Fact-checked by Grok 2 weeks ago

Rate equation

In chemical kinetics, the rate equation, also known as the rate law, is an empirical mathematical relationship that expresses the rate of a chemical reaction as a function of the concentrations of reactants, and sometimes products or catalysts. For a general reaction aA + bB \rightarrow products, it takes the form \text{rate} = k [A]^m [B]^n, where k is the rate constant, [A] and [B] are the concentrations of the reactants, and m and n are the partial reaction orders determined experimentally. The overall reaction order is the sum m + n, which classifies the reaction as zero-order, first-order, second-order, or higher. The foundational concept behind the rate equation emerged from the law of mass action, proposed by Norwegian chemists Cato Maximilian Guldberg and Peter Waage in 1864, which posits that the rate of a chemical reaction is directly proportional to the product of the concentrations of the reacting substances, each raised to a power equal to its stoichiometric coefficient for elementary steps. However, for complex, multi-step reactions, the exponents in the rate equation do not necessarily correspond to stoichiometric coefficients and must be established through experimental methods, such as the initial rates technique, where reaction rates are measured at varying initial concentrations while keeping other variables constant. This empirical nature distinguishes rate equations from balanced chemical equations, providing critical insights into the underlying reaction mechanism and the sequence of elementary steps. The rate constant k in the rate equation is temperature-dependent and follows the , k = A e^{-E_a / RT}, where A is the representing the frequency of collisions with proper orientation, E_a is the (the minimum energy barrier for the reaction), R is the , and T is the absolute in . Higher temperatures exponentially increase k by enabling more reactant molecules to overcome the activation barrier; as a rough approximation, for many reactions at around , the rate roughly doubles for every 10 °C rise. Rate equations are essential for predicting reaction behavior, optimizing industrial processes like and , and understanding phenomena in fields ranging from to biochemistry.

Fundamentals of Rate Equations

Definition and Basic Principles

In , a equation, often referred to as a rate law, is a mathematical expression that describes the relationship between the of a and the concentrations of its reactants./12:_Kinetics/12.04:_Rate_Laws) The of reaction itself is defined as the change in concentration of a reactant or product over time, typically expressed as the negative change in reactant concentration or the positive change in product concentration, adjusted for stoichiometric coefficients to ensure consistency across species. For a general reaction aA + bB \rightarrow products, the rate law takes the form \text{[rate](/page/Rate)} = k [A]^m [B]^n, where [A] and [B] are the concentrations of the reactants, m and n are the reaction orders with respect to each reactant (which may be integers, fractions, or zero), and k is the constant./12:_Kinetics/12.04:_Rate_Laws) This differential form can be written equivalently as \frac{d[\text{product}]}{dt} = k [A]^m [B]^n for product formation or -\frac{d[A]}{dt} = \frac{1}{a} k [A]^m [B]^n for reactant consumption, emphasizing that the quantifies the speed at which the reaction proceeds. The rate constant k is a fundamental parameter in the rate equation, representing the intrinsic speed of the reaction under specified conditions and incorporating factors such as , solvent, and catalysts./12:_Kinetics/12.04:_Rate_Laws) It exhibits a strong dependence on , as described by the k = A e^{-E_a / RT}, where A is the related to , E_a is the barrier, R is the , and T is the absolute ; higher temperatures exponentially increase k by providing more molecules with sufficient energy to react. While the rate of reaction measures the observable change in concentrations (e.g., in moles per liter per second), the rate law is the specific equation that mathematically models this rate as a function of concentrations, distinguishing it as an empirical tool rather than a direct measure. Rate equations are derived under key assumptions that differentiate elementary reactions from overall reactions. For an elementary reaction—a single-step process—the rate law can be directly inferred from the stoichiometry of the balanced equation, as the molecularity (number of colliding molecules) determines the order; for instance, a bimolecular elementary step yields a second-order rate law. In contrast, overall reactions, which often involve multiple elementary steps, do not permit direct rate law prediction from the net equation; instead, the observed rate law reflects the slowest (rate-determining) step or a combination of steps, requiring experimental determination to identify the effective orders. This distinction underscores that rate equations for complex mechanisms are phenomenological models, not mechanistic derivations, ensuring they accurately capture kinetic behavior without assuming a single-step pathway.

Role in Chemical Kinetics

The rate equation, foundational to , originated from the work of Maximilian Guldberg and Peter Waage, who in 1864 formulated the , proposing that the rate of a is proportional to the product of the concentrations of the reactants raised to powers equal to their stoichiometric coefficients. This insight shifted the study of chemistry from static equilibria to dynamic processes, establishing rate equations as differential expressions describing how reaction rates depend on species concentrations, thereby enabling quantitative predictions of reaction progress. Rate equations are integrated over time to derive concentration-time profiles, which mathematically express how reactant or product concentrations evolve during a , providing a direct link between kinetic parameters and observable changes. For instance, these integrated forms allow chemists to model the temporal behavior of systems, such as in batch reactions where initial concentrations and rate constants predict the full trajectory of species depletion or formation. In applications, this capability supports reactor design by informing the sizing and operational parameters of continuous systems like reactors, where rate equations determine the volume required for a target conversion based on flow rates and . Similarly, half-lives—the time for reactant concentration to halve—are derived from these profiles, aiding in forecasting durations and stability assessments without exhaustive simulations. Steady-state approximations further leverage rate equations by assuming negligible concentration changes for intermediates, simplifying the analysis of complex multi-step mechanisms to approximate overall rates. The dependence of rate equations is captured through the rate constant k, governed by the k = A e^{-E_a / RT}, where E_a is the , A is the , R is the , and T is ; higher E_a exponentially slows reactions by limiting the fraction of collisions with sufficient energy. Catalysts accelerate reactions by lowering E_a or enhancing A, thus increasing k without being consumed, as seen in enzymatic processes where stabilization reduces the energy barrier. This relationship underscores the role of rate equations in optimizing catalytic systems for industrial efficiency.

Power Law Rate Equations

General Form and Reaction Order

The power law rate equation provides a mathematical description of the reaction rate for many chemical processes, particularly elementary reactions and those approximating such behavior. For a general reaction involving reactants A, B, and others, the rate is expressed as \text{rate} = k [\ce{A}]^m [\ce{B}]^n \cdots where k is the rate constant, [\ce{A}] and [\ce{B}] are the concentrations of the reactants, and m and n are the partial reaction orders with respect to A and B, respectively. The overall reaction order is defined as the sum of these exponents, m + n + \cdots, which indicates the total dependence of the rate on reactant concentrations. This form assumes that the rate is proportional to the concentrations raised to constant powers, a simplification valid for many systems under constant conditions. Reaction order differs fundamentally from , the theoretical number of reactant molecules involved in an elementary step as per . While is an integer (unimolecular, bimolecular, etc.) deduced from the , reaction order is an experimental parameter that may be fractional or zero and does not necessarily match the . For elementary reactions, the reaction order equals the , but complex mechanisms can yield orders that deviate, highlighting the empirical nature of rate laws. The units of the rate constant k are determined by the overall reaction order n to ensure dimensional consistency, since the rate has units of concentration per time (typically \ce{M s^{-1}}). For an nth-order reaction, k has units of \ce{M^{1-n} s^{-1}}, such as \ce{s^{-1}} for first-order or \ce{M^{-1} s^{-1}} for second-order processes. This dependency arises directly from the power law structure, where balancing the equation requires k to compensate for the concentration terms. A simple example is the decomposition reaction \ce{A -> products}, where the rate law simplifies to \text{rate} = k [\ce{A}]^m and the overall order is simply m. The value of m varies by reaction type—for instance, m = 1 for many unimolecular decompositions or m = 2 for certain bimolecular processes—illustrating how the general form adapts to specific without altering its foundational structure.

Zero-Order Reactions

In zero-order reactions, the rate of reaction is independent of the concentrations of the reactants, resulting in a constant reaction rate throughout the process. The rate law for such reactions is expressed as rate = k, where k is the , indicating no dependence on reactant concentration. This form arises as a special case of the power law rate equation when the reaction order is zero, often observed under conditions where the reaction is limited by factors other than reactant availability. To derive the integrated rate law, start with the differential form -d[A]/dt = k. Integrating both sides with respect to time from t = 0 (where [A] = [A]_0) to t (where [A] = [A]) yields [A] = [A]0 - kt. This linear relationship shows that the concentration of the reactant decreases at a steady rate over time. The half-life for a zero-order reaction, the time required for the concentration to halve, is given by t{1/2} = [A]_0 / (2k), which depends on the initial concentration and inversely on the rate constant. Zero-order kinetics commonly occurs in scenarios involving saturation, such as enzyme-catalyzed reactions under the Michaelis-Menten model. When substrate concentration greatly exceeds the Michaelis constant (K_m), the enzyme active sites become fully occupied, making the rate equal to the maximum velocity V_max and independent of further substrate addition, thus zero-order in substrate. Similarly, in heterogeneous catalysis, zero-order behavior is seen in surface reactions where all catalytic sites are saturated with reactants, limiting the rate to the availability of those sites rather than gas-phase concentrations. Graphically, zero-order reactions are identified by plotting concentration [A] versus time t, which produces a straight line with a slope of -k, confirming the constant rate and allowing determination of the rate constant from experimental data.

First-Order Reactions

A first-order reaction is characterized by a rate law in which the reaction rate is directly proportional to the concentration of a single reactant, expressed as
\text{rate} = -\frac{d[\ce{A}]}{dt} = k[\ce{A}]
where k is the rate constant and [\ce{A}] is the concentration of reactant A. This form arises for elementary unimolecular processes or under conditions where the rate-determining step involves one species.
To derive the integrated rate law, separate variables and integrate:
\int_{[\ce{A}]_0}^{[\ce{A}]} \frac{d[\ce{A}]}{[\ce{A}]} = -k \int_0^t dt
yielding
\ln[\ce{A}] = \ln[\ce{A}]_0 - kt
or equivalently,
[\ce{A}] = [\ce{A}]_0 e^{-kt}.
This exponential decay describes how the concentration decreases over time. A key property is the constant half-life, t_{1/2} = \frac{\ln 2}{k}, which remains independent of the initial concentration [\ce{A}]_0, unlike higher-order reactions.
Common examples include , where the rate of disintegration is proportional to the number of undecayed nuclei, following the integrated form precisely. Another is the unimolecular decomposition of gas-phase molecules, such as the thermal of to propene, where the reaction proceeds via an energized . In multi-reactant systems, behavior can emerge under pseudo- conditions when one reactant is in large excess, though full details are covered elsewhere. Graphically, kinetics is confirmed by plotting \ln[\ce{A}] versus time, which yields a straight line with -k, allowing determination of the rate constant from experimental data. This linear relationship distinguishes it from other orders and facilitates analysis of concentration-time profiles.

Second-Order Reactions

Second-order reactions are those in which the overall reaction rate depends on the concentration of one or more reactants raised to the power of two, corresponding to an overall order of two in the rate equation. These reactions typically involve bimolecular elementary steps, where two reactant molecules collide and react. There are two primary forms: reactions involving two molecules of the same reactant (e.g., 2A → products), with the rate law rate = k [A]^2, or reactions between two different reactants (e.g., A + B → products), with rate = k [A][B]. The integrated rate law for a second-order reaction of the form 2A → products is derived by integrating the differential rate equation, yielding: \frac{1}{[A]} = \frac{1}{[A]_0} + kt where [A] is the concentration at time t, [A]_0 is the initial concentration, k is the rate constant, and t is time. This equation shows that a plot of 1/[A] versus t produces a straight line with slope equal to k, providing a graphical method to confirm second-order kinetics and determine the rate constant. For the case of A + B → products with equal initial concentrations ([A]_0 = [B]_0), the integrated form is analogous: 1/[A] = 1/[A]_0 + kt, allowing similar linear plotting of reciprocal concentrations against time. The for a second-order reaction, t_{1/2}, is the time required for the concentration of the reactant to decrease to half its initial value and is given by t_{1/2} = 1 / (k [A]_0). Unlike reactions, the half-life depends inversely on the initial concentration, meaning higher starting concentrations result in shorter half-lives. Representative examples of second-order reactions include the dimerization of (2 C_4H_6 → C_8H_{12}), which follows rate = k [C_4H_6]^2 and is often studied to illustrate hyperbolic concentration decay over time, and certain SN2 reactions, such as the reaction of methyl iodide with hydroxide ion (CH_3I + OH^- → CH_3OH + I^-), which proceeds with rate = k [CH_3I][OH^-].

Fractional and Higher-Order Reactions

Fractional-order reactions occur when the reaction order is a non-integer value, such as 1/2 or 3/2, and these typically result from complex reaction mechanisms that are not elementary steps. These mechanisms often involve multiple sequential or parallel pathways, including where reactive intermediates propagate the process, leading to rate laws that do not yield exponents upon experimental determination. For instance, a 3/2-order dependence has been observed in certain reactions, where the rate is proportional to the of one reactant's concentration combined with a linear term for another, reflecting the influence of surface effects or intermediates. Higher-order reactions, defined by an overall order n > 2, are uncommon in due to the low probability of simultaneous multi-molecular collisions required for such elementary steps. Termolecular or higher collisions demand precise alignment and energy transfer among three or more molecules, which is statistically rare compared to unimolecular or bimolecular events. The general integrated for an nth-order reaction (n ≠ 1) is derived by separating variables in the -d[A]/dt = k [A]^n, yielding: \frac{[A]^{1-n} - [A_0]^{1-n}}{1-n} = k t or equivalently, \frac{1}{[A]^{n-1}} = \frac{1}{[A_0]^{n-1}} + (n-1) k t. This form allows concentration [A] to be solved as a function of time t, with the rate constant k carrying units of (concentration)^{1-n} time^{-1}. For fractional orders like n = 3/2, the equation simplifies accordingly, though numerical methods may be needed for integration in practice. Negative orders, which imply rate inhibition by increasing reactant concentration, arise in even more intricate mechanisms and are not covered in detail here. To compare fractional and higher-order reactions with the more common integer cases, the following table summarizes key features of integrated rate laws, half-lives, and diagnostic plots for orders 0, 1, 2, and general n (n ≠ 1). Half-life t_{1/2} represents the time for [A] to reach [A_0]/2, and linear plots confirm the order by yielding a straight line with slope related to k. For zero-, first-, and second-order reactions, these expressions are exact; for general n, they extend the pattern but require caution near n = 1, where the form approaches the first-order logarithmic equation via L'Hôpital's rule. Plots for fractional or higher n use the linearized form, often requiring computational fitting for non-integer values.
Order (n)Integrated Rate LawHalf-Life (t_{1/2})Linear Plot (y vs. t)
0[A] = [A_0] - k t[A_0] / (2 k)[A] (slope = -k)
1\ln [A] = \ln [A_0] - k t\ln 2 / k (independent of [A_0])\ln [A] (slope = -k)
21/[A] = 1/[A_0] + k t1 / (k [A_0])1/[A] (slope = k)
n (>2 or fractional, n ≠ 1)1/[A]^{n-1} = 1/[A_0]^{n-1} + (n-1) k t(2^{n-1} - 1) / ((n-1) k [A_0]^{n-1})1/[A]^{n-1} (slope = (n-1) k)
These comparisons highlight how higher or fractional orders lead to more rapidly changing half-lives with initial concentration [A_0] compared to lower orders, emphasizing their sensitivity to conditions in complex systems./12%3A_Kinetics/12.05%3A_Integrated_Rate_Laws)

Pseudo-Order Approximations

In chemical kinetics, pseudo-order approximations simplify the study of multi-reactant rate laws by using a large excess of one or more reactants, rendering their concentrations effectively constant throughout the reaction. This approach reduces the apparent reaction order with respect to the limiting reactant, facilitating easier experimental analysis and integration of the rate equation. These approximations are particularly useful for reactions where direct measurement of multiple changing concentrations is challenging. A common example is the pseudo-first-order approximation applied to second-order reactions of the form → products, with rate = k [A][B]. When reactant B is present in large excess such that [B]_0 ≫ [A]_0 (typically by a factor of 10 or more), the concentration of B remains nearly constant at [B] ≈ [B]_0 during the reaction. The rate law then simplifies to rate = k' [A], where k' = k [B]_0 is the pseudo- rate constant. This transformed equation behaves like a true rate law in A, allowing the use of first-order integrated forms for data analysis, such as ln([A]/[A]_0) = -k' t. The validity of this approximation holds only while the excess condition is maintained, ensuring minimal depletion of B relative to its initial amount. Pseudo-zero-order approximations arise in scenarios where all reactants except one are maintained in large , or when mechanistic saturation makes the rate independent of the limiting reactant's concentration. For instance, in reactions following a power-law but under conditions where the limiting reactant does not influence the due to of others, the observed appear zero-order, with = k' (). This is common in catalytic or enzymatic systems where the catalyst or sites are fully saturated by , leading to a until depletion occurs. The approximation requires the excess concentrations to vastly exceed stoichiometric needs, often [excess]_0 ≫ 100 × [limiting]_0, to keep the invariant over a significant portion of the reaction progress. These approximations find widespread applications in hydrolysis reactions, such as the acid-catalyzed hydrolysis of esters like ethyl acetate (CH_3COOC_2H_5 + H_2O → CH_3COOH + C_2H_5OH), where water is in vast excess (pseudo-first-order in ester). The reaction is monitored spectroscopically, often via UV-visible absorbance changes in the products or unreacted species, allowing real-time tracking of concentration versus time under simplified kinetics. Similarly, in analytical chemistry, the reaction of Fe^{3+} with SCN^- to form the red [Fe(SCN)]^{2+} complex uses excess Fe^{3+} for pseudo-first-order conditions, enabling thiocyanate quantification through colorimetric spectroscopy. Limitations include the approximation's breakdown if the excess reactant depletes significantly (violating [excess] ≫ stoichiometry), potentially leading to non-linear kinetics and erroneous rate constants; thus, experiments must verify constancy of the excess species.

Experimental Determination of Rate Laws

Method of Initial Rates

The method of initial rates is an experimental technique in employed to determine the orders of a with respect to its reactants by measuring the instantaneous at the very beginning of the (t ≈ 0) across multiple trials with systematically varied initial concentrations. This approach relies on the power law form of the rate equation, where the initial is proportional to the initial concentrations raised to their respective orders. In the procedure, a series of experiments is performed in which the initial concentration of one reactant is varied while keeping the initial concentrations of all other reactants constant, and the initial is determined for each set of conditions, often by monitoring the change in concentration of a species (e.g., via ) over a short initial time interval. The reaction order m with respect to the varied reactant A is then calculated using the relation m = \frac{\log\left(\frac{\text{rate}_2}{\text{rate}_1}\right)}{\log\left(\frac{[\mathrm{A}]_2}{[\mathrm{A}]_1}\right)}, where rate1 and rate2 are the measured initial rates corresponding to initial concentrations [A]1 and [A]2, respectively. This process is repeated for each reactant to obtain the full rate law. The advantages of this include its simplicity, as it circumvents the need to integrate the differential rate equation—a task that becomes mathematically challenging for non-first-order or complex —and its applicability to systems where analytical is impractical or unknown. It provides a direct way to isolate the dependence on each reactant's concentration without complications from time-dependent changes. A key assumption underlying the is that during the initial measurement period, the concentrations of reactants remain essentially constant, and the buildup of products is negligible, ensuring no reverse or product inhibition interferes with the forward . This holds best for irreversible reactions or early stages far from . As an illustrative example, consider the hypothetical reaction 2A + B → products, where initial rates are measured for varied [A]0 and [B]0. The following table summarizes typical experimental data:
Experiment[A]0 (M)[B]0 (M)Initial Rate (M/s)
10.100.102.0 × 10−3
20.200.108.0 × 10−3
30.100.204.0 × 10−3
Comparing experiments 1 and 2 (constant [B]0), the order with respect to A is m = \frac{\log\left(\frac{8.0 \times 10^{-3}}{2.0 \times 10^{-3}}\right)}{\log\left(\frac{0.20}{0.10}\right)} = \frac{\log 4}{\log 2} = 2. Comparing experiments 1 and 3 (constant [A]0), the order with respect to B is n = \frac{\log\left(\frac{4.0 \times 10^{-3}}{2.0 \times 10^{-3}}\right)}{\log\left(\frac{0.20}{0.10}\right)} = \frac{\log 2}{\log 2} = 1. Thus, the rate law is rate = k [A]2 [B]. Such tabulations and calculations allow straightforward determination of orders in practice.

Integral Method

The integral method involves fitting experimental concentration-time data to the integrated form of a proposed rate law to determine the reaction order and rate constant. This approach assumes a power-law rate equation of the form rate = k [A]^n and integrates it over time, then tests for the best linear fit across the dataset to identify the order n. For integer orders, characteristic plots are constructed: for zero-order reactions, a plot of [A] versus time t is linear with slope equal to -k; for first-order, a plot of \ln [A] versus t is linear with slope -k; and for second-order, a plot of 1/[A] versus t is linear with slope k. The rate constant k is calculated directly from the slope of the linear plot, providing a quantitative measure once the order is confirmed. For reactions suspected to have non-integer or unknown orders, the integral method employs trial-and-error testing of various n values or to minimize the error between experimental and predicted concentrations from the integrated rate law. software, such as that available in tools like Excel or specialized packages, optimizes both n and k by fitting the full integrated equation to the data, offering a robust way to handle fractional orders without assuming linearity in transformed plots. This technique is particularly useful when the order deviates from simple integers, as seen in some catalytic or complex gas-phase reactions. A key advantage of the method is its utilization of the entire experimental , from initial concentrations through the reaction progress, which enhances statistical reliability compared to methods relying on limited points. Additionally, deviations from linearity in the plots can reveal changes in reaction order over time, such as those due to shifting mechanisms or approaching , allowing researchers to identify non-ideal behaviors early. As an illustrative example, consider concentration-time data from the of a reactant A. If plotting \ln [A] versus t yields a straight line, the reaction is confirmed as , with derived from the ; however, if the plot curves upward, testing 1/[A] versus t may produce linearity, indicating second-order and distinguishing the effectively using the full time course. This process confirms the and provides without needing multiple initial concentration runs.

Isolation Method

The isolation method, also known as the flooding method, is an experimental approach for determining the reaction with respect to a specific reactant in multi-reactant systems by minimizing variations in the concentrations of other . In this technique, all reactants except the one under study are introduced in large excess, typically 10 to 1000 times their stoichiometric amounts, ensuring that their concentrations remain nearly constant during the reaction. This simplification transforms the overall law into a pseudo-order form dependent only on the isolated reactant's concentration, allowing the use of standard integrated rate laws to identify the order. The process is then repeated sequentially for each reactant to construct the complete rate law. Consider a hypothetical A + B + C → products with the rate law r = k [\mathrm{A}]^m [\mathrm{B}]^n [\mathrm{C}]^p. To determine the order m with respect to A, B and C are added in large excess, rendering [\mathrm{B}] and [\mathrm{C}] effectively constant. The rate law then reduces to r = k' [\mathrm{A}]^m, where k' = k [\mathrm{B}]^n [\mathrm{C}]^p is the pseudo-rate constant. Monitoring the concentration of A over time and fitting the data to integrated rate equations for different orders (e.g., first-order: \ln[\mathrm{A}] = -k't + \ln[\mathrm{A}]_0) reveals m. This isolation is performed analogously for B and C by varying their roles. (Frost and Pearson, 1961) The method's primary advantages include its ability to reduce complex kinetics to simpler, well-characterized forms, facilitating straightforward data analysis via graphical or numerical integration, and its applicability to where substrates can be flooded relative to catalysts. It is particularly effective for reactions where direct variation of all concentrations simultaneously is impractical. However, the isolation method has limitations, such as the necessity for high concentrations of excess reactants, which may exceed limits, alter mechanisms, or promote unintended side reactions. Additionally, if the excess is insufficient, the assumption of constant concentrations may fail, leading to inaccurate order determinations.

Complex Rate Laws

Reversible Reactions

In reversible reactions, the net rate is determined by the difference between the forward and reverse reaction rates. For a general reversible process, the net rate r_{\text{net}} is expressed as r_{\text{net}} = k_f [\text{reactants}]_f - k_r [\text{products}]_r, where k_f is the forward constant, k_r is the reverse constant, and the subscripts denote the concentrations involved in each direction. This form accounts for the opposing processes that characterize equilibrium systems in . The system approaches when the net rate becomes zero, such that \frac{d[\text{product}]}{dt} = 0. At this point, the forward rate equals the reverse rate, leading to the equilibrium constant K_{\text{eq}} = \frac{k_f}{k_r} = \frac{[\text{products}]_{\text{eq}}}{[\text{reactants}]_{\text{eq}}}. This relationship, derived from the , highlights how kinetic parameters directly relate to . A simple example is the reversible first-order reaction \ce{A ⇌ B}, where the rate equation is -\frac{d[\ce{A}]}{dt} = k_1 [\ce{A}] - k_{-1} [\ce{B}], with k_1 and k_{-1} as the forward and reverse rate constants, respectively. Assuming initial conditions [\ce{A}]_0 = a and [\ce{B}]_0 = 0, and using the relation [\ce{B}] = a - [\ce{A}], the integrated rate law is [\ce{A}] = [\ce{A}]_{\text{eq}} + ([\ce{A}]_0 - [\ce{A}]_{\text{eq}}) e^{-(k_1 + k_{-1}) t}, where [\ce{A}]_{\text{eq}} = \frac{a}{1 + K_{\text{eq}}} and K_{\text{eq}} = \frac{k_1}{k_{-1}}. This exponential form describes the approach to equilibrium, with the effective rate constant k_1 + k_{-1} governing the relaxation time \tau = \frac{1}{k_1 + k_{-1}}. For multi-step reversible reactions involving intermediates, the steady-state approximation is often applied to simplify the rate equations. This assumes that the concentration of each intermediate remains nearly constant, such that its rate of formation equals its rate of consumption: \frac{d[\text{intermediate}]}{dt} \approx 0. For instance, in a mechanism with a fast reversible step followed by a slow irreversible step, solving the steady-state equation for the intermediate yields a net rate law incorporating both forward and reverse contributions from the reversible step. This approach extends the simple reversible form to complex networks while maintaining focus on the balance of opposing rates.

Consecutive Reactions

Consecutive reactions, or sequential reactions, describe a series of irreversible steps in which the product of an initial reaction becomes the reactant for a subsequent one, exemplified by the scheme A → B → C where B is an . These systems are common in multi-step processes, and their are analyzed assuming each elementary step follows rate laws. The equations for the concentrations are: \frac{d[\mathrm{A}]}{dt} = -k_1 [\mathrm{A}] \frac{d[\mathrm{B}]}{dt} = k_1 [\mathrm{A}] - k_2 [\mathrm{B}] \frac{d[\mathrm{C}]}{dt} = k_2 [\mathrm{B}] where k_1 and k_2 are the constants for the respective steps. Integrating these equations with initial conditions [\mathrm{A}]_0 = [\mathrm{A}](0), [\mathrm{B}]_0 = [\mathrm{C}]_0 = 0 yields the time-dependent concentrations: [\mathrm{A}](t) = [\mathrm{A}]_0 e^{-k_1 t} [\mathrm{B}](t) = [\mathrm{A}]_0 \frac{k_1}{k_2 - k_1} \left( e^{-k_1 t} - e^{-k_2 t} \right) These expressions hold for k_1 \neq k_2. When k_1 \approx k_2, the intermediate [B] exhibits a pronounced buildup before decaying, with the maximum concentration occurring at t_\mathrm{max} = \frac{\ln(k_2 / k_1)}{k_2 - k_1}. Such models apply to chain reactions, as seen in polymerization where sequential propagation steps mimic consecutive kinetics. Concentration-time profiles for consecutive reactions graphically illustrate the dynamics: [A] decays exponentially from its initial value, [B] rises to a reflecting the interplay of formation and rates, and [C] sigmoidally approaches [A]_0, highlighting the transient accumulation of the .

Parallel Reactions

In parallel reactions, a single reactant undergoes competing pathways to form different products simultaneously, leading to a total rate of that is the sum of the individual pathway rates. For two competing irreversible s, the total rate is given by r_{\text{total}} = r_1 + r_2, where r_1 and r_2 are the rates of the individual paths. The selectivity toward a particular product, defined as the fraction of reactant converted via that pathway, is S_1 = \frac{r_1}{r_{\text{total}}}. This is fundamental for predicting product distributions in systems where multiple channels are accessible from the same starting material. A common case involves two parallel first-order reactions, such as the decomposition of reactant A into products B and C: A → B with rate constant k_1 and A → C with rate constant k_2. The rate equations are \frac{d[A]}{dt} = -(k_1 + k_2)[A], \frac{d[B]}{dt} = k_1 [A], and \frac{d[C]}{dt} = k_2 [A], where the total rate constant is k_{\text{total}} = k_1 + k_2. Integrating these yields the concentration of B as [B] = [A]_0 \frac{k_1}{k_{\text{total}}} (1 - e^{-k_{\text{total}} t}), with a similar expression for C by substituting k_2. The selectivity S_B = \frac{k_1}{k_{\text{total}}} remains constant over time, independent of concentration, allowing straightforward prediction of product yields. When parallel pathways have different orders, such as one (r_1 = k_1 [A]) and one second-order (r_2 = k_2 [A]^2), the total rate becomes r_{\text{total}} = k_1 [A] + k_2 [A]^2, and selectivity S_1 = \frac{k_1 [A]}{k_1 [A] + k_2 [A]^2} = \frac{k_1}{k_1 + k_2 [A]} varies with the concentration of A. At high concentrations, the second-order path dominates, favoring the corresponding product, while at low concentrations, the path prevails. This concentration dependence complicates , often requiring numerical methods, but highlights how reaction conditions can tune branching ratios. In , parallel reactions are critical for controlling product distribution, as competing pathways can lead to desired versus undesired products; kinetic optimization, such as adjusting concentrations or catalysts, enhances selectivity toward target molecules, as exemplified in processes like selective oxidations or cross-coupling reactions.

Advanced Applications

Stoichiometric Networks

Stoichiometric networks describe complex systems of interconnected chemical reactions where multiple species participate in a web of transformations, often involving cycles or branches that exceed simple linear or pathways. The evolution of species concentrations in such networks is compactly represented using a formulation. Let \mathbf{c} denote the of species concentrations, \mathbf{N} the stoichiometric whose entries N_{ij} are the stoichiometric coefficients (positive for products, negative for reactants) for species i in reaction j, and \mathbf{v} the of reaction rates. The are then given by \frac{d\mathbf{c}}{dt} = \mathbf{N} \mathbf{v}, where each component of \mathbf{v} follows a rate law specific to its reaction, such as power-law or mass-action kinetics. This formulation facilitates analysis of mass balance and network structure, enabling the identification of conserved quantities through the kernel of \mathbf{N}^T and the computation of fluxes under steady-state conditions. For networks with short-lived intermediates, the steady-state approximation simplifies the model by assuming that the concentrations of these species remain nearly constant over the timescale of interest, i.e., \frac{d[\inter]}{dt} = 0 for each intermediate. This leads to a reduced algebraic system: the rows of \mathbf{N} corresponding to intermediates are set to zero, yielding \mathbf{N}_{\inter} \mathbf{v} = 0, which can be solved for the rates \mathbf{v} in terms of stable species concentrations. The approximation is valid when intermediates react much faster than stable species, minimizing their transient buildup and allowing focus on overall network fluxes. Validity conditions include separation of timescales, with error bounds derived from singular perturbation theory. A prominent application arises in catalytic cycles, such as , where the steady-state approximation yields the Michaelis-Menten rate law. Consider an E binding S to form complex ES, which then produces product P and regenerates E: the rate of product formation is v = \frac{V_{\max} [S]}{K_m + [S]}, with V_{\max} = k_2 [E]_0 the maximum rate and K_m = \frac{k_{-1} + k_2}{k_1} the Michaelis constant. This derives from setting d[ES]/dt = 0, balancing formation and decay of ES, and captures saturation behavior in biological networks. For intricate networks defying analytical solutions, numerical methods simulate the full ODE system \frac{d\mathbf{c}}{dt} = \mathbf{N} \mathbf{v}(\mathbf{c}) to predict time courses and steady states. Deterministic approaches employ integrators like explicit Runge-Kutta schemes for non-stiff systems or implicit methods (e.g., ) for stiff dominated by disparate timescales, ensuring accurate resolution of network behavior without algebraic reduction. These simulations reveal emergent properties, such as oscillations or , in large-scale models.

Unimolecular Reaction Dynamics

Unimolecular reactions involve the decomposition or isomerization of a single reactant molecule, often appearing first-order in kinetics but requiring collisional activation in gas-phase environments. The foundational Lindemann mechanism, proposed in 1922, describes this process through a two-step sequence: activation of the reactant A by collision to form an energized intermediate A*, followed by its decay to products. The mechanism is A + A ⇌ A* + A with forward and reverse rate constants k_1 and k_{-1}, and A* → products with rate constant k_2. Applying the steady-state approximation to [A*] yields the rate law: rate = \frac{k_1 k_2 [A]^2}{k_{-1} [A] + k_2}. In the high-pressure limit, where collisional deactivation dominates (k_{-1} [A] \gg k_2), the rate simplifies to a first-order form: rate ≈ k_\infty [A], with k_\infty = \frac{k_1 k_2}{k_{-1}}, reflecting rapid equilibration of A and A* before reaction. Conversely, in the low-pressure regime (k_2 \gg k_{-1} [A]), the rate becomes second-order: rate ≈ k_1 [A]^2, as every activation leads to reaction without deactivation. The fall-off regime, bridging these limits, is approximated as rate = \frac{k_\infty [A]}{1 + \frac{k_\infty}{k_\text{coll} [M]}}, where [M] is the concentration of bath gas molecules facilitating collisions, and k_\text{coll} is the collision rate constant; this captures pressure-dependent behavior observed in experiments. The Rice-Ramsperger-Kassel-Marcus (RRKM) theory extends the Lindemann framework by incorporating to compute microcanonical rate constants k(E) for energized molecules with excess E above the reaction threshold E_0. In RRKM, k(E) = \frac{N^\ddagger (E - E_0)}{h \rho(E)}, where N^\ddagger (E - E_0) is the number of vibrational states accessible at the up to E - E_0, \rho(E) is the of the reactant at E, and h is Planck's constant; this emphasizes that the depends critically on the intramolecular distribution and partitioning between reactant and . Developed in 1952, RRKM provides a quantum mechanical basis for predicting unimolecular rates across energy landscapes. Applications of these rate equations are prominent in gas-phase isomerizations, such as the thermal conversion of to propene, where the reaction follows unimolecular above ~0.01 atm at 490°C, exhibiting fall-off behavior consistent with the . Modern computational extensions, including simulations, solve time-dependent population balance equations for energy-grained species to model pressure, temperature, and bath gas effects on rate constants, enabling accurate predictions for complex systems like reactions. These simulations integrate RRKM rates with collisional energy transfer models, as reviewed in theoretical methodologies.

References

  1. [1]
    Chemical Kinetics - Purdue University
    the rate at which the reactants are transformed into the products of the reaction. The term rate is often used to describe the change in a quantity that occurs ...Chemical Kinetics · Instantaneous Rates of... · Different Ways of Expressing...
  2. [2]
    Introduction to Rate Laws
    A rate law is a means by which we can relate the rate of a chemical reaction to concentrations of the reactants. The rate law for a reaction is dependent on the ...
  3. [3]
    Chemical Kinetics: Reaction Rates
    The term reaction rate refers to how fast the chemical reaction occurs. The rate of reaction, r, describes how fast the concentrations of reactants and products ...
  4. [4]
    [PDF] Determining Rate Laws for Chemical Reactions
    Modern chemical kinetics was solidified in 1864 when two Norwegian chemists, Peter Waage and Cato Guldberg, formulated the law of mass action, which states ...Missing: history | Show results with:history<|separator|>
  5. [5]
    Arrhenius Theory and Reaction Coordinates
    The Arrhenius theory and reaction coordinates relate to a particular perspective on chemical kinetics that imagines the path taken by the reactants as they are ...
  6. [6]
    12.1 Chemical Reaction Rates - Chemistry 2e | OpenStax
    Feb 14, 2019 · The rate of a reaction may be expressed as the change in concentration of any reactant or product. For any given reaction, these rate ...<|control11|><|separator|>
  7. [7]
    Arrhenius Equation - an overview | ScienceDirect Topics
    The Arrhenius equation is a simple formula originally given for the temperature dependence of the rate constant, and therefore the rate of a chemical reaction.
  8. [8]
    Kinetics Review | Chem Lab
    Jan 17, 2012 · There is only one exception to the general rule that one cannot write a rate law directly from a balanced chemical equation and this is for an ...
  9. [9]
    The Relationship Between Elementary Steps and Rate Law - kinetics
    If all we know is the overall equation for the reaction (these reactants are converted into those products), it is difficult to tell what the rate law will be.
  10. [10]
    [PDF] Chapter 19: Chemical Kinetics
    ➢ rate law defined in terms of reactants only. ➢ exponents are usually ... Quantified by the Arrhenius Equation for the rate constant k = Ae. Ea/RT. A ...
  11. [11]
    Cato Guldberg, Peter Waage: Law of Mass Action History
    Just over 150 years ago, on 15 March 1864, Peter Waage and Cato Guldberg (Figure 1) published a paper in which they propounded what has come to be known as the ...
  12. [12]
    [PDF] Reaction Kinetics - Claire Vallance
    Chemical reaction kinetics deals with the rates of chemical processes, which are broken down into elementary steps. The rate is the change of concentration ...<|control11|><|separator|>
  13. [13]
    [PDF] Chemical Reaction Engineering
    design equations, rate constants and concentration profiles. In this section, we introduce some structure (procedures) for the design of reactors. 1. Recall ...
  14. [14]
    Activation Energy - Chemical Kinetics
    The Arrhenius equation can also be used to calculate what happens to the rate of a reaction when a catalyst lowers the activation energy.
  15. [15]
    [PDF] Understanding Catalysis - RSC Publishing
    Jun 20, 2014 · While it is common knowledge that catalysts enhance reaction rates by lowering the activation energy it is often obscure how catalysts achieve ...
  16. [16]
    The Mechanisms of Chemical Reactions
    The rate law for a reaction is a useful way of probing the mechanism of a chemical reaction but it isn't very useful for predicting how much reactant remains ...
  17. [17]
    chapter iv chemical kinetics
    Thus, for all elementary reactions, the overall order equals the molecularity. Non-elementary reactions involve a series of two or more elementary reactions.
  18. [18]
    Experimental Rate Laws - EdTech Books
    Dimensional analysis requires the rate constant unit for a reaction whose overall order is x to be L x − 1 mol 1 − x s −1 . L ...
  19. [19]
    18.3 Rate Laws – Chemistry Fundamentals - UCF Pressbooks
    The units for a rate constant will vary as appropriate to accommodate the overall order of the reaction. The unit of the rate constant for the second-order ...
  20. [20]
    Chemical Kinetics: Integrated Rate Laws
    To illustrate this point, consider the reaction. A → B. The rate of reaction, r, is given by. r = –. d [A] d t. Suppose this reaction obeys a first-order rate ...
  21. [21]
    [PDF] Kinetics: The Differential and Integrated Rate Laws in Chemistry ...
    For a reaction aA → bB + cC, the rate is given by: Rate = − 1/aa d[AA]/dd = 1/bb d[BB]/dd = 1/cc d[CC]/dd. Reactants decrease, products increase.
  22. [22]
    Integrated Rate laws
    In order to determine the rate law for a reaction from a set of data consisting of concentration (or the values of some function of concentration) versus time, ...
  23. [23]
    Thermodynamics and Kinetics
    Common examples of first order reactions are radioactive decay processes. As must be the case for first order reactions, the rate of decay depends on the amount ...
  24. [24]
    and bimolecular steps are proposed in reaction mechanisms.
    The molecularity is a way of stating exactly how many molecules are involved in the elementary step. One molecule reacting by itself is a unimolecular reaction, ...
  25. [25]
    [PDF] Reaction Kinetics - Claire Vallance
    ... rate of the reaction and/or its individual elementary ... As well as having rate laws for overall reactions, we can of course also write down individual rate.
  26. [26]
    [PDF] Chemical Kinetics Integrated Rate Laws
    Example. Page 4. Second Order, one reactant. ○ If second order kinetics apply, a plot of. 1/[A] vs. t will be a line with slope k . rate = − d[A] dt. = k [A]2.
  27. [27]
    Chemical Kinetics: Differential Rate Laws
    A rate law is a mathematical equation that describes the progress of the reaction. In general, rate laws must be determined experimentally. Unless a reaction is ...
  28. [28]
    [PDF] The rates of chemical reactions - FIUnix Faculty Sites
    Chemical kinetics deals with how rapidly reactants are consumed and products are formed, how reaction rates respond to changes in the conditions or presence of ...
  29. [29]
    [PDF] KINETICS AND SOLVENT EFFECTS IN THE SYNTHESIS OF IONIC ...
    It is widely known that the leaving group abilities highly affects rates of reaction when operating under a second-order nucleophilic substitution (SN2).
  30. [30]
    [PDF] How to Calculate and Determine Chemical Kinetics - Jurnal UPI
    Chemical kinetics studies reaction rates. This article provides a step-by-step guide to determine reaction rate and order using experimental data and methods ...
  31. [31]
    [PDF] 1 Fundamentals of Chemical Kinetics
    Notice that the rate law is a differential equation: it gives the derivative (with respect to time) of one of the concentrations in terms of all the ...
  32. [32]
    Integrated Rate Laws
    The goal of an integrated rate law to have an expression for the concentration of the reactants (or products) as a function of time.
  33. [33]
    2.8.1: Pseudo-1st-order Reactions
    ### Summary of Pseudo-1st-Order Reactions
  34. [34]
    12.4 Integrated Rate Laws - Chemistry 2e | OpenStax
    Feb 14, 2019 · According to the linear format of the first-order integrated rate law, the rate constant is given by the negative of this plot's slope.
  35. [35]
    Mechanistic Implications of Pseudo Zero Order Kinetics in Kinetic ...
    The observation of pseudo zero order substrate dependency in catalytic kinetic resolution reactions typically arises due to one of the two generic mechanisms ...
  36. [36]
    Chemical Kinetics: Method of Initial Rates
    In the Method of Initial Rates, the differential rate law is determined by varying the initial concentrations of the various reactants and observing the effect ...
  37. [37]
    Determining Reaction Rates
    The initial rate of a reaction is the instantaneous rate at the start of the reaction (i.e., when t = 0). The initial rate is equal to the negative of the slope ...
  38. [38]
    Kinetics - FSU Chemistry & Biochemistry
    Kinetics is the study of the rates of chemical processes. The rate of a reaction is defined at the change in concentration over time: Rate Equation. Rate ...
  39. [39]
    [PDF] Chemical Kinetics
    ○ Reaction order can only be found by experiments. Examples. 2 N2O5 → 4NO2 + O2 rate = k [N2O5]. ○ BUT. 2 NO2 → 2NO + O2 rate = k [NO2]2. ○ CAN'T predict ...
  40. [40]
    [PDF] Chapter 14 Chemical Kinetics - MSU chemistry
    Reaction rates are always positive quantities, no matter if a reactant consumed or a product is formed. ▫ We divide each rate by the stoichiometric factor of ...
  41. [41]
    [PDF] How to Calculate and Determine Chemical Kinetics - Jurnal UPI
    The initial rate method is an approach that is often used to determine the reaction order by measuring the reaction rate at different initial concentrations.
  42. [42]
    [PDF] Chemical Kinetics: The Method of Initial Rates
    The rate law of a chemical reaction is a mathematical equation that describes how the reaction rate depends upon the concentration of each reactant.
  43. [43]
    Chapter 13 - Kinetics: The Study of Rates of Reaction
    Apply the method of initial rates to determine the rate law for a reaction. Use the rate law and rate data to calculate a rate constant, k, or use the rate law ...<|control11|><|separator|>
  44. [44]
    ChE 344 Winter 2016 Chemical Reaction Engineering Final Exam ...
    • Integral method: why and how is it used? • What is the differential method for constant volume systems? • Nonlinear regression → thanks Excel. • How would ...
  45. [45]
    A general integral method for the determination of empirical reaction ...
    A general procedure for the determination of empirical rates and kinetic parameters of irreversible, constant volume reactions is presented which is clearly ...
  46. [46]
  47. [47]
    Chemical Reactions and Kinetics
    To illustrate the power of the integrated form of the rate law for a reaction, let's use this equation to calculate how long it would take for the 14C in a ...
  48. [48]
    [PDF] Introduction to Kinetics and Equilibrium - UCI Department of Chemistry
    >> rate reverse. As products form, the rate of the reverse reaction increases: CO. (g). + Cl. 2(g). ← COCl. 2(g). 31 rate reverse. = k r x [COCl. 2. ] Page 20 ...
  49. [49]
    [PDF] 75 Chapter 3: Kinetics The hydrolysis of ethylacetate by sodium ...
    Apr 3, 2021 · We first consider the general case and then consider the integrated rate law for a reversible first-order reaction. The concept of chemical ...
  50. [50]
    [PDF] Lecture 32: Kinetics: Reaction Mechanisms - MIT OpenCourseWare
    Dec 1, 2014 · Steady-state approximation: the rate of formation of intermediates of decay of intermediates. The slowest elementary step in a sequence of ...
  51. [51]
    [PDF] Rate Law for Consecutive & Parallel Reactions of Ist Order Reactions
    Consecutive reactions are multiple steps. The rate of disappearance of reactant A is given by -d[CA]/dt = k1CA. The rate of formation of intermediate B is ...
  52. [52]
    [PDF] Chapter 20: Rates of Chemical Reactions - Chemistry
    Oct 16, 2010 · Rate Law: mathematical statement of how the reaction rate depends on concentration. Important because it is a guide to the reaction ...
  53. [53]
  54. [54]
    [PDF] AN INTRODUCTION TO CHEMICAL ENGINEERING KINETICS ...
    Feb 6, 2025 · ... PARALLEL REACTIONS. The term parallel reactions describes situations in which reactants can undergo two or more reactions independently and ...
  55. [55]
    A Brief Introduction to Chemical Reaction Optimization
    Feb 23, 2023 · This review highlights the basics, and the cutting-edge, of modern chemical reaction optimization as well as its relation to process scale-up.
  56. [56]
    Basic concepts and principles of stoichiometric modeling of ... - NIH
    Jul 29, 2013 · The r × 1 rate vector, v, contains the rate equations of the r reactions in the network, which are typically expressed in terms of enzyme ...
  57. [57]
    Steady state and equilibrium approximations in reaction kinetics
    Some considerations on the fundamentals of chemical kinetics: Steady state, quasi-equilibrium, and transition state theory.
  58. [58]
    [PDF] The QSSA in Chemical Kinetics: As Taught and as Practiced
    The evolution of species concentrations is usually described by sets of nonlinear ordinary differential equations (ODEs) that must be solved numerically.
  59. [59]
    A guide to the Michaelis–Menten equation: steady state and beyond
    Jul 16, 2021 · The article talks about the primacy of the Michaelis–Menten equation as a framework for the quantitative understanding of enzyme kinetics.Assumptions and... · Kinetic isotope effects and... · Looking ahead and moving...
  60. [60]
    Modeling and Simulating Chemical Reactions | SIAM Review
    This article introduces some of the basic concepts in an accessible manner and points to some challenges that currently occupy researchers in this area.
  61. [61]