Rate-determining step

Last updated

In chemical kinetics, the overall rate of a reaction is often approximately determined by the slowest step, known as the rate-determining step (RDS or RD-step [1] or r/d step [2] [3] ) or rate-limiting step. For a given reaction mechanism, the prediction of the corresponding rate equation (for comparison with the experimental rate law) is often simplified by using this approximation of the rate-determining step.

Contents

In principle, the time evolution of the reactant and product concentrations can be determined from the set of simultaneous rate equations for the individual steps of the mechanism, one for each step. However, the analytical solution of these differential equations is not always easy, and in some cases numerical integration may even be required. [4] The hypothesis of a single rate-determining step can greatly simplify the mathematics. In the simplest case the initial step is the slowest, and the overall rate is just the rate of the first step.

Also, the rate equations for mechanisms with a single rate-determining step are usually in a simple mathematical form, whose relation to the mechanism and choice of rate-determining step is clear. The correct rate-determining step can be identified by predicting the rate law for each possible choice and comparing the different predictions with the experimental law, as for the example of NO2 and CO below.

The concept of the rate-determining step is very important to the optimization and understanding of many chemical processes such as catalysis and combustion.

Example reaction: NO2 + CO

As an example, consider the gas-phase reaction NO2 + CO → NO + CO2. If this reaction occurred in a single step, its reaction rate (r) would be proportional to the rate of collisions between NO2 and CO molecules: r = k[NO2][CO], where k is the reaction rate constant, and square brackets indicate a molar concentration. Another typical example is the Zel'dovich mechanism.

First step rate-determining

In fact, however, the observed reaction rate is second-order in NO2 and zero-order in CO, [5] with rate equation r = k[NO2]2. This suggests that the rate is determined by a step in which two NO2 molecules react, with the CO molecule entering at another, faster, step. A possible mechanism in two elementary steps that explains the rate equation is:

  1. NO2 + NO2 → NO + NO3(slow step, rate-determining)
  2. NO3 + CO → NO2 + CO2(fast step)

In this mechanism the reactive intermediate species NO3 is formed in the first step with rate r1 and reacts with CO in the second step with rate r2. However, NO3 can also react with NO if the first step occurs in the reverse direction (NO + NO3 → 2 NO2) with rate r−1, where the minus sign indicates the rate of a reverse reaction.

The concentration of a reactive intermediate such as [NO3] remains low and almost constant. It may therefore be estimated by the steady-state approximation, which specifies that the rate at which it is formed equals the (total) rate at which it is consumed. In this example NO3 is formed in one step and reacts in two, so that

The statement that the first step is the slow step actually means that the first step in the reverse direction is slower than the second step in the forward direction, so that almost all NO3 is consumed by reaction with CO and not with NO. That is, r−1r2, so that r1r2 ≈ 0. But the overall rate of reaction is the rate of formation of final product (here CO2), so that r = r2r1. That is, the overall rate is determined by the rate of the first step, and (almost) all molecules that react at the first step continue to the fast second step.

Pre-equilibrium: if the second step were rate-determining

The other possible case would be that the second step is slow and rate-determining, meaning that it is slower than the first step in the reverse direction: r2r−1. In this hypothesis, r1 − r−1 ≈ 0, so that the first step is (almost) at equilibrium. The overall rate is determined by the second step: r = r2r1, as very few molecules that react at the first step continue to the second step, which is much slower. Such a situation in which an intermediate (here NO3) forms an equilibrium with reactants prior to the rate-determining step is described as a pre-equilibrium [6] For the reaction of NO2 and CO, this hypothesis can be rejected, since it implies a rate equation that disagrees with experiment.

  1. NO2 + NO2 → NO + NO3(fast step)
  2. NO3 + CO → NO2 + CO2(slow step, rate-determining)

If the first step were at equilibrium, then its equilibrium constant expression permits calculation of the concentration of the intermediate NO3 in terms of more stable (and more easily measured) reactant and product species:

The overall reaction rate would then be

which disagrees with the experimental rate law given above, and so disproves the hypothesis that the second step is rate-determining for this reaction. However, some other reactions are believed to involve rapid pre-equilibria prior to the rate-determining step, as shown below.

Nucleophilic substitution

Another example is the unimolecular nucleophilic substitution (SN1) reaction in organic chemistry, where it is the first, rate-determining step that is unimolecular. A specific case is the basic hydrolysis of tert-butyl bromide (t-C
4
H
9
Br
) by aqueous sodium hydroxide. The mechanism has two steps (where R denotes the tert-butyl radical t-C
4
H
9
):

  1. Formation of a carbocation R−Br → R+
    + Br
    .
  2. Nucleophilic attack by hydroxide ion R+
    + OH
    → ROH.

This reaction is found to be first-order with r = k[R−Br], which indicates that the first step is slow and determines the rate. The second step with OH is much faster, so the overall rate is independent of the concentration of OH.

In contrast, the alkaline hydrolysis of methyl bromide (CH
3
Br
) is a bimolecular nucleophilic substitution (SN2) reaction in a single bimolecular step. Its rate law is second-order: r = k[R−Br][OH
].

Composition of the transition state

A useful rule in the determination of mechanism is that the concentration factors in the rate law indicate the composition and charge of the activated complex or transition state. [7] For the NO2–CO reaction above, the rate depends on [NO2]2, so that the activated complex has composition N
2
O
4
, with 2 NO2 entering the reaction before the transition state, and CO reacting after the transition state.

A multistep example is the reaction between oxalic acid and chlorine in aqueous solution: H
2
C
2
O
4
+ Cl
2
→ 2 CO2 + 2 H+
+ 2 Cl
. [7] The observed rate law is

which implies an activated complex in which the reactants lose 2H+
+ Cl
before the rate-determining step. The formula of the activated complex is Cl
2
+ H
2
C
2
O
4
− 2 H+
Cl
+ x H2O , or C
2
O
4
Cl(H
2
O)
x
(an unknown number of water molecules are added because the possible dependence of the reaction rate on H2O was not studied, since the data were obtained in water solvent at a large and essentially unvarying concentration).

One possible mechanism in which the preliminary steps are assumed to be rapid pre-equilibria occurring prior to the transition state is [7]

Cl
2
+ H2O HOCl + Cl
+ H+
H
2
C
2
O
4
H+
+ HC
2
O
4
HOCl + HC
2
O
4
H2O + Cl
+ 2 CO2

Reaction coordinate diagram

In a multistep reaction, the rate-determining step does not necessarily correspond to the highest Gibbs energy on the reaction coordinate diagram. [8] [6] If there is a reaction intermediate whose energy is lower than the initial reactants, then the activation energy needed to pass through any subsequent transition state depends on the Gibbs energy of that state relative to the lower-energy intermediate. The rate-determining step is then the step with the largest Gibbs energy difference relative either to the starting material or to any previous intermediate on the diagram. [8] [9]

Also, for reaction steps that are not first-order, concentration terms must be considered in choosing the rate-determining step. [8] [6]

Chain reactions

Not all reactions have a single rate-determining step. In particular, the rate of a chain reaction is usually not controlled by any single step. [8]

Diffusion control

In the previous examples the rate determining step was one of the sequential chemical reactions leading to a product. The rate-determining step can also be the transport of reactants to where they can interact and form the product. This case is referred to as diffusion control and, in general, occurs when the formation of product from the activated complex is very rapid and thus the provision of the supply of reactants is rate-determining.

See also

Related Research Articles

In a chemical reaction, chemical equilibrium is the state in which both the reactants and products are present in concentrations which have no further tendency to change with time, so that there is no observable change in the properties of the system. This state results when the forward reaction proceeds at the same rate as the reverse reaction. The reaction rates of the forward and backward reactions are generally not zero, but they are equal. Thus, there are no net changes in the concentrations of the reactants and products. Such a state is known as dynamic equilibrium.

<span class="mw-page-title-main">Chemical reaction</span> Process that results in the interconversion of chemical species

A chemical reaction is a process that leads to the chemical transformation of one set of chemical substances to another. Classically, chemical reactions encompass changes that only involve the positions of electrons in the forming and breaking of chemical bonds between atoms, with no change to the nuclei, and can often be described by a chemical equation. Nuclear chemistry is a sub-discipline of chemistry that involves the chemical reactions of unstable and radioactive elements where both electronic and nuclear changes can occur.

<span class="mw-page-title-main">Stoichiometry</span> Calculation of relative weights of reactants and products in chemical reactions

Stoichiometry is the relationship between the weights of reactants and products before, during, and following chemical reactions.

A chemical equation is the symbolic representation of a chemical reaction in the form of symbols and chemical formulas. The reactant entities are given on the left-hand side and the product entities are on the right-hand side with a plus sign between the entities in both the reactants and the products, and an arrow that points towards the products to show the direction of the reaction. The chemical formulas may be symbolic, structural, or intermixed. The coefficients next to the symbols and formulas of entities are the absolute values of the stoichiometric numbers. The first chemical equation was diagrammed by Jean Beguin in 1615.

In chemistry, a nucleophilic substitution is a class of chemical reactions in which an electron-rich chemical species replaces a functional group within another electron-deficient molecule. The molecule that contains the electrophile and the leaving functional group is called the substrate.

<span class="mw-page-title-main">Reaction rate</span> Speed at which a chemical reaction takes place

The reaction rate or rate of reaction is the speed at which a chemical reaction takes place, defined as proportional to the increase in the concentration of a product per unit time and to the decrease in the concentration of a reactant per unit time. Reaction rates can vary dramatically. For example, the oxidative rusting of iron under Earth's atmosphere is a slow reaction that can take many years, but the combustion of cellulose in a fire is a reaction that takes place in fractions of a second. For most reactions, the rate decreases as the reaction proceeds. A reaction's rate can be determined by measuring the changes in concentration over time.

Chemical kinetics, also known as reaction kinetics, is the branch of physical chemistry that is concerned with understanding the rates of chemical reactions. It is different from chemical thermodynamics, which deals with the direction in which a reaction occurs but in itself tells nothing about its rate. Chemical kinetics includes investigations of how experimental conditions 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 also can describe the characteristics of a chemical reaction.

In chemistry, a reaction mechanism is the step by step sequence of elementary reactions by which overall chemical reaction occurs.

In physical organic chemistry, a kinetic isotope effect (KIE) is the change in the reaction rate of a chemical reaction when one of the atoms in the reactants is replaced by one of its isotopes. Formally, it is the ratio of rate constants for the reactions involving the light (kL) and the heavy (kH) isotopically substituted reactants (isotopologues):

In chemistry, the rate equation is an empirical differential mathematical expression for the reaction rate of a given reaction in terms of concentrations of chemical species and constant parameters only. For many reactions, the initial rate is given by a power law such as

In chemistry, molecularity is the number of molecules that come together to react in an elementary (single-step) reaction and is equal to the sum of stoichiometric coefficients of reactants in the elementary reaction with effective collision and correct orientation. Depending on how many molecules come together, a reaction can be unimolecular, bimolecular or even trimolecular.

<span class="mw-page-title-main">Acid catalysis</span> Chemical reaction

In acid catalysis and base catalysis, a chemical reaction is catalyzed by an acid or a base. By Brønsted–Lowry acid–base theory, the acid is the proton (hydrogen ion, H+) donor and the base is the proton acceptor. Typical reactions catalyzed by proton transfer are esterifications and aldol reactions. In these reactions, the conjugate acid of the carbonyl group is a better electrophile than the neutral carbonyl group itself. Depending on the chemical species that act as the acid or base, catalytic mechanisms can be classified as either specific catalysis and general catalysis. Many enzymes operate by general catalysis.

In chemistry, a steady state is a situation in which all state variables are constant in spite of ongoing processes that strive to change them. For an entire system to be at steady state, i.e. for all state variables of a system to be constant, there must be a flow through the system. A simple example of such a system is the case of a bathtub with the tap running but with the drain unplugged: after a certain time, the water flows in and out at the same rate, so the water level stabilizes and the system is in a steady state.

Reactions on surfaces are reactions in which at least one of the steps of the reaction mechanism is the adsorption of one or more reactants. The mechanisms for these reactions, and the rate equations are of extreme importance for heterogeneous catalysis. Via scanning tunneling microscopy, it is possible to observe reactions at the solid gas interface in real space, if the time scale of the reaction is in the correct range. Reactions at the solid–gas interface are in some cases related to catalysis.

In chemistry, a reaction intermediate, or intermediate, is a molecular entity arising within the sequence of a stepwise chemical reaction. It is formed as the reaction product of an elementary step, from the reactants and/or preceding intermediates, but is consumed in a later step. It does not appear in the chemical equation for the overall reaction.

<span class="mw-page-title-main">Transition state theory</span> Theory describing the reaction rates of elementary chemical reactions

In chemistry, transition state theory (TST) explains the reaction rates of elementary chemical reactions. The theory assumes a special type of chemical equilibrium (quasi-equilibrium) between reactants and activated transition state complexes.

In chemical kinetics, the Lindemann mechanism is a schematic reaction mechanism for unimolecular reactions. Frederick Lindemann and J. A. Christiansen proposed the concept almost simultaneously in 1921, and Cyril Hinshelwood developed it to take into account the energy distributed among vibrational degrees of freedom for some reaction steps.

In chemistry, hyponitrite may refer to the anion N
2
O2−
2
([ON=NO]2−), or to any ionic compound that contains it. In organic chemistry, it may also refer to the group −O−N=N−O−, or any organic compound with the generic formula R1−O−N=N−O−R2, where R1 and R2 are organic groups. Such compounds can be viewed as salts and esters of respectively hyponitrous acid H
2
N
2
O
2
or HON=NOH.

In chemistry, dissociative substitution describes a reaction pathway by which compounds interchange ligands. The term is typically applied to coordination and organometallic complexes, but resembles the SN1 mechanism in organic chemistry. This pathway can be well described by the cis effect, or the labilization of CO ligands in the cis position. The opposite pathway is associative substitution, being analogous to SN2 pathway. Pathways that are intermediate between the pure dissociative and pure associative pathways are called interchange mechanisms.

Chemical reaction network theory is an area of applied mathematics that attempts to model the behaviour of real-world chemical systems. Since its foundation in the 1960s, it has attracted a growing research community, mainly due to its applications in biochemistry and theoretical chemistry. It has also attracted interest from pure mathematicians due to the interesting problems that arise from the mathematical structures involved.

References

  1. Kozuch, Sebastian; Martin, Jan (June 2011). "The Rate-Determining Step is Dead. Long Live the Rate-Determining State". ChemPhysChem DOI:10.1002/CPHC.201100137. 12 (8): 1413–1418. doi:10.1002/cphc.201100137. PMID   21523880 via PubMed.
  2. {{Organic Chemistry, Volume 1, 6/E By Finar}}
  3. {{Aliphatic Organic Chemistry By Amit Arora}}
  4. Steinfeld J. I., Francisco J. S., Hase W. L. Chemical Kinetics and Dynamics (2nd ed., Prentice-Hall 1999) ch. 2.
  5. Whitten K. W., Galley K. D., Davis R. E. General Chemistry (4th edition, Saunders 1992), p. 638–639.
  6. 1 2 3 Peter Atkins and Julio de Paula, Physical Chemistry (8th ed., W. H. Freeman 2006) p. 814–815. ISBN   0-7167-8759-8.
  7. 1 2 3 Espenson, J. H. (2002). Chemical Kinetics and Reaction Mechanisms (2nd ed.). McGraw-Hill. pp. 127–132. ISBN   0072883626.
  8. 1 2 3 4 Keith J. Laidler. Chemical Kinetics (3rd ed., Harper and Row 1987) p. 283–285. ISBN   0-06-043862-2.
  9. Murdoch, Joseph R. (1981). "What is the rate-limiting step of a multistep reaction?". Journal of Chemical Education. 58 (1): 32–36. Bibcode:1981JChEd..58...32M. doi:10.1021/ed058p32.