APPENDIX I HOW TO USE THE CRE WEB RESOURCES 931
E. Biomass Reactions
Substrate (S) + Cells (C) → More Cells + Product
Note: The rate constants, k, and activation energies for a number of the reactions in these exam- ples are given in the Databases in Appendix E.
TABLE 3-1 EXAMPLESOF REACTION RATE LAWS (CONTINUED)
NO2
Cl+ 2NH3
NO2 NH2
+ NH4Cl rONCB kONCBCONCBCNH 3
CNBr CH NH+ 3 2⎯ →⎯ CH Br NCNH3 + 2 rCNBr kCCNBrCCH
3NH2
CH3COOC2H5C4H9OH ⎯→←⎯ CH3COOC4H9C2H5OH ⎯→←⎯
A + BB C + D rAk C[ ACBCCCD/KC]
CH CHO3 ⎯ →cat⎯ CH4+CO rCH
3CHO
kC3 2CH3CHO
CH(CH3)2
+ C3H6 Cumene (C) Benzene (B) + Propylene (P)
cat
rC
k P[ CPBPPKP] 1KBPBKCPC ---
H2O
⎯⎯→ rU kCU KMCU ---
rS
kCSCC KSCS ---
CO C1 2→COC12
in which the kinetic rate law is
This reaction is first order with respect to carbon monoxide, three-halves order with respect to chlorine, and five-halves order overall.
Sometimes reactions have complex rate expressions that cannot be sepa- rated into solely temperature-dependent and concentration-dependent portions.
In the decomposition of nitrous oxide,
the kinetic rate law is
Both and are strongly temperature dependent. When a rate expression such as the one given above occurs, we cannot state an overall reaction order.
Here, we can only speak of reaction orders under certain limiting conditions.
For example, at very low concentrations of oxygen, the second term in the denominator would be negligible with respect to 1 (1 >> ), and the reac- tion would be “apparent” first order with respect to nitrous oxide and first order overall. However, if the concentration of oxygen were large enough so that the number 1 in the denominator were insignificant in comparison with the second term, ( >> 1), the apparent reaction order would be –1 with respect to oxygen and first order with respect to nitrous oxide, giving an overall apparent zero order. Rate expressions of this type are very common for liquid and gaseous reactions promoted by solid catalysts (see Chapter 10).
They also occur in homogeneous reaction systems with reactive intermediates (see Chapter 9).
It is interesting to note that although the reaction orders often correspond to the stoichiometric coefficients, as evidenced for the reaction between hydro- gen and iodine, just discussed to form HI, the rate expression for the reaction between hydrogen and another halogen, bromine, is quite complex. This non- elementary reaction
proceeds by a free-radical mechanism, and its reaction rate law is
(3-8)
Apparent First-Order Reactions. Because the law of mass action in collision theory shows that two molecules must collide giving a second-order dependence on the rate, you are probably wondering how rate laws such as Equation (3-8) as well as the rate law for first-order reactions come about. An example of first-order reaction not involving radioactive decay is the decomposition of eth- anol to form ethylene and hydrogen.
rCO
kCCOCC1
2 3 2
2N2O→2N2O2
rN
2O
kN
2OCN
2O
1 kCO 2
---
kN
2O k
kCO
2
kCO
2 kCO
2
H2Br2→2HBr
rBr2
kBr
2CH
2CBr
2 1 2
k' CHBr CBr 2
---
In terms of symbols
Rate laws of this form usually involve a number of elementary reactions and at least one active intermediate. An active intermediate is a high-energy molecule that reacts virtually as fast as it is formed. As a result, it is present in very small concentrations. Active intermediates (e.g., A*) can be formed by colli- sion or interaction with other molecules (M) such as inerts or reactants
Here, the activation occurs when translational kinetic energy is transferred into energy stored in internal degrees of freedom, particularly vibrational degrees of freedom.3 An unstable molecule (i.e., active intermediate) is not formed solely as a consequence of the molecule moving at a high velocity (high-translational kinetic energy). The translational kinetic energy must be absorbed into the chemical bonds where high-amplitude oscillations will lead to bond ruptures, molecular rearrangement, and decomposition. In the absence of photochemical effects or similar phenomena, the transfer of translational energy to vibrational energy to produce an active intermediate can occur only as a consequence of molecular collision or interaction. Collision theory is discussed in the Profes- sional Reference Shelf on the CRE Web site for Chapter 3.
As will be shown in Chapter 9, the mole A becomes activated to A* by collision with another molecule M. The activated molecule can become deacti- vated by collision with another molecule or the activated molecule can decom- pose to B and C.
Using this mechanism, we show in Section 9.1.1 that at high concentrations of M the rate law for this mechanism becomes
In Chapter 9, we will discuss reaction mechanisms and pathways that lead to nonelementary rate laws, such as the rate of formation of HBr shown in Equation (3-8).
3W. J. Moore, Physical Chemistry (Reading, MA: Longman Publishing Group, 1998).
⎯ →⎯ + C H2 6 C H2 4 H2
rC
2H6
kCC
2H6
A ⎯⎯→ B + C rA
kCA
+ ⎯ →⎯ + A M k1 A* M
+ ← ⎯⎯ →⎯⎯ +
↓ +
k
A M A M
B C
k k
*
3
1 2
rA
kACA
Heterogeneous Reactions. Historically, it has been the practice in many gas-solid catalyzed reactions to write the rate law in terms of partial pressures rather than concentrations. In heterogeneous catalysis it is the weight of cata- lyst that is important, rather than the reactor volume. Consequently, we use in order to write the rate law in terms of mol per kg of catalyst per time in order to design PBRs. An example of a heterogeneous reaction and corre- sponding rate law is the hydrodemethylation of toluene (T) to form benzene (B) and methane (M) carried out over a solid catalyst
The rate of disappearance of toluene per mass of catalyst, , i.e., (mol/mass/time) follows Langmuir-Hinshelwood kinetics (discussed in Chapter 10), and the rate law was found experimentally to be
where the prime in notes typical units are in per kilogram of catalyst (mol/kg-cat/s), PT, , and PB are partial pressures of toluene, hydrogen, and benzene in (kPa or atm), and KB and KT are the adsorption constants for ben- zene and toluene respectively, with units of kPa–1 (or atm–1). The specific reac- tion rate k has units of
You will find that almost all heterogeneous catalytic reactions will have a term such as (1 + KAPA + … ) or (1 + KAPA + … )2 in the denominator of the rate law (cf. Chapter 10).
To express the rate of reaction in terms of concentration rather than par- tial pressure, we simply substitute for Pi using the ideal gas law
(3-9) The rate of reaction per unit weight (i.e., mass) catalyst, (e.g., ), and the rate of reaction per unit volume, , are related through the bulk density ρb (mass of solid/volume) of the catalyst particles in the fluid media:
In fluidized catalytic beds, the bulk density, ρb, is normally a function of the volumetric flow rate through the bed.
Consequently, using the above equations for Pi and we can write the rate law for the hydromethylation of toluene in terms of concentration and in (mole/dm3) and the rate, in terms of reactor volume, i.e.,
rA
C6H5CH3H2 cat C→ 6H6CH4
rT
rT
kPH
2PT 1KBPBKTPT ---
rT
PH
2
k
{ } mol toluene kg-cat s kPa2 ---
PiCiRT
rA
rT
rA
rA
( )b (rA)
Relating rate per unit volume and rate by per unit mass of catalyst
moles time volume
--- mass volume ---
⎝ ⎠
⎛ ⎞ moles
time mass ---
⎝ ⎠
⎛ ⎞
rT rT
or as we will see in Chapter 4 leave it in terms of partial pressures.
In summary on reaction orders, they cannot be deduced from reaction stoichiometry. Even though a number of reactions follow elementary rate laws, at least as many reactions do not. One must determine the reaction order from the literature or from experiments.
3.2.3 Reversible Reactions
All rate laws for reversible reactions must reduce to the thermodynamic rela- tionship relating the reacting species concentrations at equilibrium. At equilib- rium, the rate of reaction is identically zero for all species (i.e., ).
That is, for the general reaction
(2-1) the concentrations at equilibrium are related by the thermodynamic relation- ship for the equilibrium constant KC (see Appendix C).
(3-10) The units of the thermodynamic equilibrium constant, KC, are . To illustrate how to write rate laws for reversible reactions, we will use the combination of two benzene molecules to form one molecule of hydrogen and one of diphenyl. In this discussion, we shall consider this gas-phase reac- tion to be elementary and reversible
or, symbolically,
The forward and reverse specific reaction rate constants, and , respectively, will be defined with respect to benzene.
Benzene (B) is being depleted by the forward reaction
in which the rate of disappearance of benzene is
ρ
{ }
( )
− = ⎡⎣ ′ ⎤⎦
+ + ⎛ •
⎝⎜ ⎞
⎠⎟
= •
⎛
⎝⎜
⎞
⎠⎟
r k RT C C
K RTC K RTC s
k s
1
mol dm dm
mol
b k
T
2
H T
B B T T
3
3
2
rA0 aAbB ⎯⎯→←⎯⎯ cCdD
Thermodynamic equilibrium
relationship KC CCec CDed
CAea CBeb ---
mol/dm3
( )dcba
2C6H6 ⎯⎯→←⎯⎯ kB C12H10H2
kB
2B ⎯⎯→←⎯⎯ kB D H 2
kB
kB kB
2C6H6 ⎯⎯→ kB C12H10H2
rB forward,
kBCB2
If we multiply both sides of this equation by 1, we obtain the expression for the rate of formation of benzene for the forward reaction
(3-11) For the reverse reaction between diphenyl (D) and hydrogen ( ),
the rate of formation of benzene is given as
(3-12) Again, both the rate constants kB and k–B are defined with respect to benzene!!!
The net rate of formation of benzene is the sum of the rates of formation from the forward reaction [i.e., Equation (3-11)] and the reverse reaction [i.e., Equation (3-12)]
(3-13) Multiplying both sides of Equation (3-13) by 1, and then factoring out kB, we obtain the rate law for the rate of disappearance of benzene,
Replacing the ratio of the reverse to forward rate law constants by the recipro- cal of the concentration equilibrium constant, KC, we obtain
(3-14) where
The equilibrium constant decreases with increasing temperature for exothermic reactions and increases with increasing temperature for endothermic reactions.
Let’s write the rate of formation of diphenyl, rD, in terms of the concen- trations of hydrogen, H2, diphenyl, D, and benzene, B. The rate of formation of diphenyl, rD, must have the same functional dependence on the reacting species concentrations as does the rate of disappearance of benzene, –rB. The rate of formation of diphenyl is
(3-15) rB forward, kBCB2
H2 C12H10H2 ⎯⎯→ kB 2C6H6
The specific reaction rate constant, ki, must be defined with respect to a particular species.
rB, reverse kBCDCH 2
Net rate rBrB, netrB, forwardrB, reverse rB kBCB2 kBCDCH
2
rB
Elementary reversible A ⎯→←⎯ B rA
k CA CB KC ---
⎝ ⎠
⎛ ⎞
rB
kBCB2 kBCDCH
2 kB CB2 kB kB --- CDCH 2
⎝ ⎠
⎜ ⎟
⎛ ⎞
rB
kB CB2 CDCH
2
KC ---
⎝ ⎠
⎛ ⎞
kB kB
---KCConcentration equilibrium constant
rD kD CB2 CDCH
2
KC ---
⎝ ⎠
⎛ ⎞
Using the relationship given by Equation (3-1) for the general reaction (3-1) we can obtain the relationship between the various specific reaction rates, ,
(3-16) Comparing Equations (3-15) and (3-16), we see the relationship between the specific reaction rate with respect to diphenyl, kD, and the specific reaction rate with respect to benzene, kB, is
Consequently, we see the need to define the rate constant, k, with respect to a particular species.
Finally, we need to check to see if the rate law given by Equation (3-14) is thermodynamically consistent at equilibrium. Applying Equation (3-10) (and Appendix C) to the diphenyl reaction and substituting the appropriate species concentration and exponents, thermodynamics tells us that
(3-17) Now let’s look at the rate law. At equilibrium, –rB ≡ 0, and the rate law given by Equation (3-14) becomes
Rearranging, we obtain, as expected, the equilibrium expression
that is identical to Equation (3-17) obtained from thermodynamics.
From Appendix C, Equation (C-9), we know that when there is no change in the total number of moles and the heat capacity term, ΔCP = 0, the temperature dependence of the concentration equilibrium constant is
(C-9)
Relative rates rA
a --- rB
b --- rC
---c rD ---d
kB kD rD
---1 rB 2
--- kB CB2 CDCH 2KC
[ ]
2
--- kB
--- C2 B2 CDCH
2
KC ---
kD kB ---2
KC CDeCH
2e
C2Be
---
At equilibrium, the rate law must reduce to an equation consistent wth thermodynamic
equilibrium. rB0 kB C2Be
CDeCH
2e
KC ---
KC CDeCH
2e
C2Be
---
KC( )T KC( )T1 exp HRx
---R 1 T1 --- 1
T---
⎝ ⎠
⎛ ⎞
Therefore, if we know the equilibrium constant at one temperature, T1 [i.e., KC (T1)], and the heat of reaction, , we can calculate the equilibrium constant at any other temperature T. For endothermic reactions, the equilibrium constant, KC, increases with increasing temperature; for exothermic reactions, KC decreases with increasing temperature. A further discussion of the equilibrium constant and its thermodynamic relationship is given in Appendix C. For large values of the equilibrium constant, KC, the reaction behaves as if it were irre- versible.
3.3 Rates and the Reaction Rate Constant
There are three molecular concepts relating to the rate of reaction we will discuss:
Concept 1. Law of Mass Action. The rate of reaction increases with increasing concentration of reactants owing to the increased number of molec- ular collisions at the higher reactant concentrations. In Section 3.2, we have just discussed Concept 1: rate laws, and the dependence of the rate reactant concentration. We now discuss Concept 2, potential energy surfaces and energy barriers, and Concept 3, the energy needed for crossing the barriers.
3.3.1 The Rate Constant k
The reaction rate constant k is not truly a constant; it is merely independent of the concentrations of the species involved in the reaction. The quantity k is referred to as either the specific reaction rate or the rate constant. It is almost always strongly dependent on temperature. It also depends on whether or not a catalyst is present, and in gas-phase reactions, it may be a function of total pressure. In liquid systems it can also be a function of other parameters, such as ionic strength and choice of solvent. These other variables normally exhibit much less effect on the specific reaction rate than does temperature, with the exception of supercritical solvents, such as supercritical water.
Consequently, for the purposes of the material presented here, it will be assumed that kA depends only on temperature. This assumption is valid in most laboratory and industrial reactions, and seems to work quite well.
It was the great Nobel Prize–winning Swedish chemist Svante Arrhenius (1859–1927) who first suggested that the temperature dependence of the spe- cific reaction rate, kA, could be correlated by an equation of the type
(3-18) where A = pre-exponential factor or frequency factor
E = activation energy, J/mol or cal/mol
R = gas constant = 8.314 J/mol • K = 1.987 cal/mol • K T = absolute temperature, K
Equation (3-18), known as the Arrhenius equation, has been verified empiri- cally to give the correct temperature behavior of most reaction rate constants within experimental accuracy over fairly large temperature ranges. The Arrhe- nius equation is derived in the Professional Reference Shelf R3.A: Collision Theory on the CRE Web site.
HRx
Arrhenius equation kA( )T AeE/RT
T(K) k
Additionally, one can view the activation energy in terms of collision theory (Professional Reference Shelf R3.1). By increasing the temperature, we increase the kinetic energy of the reactant molecules. This kinetic energy can in turn be transferred through molecular collisions to internal energy to increase the stretching and bending of the bonds, causing them to reach an activated state, vulnerable to bond breaking and reaction.
Why is there an activation energy? If the reactants are free radicals that essentially react immediately on collision, there usually isn’t an activation energy. However, for most atoms and molecules undergoing reaction, there is an activation energy. A couple of the reasons are that in order to react
1. The molecules need energy to distort or stretch their bonds so that they break and now can form new bonds.
2. The molecules need energy to overcome the steric and electron repul- sive forces as they come close together.
The activation energy can be thought of as a barrier to energy transfer (from kinetic energy to potential energy) between reacting molecules that must be overcome. The activation energy is the minimum increase in potential energy of the reactants that must be provided to transform the reactants into products. This increase can be provided by the kinetic energy of the colliding molecules.
In addition to the concentrations of the reacting species, there are two other factors that affect the rate of reaction,
• the height of the barrier, i.e., activation energy, and
• the fraction of molecular collisions that have sufficient energy to cross over the barrier (i.e., react when the molecules collide).
If we have a small barrier height, the molecules colliding will need only low kinetic energies to cross over the barrier. For reactions of molecules with small barrier heights occurring at room temperatures, a greater fraction of molecules will have this energy at low temperatures. However, for larger barrier heights, we require higher temperatures where a higher fraction of colliding molecules will have the necessary energy to cross over the barrier and react. We will dis- cuss each of these concepts separately.
Concept 2. Potential Energy Surfaces and Energy Barriers. One way to view the barrier to a reaction is through the use of potential energy surfaces and the reaction coordinates. These coordinates denote the minimum potential energy of the system as a function of the progress along the reaction path as we go from reactants to an intermediate to products. For the exothermic reaction
A + BC A – B – C ⎯→ AB + C
the potential energy surface and the reaction coordinate are shown in Figures 3-1 and 3-2. Here EA, EC, EAB, and EBC are the potential energy surface energies of the reactants A, BC and product molecules (AB and C), and EABC is the energy of the complex A–B–C at the top of the barrier.
Figure 3-1(a) shows the 3–D plot of the potential energy surface, which is analogous to a mountain pass where we start out in a valley and then climb up
⎯→←⎯
to pass over the top of the pass, i.e., the col or saddle point, and proceed down into the next valley. Figure 3-1(b) shows a contour plot of the pass and valleys and the reaction coordinate as we pass over the col from valley to valley.
Energy changes as we move within the potential energy surfaces.
At point X in Figure 3-1(b), species A and BC are far apart and are not inter- acting; RBC is just the equilibrium bond length. Consequently, the potential energy is just the BC bond energy. Here, A and BC are in their minimum potential energy in the valley and the steep rise up from the valley to the col from X would correspond to increases in the potential energy as A comes close to BC. If the BC bond is stretched from its equilibrium position at X, the
Energy
Transition state
3 1
RBC, B–C Distance (angstroms) 1.5
2 2.5
1 1.5 2 2.5 3
RAB, A–B Distance (angstroms) (b)
(a)
X YRAB
, A–B Distance RBC
, B–C Di stance
X
Y
Figure 3-1 A potential energy surface for the H + CH3OH ⎯→H2 + CH2OH from the calculations of Blowers and Masel. The lines in the figure are contours of constant energy.
The lines are spaced 5 kcal/mol. Richard I. Masel, Chemical Kinetics and Catalysis, p. 370, Fig.7.6 (Wiley, 2001).
Transition state Products
Reactants
CH3 -I Bond distance in angstroms
Ground state energy
CH3I + Cl CH3CI + I
1.9 2.1 2.3 2.5 2.7 23
25 27 29 31 33 36 39 41
Potential energy
Reaction coordinate Products Reactants
X Y
Energy barrier A-B-C
A+BC
AB+C EB
HRX
(EAB + EC) (EA + EBC) EA-B-C
(a) (b)
Energy (kcal/mol)
RBC
(
A○)
Equilibrium BC position
(c)
Figure 3-2 Progress along reaction path. (a) Symbolic reaction; (b) Calculated from computational software on the CRE Web site, Chapter 3 Web Module. (c) Side view at point X in Figure 3-1(b) showing the valley.
(c)
potential energy increases up one side of the valley hill (RBC increases). The potential energy is now greater because of the attractive forces trying to bring the B–C distance back to its equilibrium position. If the BC bond is compressed at X from its equilibrium position, the repulsive forces cause the potential energy to increase on the other side of the valley hill, i.e., RBC decreases at X.
At the end of the reaction, point Y, the products AB and C are far apart on the valley floor and the potential energy is just the equilibrium AB bond energy.
The sides of the valley at Y represent the cases where AB is either compressed or stretched causing the corresponding increases in potential energy at point Y and can be described in an analogous manner to BC at point X.
We want the minimum energy path across the barrier for converting the kinetic energy of the molecules into potential energy. This path is shown by the curve X→Y in Figure 3-1 and also by Figure 3-2(a). As we move along the A–B distance axis in Figure 3-2(a), A comes closer to BC and B begins to bond with A and to push BC apart such that the potential energy of the reac- tion pair continues to increase until we arrive at the top of the energy barrier, which is the transition state. In the transition state the molecular distances between A and B and between B and C are close. As a result, the potential energy of the three atoms (molecules) is high. As we proceed further along the arc length of the reaction coordinate depicted in Figure 3-1(a), the AB bond strengthens and the BC bond weakens and C moves away from AB and the energy of the reacting pair decreases until we arrive at the valley floor where AB is far apart from C. The reaction coordinate quantifies how far the reaction has progressed. The commercial software available to carry out calculations for the transition state for the real reaction
CH3I Cl → CH3Cl I
shown in Figure 3-2(b) is discussed in the Web Module Molecular Modeling in Chemical Reaction Engineering on the CRE Web site.
We next discuss the pathway over the barrier shown along the line Y–X.
We see that for the reaction to occur, the reactants must overcome the mini- mum energy barrier, EB, shown in Figure 3-2(a). The energy barrier, EB, is related to the activation energy, E. The energy barrier height, EB, can be calcu- lated from differences in the energies of formation of the transition-state mol- ecule and the energy of formation of the reactants; that is,
(3-19) The energy of formation of the reactants can be found in the literature or cal- culated from quantum mechanics, while the energy of formation of the transi- tion state can also be calculated from quantum mechanics using a number of software packages, such as Gaussian (http://www.gaussian.com/) and Dacapo (https://wiki.fysik.dtu.dk/dacapo). The activation energy, E, is often approxi- mated by the barrier height, EB, which is a good approximation in the absence of quantum mechanical tunneling.
Now that we have the general idea for a reaction coordinate, let’s con- sider another real reaction system
H• + C2H6→ H2 + C2H5•
The energy-reaction coordinate diagram for the reaction between a hydrogen atom and an ethane molecule is shown in Figure 3-3 where the bond distor-
EBEfAB C (EfAEfB C )