Fact-checked by Grok 2 weeks ago

Collision theory

Collision theory is a model in that describes the rates of bimolecular chemical reactions as arising from collisions between reactant molecules. Successful reactions require collisions with sufficient to overcome the barrier E_a and proper molecular orientations. The theory treats molecules as that interact only upon contact, with the proportional to the frequency of effective collisions. Independently proposed by Max Trautz in 1916 and William Lewis in 1918, the theory applies concepts from the to reaction dynamics, linking temperature effects to the Maxwell-Boltzmann distribution of molecular speeds. The rate constant for a second-order reaction is given by k = p Z e^{-E_a / [RT](/page/RT)}, where Z is the collision frequency factor (derived from molecular radii and relative velocities), p is the accounting for orientation (often p < 1), and the exponential term represents the fraction of collisions with energy exceeding E_a. This explains increases in reaction rates with concentration and temperature, primarily for gas-phase reactions and simple systems. Though limited by neglecting quantum effects, solvent influences, and complex mechanisms, collision theory provides a foundational understanding of chemical reactivity and influenced later theories like .

Fundamentals

Historical Development

Collision theory in originated in the early as chemists sought a molecular-level explanation for reaction rates, building on the established . Max Trautz, a , first proposed the theory in 1916, suggesting that reaction rates are determined by the frequency of collisions between reactant molecules, with only a fraction leading to products based on their and . His seminal paper, titled "Das Gesetz der Reaktionsgeschwindigkeit und der Gleichgewichte in Gasen. Bestätigung der Additivität von C_v-3/2R. Neue Bestimmung der Integrationskonstanten und der Moleküldurchmesser," appeared in the Zeitschrift für anorganische und allgemeine Chemie, where he derived rate expressions linking collision numbers to experimental observations. Independently, British chemist William Lewis developed a similar framework in 1918, emphasizing the role of molecular collisions in gas-phase reactions and integrating concepts from the kinetic theory. Lewis's work, published in the Journal of the Chemical Society under the title "Studies in catalysis. Part IX. The calculation in absolute measure of velocity constants and equilibrium constants in gaseous systems," provided quantitative predictions for bimolecular rate constants by considering average molecular speeds and cross-sections. This theory was firmly rooted in the Maxwell-Boltzmann distribution, which describes the statistical distribution of molecular speeds and energies in gases, allowing Trautz and Lewis to explain temperature dependence through the proportion of high-energy collisions. The original formulations by Trautz and already incorporated energy thresholds, recognizing that only collisions exceeding a minimum could result in reaction. By the 1920s, collision theory was refined by Hinshelwood and others, extending the framework to account for energy distribution among molecular and applying it to more complex systems, marking key milestones in the theory's maturation.

Basic Principles

Collision theory provides a foundational framework for understanding rates by emphasizing the molecular-level interactions required for reactions to occur. It posits that for a to take place, reactant molecules must collide with one another, and the rate of the reaction is proportional to the of these collisions that are effective in producing products. A key element of collision theory is the concept of , also known as (E_a), which represents the minimum that colliding molecules must possess to overcome the energy barrier and form the leading to products. Collisions with below E_a are insufficient to break or form bonds, resulting in molecules simply bouncing off each other without reaction. For instance, the of requires an of approximately 180 kJ/mol, illustrating how this energy threshold determines reaction feasibility. Temperature plays a crucial role in collision theory, as described by the kinetic molecular theory of gases. Higher temperatures increase the average of molecules, thereby raising both the frequency of collisions and the fraction of molecules that have energies exceeding E_a. This dual effect explains the observed exponential increase in reaction rates with temperature; for many reactions, a 10°C rise approximately doubles the rate due to a greater proportion of successful collisions. The theory distinguishes between total collisions, which occur whenever molecules come into contact, and effective (or reactive) collisions, which are a small subset that satisfy the conditions of sufficient and proper for rearrangement. While total collisions can be frequent, only effective ones contribute to the reaction progress, accounting for why reaction rates are often much slower than expected from collision frequencies alone.

Mathematical Formulation

Collision Frequency

In collision theory, the collision frequency Z quantifies the rate at which reactant molecules collide in an , serving as a foundational element for predicting bimolecular reaction rates based on kinetic theory. This frequency represents the total number of collisions per unit volume per unit time between molecules of types A and B, assuming random motion and encounters without from other factors. The concept originates from the application of gas kinetic theory to , as independently proposed by Max Trautz and William Lewis in the early . The derivation of Z relies on the hard-sphere model, where molecules are idealized as rigid spheres of effective diameter d (typically the average of the molecular diameters of A and B), colliding when their centers approach within this distance. To compute Z, the relative motion of the molecules is considered: a single molecule of A sweeps out a cylindrical volume per unit time with cross-sectional area \pi d^2 and length equal to the average relative speed \langle v \rangle. The number density of B molecules is n_B = N_A [B], where N_A is Avogadro's number and [B] is the (in mol m^{-3}). Thus, the collision rate for one A molecule with all B molecules is \pi d^2 \langle v \rangle n_B, and multiplying by the number density of A molecules n_A = N_A [A] yields the total collision frequency per unit volume: Z = N_A [A] \cdot \pi d^2 \langle v \rangle \cdot N_A [B] = N_A^2 \pi d^2 \langle v \rangle [A][B]. In standard notation with concentrations in ^{-3} (M), the number densities are n_A = 10^3 N_A [A] and n_B = 10^3 N_A [B] (molecules m^{-3}), so the collision frequency per m^3 is Z = 10^6 N_A^2 \pi d^2 \langle [v](/page/V.) \rangle [A][B] (collisions m^{-3} s^{-1}). The corresponding bimolecular collision rate constant (in ^3 ^{-1} s^{-1}) is z = 10^3 N_A \pi d^2 \langle [v](/page/V.) \rangle, accounting for conversion from m^3 to ^3 and molecules to moles. This expression highlights how Z scales with the product of concentrations, reflecting the probabilistic nature of encounters in dilute gases. The average relative speed \langle v \rangle is derived from the Maxwell-Boltzmann distribution of velocities, representing the mean magnitude of the velocity difference between A and B molecules. For unlike molecules, it is given by \langle v \rangle = \sqrt{\frac{8 k_B T}{\pi \mu}}, where k_B is the Boltzmann constant, T is the absolute temperature, and \mu = \frac{m_A m_B}{m_A + m_B} is the reduced mass of the two species. This formula emerges from integrating the relative velocity distribution, assuming isotropic and uncorrelated motions. The hard-sphere assumption simplifies interactions to geometric collisions, neglecting potential energy effects or molecular flexibility, which holds well for low-density gases at moderate temperatures. This collision frequency framework extends naturally to the total number of collisions in a given volume and time by multiplying Z by the volume and duration, incorporating N_A to bridge microscopic number densities with macroscopic concentrations. For identical molecules (A = B), the formula adjusts by a factor of $1/2 to avoid double-counting, but the bimolecular case for distinct uses the form above. These derivations underpin the quantitative predictions of collision theory while relying on the ideal gas approximations of point-like particles except for collision geometry.

Rate Equations and Derivation

In collision theory, the rate of a bimolecular gas-phase reaction between species A and is expressed as the product of the Z_{AB} and the fraction of those collisions possessing sufficient energy to surmount the E_a. The Z_{AB}, detailed in the preceding section, represents the total number of collisions between A and molecules per unit volume per unit time and is proportional to the concentrations [A] and [B], as well as the average relative speed and collision cross-section of the molecules. The key step in the derivation involves determining the fraction of collisions with relative translational energy exceeding E_a. According to the Maxwell-Boltzmann distribution of molecular speeds, the probability that the relative kinetic energy along the line of centers is greater than E_a is given by the integral over the high-energy tail of the distribution, which approximates to e^{-E_a / RT} for E_a \gg RT, where R is the gas constant and T is the absolute temperature. Thus, the reaction rate is \text{rate} = Z_{AB} \cdot e^{-E_a / RT}. Since Z_{AB} is linearly dependent on [A] and [B], the rate can be rewritten in the standard second-order form \text{rate} = k [A][B], where the rate constant k takes the Arrhenius form k = A e^{-E_a / RT}. Here, the A is directly related to the parameters, specifically A = N_A \sigma_{AB} \sqrt{\frac{8 k_B T}{\pi \mu_{AB}}} (in m^3 molecule^{-1} s^{-1} times N_A \times 10^3 for dm^3 mol^{-1} s^{-1}, assuming unit steric efficiency), with N_A Avogadro's number, \sigma_{AB} the collision cross-section, k_B Boltzmann's constant, and \mu_{AB} the . This derivation bridges the microscopic to the empirical , explaining the exponential temperature dependence of reaction rates. A classic application is the thermal decomposition of hydrogen iodide, $2\text{HI}(g) \to \text{H}_2(g) + \text{I}_2(g), a second-order reaction with rate = k [\text{HI}]^2. Collision theory predicts the observed Arrhenius behavior, with experimental values of E_a \approx 184 \, \text{kJ/mol} and A consistent with estimated collision parameters for HI molecules when including a steric factor.

Validity and Limitations

Assumptions and Steric Factor

Collision theory is predicated on several fundamental assumptions rooted in the kinetic theory of gases. It models reactant molecules as rigid, hard spheres that interact solely through elastic collisions, neglecting any attractive or repulsive forces beyond the point of contact. Collisions are assumed to occur randomly with respect to molecular orientations, and the energy necessary to surmount the activation energy barrier E_a is provided exclusively by the translational kinetic energy of the colliding molecules, following the Maxwell-Boltzmann distribution. The theory further presumes ideal gas behavior, where molecules move independently without significant intermolecular interactions, and reaction rates are determined primarily by the frequency of effective collisions in the gas phase. A key limitation of these assumptions arises from the random postulate, which implies that all collisions exceeding E_a should lead to reaction, often overestimating actual for reactions requiring specific geometries. To rectify this, the P (where $0 < P \leq 1) is incorporated as an empirical correction for the fraction of collisions with favorable orientations. The adjusted rate expression then becomes \text{[rate](/page/Rate)} = [P](/page/P′′) \cdot [Z](/page/Z) \cdot e^{-E_a / RT}, where Z represents the and the term accounts for the energy requirement (as derived in the basic formulation without P). This modification acknowledges that only a subset of energetic collisions align properly for bond breaking and formation. The value of P is influenced by molecular geometry, size, and the underlying reaction mechanism, typically approaching 1 for simple systems like atomic recombination reactions (e.g., \text{Na} + \text{Cl} \rightarrow \text{NaCl}), where minimal orientational specificity is needed. In contrast, for complex polyatomic molecules involving intricate transition states, such as the addition of HCl to ethylene requiring perpendicular approach of the hydrogen end to the double bond, P can be significantly smaller, often on the order of $10^{-2} to $10^{-6}, reflecting stringent steric demands. Despite these refinements, collision theory's assumptions lead to validity constraints. It tends to overestimate rates for reactions with high energies or pronounced steric hindrance, as the hard-sphere model underestimates orientational barriers and ignores quantum effects or contributions. The theory is inapplicable to unimolecular reactions, which do not require collisions, and to reactions, where steps deviate from simple collision-dominated .

Extensions for Solutions

In solutions, particularly dilute ones, collision theory must account for solvent-mediated effects that alter molecular encounters compared to gas-phase reactions. The presence of solvent molecules reduces effective collision frequencies through phenomena such as caging, where reactant pairs are temporarily trapped within a solvent shell, and increased , which hinders diffusive motion. These factors shift the toward diffusion-controlled rates, where the rate-limiting step is the transport of reactants to form an encounter complex rather than the collision itself. A foundational extension is the Smoluchowski equation, which describes the encounter rate constant k_{\text{diff}} for bimolecular reactions in dilute solutions as k_{\text{diff}} = 4 \pi (D_A + D_B) (r_A + r_B), where D_A and D_B are the diffusion coefficients of reactants A and B, and r_A and r_B are their effective radii. This expression, derived from solving the diffusion equation with absorbing boundary conditions at the reaction distance, replaces the gas-phase collision frequency with a diffusion flux term, emphasizing that encounters occur when molecules diffuse within a critical separation. The diffusion coefficients themselves depend on solvent viscosity \eta via the Stokes-Einstein relation D = kT / (6 \pi \eta r), highlighting how solvent properties directly influence reactivity. Reactions in solutions can transition from collision-limited (where every encounter leads to reaction) to diffusion-limited regimes as dominate, particularly for highly exothermic or barrierless processes. In the diffusion-limited case, the observed rate constant approaches k_{\text{diff}}, while in the collision-limited regime, it is lower due to incomplete reactivity upon encounter. A classic example is fast ion recombination, such as the neutralization of oppositely charged ions in , where governs the rate and yields are often near the Smoluchowski prediction of approximately $10^{10} M^{-1} s^{-1} for typical small ions. This transition is probed experimentally by varying or to modulate rates. To address incomplete reactivity, collision theory in solutions incorporates steric-like factors adapted for solvated species, accounting for orientation requirements and solvation layers. For solvated molecules, an effective "steric factor" p modifies the rate as k = p k_{\text{diff}}, where p < 1 reflects the probability of upon , influenced by shells that impose additional barriers to proper . These factors are modeled using partial absorption boundary conditions, such as the radiation boundary condition D \frac{\partial c}{\partial r} |_{r=R} = k_f c(R), where k_f is the intrinsic at , effectively capturing -induced steric hindrance. For ions or polar molecules, shells can reduce p by factors of 0.1 to 0.01 compared to gas-phase estimates, underscoring the role of solvent structure in selectivity.

Applications and Comparisons

Relation to Other Theories

Collision theory, as a foundational model in , posits that chemical reactions occur through direct collisions between reactant molecules possessing sufficient energy and proper orientation. In contrast, (), developed later by Eyring and others, describes reactions as proceeding via the formation of a high-energy or in quasi-equilibrium with the reactants, where the rate is determined by the barrier rather than solely collision dynamics. This shift from a mechanical collision perspective to a statistical thermodynamic framework allows to account for the more comprehensively, particularly for reactions involving complex intermediates. The A in the , which collision theory interprets as related to modulated by a , finds a deeper theoretical basis in theory (a variant of ). Here, A emerges from the partition functions of the reactants and the , providing a molecular-level justification for the frequency factor that collision theory treats empirically. This connection highlights how collision theory's classical approach serves as an approximation to the more rigorous underlying TST, especially for gas-phase reactions where molecular velocities align with Maxwell-Boltzmann distributions. Subsequent advancements in kinetic modeling have evolved beyond classical collision theory by incorporating quantum mechanical effects, such as tunneling through energy barriers, which the model inherently overlooks due to its reliance on classical trajectories. For instance, variational and methods extend TST to capture these non-classical phenomena, offering improved accuracy for reactions at low temperatures or with light atoms. Collision theory remains sufficient, however, for simple bimolecular gas-phase reactions where steric factors adequately adjust for orientation, whereas complex mechanisms involving multi-step pathways or solution-phase dynamics are better addressed by TST and its derivatives.

Experimental Validation

Classic experiments have provided empirical support for collision theory by demonstrating the temperature dependence of reaction rates consistent with the predicted Arrhenius form, which arises from the fraction of collisions possessing sufficient energy. A notable example is the gas-phase reaction of hydrogen and iodine, H₂ + I₂ → 2HI, where measurements over temperatures from 500 K to 700 K show a rate constant following k ≈ 10^{10} exp(-E_a / RT) L mol⁻¹ s⁻¹ with E_a ≈ 150 kJ/mol, confirming the exponential increase with temperature derived from collision energetics. Measurements of steric factors have further validated the role of molecular orientation in effective collisions, particularly through experiments on reactions like K + CH₃I → KI + CH₃. Using oriented molecular beams, where CH₃I molecules are aligned and potassium atoms directed at specific ends, researchers observed a strong dependence of reactivity on approach angle, with the iodine end facing the atom yielding higher reaction probabilities; this directly quantifies the as the ratio of reactive to total collision cross-sections, often approaching unity for favorable orientations. Complementary techniques, such as ultraviolet , have been employed to probe product formation and orientation effects in these systems by monitoring absorption spectra of reaction intermediates or products. Deviations from simple collision predictions highlight limitations, particularly in complex environments where observed rates are lower than expected from collision frequencies alone, necessitating the introduction of steric factors P ranging from 10⁻⁶ for intricate gas-phase reactions to near 1 for simple atom-diatom encounters. In , collision theory overpredicts rates for substrate-enzyme encounters due to stringent requirements and barriers beyond simple impacts, resulting in effective P values much less than 1 and rates governed more by limits than pure collisions. Similarly, in heterogeneous surface reactions, such as catalytic processes on metal surfaces, overprediction occurs because of site-specific adsorption and reduced mobility, with P quantified as low as 10⁻⁶ to account for the discrepancy between bulk collision estimates and measured turnover frequencies. Modern techniques, including crossed molecular beam experiments, have precisely validated collision frequencies in the gas phase by measuring differential cross-sections for reactions like atom-halide alkyl systems. These setups allow isolation of single collisions under controlled velocities, confirming that observed rates align with theoretical collision rates Z when adjusted for and thresholds, providing direct empirical tests of the theory's core assumptions without interference from multiple collisions.

References

  1. [1]
    [PDF] Basic Chemical Kinetic Principles I
    Simple collision theory: A + B → products. Assumptions? – molecules hard-spheres. – every collision is reactive. Calculate rate of collision ⇒ rate of reaction.
  2. [2]
    Collision Theory – Chemistry - JMU Libraries Pressbooks
    Chemical reactions typically require collisions between reactant species. These reactant collisions must be of proper orientation and sufficient energy in order ...Missing: principles | Show results with:principles
  3. [3]
    A BRIEF HISTORY OF CHEMICAL KINETICS
    Jan 31, 2001 · 1917: Trautz (Germany) and Lewis (UK) independently proposed that the rate of reaction is determined by the frequency of molecular collisions.
  4. [4]
    A Brief Introduction to the History of Chemical Kinetics - IntechOpen
    The following text provides brief historical background to chemical kinetics, lays the foundation of transition state theory (TST), and reaction thermodynamics.<|control11|><|separator|>
  5. [5]
    [PDF] Cyril N. Hinshelwood - Nobel Lecture
    First, there was the unmistakable evidence that molecular collisions play the all- important role in communicating the activation energy to the molecules which ...
  6. [6]
    The Collision Model of Chemical Kinetics
    According to the collision model, a chemical reaction can occur only when the reactant molecules, atoms, or ions collide with more than a certain amount of ...
  7. [7]
    [PDF] Kinetic Theory of Gases
    Z = collision frequency = σ<v>N. But we should take into account that other molecules are moving. When this is considered <v> should be replaced by . Also we ...
  8. [8]
    [PDF] Ch. 27: Kinetic Theory of Gases
    The collision frequency is thur z = op <Vel>. 1. Eq. = √26p <v> for like partides G 27.50. The mean free path is the distance a molecule travels between ...
  9. [9]
    Das Gesetz der Reaktionsgeschwindigkeit und der Gleichgewichte ...
    References · p1_1 M. Trautz, Die Theorie der chemischen Reaktionssgeschwindigkeit und ein neues Grenzgesetz für ideale Gase: Die Additivität der inneren ...
  10. [10]
    8.7: Theories of Reaction Rates - Chemistry LibreTexts
    Mar 26, 2025 · The steric factor, ρ is then introduced to represent is the probability of the reactant molecules colliding with the right orientation and ...Missing: recombination | Show results with:recombination
  11. [11]
    [PDF] Collision Theory of Reaction Rates and Its Limitations | Dalal Institute
    In 1916, a German chemist Max Trautz proposed a theory based on the collisions of reacting molecules to explain reaction kinetics. Two years later ...
  12. [12]
    [PDF] CHEMICAL ENGINEERING KINETICS | Rosen Review
    Collision theory can be used to estimate kinetic parameters from first-principles. ... Here, p is a steric factor which is effectively a correction factor ...
  13. [13]
    [PDF] 16.5 Theories of Chemical Kinetics The Effect of Temperature
    • Collision theory – molecules must collide in order to react ... ➢Steric factor (p) – the colliding molecules must have proper orientation ...
  14. [14]
    [PDF] On Rate Constants: Simple Collision Theory, Arrhenius Behavior ...
    Feb 17, 2010 · • For steric/orientational nature of collision, we introduce a steric factor. (empirically), p. • p < 1 generally. Thus, we can write a more ...
  15. [15]
    Collision Theory of Chemical Reactions | CK-12 Foundation
    ... Max Trautz and William Lewis (1916-1918), offers a more mechanistic insight into how reactions occur. It is based on the kinetic theory of gases and ...
  16. [16]
    [PDF] Theories of Reaction Rates
    Collision Theory. The so-called collision theory treats Ea as a potential energy barrier and then develops a molecular model for the frequency factor A ...Missing: threshold | Show results with:threshold
  17. [17]
  18. [18]
  19. [19]
  20. [20]
    Gas-phase oxidation of nitric oxide: chemical kinetics and rate ...
    This review describes the chemical kinetics and rate constant values k for the reaction 2NO + O2 = 2NO2. This reaction has been well established as a third- ...Missing: dependence Bodenstein
  21. [21]
    Reactions of Oriented Molecules | Science
    Reactions of Oriented Molecules: Molecules can be oriented in molecular beams and their reactions show some unexpected steric effects. Philip R. BrooksAuthors ...
  22. [22]
    Steric hindrance in potassium atom-oriented molecule reactions ...
    Analysis of the steric dependence of the CH3I + Rb reaction using a legendre expansion technique. ... Focusing, rotational state selection and orientation of CH3I ...
  23. [23]
    [PDF] CHEMICAL REACTION RATES Dr M. BROUARD Trinity Term 2003 ...
    Simple collision theory states reactions occur by collisions; rate is collision rate times fraction of collisions with energy exceeding the barrier height. The ...
  24. [24]
    Parameter Reliability and Understanding Enzyme Function - PMC
    Jan 1, 2022 · Whereas the kinetics of enzyme-substrate interactions has often assumed simple diffusion limited, or controlled, collision theory [71], the ...
  25. [25]
    Molecular Beam Study of the K+CH3I Reaction: Energy ...
    Dec 15, 1972 · A molecular beam study has been carried out on the reaction of a velocity‐selected K beam with a crossed beam of CH3I (from a glass capillary‐ ...Missing: steric factor