ⎠⎝ ⎠<br />

where x is the total reaction order, or the sum <strong>of</strong> the individual reaction orders.<br />

Integrated <strong>Rate</strong> <strong>Laws</strong>:


Zero Order <strong>Reaction</strong>s:<br />

Lets say the reaction A → B has the rate law<br />

rate k A<br />

dA [ ]<br />

− = rate = k<br />

dt<br />

0<br />

= [ ] . Since<br />

<strong>we</strong> can solve the differential equation for the concentration <strong>of</strong> A with time:<br />

if <strong>we</strong> define [A] = [A]0 at time t = 0, then:<br />

[ A] dA [ ] = − kdt<br />

∫ dA [ ] =−∫<br />

kdt<br />

− [ A] =−k( t − t )<br />

f i f i<br />

[ A] = [ A] − kt<br />

A similar procedure can be done to determine the appearance <strong>of</strong> B with respect to time.<br />

First Order <strong>Reaction</strong>s:<br />

Lets say the reaction A→ B has the rate law<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 8 <strong>of</strong> 33<br />

0<br />

rate k A<br />

dA [ ]<br />

− = kA [ ]<br />

dt<br />

1<br />

= [ ] . Since<br />

<strong>we</strong> can solve the differential equation for the concentration <strong>of</strong> A with time:<br />

if <strong>we</strong> define [A] = [A]0 at time t = 0, then:<br />

In this reaction, [A]0 + [B]0 = [A] +<br />

dA [ ]<br />

= − kdt<br />

[ A ]<br />

dA [ ]<br />

kdt<br />

[ A ]<br />

=− ∫ ∫<br />

[ A]<br />

f<br />

ln = −kt ( f − ti)<br />

[ A ]<br />

i<br />

[ A] [ A] e −<br />

=<br />

0<br />

kt


[B], so if [B] = 0 at time t = 0,<br />

Second Order <strong>Reaction</strong>s:<br />

Type I:<br />

kt ( )<br />

[ B] [ A] 1 e −<br />

= −<br />

Lets say the reaction 2A→ B has the rate law<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 9 <strong>of</strong> 33<br />

0<br />

rate k A<br />

1 dA [ ]<br />

− = kA [ ]<br />

2 dt<br />

2<br />

= [ ] . Since<br />

<strong>we</strong> can solve the differential equation for the concentration <strong>of</strong> A with time:<br />

if <strong>we</strong> define [A] = [A]0 at time t = 0, then:<br />

Type II:<br />

dA [ ]<br />

2kdt<br />

2<br />

[ A ]<br />

=−<br />

dA [ ]<br />

2kdt<br />

2<br />

[ A ]<br />

=−<br />

∫ ∫<br />

1 1<br />

= + 2kt<br />

[ A] [ A]<br />

Lets say the reaction A + B→ C has the rate law rate = k[ A][ B]<br />

. Since<br />

dA [ ]<br />

− = kA [ ][ B]<br />

dt<br />

if <strong>we</strong> let [ A] = [ A] 0 − x <strong>and</strong> [ B] = [ B] 0 − x,<br />

then <strong>we</strong> can perform a change <strong>of</strong> variables:<br />

<strong>and</strong><br />

dx<br />

rate = = k[ A][ B] = k [ A] 0 −x[ B] 0 − x<br />

dt<br />

dx<br />

[ A] x [ B] x =<br />

( − )( −<br />

)<br />

0 0<br />

0<br />

2<br />

( )( )<br />

kdt


<strong>So</strong>lving using partial fractions for [ A] 0 ≠ [ B]<br />

0<br />

finally<br />

If 0 0<br />

( − )<br />

( )<br />

1 [ B] 0 [ A] 0 x<br />

ln<br />

[ A] −[ B] [ A] [ B] −x<br />

0 0 0 0<br />

1 [ B] [ A]<br />

[ ] [ ] [ ] [ ]<br />

0 ln<br />

A 0 B 0 A 0 B =<br />

−<br />

[ A] = [ B]<br />

, then the expression reduces to Type I.<br />

n th Order <strong>Reaction</strong>s:<br />

= kt<br />

Lets say the reaction nA → B has the rate law [ ] n<br />

rate = k A . Then<br />

IV. Mechanisms<br />

Goals:<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 10 <strong>of</strong> 33<br />

kt<br />

1 ⎛ 1 1 ⎞<br />

⎜ ⎟<br />

n−1 ⎝[ A] [ A]<br />

⎠<br />

− n−1 n−1 0<br />

= nkt<br />

We <strong>have</strong> previously discussed reactions that occur in one step. We now want to turn our<br />

attention to more complex reaction mechanisms that occur over many steps. The set <strong>of</strong> steps<br />

in the reaction is called the reaction mechanism. The reaction mechanism consists <strong>of</strong> a<br />

number <strong>of</strong> elementary reactions.<br />

The eventual goal in this process is to propose a mechanism where the rate law is consistent<br />

with experimental data. It is useful to point out on the onset that the rate law does not prove<br />

a particular mechanism, but it can disprove the mechanism. In other words, two different<br />

mechanisms may result in the same rate law.<br />

Parallel <strong>Reaction</strong>s:<br />

Consider the following reaction, where the reactant can form either product B or product C:<br />

k1<br />

A⎯⎯→ B<br />

k2<br />

A⎯⎯→C


We can express the rate <strong>of</strong> laws for each component:<br />

dB [ ]<br />

k [ A]<br />

1<br />

dt = , 2<br />

dt<br />

dC [ ]<br />

−dA<br />

[ ]<br />

= k [ A]<br />

<strong>and</strong> = k1[ A] + k2[ A] = ( k1+ k2) [ A]<br />

dt<br />

We can integrate the last expression to get the concentration pr<strong>of</strong>ile <strong>of</strong> [A]:<br />

Now:<br />

dB [ ]<br />

dt<br />

[ A] = [ A] e<br />

− ( k1+ k2) t<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 11 <strong>of</strong> 33<br />

0<br />

dC [ ]<br />

dt<br />

− ( k1+ k2) t<br />

− ( k1+ k2) t<br />

= k1[ A] = k1[ A] 0e<br />

<strong>and</strong> = k2[ A] = k2[ A] 0e<br />

If <strong>we</strong> assume that at t = 0, [B]0 = [C]0 = 0, then <strong>we</strong> can integrate the above expressions to<br />

obtain the concentration pr<strong>of</strong>ile <strong>of</strong> [B] <strong>and</strong> [C]:<br />

Note that for all times:<br />

Equilibrium <strong>Reaction</strong>s:<br />

k [ A]<br />

B = −e<br />

1 0 [ ] 1<br />

k1+ k2<br />

− ( k1+ k2) t { }<br />

[ B] k<br />

=<br />

[ C] k<br />

Consider the equilibrium process bet<strong>we</strong>en A <strong>and</strong> B:<br />

k [ A]<br />

k + k<br />

2 0 − ( k1+ k2) t<br />

<strong>and</strong> [ C] = { 1−e<br />

}<br />

1<br />

2<br />

A←⎯⎯ ⎯⎯→ B<br />

k1<br />

k−1<br />

1 2<br />

In this reaction, the reactant A forms B, but then B converts back to A. We denote the rate<br />

constant for a reverse reaction with a negative subscript. Consider the rate law for [A]:<br />

−dA<br />

[ ]<br />

rate = = k1[ A] − k−1[ B]<br />

dt<br />

In this expression, the reaction k1 results in a disappearance <strong>of</strong> [A] (negative derivative),<br />

while the reaction k-1 results in formation <strong>of</strong> [A] (positive derivative).<br />

−dA<br />

[ ]<br />

When the reaction reaches equilibrium, <strong>we</strong> know that = 0 , i.e.: there is no net<br />

dt<br />

formation or consumption <strong>of</strong> [A]. Relating this to the rate law:


But <strong>we</strong> also know that [ B] /[ A] = K , so:<br />

eq eq eq<br />

[ B] eq k1<br />

=<br />

[ A] k− K<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 12 <strong>of</strong> 33<br />

eq<br />

eq<br />

k<br />

=<br />

k<br />

1<br />

forward<br />

reverse<br />

I smell a relationship with thermodynamics… Don’t get too excited.<br />

Consecutive <strong>Reaction</strong>s:<br />

Now consider a reaction that occurs in two consecutive steps:<br />

k1 k2<br />

A⎯⎯→B⎯⎯→ C<br />

The rate laws for each expression are as follows:<br />

−dA<br />

[ ]<br />

dt<br />

dB [ ]<br />

k [ A] k [ B]<br />

dt = − <strong>and</strong> dC [ ]<br />

k2[ B]<br />

dt =<br />

= k1[ A]<br />

, 1 2<br />

In the above expressions, [A] disappears with rate constant k1, [B] is formed as k1, but<br />

disappears (negative derivative) as k2, <strong>and</strong> [C] is formed as k2.<br />

We can integrate these rate laws to get the respective concentration pr<strong>of</strong>iles (assuming [B]<br />

<strong>and</strong> [C] are zero initially):<br />

Now,<br />

[ A] [ A] e −<br />

=<br />

0<br />

kt 1<br />

dB [ ]<br />

−kt<br />

1<br />

= k1[ A] − k2[ B] = k1[ A] 0e − k2[ B]<br />

dt<br />

dB [ ]<br />

−kt<br />

1<br />

+ k2[ B] = k1[ A] 0e<br />

[1]<br />

dt<br />

To solve this equation, <strong>we</strong> first make an assumption as to the functional form <strong>of</strong> [B]:<br />

[ B] yte ( )<br />

−<br />

kt 2<br />

t = [2]<br />

This basically says that [B] decays exponentially, but with a time-dependent amplitude. The<br />

‘dummy’ function y(t) basically accounts for the buildup <strong>of</strong> [B] from reaction 1, but <strong>we</strong> <strong>have</strong>


to find what y(t) is. With [2] <strong>and</strong> the chain rule, <strong>we</strong> can take the time derivative <strong>of</strong> this form<br />

for [B]:<br />

dB [ ]<br />

−kt 2 −kt⎛dy⎞ 2<br />

=− yt () ( ke 2 ) + e ⎜ ⎟<br />

dt ⎝ dt ⎠<br />

But the first term is simply equal to –k2[B]t, so moving to the other side:<br />

Equating [1] <strong>and</strong> [3]:<br />

dB [ ]<br />

−kt⎛dy⎞ 2<br />

+ k2[ B] t = e ⎜ ⎟<br />

dt ⎝ dt ⎠ [3]<br />

⎛dy ⎞ −kt 2 −kt<br />

1<br />

⎜ ⎟e<br />

= k1[ A] 0e<br />

⎝ dt ⎠<br />

⎛dy ⎞<br />

⎜ ⎟=<br />

k [ A] e<br />

⎝ dt ⎠<br />

( e ) = k [ A] e<br />

− kt 1 + k2t ( k2−k1) t<br />

1 0 1 0<br />

k[ A]<br />

y() t = e + C<br />

k − k<br />

1 0 ( k2−k1) t<br />

Integrating:<br />

2 1<br />

k1[ A]<br />

0<br />

We know that at t = 0, y = 0, so: C =−<br />

k − k<br />

k [ A]<br />

k − k<br />

2 1<br />

1 0 ( k2−k1) t<br />

<strong>So</strong>: yt () = ( e −1)<br />

kt 2<br />

Finally, since [ B] yte ( )<br />

−<br />

= :<br />

2 1<br />

k [ A]<br />

[ ] = −<br />

−kt 1 −k2t<br />

( )<br />

1 0 B e e<br />

k2 − k1<br />

To get [C], <strong>we</strong> know that [A] + [B] + [C] = [A]0, so:<br />

⎡ ke − ke ⎤<br />

[ C] = [ A]<br />

1−<br />

−kt 1 −k2t<br />

2 1<br />

0 ⎢ ⎥<br />

⎣ k2 − k1<br />

⎦<br />

This is the last differential equation <strong>we</strong> will solve for a while…<br />

Let us consider two extremes in this 2-step mechanism:<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 13 <strong>of</strong> 33


(1) First step is fast ( k1 k2)<br />

[ ]<br />

k[ A]<br />

−kt 1 −k2t<br />

( )<br />

1 0 B = e −e<br />

k2 − k1<br />

−kt 1 −k2t<br />

2 1<br />

0 ⎢ ⎥<br />

⎣ k2 − k1<br />

⎦<br />

k<br />

[ B] = [ A] 0 − e = [ A] e<br />

−k<br />

1 → ( )<br />

−kt 2 −kt<br />

2<br />

0 0<br />

−kt<br />

2<br />

⎡ ke − ke ⎤<br />

⎡ 0 − ke ⎤ 1<br />

−kt<br />

2<br />

[ C] = [ A]<br />

1−<br />

→ [ C] = [ A] 0 ⎢1 − ⎥ = [ A] ⎡ 0 1−e<br />

⎤<br />

k ⎣ ⎦<br />

⎣ − 1 ⎦<br />

Consider this graphically:<br />

(2) Second step is fast ( k2 k1)<br />

:<br />

k[ A]<br />

[ ] = −<br />

−kt 1 −k2t<br />

( )<br />

1 0 B e e<br />

k2 − k1<br />

−kt 1 −k2t<br />

2 1<br />

0 ⎢ ⎥<br />

⎣ k2 − k1<br />

⎦<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 14 <strong>of</strong> 33<br />

1<br />

The reaction be<strong>have</strong>s as if the first<br />

step did not exist – as if [B] is the<br />

initial reactant (with a concentration<br />

equal to [A]0) <strong>and</strong> the net reaction is<br />

first order in [B]. Since the second<br />

step dictates the kinetics <strong>of</strong> the entire<br />

reaction, it is said to be the rate<br />

determining (or rate limiting) step.<br />

k[ A] k[ A]<br />

B = e − = e<br />

k k<br />

→ [ ] ( 0)<br />

1 0 −kt 1 1 0 −kt<br />

1<br />

2 2<br />

−kt<br />

1<br />

⎡ ke − ke ⎤<br />

⎡ ke ⎤ 2<br />

−kt<br />

1<br />

[ C] = [ A]<br />

1−<br />

→ [ C] = [ A] 0 ⎢1 − ⎥ = [ A] ⎡ 0 1−e<br />

⎤<br />

k ⎣ ⎦<br />

⎣ 2 ⎦<br />

Consider this graphically:<br />

This is equivalent to the<br />

concentration pr<strong>of</strong>iles <strong>we</strong> would<br />

obtain if the reaction <strong>we</strong>re one-step<br />

A → C . The concentration <strong>of</strong> [C]<br />

be<strong>have</strong>s as if [B] <strong>we</strong>re not there. In<br />

this reaction, the first step is the rate<br />

determining step.


Take home message: The overall rate <strong>of</strong> reaction is equivalent to the rate for the<br />

rate limiting step!<br />

We can use this to determine the rates for more complex reaction mechanisms.<br />

Elementary <strong>Reaction</strong>s<br />

We <strong>have</strong> already had a taste <strong>of</strong> mechanisms. Now <strong>we</strong> can consider consecutive,<br />

equilibrium <strong>and</strong> parallel reactions all at the same time. First, ho<strong>we</strong>ver, <strong>we</strong> <strong>have</strong> to<br />

consider how the elementary reactions translate to rate laws. We do this by<br />

determining the molecularity <strong>of</strong> the step.<br />

The molecularity <strong>of</strong> a reaction is the number <strong>of</strong> molecules coming together to react.<br />

For example, if <strong>we</strong> consider the isomerization <strong>of</strong> cyclopropane to propene:<br />

We notice that a single molecule breaks itself into the new arrangement. A reaction<br />

with a single molecule as a reactant is called a unimolecular reaction.<br />

In a bimolecular reaction, a pair <strong>of</strong> molecules collide <strong>and</strong> exchange energy, such as<br />

in the reaction <strong>of</strong> a hydrogen radical with gaseous bromine:<br />

H i+ Br → HBr + Bri<br />

2<br />

The rate law for an elementary reaction simply takes the molecularity <strong>of</strong> the reaction<br />

as the reaction order. For example, the rate law for the unimolecular reaction (first<br />

reaction above) would be rate = k[ cyclopropane]<br />

<strong>and</strong> the bimolecular reaction would<br />

be rate = k[ H i ][ Br2<br />

] . Note that <strong>we</strong> cannot do this for the net reaction, only for the<br />

elementary reactions!<br />

Steady State Approximation <strong>and</strong> Pre-equilibrium:<br />

Consider the following 3-step mechanism for the net reaction A+ B+ D→ P:<br />

A+ B←⎯⎯ ⎯⎯→ X<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 15 <strong>of</strong> 33<br />

k1<br />

k−1<br />

k2<br />

X + D⎯⎯→ P<br />

We do not necessarily know which step is rate limiting, nonetheless <strong>we</strong> can write the<br />

rate law in terms <strong>of</strong> the formation <strong>of</strong> product P:


dP [ ]<br />

rate = = k2[ X ][ D]<br />

dt<br />

Unfortunately, this rate law depends on the intermediate [X]. An intermediate is a<br />

substance that is a product <strong>of</strong> an earlier step <strong>and</strong> a reactant in a later step.<br />

Intermediates cannot appear in the overall rate law!<br />

To eliminate the intermediate concentration from the rate law, <strong>we</strong> can use the steady<br />

state approximation (SSA). If <strong>we</strong> consider the concentration <strong>of</strong> [X] as the reaction<br />

proceeds, [X]0 = 0, but increases as the reaction begins. Once it is formed, it becomes<br />

consumed in the second reaction. For most <strong>of</strong> the reaction, [X] is at a steady state<br />

(constant) value. This means that the rate <strong>of</strong> formation for [X] is equal to zero! Now:<br />

dX [ ]<br />

≈ 0<br />

[SSA]<br />

dt<br />

dX [ ]<br />

= k1[ A][ B] −k−1[ X] − k2[ X][ D]<br />

= 0<br />

dt<br />

k1[ A][ B] = k−1[ X] + k2[ X][ D]<br />

k1[ A][ B]<br />

[ X ] =<br />

k + k [ D]<br />

−1<br />

2<br />

We can use this steady state concentration <strong>of</strong> X to plug into our original rate law:<br />

dP [ ] ⎛ k1[ A][ B] ⎞ kk 1 2[<br />

A][ B][ D]<br />

rate = = k2[ X ][ D] = k2⎜ ⎟[<br />

D]<br />

=<br />

dt ⎝k−1+ k2[ D] ⎠ k−1+ k2[ D]<br />

We now <strong>have</strong> a rate law in terms <strong>of</strong> stoichiometric products <strong>and</strong> reactants! This rate<br />

law is true independent <strong>of</strong> the relative rates <strong>of</strong> each step<br />

But… Consider the case where ( k−1 k2)<br />

<br />

kk[ A][ B][ D]<br />

1 2 rate = = k1 A B<br />

k2[ D]<br />

[ ][ ]<br />

This is consistent with the rate law as if the mechanism <strong>we</strong>re:<br />

k1<br />

A+ B⎯⎯→ P<br />

<strong>So</strong> the rate law does not see the second step.<br />

Now consider the case where ( k−1 k2)<br />

:<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 16 <strong>of</strong> 33


kk[ A][ B][ D] kk<br />

= = [ ][ ][ ] = [ ][ ][ ]<br />

1 2 1 2<br />

rate A B D kobs A B D<br />

k−1 k−1<br />

If <strong>we</strong> think about what is physically happening here, the reverse step <strong>of</strong> the first<br />

reaction is large, so much so that the first reaction reaches equilibrium before the<br />

second step begins!<br />

Consider another way. We mentioned previously that<br />

dP [ ]<br />

rate = = k2[ X ][ D]<br />

dt<br />

If the first step reaches equilibrium rapidly (i.e.: before the second step begins), then:<br />

K<br />

1<br />

[ X ]<br />

= , so [ X ] K1[ A][ B]<br />

[ A][ B]<br />

= <strong>and</strong> 2 1 =<br />

rate k K [ A][ B][ D]<br />

This is consistent with the general expression above since K1 = k1/ k−1. This is called<br />

the pre-equilibrium approximation. It is used whenever a mechanism involves a fast<br />

initial equilibrium step.<br />

Example: Determine the rate law for the following mechanism:<br />

A←⎯⎯ ⎯⎯→ B+ C<br />

(fast equilibrium)<br />

k1<br />

k−1<br />

k2<br />

k−2<br />

k3<br />

A+ C←⎯⎯ ⎯⎯→ E<br />

(fast equilibrium)<br />

E⎯⎯→ P<br />

(slow)<br />

From above, the net reaction is 2A → B+ P,<br />

so C <strong>and</strong> E are intermediates. Since<br />

step 3 is rate limiting:<br />

dP [ ]<br />

rate = = k3[ E]<br />

dt<br />

Using the pre-equilibrium approximation:<br />

<strong>So</strong>:<br />

[ E]<br />

K<br />

2<br />

1 2 = <strong>and</strong><br />

1<br />

KK[ A]<br />

[ B]<br />

[ B][ C]<br />

[ E]<br />

= <strong>and</strong> K2<br />

=<br />

[ A]<br />

[ A][ C]<br />

kKK[ A] [ A]<br />

[ B] [ B]<br />

2 2<br />

3 1 2 rate = =<br />

kobs<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 17 <strong>of</strong> 33


Principles <strong>of</strong> Microscopic Reversibility <strong>and</strong> Detailed Balance:<br />

For a series <strong>of</strong> equilibrium reactions, the net equilibrium constant is the product <strong>of</strong> the<br />

elementary reaction equilibrium constants. This is also the product <strong>of</strong> the forward<br />

rate constants divided by the product <strong>of</strong> the reverse rate constants. The Principle <strong>of</strong><br />

Microscopic Reversibility (Tolman) states that, “…in a system at equilibrium, any<br />

molecular process <strong>and</strong> the reverse <strong>of</strong> that process occur, on average, at the same rate.”<br />

A closely related principle states that at equilibrium, every molecular collision has an<br />

exact counterpart in the reverse direction such that the rate in the forward direction is<br />

balanced by the reverse reaction. This is called the Principle <strong>of</strong> Detailed Balance.<br />

Note that both <strong>of</strong> these principles apply to equilibrium systems.<br />

Checking for Consistency:<br />

Armed with a procedure to determine rate laws for composite reactions, <strong>we</strong> can<br />

consider how a proposed mechanism can be consistent (or not) with an<br />

experimentally determined rate law.<br />

+ −<br />

Example: The reaction Cl2 + H2S→ S + 2H+ 2Cl<br />

[ Cl2][ H2S] Has the experimentally determined rate law: rate k<br />

[ H ]<br />

+ =<br />

Two mechanisms <strong>have</strong> been proposed. Which one is consistent with the observed<br />

rate law?<br />

Mechanism 1:<br />

Mechanism 2:<br />

←⎯⎯ ⎯⎯→ +<br />

[fast equilibrium]<br />

HS 2<br />

k1<br />

k−1<br />

−<br />

HS<br />

+<br />

H<br />

k2<br />

Cl ⎯⎯→ − +<br />

2 Cl + Cl<br />

k−2<br />

+ − k3<br />

+ −<br />

←⎯⎯ [fast equilibrium]<br />

Cl + HS ⎯⎯→+ H + Cl + S<br />

[slow]<br />

←⎯⎯ ⎯⎯→ +<br />

[fast equilibrium]<br />

HS + Cl ⎯⎯→ Cl + S + H<br />

[slow]<br />

HS 2<br />

k1<br />

k−1<br />

−<br />

HS<br />

+<br />

H<br />

−<br />

2<br />

k2<br />

2<br />

− +<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 18 <strong>of</strong> 33


Specific Mechanisms:<br />

Lindemann – Hinshelwood mechanism:<br />

Many gas phase reactions are unimolecular, for example the isomerization <strong>of</strong><br />

cyclopropane mentioned previously. The issue with first order reactions is that the<br />

molecule somehow has to obtain enough energy (from collisions with other<br />

molecules) for the reaction to occur (<strong>we</strong> will discuss activation energies soon!) But<br />

collisions are bimolecular by nature, so one would think that they should not result in<br />

a first order reaction. The mechanism for these type <strong>of</strong> reactions was provided by<br />

Lindemann, then elaborated by Hinshelwood. In the mechanism, two reactant<br />

molecules collide to form an activated molecule:<br />

k1<br />

A+ A⎯⎯→ A* + A<br />

The activated molecule can suffer one <strong>of</strong> two fates. It may lose its energy through<br />

another collision with an A molecule:<br />

Or is can fragment into products:<br />

k2<br />

A* + A⎯⎯→ 2A<br />

k3<br />

A* ⎯⎯→ P<br />

Of course, the first two reactions are reverse processes, so the net mechanism can be<br />

written as:<br />

A+ A←⎯⎯ ⎯⎯→ A* + A<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 19 <strong>of</strong> 33<br />

k1<br />

k2<br />

k3<br />

A* ⎯⎯→ P<br />

We do not <strong>have</strong> any information about the relative rates (at this point), so <strong>we</strong> cannot<br />

use the pre-equilibrium approximation, but <strong>we</strong> can use the SSA:<br />

so:<br />

dP<br />

rate = = k3[ A*]<br />

dt<br />

2<br />

k1[ A]<br />

[ A*]<br />

=<br />

k3 + k2[ A]<br />

2<br />

dP kk 1 3[<br />

A]<br />

rate = =<br />

dt k + k [ A]<br />

3 2<br />

This mechanism certainly does not appear to be first order in [A], but <strong>we</strong> can consider<br />

two extremes for the disintegration <strong>of</strong> [A*]:


Consider the case where the deactivation process is much faster than the<br />

2<br />

fragmentation ( k3 k2 [ A]<br />

) . Then the rate becomes: rate = k1[ A]<br />

But if the product 2 [ ] k A is much larger than k 3 :<br />

kk 1 3<br />

rate = kobs[ A]<br />

where kobs<br />

k<br />

First order in [A]! It has been shown experimentally that many gas phase reactions<br />

are second order in reactant at low concentration <strong>and</strong> first order in reactant at high<br />

concentrations.<br />

Chain reactions – the hydrogen-bromine reaction:<br />

Consider the gas-phase reaction bet<strong>we</strong>en hydrogen <strong>and</strong> bromine to form HBr:<br />

H2 + Br2 → 2HBr<br />

This reaction involves the formation <strong>of</strong> radicals, which are very reactive. One aspect<br />

<strong>of</strong> radical reactions is that radicals beget radicals, meaning that the products <strong>of</strong> the<br />

elementary reactions are typically radicals. These radicals continue to propagate until<br />

a termination, typically by radical recombination. As a result, the mechanism for this<br />

reaction can be expressed as a chain reaction:<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 20 <strong>of</strong> 33<br />

2<br />

k1<br />

Br2⎯⎯→ 2Br<br />

[initiation]<br />

k2<br />

Br+ H2⎯⎯→ HBr+ H<br />

[propagation]<br />

k3<br />

H + Br2 ⎯⎯→ HBr + Br<br />

[propagation]<br />

k−2<br />

H + HBr ⎯⎯→ Br + H 2<br />

[retardation]<br />

k−1<br />

2Br⎯⎯→<br />

Br<br />

[termination]<br />

The rate can be expressed as the formation <strong>of</strong> HBr, but because HBr is in many steps<br />

in the mechanism:<br />

2<br />

dHBr [ ]<br />

rate = = k2[ Br][ H 2] + k3[ H ][ Br2] − k−2[ H ][ HBr]<br />

[1]<br />

dt<br />

We should go immediately to the SSA for the intermediates Br <strong>and</strong> H:


<strong>and</strong>:<br />

dH [ ]<br />

= k2[ Br][ H2] −k3[ H][ Br2] − k−2[ H][ HBr]<br />

= 0<br />

[2]<br />

dt<br />

dBr [ ]<br />

2<br />

= k1[ Br2] − k2[ Br][ H2] + k3[ H][ Br2] + k−2[ H][ HBr] − k−1[ Br]<br />

= 0 [3]<br />

dt<br />

We <strong>have</strong> two equations <strong>and</strong> two unknowns. If <strong>we</strong> add [2] <strong>and</strong> [3]:<br />

so:<br />

k Br k Br<br />

2<br />

1[ 2] − −1[<br />

] = 0<br />

⎛ k ⎞<br />

= ⎜ ⎟<br />

⎝ ⎠<br />

1 [ Br] [ Br2<br />

]<br />

k−1 We can plug this into [2] <strong>and</strong> rearrange to get [H]:<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 21 <strong>of</strong> 33<br />

1/2<br />

1/2<br />

⎛ k ⎞<br />

1<br />

2⎜ 2 ⎟ 2<br />

k−1<br />

k [ Br ] [ H ]<br />

[ H ] =<br />

⎝ ⎠<br />

k [ Br ] + k [ HBr]<br />

3 2 −2<br />

Now, subtracting [2] from [1], <strong>and</strong> substituting the expression for [H], <strong>we</strong> obtain:<br />

3/2<br />

dHBr [ ]<br />

⎛ k ⎞ 1 2 kk 3 2[ H2][ Br2]<br />

rate = = [ H ][ Br2<br />

] =⎜ ⎟<br />

dt ⎝k−1⎠ k3[ Br2] + k−2[ HBr]<br />

This can be rearranged to obtain:<br />

where<br />

2<br />

1/2<br />

3/2<br />

kH [ 2][ Br2]<br />

rate =<br />

[ Br ] + k'[ HBr]<br />

= 2<br />

⎛ k1<br />

2 ⎜<br />

1<br />

1/2<br />

⎞<br />

⎟<br />

k k<br />

⎝k−⎠ <strong>and</strong><br />

k<br />

k ' =<br />

k<br />

Consistent with the experimentally observed rate law.<br />

−2<br />

3


V. Analysis <strong>of</strong> Kinetic Data<br />

Differential <strong>Rate</strong> Method/Initial <strong>Rate</strong> Method:<br />

As the name implies, this method is based on the idea that the rate can be expressed<br />

as the time derivative <strong>of</strong> concentration for a particular component.<br />

1 [ ]<br />

[ ] n<br />

dX<br />

rate = = k X<br />

x dt<br />

<strong>So</strong> if you take the derivative <strong>of</strong> [X] with respect to time, you will obtain an<br />

instantaneous rate. If you extrapolate this to the initial rate, you can compare the<br />

initial rates <strong>of</strong> reaction for various initial concentrations <strong>of</strong> species that appear in the<br />

rate law.<br />

m n<br />

Consider the generic reaction A+ B→ P given by the rate law kA [ ] [ B ] . If <strong>we</strong> run<br />

the reaction under different sets <strong>of</strong> conditions, <strong>we</strong> will obtain different initial rates.<br />

We can then compare these initial rates to obtain more information about the reaction<br />

orders <strong>and</strong> the rate constant. For example, if <strong>we</strong> run two trials holding [B] constant<br />

<strong>and</strong> varying [A], the initial rates are related by:<br />

m n<br />

rate1 k[ A] 1[ B] ⎛[ A]<br />

⎞ 1<br />

= =⎜ m n ⎟<br />

rate2 k[ A] 2[ B] ⎝[ A]<br />

2 ⎠<br />

This procedure is then done for each component in the rate law.<br />

Example: Consider the reaction <strong>of</strong> pyridine <strong>and</strong> methyl iodide:<br />

C6H5N+ CH3I→ products<br />

Initial rate data <strong>we</strong>re taken at various concentrations <strong>of</strong> the reactants below:<br />

Trial [ CHN] 6 5 (M) [ CH3I ] (M) Initial rate (M/s)<br />

1<br />

2<br />

3<br />

1.00 10 −<br />

×<br />

2.00 10 −<br />

×<br />

2.00 10 −<br />

×<br />

4<br />

4<br />

4<br />

1.00 10 −<br />

×<br />

2.00 10 −<br />

×<br />

4.00 10 −<br />

×<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 22 <strong>of</strong> 33<br />

4<br />

4<br />

4<br />

m<br />

7.5 10 −<br />

×<br />

3.0 10 −<br />

×<br />

6.0 10 −<br />

×<br />

From trials 2 <strong>and</strong> 3, if [A] remains constant <strong>and</strong> [B] doubles, the rate doubles, so the<br />

reaction must be first order in [B]. From trials 1 <strong>and</strong> 2 if both initial concentrations<br />

double, the rate quadruples, so the reaction must be first order in [A] as <strong>we</strong>ll.<br />

Now that <strong>we</strong> <strong>have</strong> the reaction orders, <strong>we</strong> can calculate the rate constant:<br />

7<br />

6<br />

6


ate<br />

k =<br />

[ A] [ B]<br />

1 1<br />

0 0<br />

All three trials should yield the same rate constant, but to be sure <strong>we</strong> average over all<br />

three trials. Finally:<br />

−1 −1<br />

( )<br />

rate = 75 M s [ A][ B]<br />

<strong>So</strong>me experimental error may be present in real data, so in the literature the rate<br />

constants <strong>and</strong> reaction orders include error bars, i.e.: n = 1.02 ± 0.1 would correspond<br />

to a reaction order <strong>of</strong> 1.<br />

Integral <strong>Rate</strong> Method:<br />

The integral rate method assumes you <strong>have</strong> data for [A] as a function <strong>of</strong> time. You<br />

then guess the reaction order with respect to a particular reactant (or species in the<br />

rate law) <strong>and</strong> plot the data in form consistent with the integrated rate law to yield a<br />

linear fit.<br />

1. Zero order reactions:<br />

From the integrated rate law:<br />

[ A] = [ A] − kt<br />

A plot <strong>of</strong> [A] vs. time should yield a linear relationship where the slope <strong>of</strong> the line<br />

is equal to − k <strong>and</strong> the intercept is equal to [ A ] 0 . If there is not a linear<br />

relationship in the data, the reaction is not zero order with respect to [A].<br />

2. First order reactions:<br />

From the integrated rate law:<br />

This can be rearranged to:<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 23 <strong>of</strong> 33<br />

0<br />

[ A] [ A] e −<br />

=<br />

0<br />

0<br />

kt<br />

ln[ A] = ln[ A] −<br />

kt


If the reaction is first order in A, A plot <strong>of</strong> ln[ A ] vs. time will yield a straight line<br />

with a slope equal to k<br />

ln[ A ] .<br />

3. Second order reactions – Type I:<br />

From the integrated rate law:<br />

− <strong>and</strong> an intercept equal to 0<br />

1 1<br />

= + 2kt<br />

[ A] [ A]<br />

If the reaction is second order in [A], a plot <strong>of</strong> 1/[ A ] vs. time will yield a linear<br />

plot with a slope <strong>of</strong> 2k <strong>and</strong> an intercept <strong>of</strong> 1/[ A ] 0 . Note that some treatments<br />

group the ‘2’ (from the stoichiometry <strong>of</strong> the reaction) into the rate constant.<br />

4. Second order reactions – Type II:<br />

From the integrated rate law:<br />

This can be rearranged into:<br />

1 [ B] [ A]<br />

ln<br />

[ ] [ ] [ ] [ ]<br />

0<br />

A 0 B 0 A 0 B =<br />

−<br />

0<br />

ln ⎜ ⎟ ln ⎜ ⎟ [ ] 0 [ ] 0<br />

[ B] [ B]<br />

0<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 24 <strong>of</strong> 33<br />

0<br />

kt<br />

⎛[ A]<br />

⎞ ⎛[ A]<br />

⎞<br />

= + k( A − B ) t<br />

⎝ ⎠ ⎝ ⎠<br />

A plot <strong>of</strong> ln ( [ A]/[ B ] ) vs. time will yield a linear plot with a slope <strong>of</strong><br />

( [ A] − [ B] ) k <strong>and</strong> an intercept <strong>of</strong> ln ( [ A] /[ B ] ) .<br />

0 0<br />

5. n th order reactions:<br />

0 0<br />

From the integrated rate law (grouping stoichiometric coefficients together):<br />

1 1<br />

[ A] [ A]<br />

− = kt<br />

n−1 n−1 0<br />

If the reaction is n th 1<br />

order in [A], a plot <strong>of</strong> 1/[ ] n<br />

A − vs. time will yield a linear plot<br />

1<br />

with a slope <strong>of</strong> k <strong>and</strong> an intercept <strong>of</strong> A − .<br />

1/[ ] 0<br />

n


Flooding/Isolation Method:<br />

Consider the reaction A + B→ C where rate = k[ A][ B]<br />

. The data for [A] vs. time<br />

<strong>and</strong> [B] vs. time may be difficult to analyze. The integrated rate law is:<br />

⎛[ A]<br />

⎞ ⎛[ A]<br />

⎞<br />

= + k( A − B ) t<br />

⎝ ⎠ ⎝ ⎠<br />

0<br />

ln ⎜ ⎟ ln ⎜ ⎟ [ ] 0 [ ] 0<br />

[ B] [ B]<br />

0<br />

Now let us consider running this reaction with the condition that [ A] 0 [ B]<br />

0 such<br />

that the concentration <strong>of</strong> A remains nearly constant throughout the reaction. The<br />

integrated rate law can be rearranged:<br />

⎛ [ A] ⎞ ⎛ [ B]<br />

⎞<br />

ln⎜ ⎟− ln ⎜ ⎟=<br />

k( [ A] 0 −[<br />

B] 0)<br />

t<br />

[ A] [ B]<br />

⎝ 0 ⎠ ⎝ 0 ⎠<br />

But [ A]/[ A] 0 ≈ 1 <strong>and</strong> [ A] 0 −[ B] 0 ≈ [ A]<br />

0.<br />

Now:<br />

And:<br />

⎛ [ B]<br />

⎞<br />

ln ⎜ ⎟=−kA<br />

[ ] 0t<br />

[ B]<br />

⎝ 0 ⎠<br />

[ B] [ B] e −<br />

=<br />

k[ A] 0 t<br />

Now the reaction looks first order in [B]. This technique <strong>of</strong> isolation results in a<br />

pseudo-first order reaction. Physically, one component [A] is so large that its<br />

concentration stays constant over the course <strong>of</strong> the entire reaction, so [ A] 0 ≈ [ A]<br />

t .<br />

This isolates the other component [B].<br />

Now, a plot <strong>of</strong> ln[ B ] vs. time will yield a straight line with an intercept equal to<br />

ln[ B ] 0 <strong>and</strong> a slope equal to − k ',<br />

the pseudo-first order rate constant. The true second<br />

order rate constant is simply k =<br />

k'/[ A]<br />

0<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 25 <strong>of</strong> 33<br />

0


More complicated techniques – Kezdy-Swinbourne <strong>and</strong> Guggenheim Methods:<br />

The analysis <strong>of</strong> kinetic data relies on the ability to measure the concentration <strong>of</strong> a<br />

particular component in the reaction. <strong>So</strong>metimes this is difficult to do. Consider the<br />

thermal decomposition <strong>of</strong> chlorocyclohexane (CXCl):<br />

CH Clg ( ) → CH ( g) + HClg ( )<br />

6 11 6 10<br />

In order to obtain the rate constant for this reaction, <strong>we</strong> would <strong>have</strong> to measure the<br />

pressure <strong>of</strong> CXCl. This is extremely difficult to do since the reaction vessel contains<br />

a mixture <strong>of</strong> gases. On the other h<strong>and</strong>, measuring the total pressure is relatively easy.<br />

We need to develop a method to relate the rate constant <strong>of</strong> the reaction to the relative<br />

total pressures at various points in time. In order to do this, <strong>we</strong> <strong>have</strong> to make<br />

assumptions as to the reaction order with respect to the reactant. If <strong>we</strong> assume the<br />

above reaction is first order in [CXCl], <strong>we</strong> can define some boundary conditions:<br />

Let z t be the total pressure at some time t . Initially, zt= z0,<br />

<strong>and</strong> at the end <strong>of</strong> the<br />

reaction, zt= z∞. We now need a functional form <strong>of</strong> z t that satisfies these boundary<br />

conditions while incorporating a single exponential decay characteristic <strong>of</strong> our first<br />

kt<br />

order reaction ( [ A] [ A] 0e<br />

−<br />

= ):<br />

t<br />

( )<br />

z z z z e −<br />

= + −<br />

∞ 0 ∞<br />

We do not necessarily know where z∞ is because <strong>of</strong> equilibrium conditions, unless<br />

<strong>we</strong> do a good thermodynamic analysis, but <strong>we</strong> can relate the total pressure z t at some<br />

time t with the total pressure zt+ τ at some later time t + τ .<br />

Now, subtracting [1] from [2];<br />

t<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 26 <strong>of</strong> 33<br />

kt<br />

( 0 ) kt<br />

( 0 ) − k( t+<br />

τ )<br />

( ) −kt −k<br />

z z z z e −<br />

= + − [1]<br />

∞ ∞<br />

z = z + z − z e<br />

t+<br />

τ ∞ ∞<br />

z z z z e e τ<br />

= + − [2]<br />

t+<br />

τ ∞ 0 ∞<br />

t+ τ t ∞<br />

−kt −k<br />

( 0 ) ( 1 )<br />

z z z z e e τ<br />

− = − − [3]<br />

⎡ ⎤<br />

⎣ ⎦<br />

k<br />

<strong>and</strong> ln ( zt zt) kt ln ( z z0)( 1 e ) τ −<br />

+ τ − =− + ∞ − −<br />

Now <strong>we</strong> can plot ( z z )<br />

a Guggenheim Plot.<br />

ln t+ τ t − vs. time <strong>and</strong> expect a linear relationship. This is called


From this plot, the slope is equal to − k <strong>and</strong> the intercept (t = 0) is equal to<br />

ln z − z . Thus <strong>we</strong> can extract the initial pressure. Now, the intercept is also equal<br />

( )<br />

τ<br />

0<br />

⎡ ⎤<br />

⎣ ⎦ , <strong>and</strong> from this <strong>we</strong> obtain z∞ .<br />

k<br />

to ln ( z z0)( 1 e ) τ −<br />

∞ − −<br />

In the Kezdy-Swinbourne formalism, dividing equation [1] by equation [2] gives:<br />

zt−z∞ z − z<br />

t+<br />

τ ∞<br />

<strong>and</strong> ( 1 )<br />

A plot <strong>of</strong> z t vs. zt+ τ gives a slope <strong>of</strong><br />

(for z∞ ).<br />

Note that zt zt+ τ<br />

with the 45 o line, this corresponds to z∞ .<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 27 <strong>of</strong> 33<br />

= e<br />

kτ<br />

z = z − e + z e<br />

kτkτ t ∞ t+<br />

τ<br />

k<br />

e τ k<br />

(to extract k ) <strong>and</strong> an intercept <strong>of</strong> z ( 1 e ) τ<br />

∞ −<br />

= at t =∞ (equilibrium). <strong>So</strong> when the experimental data intersects<br />

In general, for either method, you want to use a τ value that is 2 – 3 half lives


Half-lives:<br />

One relevant parameter in the interpretation <strong>of</strong> kinetic data is called the half-life. The<br />

[ A] = [ A]<br />

/2.<br />

half-life τ 1/2 is defined as the time when 0<br />

Zero order reaction:<br />

From the integrated rate law: [ A] = [ A] 0 − kt<br />

The half-life is: [ A] 0 /2 = [ A] 0 − kτ1/2<br />

<strong>So</strong>:<br />

First order reaction:<br />

[ A]<br />

τ 1/2 =<br />

2k<br />

kt<br />

From the integrated rate law: [ A] [ A] 0e<br />

−<br />

=<br />

k<br />

[ A] /2 [ A] e τ −<br />

=<br />

<strong>So</strong>: 1/2<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 28 <strong>of</strong> 33<br />

0<br />

0 0<br />

− ln 2 =− kτ<br />

τ =<br />

ln 2<br />

k<br />

Note that the first order half-life does not depend on [A]0.<br />

Second order reaction – Type I:<br />

From the integrated rate law:<br />

1/2<br />

1 1<br />

=<br />

[ A] [ A]<br />

0<br />

+ kt<br />

2<br />

[ A] 1<br />

=<br />

[ A]<br />

+ kτ<br />

Now: 1/2<br />

0 0<br />

<strong>So</strong>: 1/2<br />

Pseudo-first order reaction:<br />

1<br />

τ =<br />

[ A] k<br />

ln 2 ln 2<br />

Similar to the first order reaction: τ 1/2 = =<br />

k' [ A] 0 k<br />

Where k ' is the pseudo-first order rate constant <strong>and</strong> k is the second order rate<br />

constant.<br />

0<br />

1/2


VI. Thermodynamic vs. kinetic control:<br />

There will be a question on the exam regarding this material…<br />

Consider the reaction:<br />

When run under different conditions, different product ratios are obtained:<br />

Why?<br />

HBr<br />

⎯⎯⎯→<br />

Recall parallel reactions:<br />

Conditions: 1-2 addition 1-4 addition<br />

-80 o C 81% 19%<br />

Room Temp 44% 56%<br />

45 o C 15% 85%<br />

k1<br />

A⎯⎯→ B<br />

k2<br />

A⎯⎯→C In reality, ho<strong>we</strong>ver, these are really equilibrium processes:<br />

A←⎯⎯ ⎯⎯→ B<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 29 <strong>of</strong> 33<br />

k1<br />

k−1<br />

k2<br />

k−2<br />

A←⎯⎯ ⎯⎯→ C<br />

We can now express the rate laws for each component:<br />

Br<br />

−dA<br />

[ ]<br />

= k1[ A] + k2[ A] −k−1[ B] − k−2[ C]<br />

dt<br />

dB [ ]<br />

= k1[ A] − k−1[ B]<br />

dt<br />

dC [ ]<br />

= k2[ A] − k−2[ C]<br />

dt<br />

If <strong>we</strong> consider very short times (i.e.: [B] <strong>and</strong> [C] are so small that the reverse rates do<br />

not occur), <strong>we</strong> need only consider the forward rates. We obtain the expressions for<br />

[B] <strong>and</strong> [C] as a function <strong>of</strong> time:<br />

+<br />

Br


k [ A]<br />

B = −e<br />

1 0 [ ] 1<br />

k1+ k2<br />

− ( k1+ k2) t { }<br />

This is from before. We also found that:<br />

finally:<br />

[ B] k<br />

[ C] k<br />

1 = , so<br />

2<br />

[ B]<br />

= e<br />

[ C]<br />

k [ A]<br />

k + k<br />

2 0 − ( k1+ k2) t<br />

<strong>and</strong> [ C] = { 1−e<br />

}<br />

[ B] Ae<br />

=<br />

[ C] Ae<br />

−ΔEa/<br />

RT<br />

1 2<br />

−Ea(<br />

B)/ RT<br />

−Ea(<br />

C)/ RT<br />

For short times, the ratio <strong>of</strong> products depends only on the difference in activation<br />

energies! The product with the larger forward rate constant is called the kinetic<br />

product, <strong>and</strong> its concentration is dominant for short times in the reaction.<br />

Now consider the reaction at long (equilibrium) times. We do not <strong>have</strong> to solve the<br />

rate laws, just look at the equilibrium conditions:<br />

K<br />

1<br />

[ B]<br />

[ C]<br />

[ B] K1<br />

= <strong>and</strong> K2<br />

= so =<br />

[ A]<br />

[ A]<br />

[ C] K<br />

<strong>So</strong> the ratio depends on the relative equilibrium constants.<br />

[ B] K e<br />

[ C] K e<br />

−ΔG(<br />

B)/ RT<br />

= 1 =<br />

2<br />

<strong>and</strong><br />

−ΔGC<br />

( )/ RT<br />

[ B]<br />

= e<br />

[ C]<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 30 <strong>of</strong> 33<br />

2<br />

−ΔΔG/<br />

RT<br />

At equilibrium, the ratio <strong>of</strong> products depends on the difference in Δ G values for each<br />

product. The product with the lo<strong>we</strong>r Δ G is called the thermodynamic product, <strong>and</strong><br />

its concentration is dominant at long times in the reaction.


Since Ea(1− 2) < Ea(1−<br />

4) , k(1− 2) > k(1−<br />

4) , thus 1-2 addition product formed faster<br />

initially. The 1-2 addition is the kinetic product.<br />

But since ΔG(1− 2) >ΔG(1− 4) , the 1-4 addition is favored at equilibrium. The 1-4<br />

addition is the thermodynamic product.<br />

Remember – a faster reaction does not imply a lo<strong>we</strong>r Δ G !!<br />

Also – one product can be favored both initially <strong>and</strong> at equilibrium (depending on the<br />

relative activation energies <strong>and</strong> G<br />

Δ ’s.<br />

VII. Perturbation/Relaxation Kinetics<br />

Imagine you want to study the kinetics <strong>of</strong> a reaction that reaches equilibrium in a very<br />

short amount <strong>of</strong> time. The specific information <strong>of</strong> concentration vs. time may not be<br />

experimentally accessible. In such an experiment, you can perform a perturbation<br />

experiment, where the equilibrium conditions are altered by a sudden change in some<br />

external variable such as temperature or pressure. Then, the rate <strong>of</strong> adjustment to the<br />

new equilibrium concentrations is measured.<br />

The key assumption is that the time required to apply the perturbation is much smaller<br />

than the reaction time.<br />

Consider the first order reaction:<br />

A←⎯⎯ ⎯⎯→ B<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 31 <strong>of</strong> 33<br />

k1<br />

k−1<br />

−dA<br />

[ ]<br />

rate = = k1[ A] − k−1[ B]<br />

dt<br />

[ B]<br />

eq<br />

At equilibrium: = K1<br />

[ A]<br />

eq<br />

Now <strong>we</strong> apply a perturbation (i.e.: T-jump) to the equilibrium system <strong>and</strong> monitor the<br />

kinetics as the system relaxes to the new equilibrium.<br />

We can express the concentrations <strong>of</strong> each component as the sum <strong>of</strong> the initial<br />

equilibrium concentration <strong>and</strong> some change in concentration:<br />

After the perturbation:<br />

[ A] = [ A] +Δ<br />

[ A]<br />

eq


[ B] = [ B] +Δ [ B]<br />

We now take these two concentrations <strong>and</strong> plug them into the rate expression above:<br />

now:<br />

( [ ] eq [ ] )<br />

− d A +Δ A<br />

rate = = k1 A +Δ A − k 1 B +Δ B<br />

dt<br />

( )<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 32 <strong>of</strong> 33<br />

eq<br />

( [ ] eq [ ] ) − ( [ ] eq [ ] )<br />

− d [ A] eq +Δ[ A] −dA [ ] eq dΔ[ A]<br />

= − = k1[ A] eq + k1Δ[ A] −k−1[ B] eq −k−1Δ [ B]<br />

dt dt dt<br />

but <strong>we</strong> know that<br />

dA [ ] eq<br />

− = 0<br />

dt<br />

We can now represent the change in concentration for each component in terms <strong>of</strong> a<br />

common factor Δ [ X ] =−Δ [ A] =Δ [ B]<br />

. This comes from the stoichiometry, <strong>and</strong> the<br />

choice to <strong>have</strong> the Δ [ A]<br />

term negative is arbitrary, as long as the signs for Δ [ A]<br />

<strong>and</strong><br />

Δ [ B]<br />

are different. We now express our rate in terms <strong>of</strong> this new variable:<br />

But <strong>we</strong> know that at equilibrium:<br />

dΔ[ X]<br />

rate = = k1[ A] eq −k1Δ[ X ] −k−1[ B] eq −k−1Δ [ X ]<br />

dt<br />

k [ A] = k [ B]<br />

[forward rate equals reverse rate]<br />

1 eq −1<br />

eq<br />

dΔ[ X]<br />

rate = =− k1+ k−1 Δ [ X ]<br />

dt<br />

Now: ( )<br />

We can solve this differential equation by separating variables:<br />

dΔ[ X]<br />

=− ( k1+ k−1) dt<br />

Δ[<br />

X ]<br />

0<br />

( k k )<br />

−<br />

Δ [ X] =Δ [ X] e<br />

− 1+ 1<br />

Define the relaxation time τ as the time required for<br />

point in the decay). For this particular reaction:<br />

1<br />

τ =<br />

k +<br />

k<br />

1 −1<br />

[ ]<br />

Δ[<br />

X ]<br />

e<br />

0<br />

Δ X = (i.e.: the 1/e


For other reactions:<br />

Key points:<br />

Mechanism 1/τ<br />

A←⎯⎯→ ⎯ B<br />

k1+ k−1 A B⎯⎯→ P<br />

k [ A] eq + [ B] eq + k− + ← ⎯ 1( ) 1<br />

+ + ← ⎯ ( [ ] eq[ ] eq [ ] eq[ ] eq [ ] eq[ ] eq )<br />

+ ←⎯⎯→ ⎯ +<br />

k1( [ A] eq + [ B] eq ) + k−1( [ P] eq + [ Q]<br />

eq )<br />

A B C ⎯⎯→ P<br />

A B P Q<br />

2A⎯⎯→ A<br />

k A B + B C + A C + k− 1 1<br />

4 k [ A] eq + k− ← ⎯ 2<br />

1 1<br />

(1) Start with rate laws<br />

(2) Apply perturbation, express in terms <strong>of</strong> Δ [ X ] (from stoichiometry)<br />

2<br />

(3) All Δ [ X ] eq terms cancel out, all Δ [ X ] <strong>and</strong> all dΔ [ X]/ dt terms go to zero<br />

dΔ[ X]<br />

(4) You are left with = ( stuff ) Δ [ X ] (first order behavior)<br />

dt<br />

(5) Integrate <strong>and</strong> solve for 1/e point.<br />

Note that <strong>we</strong> <strong>have</strong> not considered coupled equilibria..<br />

D. Savin © 2009 PSC 480 Kinetics Supplement Page 33 <strong>of</strong> 33

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!