Download as pdf or txt
Download as pdf or txt
You are on page 1of 5

Activation energy

In chemistry and physics, activation energy is the minimum amount of


energy that must be provided to compounds to result in a chemical reaction.[1]
The activation energy (Ea) of a reaction is measured in joules per mole
(J/mol), kilojoules per mole (kJ/mol) or kilocalories per mole (kcal/mol).[2]
Activation energy can be thought of as the magnitude of the potential barrier
(sometimes called the energy barrier) separating minima of the potential
energy surface pertaining to the initial and final thermodynamic state. For a
chemical reaction to proceed at a reasonable rate, the temperature of the
system should be high enough such that there exists an appreciable number of
molecules with translational energy equal to or greater than the activation
energy. The term Activation Energy was introduced in 1889 by the Swedish
scientist Svante Arrhenius.[3]

The sparks created by


Contents striking steel against a piece
of flint provide the activation
Other uses energy to initiate
Temperature dependence and the relation to the Arrhenius combustion in this Bunsen
equation burner. The blue flame
sustains itself after the
Catalysts
sparks stop because the
Relationship with Gibbs energy of activation continued combustion of the
flame is now energetically
Negative activation energy
favorable.
Activation energy in a 2D Potential Energy Surface
See also
References

Other uses
Although less commonly used, activation energy also applies to nuclear reactions[4][5] and various other
physical phenomena.[6][7][8][9]

Temperature dependence and the relation to the Arrhenius


equation
The Arrhenius equation gives the quantitative basis of the relationship between the activation energy and the
rate at which a reaction proceeds. From the equation, the activation energy can be found through the relation
where A is the pre-exponential factor for the reaction, R is the universal gas constant, T is the absolute
temperature (usually in kelvins), and k is the reaction rate coefficient. Even without knowing A, Ea can be
evaluated from the variation in reaction rate coefficients as a function of temperature (within the validity of the
Arrhenius equation).

At a more advanced level, the net Arrhenius activation energy term from the Arrhenius equation is best
regarded as an experimentally determined parameter that indicates the sensitivity of the reaction rate to
temperature. There are two objections to associating this activation energy with the threshold barrier for an
elementary reaction. First, it is often unclear as to whether or not reaction does proceed in one step; threshold
barriers that are averaged out over all elementary steps have little theoretical value. Second, even if the reaction
being studied is elementary, a spectrum of individual collisions contributes to rate constants obtained from bulk
('bulb') experiments involving billions of molecules, with many different reactant collision geometries and
angles, different translational and (possibly) vibrational energies—all of which may lead to different
microscopic reaction rates.

Catalysts
A substance that modifies
the transition state to
lower the activation
energy is termed a
catalyst; a catalyst
composed only of protein
and (if applicable) small
molecule cofactors is
termed an enzyme. A
Example of an enzyme-catalysed catalyst increases the rate
exothermic reaction of reaction without being The relationship between activation
consumed in the energy ( ) and enthalpy of formation
reaction. [10] In addition, (ΔH) with and without a catalyst, plotted
the catalyst lowers the activation energy, but it does not change the against the reaction coordinate. The
energies of the original reactants or products, and so does not highest energy position (peak position)
change equilibrium.[11] Rather, the reactant energy and the product represents the transition state. With the
energy remain the same and only the activation energy is altered catalyst, the energy required to enter
(lowered). transition state decreases, thereby
decreasing the energy required to initiate
A catalyst is able to reduce the activation energy by forming a the reaction.
transition state in a more favorable manner. Catalysts, by nature,
create a more "comfortable" fit for the substrate of a reaction to
progress to a transition state. This is possible due to a release of energy that occurs when the substrate binds to
the active site of a catalyst. This energy is known as Binding Energy. Upon binding to a catalyst, substrates
partake in numerous stabilizing forces while within the active site (i.e. Hydrogen bonding, van der Waals
forces). Specific and favorable bonding occurs within the active site until the substrate forms to become the
high-energy transition state. Forming the transition state is more favorable with the catalyst because the
favorable stabilizing interactions within the active site release energy. A chemical reaction is able to
manufacture a high-energy transition state molecule more readily when there is a stabilizing fit within the
active site of a catalyst. The binding energy of a reaction is this energy released when favorable interactions
between substrate and catalyst occur. The binding energy released assists in achieving the unstable transition
state. Reactions otherwise without catalysts need a higher input of energy to achieve the transition state. Non-
catalyzed reactions do not have free energy available from active site stabilizing interactions, such as catalytic
enzyme reactions.[12]
Relationship with Gibbs energy of activation
In the Arrhenius equation, the term activation energy (Ea) is used to describe the energy required to reach the
transition state, and the exponential relationship k = A exp(-Ea/RT) holds. In transition state theory, a more
sophisticated model of the relationship between reaction rates and the transition state, a superficially similar
mathematical relationship, the Eyring equation, is used to describe the rate of a reaction: k = (kBT / h) exp(–
ΔG ‡ / RT). However, instead of modeling the temperature dependence of reaction rate phenomenologically,
the Eyring equation models individual elementary steps of a reaction. Thus, for a multistep process, there is no
straightforward relationship between the two models. Nevertheless, the functional forms of the Arrhenius and
Eyring equations are similar, and for a one-step process, simple and chemically meaningful correspondences
can be drawn between Arrhenius and Eyring parameters.

Instead of also using Ea, the Eyring equation uses the concept of Gibbs energy and the symbol ΔG ‡ to denote
the Gibbs energy of activation to achieve the transition state. In the equation, kB and h are the Boltzmann and
Planck constants, respectively. Although the equations look similar, it is important to note that the Gibbs
energy contains an entropic term in addition to the enthalpic one. In the Arrhenius equation, this entropic term
is accounted for by the pre-exponential factor A. More specifically, we can write the Gibbs free energy of
activation in terms of enthalpy and entropy of activation: ΔG ‡ = ΔH ‡ – T ΔS ‡ . Then, for a unimolecular, one-
step reaction, the approximate relationships Ea = ΔH ‡ + RT and A = (kBT/h) exp(1 + ΔS ‡ /R) hold. Note,
however, that in Arrhenius theory proper, A is temperature independent, while here, there is a linear
dependence on T. For a one-step unimolecular process whose half-life at room temperature is about 2 hours,
ΔG ‡ is approximately 23 kcal/mol. This is also the roughly the magnitude of Ea for a reaction that proceeds
over several hours at room temperature. Due to the relatively small magnitude of TΔS ‡ and RT at ordinary
temperatures for most reactions, in sloppy discourse, Ea, ΔG ‡ , and ΔH ‡ are often conflated and all referred to
as the "activation energy".

The enthalpy, entropy and Gibbs energy of activation are more correctly written as Δ ‡ Ho , Δ ‡ So and Δ ‡ Go
respectively, where the o indicates a quantity evaluated between standard states.[13][14] However, some
authors omit the o in order to simplify the notation.[15][16]

The total free energy change of a reaction is independent of the activation energy however. Physical and
chemical reactions can be either exergonic or endergonic, but the activation energy is not related to the
spontaneity of a reaction. The overall reaction energy change is not altered by the activation energy.

Negative activation energy


In some cases, rates of reaction decrease with increasing temperature. When following an approximately
exponential relationship so the rate constant can still be fit to an Arrhenius expression, this results in a negative
value of Ea. Elementary reactions exhibiting these negative activation energies are typically barrierless
reactions, in which the reaction proceeding relies on the capture of the molecules in a potential well. Increasing
the temperature leads to a reduced probability of the colliding molecules capturing one another (with more
glancing collisions not leading to reaction as the higher momentum carries the colliding particles out of the
potential well), expressed as a reaction cross section that decreases with increasing temperature. Such a
situation no longer leads itself to direct interpretations as the height of a potential barrier.[17]

Activation energy in a 2D Potential Energy Surface


Activation energy can be represented in 2D Potential Energy Surfaces (PES), where the relation between the
geometry of the reactants and the energy involved is represented as a topographic map.
In the following graphic there is a representation
of a reaction between hydrogen in the gas phase
and a metal: tungsten. The potential energy is
obtained with PES calculations and consistent
with the position of H from the NEB method
calculations. A 2-dimensional interpolation with
the spline method can be used to evaluate the
potential energy at these positions.[18] Products
and reactants can be found in the blue surface,
however the red surface corresponds to the
steady-state approximation.
These are PES profiles for the reaction of hydrogen gas
The depics correspond to the trajectories. The and tungsten. The reaction profile in the left shows a 2
bluer the surface, the stronger the hydrogen dimensional pathway of the procedure. On the right side it
bonds, so blue colors represent minima energy is shown a 1 dimension perspective of the same trajectory.
and red colors are maxima. Tungsten’s PES is
symmetric, and has a dip at the bridge site, this
dip corresponds to the change of color in the center of the depic.

The bluer the surface between the energy minima, the lower the energy barriers, and therefore the more easily
hydrogen travels along the surfaces.

See also
Activation energy asymptotics
Chemical kinetics
Fire point
Mean kinetic temperature
Quantum tunnelling
Hydrogen safety
Dust explosion
Spark plug

References
1. "Activation Energy" (https://web.archive.org/web/20161207175516/http://www.chem.fsu.edu/ch
emlab/chm1046course/activation). www.chem.fsu.edu. Archived from the original (http://www.c
hem.fsu.edu/chemlab/chm1046course/activation) on 2016-12-07. Retrieved 2017-01-13.
2. Espenson, James (1995). Chemical Kinetics and Reaction Mechanisms. McGraw-Hill.
ISBN 0070202605.
3. "Activation Energy and the Arrhenius Equation – Introductory Chemistry- 1st Canadian Edition"
(https://opentextbc.ca/introductorychemistry/chapter/activation-energy-and-the-arrhenius-equati
on-2/). opentextbc.ca. Retrieved 2018-04-05.
4. http://www.physics.ohio-state.edu/~kagan/phy367/Lectures/P367_lec_14.html
5. "Lecture XIV" (https://www.asc.ohio-state.edu/kagan.1/phy367/Lectures/P367_lec_14.html).
www.asc.ohio-state.edu. Retrieved 2019-03-22.
6. Pratt, Thomas H. "Electrostatic Ignitions of Fires and Explosions" Wiley-AIChE (July 15, 1997)
Center for Chemical Process Safety
7. Wang, Jenqdaw; Raj, Rishi (1990). "Estimate of the Activation Energies for Boundary Diffusion
from Rate-Controlled Sintering of Pure Alumina, and Alumina Doped with Zirconia or Titania".
Journal of the American Ceramic Society. 73 (5): 1172. doi:10.1111/j.1151-
2916.1990.tb05175.x (https://doi.org/10.1111%2Fj.1151-2916.1990.tb05175.x).
8. Kiraci, A; Yurtseven, H (2012). "Temperature Dependence of the Raman Frequency, Damping
Constant and the Activation Energy of a Soft-Optic Mode in Ferroelectric Barium Titanate".
Ferroelectrics. 432: 14–21. doi:10.1080/00150193.2012.707592 (https://doi.org/10.1080%2F00
150193.2012.707592). S2CID 121142463 (https://api.semanticscholar.org/CorpusID:12114246
3).
9. Terracciano, Anthony C; De Oliveira, Samuel; Vazquez-Molina, Demetrius; Uribe-Romo,
Fernando J; Vasu, Subith S; Orlovskaya, Nina (2017). "Effect of catalytically active Ce 0.8 Gd
0.2 O 1.9 coating on the heterogeneous combustion of methane within MgO stabilized ZrO 2
porous ceramics". Combustion and Flame. 180: 32–39.
doi:10.1016/j.combustflame.2017.02.019 (https://doi.org/10.1016%2Fj.combustflame.2017.02.0
19).
10. "General Chemistry Online: FAQ: Chemical change: What are some examples of reactions that
involve catalysts?" (http://antoine.frostburg.edu/chem/senese/101/reactions/faq/examples-of-cat
alysts.shtml). antoine.frostburg.edu. Retrieved 2017-01-13.
11. Bui, Matthew. "The Arrhenius Law: Activation Energies" (https://chem.libretexts.org/Core/Physi
cal_and_Theoretical_Chemistry/Kinetics/Modeling_Reaction_Kinetics/Temperature_Depende
nce_of_Reaction_Rates/The_Arrhenius_Law/The_Arrhenius_Law%3A_Activation_Energies).
Chemistry LibreTexts. UC Davis. Retrieved February 17, 2017.
12. Berg, Jeremy (2019). Biochemistry - Ninth Edition. New York, NY: WH Freeman and Company.
pp. 240–244. ISBN 978-1-319-11467-1.
13. "Enthalpy of activation" (http://goldbook.iupac.org/terms/view/E02142). IUPAC Gold Book (2nd
edition, on-line version). IUPAC (International Union of Pure and Applied Chemistry). 2019.
Retrieved 10 May 2020.
14. Steinfeld, Jeffrey I.; Francisco, Joseph S.; Hase, William L. (1999). Chemical Kinetics and
Dynamics (2nd ed.). Prentice Hall. p. 301. ISBN 0-13-737123-3.
15. Atkins, Peter; de Paula, Julio (2006). Atkins' Physical Chemistry (https://archive.org/details/phy
sicalchemistr00atki_572) (8th ed.). W.H.Freeman. p. 883 (https://archive.org/details/physicalche
mistr00atki_572/page/n914). ISBN 0-7167-8759-8. "... but we shall omit the standard state sign
to avoid overburdening the notation."
16. Laidler, Keith J.; Meiser, John H. (1982). Physical Chemistry. Benjamin/Cummings. p. 381.
ISBN 0-8053-5682-7.
17. Mozurkewich, Michael; Benson, Sidney (1984). "Negative activation energies and curved
Arrhenius plots. 1. Theory of reactions over potential wells". J. Phys. Chem. 88 (25): 6429–
6435. doi:10.1021/j150669a073 (https://doi.org/10.1021%2Fj150669a073).
18. Kristinsdóttir, Lilja; Skúlason, Egill (2012-09-01). "A systematic DFT study of hydrogen diffusion
on transition metal surfaces" (http://www.sciencedirect.com/science/article/pii/S003960281200
1604). Surface Science. 606 (17): 1400–1404. doi:10.1016/j.susc.2012.04.028 (https://doi.org/1
0.1016%2Fj.susc.2012.04.028). ISSN 0039-6028 (https://www.worldcat.org/issn/0039-6028).

Retrieved from "https://en.wikipedia.org/w/index.php?title=Activation_energy&oldid=1031940175"

This page was last edited on 4 July 2021, at 16:20 (UTC).

Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. By using this
site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of the Wikimedia
Foundation, Inc., a non-profit organization.

You might also like