Langmuir adsorption model

Last updated
A schematic showing equivalent sites, occupied (blue) and unoccupied (red), clarifying the basic assumptions used in the model. The adsorption sites (heavy dots) are equivalent and can have unit occupancy. Also, the adsorbates are immobile on the surface. Langmuir Adsorption Model.jpg
A schematic showing equivalent sites, occupied (blue) and unoccupied (red), clarifying the basic assumptions used in the model. The adsorption sites (heavy dots) are equivalent and can have unit occupancy. Also, the adsorbates are immobile on the surface.

The Langmuir adsorption model explains adsorption by assuming an adsorbate behaves as an ideal gas at isothermal conditions. According to the model, adsorption and desorption are reversible processes. This model even explains the effect of pressure i.e. at these conditions the adsorbate's partial pressure is related to its volume V adsorbed onto a solid adsorbent. The adsorbent, as indicated in the figure, is assumed to be an ideal solid surface composed of a series of distinct sites capable of binding the adsorbate. The adsorbate binding is treated as a chemical reaction between the adsorbate gaseous molecule and an empty sorption site S. This reaction yields an adsorbed species with an associated equilibrium constant :

Contents

.

From these basic hypotheses the mathematical formulation of the Langmuir adsorption isotherm can be derived in various independent and complementary ways: by the kinetics, the thermodynamics, and the statistical mechanics approaches respectively (see below for the different demonstrations).

The Langmuir adsorption equation is

where is the fractional occupancy of the adsorption sites, i.e., the ratio of the volume V of gas adsorbed onto the solid to the volume of a gas molecules monolayer covering the whole surface of the solid and completely occupied by the adsorbate. A continuous monolayer of adsorbate molecules covering a homogeneous flat solid surface is the conceptual basis for this adsorption model. [1]

Background and experiments

In 1916, Irving Langmuir presented his model for the adsorption of species onto simple surfaces. Langmuir was awarded the Nobel Prize in 1932 for his work concerning surface chemistry. He hypothesized that a given surface has a certain number of equivalent sites to which a species can “stick”, either by physisorption or chemisorption. His theory began when he postulated that gaseous molecules do not rebound elastically from a surface, but are held by it in a similar way to groups of molecules in solid bodies. [2]

Langmuir published two papers that confirmed the assumption that adsorbed films do not exceed one molecule in thickness. The first experiment involved observing electron emission from heated filaments in gases. [3] The second, a more direct evidence, examined and measured the films of liquid onto an adsorbent surface layer. He also noted that generally the attractive strength between the surface and the first layer of adsorbed substance is much greater than the strength between the first and second layer. However, there are instances where the subsequent layers may condense given the right combination of temperature and pressure. [4]

Basic assumptions of the model

Inherent within this model, the following assumptions [5] are valid specifically for the simplest case: the adsorption of a single adsorbate onto a series of equivalent sites onto the surface of the solid.

  1. The surface containing the adsorbing sites is a perfectly flat plane with no corrugations (assume the surface is homogeneous). However, chemically heterogeneous surfaces can be considered to be homogeneous if the adsorbate is bound to only one type of functional groups on the surface.
  2. The adsorbing gas adsorbs into an immobile state.
  3. All sites are energetically equivalent, and the energy of adsorption is equal for all sites.
  4. Each site can hold at most one molecule (mono-layer coverage only).
  5. No (or ideal) interactions between adsorbate molecules on adjacent sites. When the interactions are ideal, the energy of side-to-side interactions is equal for all sites regardless of the surface occupancy.

Derivations of the Langmuir adsorption isotherm

The mathematical expression of the Langmuir adsorption isotherm involving only one sorbing species can be demonstrated in different ways: the kinetics approach, the thermodynamics approach, and the statistical mechanics approach respectively. In case of two competing adsorbed species, the competitive adsorption model is required, while when a sorbed species dissociates into two distinct entities, the dissociative adsorption model need to be used.

Kinetic derivation

This section [5] provides a kinetic derivation for a single-adsorbate case. The kinetic derivation applies to gas-phase adsorption. However, it has been mistakenly applied to solutions. The multiple-adsorbate case is covered in the competitive adsorption sub-section. The model assumes adsorption and desorption as being elementary processes, where the rate of adsorption rad and the rate of desorption rd are given by

where pA is the partial pressure of A over the surface, [S] is the concentration of free sites in number/m2, [Aad] is the surface concentration of A in molecules/m2 (concentration of occupied sites), and kad and kd are constants of forward adsorption reaction and backward desorption reaction in the above reactions.

At equilibrium, the rate of adsorption equals the rate of desorption. Setting rad = rd and rearranging, we obtain

The concentration of sites is given by dividing the total number of sites (S0) covering the whole surface by the area of the adsorbent (a):

We can then calculate the concentration of all sites by summing the concentration of free sites [S] and occupied sites:

Combining this with the equilibrium equation, we get

We define now the fraction of the surface sites covered with A as

This, applied to the previous equation that combined site balance and equilibrium, yields the Langmuir adsorption isotherm:

Thermodynamic derivation

In condensed phases (solutions), adsorption to a solid surface is a competitive process between the solvent (A) and the solute (B) to occupy the binding site. The thermodynamic equilibrium is described as

Solvent (bound) + Solute (free) ↔ Solvent (free) + Solute (bound).

If we designate the solvent by the subscript "1" and the solute by "2", and the bound state by the superscript "s" (surface/bound) and the free state by the "b" (bulk solution / free), then the equilibrium constant can be written as a ratio between the activities of products over reactants:

For dilute solutions the activity of the solvent in bulk solution and the activity coefficients () are also assumed to ideal on the surface. Thus, , and where are mole fractions. Re-writing the equilibrium constant and solving for yields

Note that the concentration of the solute adsorbate can be used instead of the activity coefficient. However, the equilibrium constant will no longer be dimensionless and will have units of reciprocal concentration instead. The difference between the kinetic and thermodynamic derivations of the Langmuir model is that the thermodynamic uses activities as a starting point while the kinetic derivation uses rates of reaction. The thermodynamic derivation allows for the activity coefficients of adsorbates in their bound and free states to be included. The thermodynamic derivation is usually referred to as the "Langmuir-like equation". [6] [7]

Statistical mechanical derivation

This derivation [8] [9] based on statistical mechanics was originally provided by Volmer and Mahnert [10] in 1925. The partition function of the finite number of adsorbents adsorbed on a surface, in a canonical ensemble, is given by

where is the partition function of a single adsorbed molecule, is the number of adsorption sites (both occupied and unoccupied), and is the number of adsorbed molecules which should be less than or equal to . The terms in the bracket give the total partition function of the adsorbed molecules by taking a product of the individual partition functions (refer to Partition function of subsystems). The factor accounts for the overcounting arising due to the indistinguishable nature of the adsorbates. The grand canonical partition function is given by

is the chemical potential of an adsorbed molecule. As it has the form of binomial series, the summation is reduced to

where

The grand canonical potential is

based on which the average number of occupied sites is calculated

which gives the coverage

Now, invoking the condition that the system is in equilibrium, that is, the chemical potential of the adsorbed molecules is equal to that of the molecules in gas phase, we have

An example plot of the surface coverage thA = P/(P + P0) with respect to the partial pressure of the adsorbate. P0 = 100 mTorr. The graph shows levelling off of the surface coverage at pressures higher than P0. Surface coverage plot.gif
An example plot of the surface coverage θA = P/(P + P0) with respect to the partial pressure of the adsorbate. P0 = 100 mTorr. The graph shows levelling off of the surface coverage at pressures higher than P0.

The chemical potential of an ideal gas is

where is the Helmholtz free energy of an ideal gas with its partition function

is the partition function of a single particle in the volume of (only consider the translational freedom here).

We thus have , where we use Stirling's approximation.

Plugging to the expression of , we have

which gives the coverage

By defining

and using the identity , finally, we have

It is plotted in the figure alongside demonstrating that the surface coverage increases quite rapidly with the partial pressure of the adsorbants, but levels off after P reaches P0.

Competitive adsorption

The previous derivations assumed that there is only one species, A, adsorbing onto the surface. This section [11] considers the case when there are two distinct adsorbates present in the system. Consider two species A and B that compete for the same adsorption sites. The following hypotheses are made here:

  1. All the sites are equivalent.
  2. Each site can hold at most one molecule of A, or one molecule of B, but not both simultaneously.
  3. There are no interactions between adsorbate molecules on adjacent sites.

As derived using kinetic considerations, the equilibrium constants for both A and B are given by

and

The site balance states that the concentration of total sites [S0] is equal to the sum of free sites, sites occupied by A and sites occupied by B:

Inserting the equilibrium equations and rearranging in the same way we did for the single-species adsorption, we get similar expressions for both θA and θB:

Dissociative adsorption

The other case of special importance is when a molecule D2 dissociates into two atoms upon adsorption. [11] Here, the following assumptions would be held to be valid:

  1. D2 completely dissociates to two molecules of D upon adsorption.
  2. The D atoms adsorb onto distinct sites on the surface of the solid and then move around and equilibrate.
  3. All sites are equivalent.
  4. Each site can hold at most one atom of D.
  5. There are no interactions between adsorbate molecules on adjacent sites.

Using similar kinetic considerations, we get

The 1/2 exponent on pD2 arises because one gas phase molecule produces two adsorbed species. Applying the site balance as done above,

Entropic considerations

The formation of Langmuir monolayers by adsorption onto a surface dramatically reduces the entropy of the molecular system.

To find the entropy decrease, we find the entropy of the molecule when in the adsorbed condition. [12]

Using Stirling's approximation, we have

On the other hand, the entropy of a molecule of an ideal gas is

where is the thermal de Broglie wavelength of the gas molecule.

Limitations of the model

The Langmuir adsorption model deviates significantly in many cases, primarily because it fails to account for the surface roughness of the adsorbent. Rough inhomogeneous surfaces have multiple site types available for adsorption, with some parameters varying from site to site, such as the heat of adsorption. Moreover, specific surface area is a scale-dependent quantity, and no single true value exists for this parameter. [1] Thus, the use of alternative probe molecules can often result in different obtained numerical values for surface area, rendering comparison problematic.

The model also ignores adsorbate–adsorbate interactions. Experimentally, there is clear evidence for adsorbate–adsorbate interactions in heat of adsorption data. There are two kinds of adsorbate–adsorbate interactions: direct interaction and indirect interaction. Direct interactions are between adjacent adsorbed molecules, which could make adsorbing near another adsorbate molecule more or less favorable and greatly affects high-coverage behavior. In indirect interactions, the adsorbate changes the surface around the adsorbed site, which in turn affects the adsorption of other adsorbate molecules nearby.

Modifications

The modifications try to account for the points mentioned in above section like surface roughness, inhomogeneity, and adsorbate–adsorbate interactions.

Two-mechanism Langmuir-like equation (TMLLE)

Also known as the two-site Langmuir equation. This equation describes the adsorption of one adsorbate to two or more distinct types of adsorption sites. Each binding site can be described with its own Langmuir expression, as long as the adsorption at each binding site type is independent from the rest.

where

– total amount adsorbed at a given adsorbate concentration,
– maximum capacity of site type 1,
– maximum capacity of site type 2,
– equilibrium (affinity) constant of site type 1,
 – equilibrium (affinity) constant of site type 2,
– adsorbate activity in solution at equilibrium

This equation works well for adsorption of some drug molecules to activated carbon in which some adsorbate molecules interact with hydrogen bonding while others interact with a different part of the surface by hydrophobic interactions (hydrophobic effect). The equation was modified to account for the hydrophobic effect (also known as entropy-driven adsorption): [13]

The hydrophobic effect is independent of concentration, since Therefore, the capacity of the adsorbent for hydrophobic interactions can obtained from fitting to experimental data. The entropy-driven adsorption originates from the restriction of translational motion of bulk water molecules by the adsorbate, which is alleviated upon adsorption.

Freundlich adsorption isotherm

The Freundlich isotherm is the most important multi-site adsorption isotherm for rough surfaces.

where αF and CF are fitting parameters. [14] This equation implies that if one makes a log–log plot of adsorption data, the data will fit a straight line. The Freundlich isotherm has two parameters, while Langmuir's equations has only one: as a result, it often fits the data on rough surfaces better than the Langmuir's equations. However, the Freundlich equation is unique; consequently, if the data fit the equation, it is only likely, but not proved, that the surface is heterogeneous. The heterogeneity of the surface can be confirmed with calorimetry. Homogeneous surfaces (or heterogeneous surfaces that exhibit homogeneous adsorption (single-site)) have a constant of adsorption as a function of the occupied-sites fraction. On the other hand, heterogeneous adsorption (multi-site) have a variable of adsorption depending on the sites occupation. When the adsorbate pressure (or concentration) are low, high-energy sites are occupied, and as the pressure (or concentration) increases, the lesser-energy sites become occupied, resulting in a lower of adsorption. [15]

A related equation is the Toth equation. Rearranging the Langmuir equation, one can obtain

Toth [16] modified this equation by adding two parameters αT0 and CT0 to formulate the Toth equation:

Temkin adsorption isotherm

This isotherm takes into account indirect adsorbate–adsorbate interactions on adsorption isotherms. Temkin [17] noted experimentally that heats of adsorption would more often decrease than increase with increasing coverage.

The heat of adsorption ΔHad is defined as

He derived a model assuming that as the surface is loaded up with adsorbate, the heat of adsorption of all the molecules in the layer would decrease linearly with coverage due to adsorbate–adsorbate interactions:

where αT is a fitting parameter. Assuming the Langmuir adsorption isotherm still applied to the adsorbed layer, is expected to vary with coverage as follows:

Langmuir's isotherm can be rearranged to

Substituting the expression of the equilibrium constant and taking the natural logarithm:

BET equation

Brunauer, Emmett and Teller (BET) model of multilayer adsorption, that is, a random distribution of sites covered by one, two, three, etc., adsorbate molecules. BET Multilayer Adsorption.svg
Brunauer, Emmett and Teller (BET) model of multilayer adsorption, that is, a random distribution of sites covered by one, two, three, etc., adsorbate molecules.

Brunauer, Emmett and Teller (BET) [18] derived the first isotherm for multilayer adsorption. It assumes a random distribution of sites that are empty or that are covered with by one monolayer, two layers and so on, as illustrated alongside. The main equation of this model is

where

and [A] is the total concentration of molecules on the surface, given by

where

in which [A]0 is the number of bare sites, and [A]i is the number of surface sites covered by i molecules.

Adsorption of a binary liquid on a solid

This section describes the surface coverage when the adsorbate is in liquid phase and is a binary mixture. [19]

For ideal both phases  no lateral interactions, homogeneous surface  the composition of a surface phase for a binary liquid system in contact with solid surface is given by a classic Everett isotherm equation (being a simple analogue of Langmuir equation), where the components are interchangeable (i.e. "1" may be exchanged to "2") without change of equation form:

where the normal definition of multi-component system is valid as follows:

By simple rearrangement, we get

This equation describes competition of components "1" and "2".

See also

Related Research Articles

<span class="mw-page-title-main">Fick's laws of diffusion</span> Mathematical descriptions of molecular diffusion

Fick's laws of diffusion describe diffusion and were first posited by Adolf Fick in 1855 on the basis of largely experimental results. They can be used to solve for the diffusion coefficient, D. Fick's first law can be used to derive his second law which in turn is identical to the diffusion equation.

<span class="mw-page-title-main">Kinetic theory of gases</span> Understanding of gas properties in terms of molecular motion

The kinetic theory of gases is a simple classical model of the thermodynamic behavior of gases. It treats a gas as composed of numerous particles, too small to see with a microscope, which are constantly in random motion. Their collisions with each other and with the walls of their container are used to explain physical properties of the gas—for example the relationship between its temperature, pressure and volume. The particles are now known to be the atoms or molecules of the gas.

Physisorption, also called physical adsorption, is a process in which the electronic structure of the atom or molecule is barely perturbed upon adsorption.

<span class="mw-page-title-main">Gamma distribution</span> Probability distribution

In probability theory and statistics, the gamma distribution is a versatile two-parameter family of continuous probability distributions. The exponential distribution, Erlang distribution, and chi-squared distribution are special cases of the gamma distribution. There are two equivalent parameterizations in common use:

  1. With a shape parameter k and a scale parameter θ
  2. With a shape parameter and an inverse scale parameter , called a rate parameter.
<span class="mw-page-title-main">Adsorption</span> Phenomenon of surface adhesion

Adsorption is the adhesion of atoms, ions or molecules from a gas, liquid or dissolved solid to a surface. This process creates a film of the adsorbate on the surface of the adsorbent. This process differs from absorption, in which a fluid is dissolved by or permeates a liquid or solid. While adsorption does often precede absorption, which involves the transfer of the absorbate into the volume of the absorbent material, alternatively, adsorption is distinctly a surface phenomenon, wherein the adsorbate does not penetrate through the material surface and into the bulk of the adsorbent. The term sorption encompasses both adsorption and absorption, and desorption is the reverse of sorption.

Temperature programmed desorption (TPD) is the method of observing desorbed molecules from a surface when the surface temperature is increased. When experiments are performed using well-defined surfaces of single-crystalline samples in a continuously pumped ultra-high vacuum (UHV) chamber, then this experimental technique is often also referred to as thermal desorption spectroscopy or thermal desorption spectrometry (TDS).

Desorption is the physical process where adsorbed atoms or molecules are released from a surface into the surrounding vacuum or fluid. This occurs when a molecule gains enough energy to overcome the activation barrier and the binding energy that keep it attached to the surface.

<span class="mw-page-title-main">Wetting</span> Ability of a liquid to maintain contact with a solid surface

Wetting is the ability of a liquid to maintain contact with a solid surface, resulting from intermolecular interactions when the two are brought together. This happens in presence of a gaseous phase or another liquid phase not miscible with the first one. The degree of wetting (wettability) is determined by a force balance between adhesive and cohesive forces. There are two types of wetting: non-reactive wetting and reactive wetting.

Brunauer–Emmett–Teller (BET) theory aims to explain the physical adsorption of gas molecules on a solid surface and serves as the basis for an important analysis technique for the measurement of the specific surface area of materials. The observations are very often referred to as physical adsorption or physisorption. In 1938, Stephen Brunauer, Paul Hugh Emmett, and Edward Teller presented their theory in the Journal of the American Chemical Society. BET theory applies to systems of multilayer adsorption that usually utilizes a probing gas (called the adsorbate) that does not react chemically with the adsorptive (the material upon which the gas attaches to) to quantify specific surface area. Nitrogen is the most commonly employed gaseous adsorbate for probing surface(s). For this reason, standard BET analysis is most often conducted at the boiling temperature of N2 (77 K). Other probing adsorbates are also utilized, albeit less often, allowing the measurement of surface area at different temperatures and measurement scales. These include argon, carbon dioxide, and water. Specific surface area is a scale-dependent property, with no single true value of specific surface area definable, and thus quantities of specific surface area determined through BET theory may depend on the adsorbate molecule utilized and its adsorption cross section.

The sticking probability is the probability that molecules are trapped on surfaces and adsorb chemically. From Langmuir's adsorption isotherm, molecules cannot adsorb on surfaces when the adsorption sites are already occupied by other molecules, so the sticking probability can be expressed as follows:

<span class="mw-page-title-main">Hill equation (biochemistry)</span> Diagram showing the proportion of a receptor bound to a ligand

In biochemistry and pharmacology, the Hill equation refers to two closely related equations that reflect the binding of ligands to macromolecules, as a function of the ligand concentration. A ligand is "a substance that forms a complex with a biomolecule to serve a biological purpose", and a macromolecule is a very large molecule, such as a protein, with a complex structure of components. Protein-ligand binding typically changes the structure of the target protein, thereby changing its function in a cell.

The principle of detailed balance can be used in kinetic systems which are decomposed into elementary processes. It states that at equilibrium, each elementary process is in equilibrium with its reverse process.

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

<span class="mw-page-title-main">Pendulum (mechanics)</span> Free swinging suspended body

A pendulum is a body suspended from a fixed support so that it swings freely back and forth under the influence of gravity. When a pendulum is displaced sideways from its resting, equilibrium position, it is subject to a restoring force due to gravity that will accelerate it back towards the equilibrium position. When released, the restoring force acting on the pendulum's mass causes it to oscillate about the equilibrium position, swinging it back and forth. The mathematics of pendulums are in general quite complicated. Simplifying assumptions can be made, which in the case of a simple pendulum allow the equations of motion to be solved analytically for small-angle oscillations.

The Freundlich equation or Freundlich adsorption isotherm, an adsorption isotherm, is an empirical relationship between the quantity of a gas adsorbed into a solid surface and the gas pressure. The same relationship is also applicable for the concentration of a solute adsorbed onto the surface of a solid and the concentration of the solute in the liquid phase. In 1909, Herbert Freundlich gave an expression representing the isothermal variation of adsorption of a quantity of gas adsorbed by unit mass of solid adsorbent with gas pressure. This equation is known as Freundlich adsorption isotherm or Freundlich adsorption equation. As this relationship is entirely empirical, in the case where adsorption behavior can be properly fit by isotherms with a theoretical basis, it is usually appropriate to use such isotherms instead. The Freundlich equation is also derived (non-empirically) by attributing the change in the equilibrium constant of the binding process to the heterogeneity of the surface and the variation in the heat of adsorption.

Sticking coefficient is the term used in surface physics to describe the ratio of the number of adsorbate atoms that adsorb, or "stick", to a surface to the total number of atoms that impinge upon that surface during the same period of time. Sometimes the symbol Sc is used to denote this coefficient, and its value is between 1 and 0. The coefficient is a function of surface temperature, surface coverage (θ) and structural details as well as the kinetic energy of the impinging particles. The original formulation was for molecules adsorbing from the gas phase and the equation was later extended to adsorption from the liquid phase by comparison with molecular dynamics simulations. For use in adsorption from liquids the equation is expressed based on solute density rather than the pressure.

In materials science, segregation is the enrichment of atoms, ions, or molecules at a microscopic region in a materials system. While the terms segregation and adsorption are essentially synonymous, in practice, segregation is often used to describe the partitioning of molecular constituents to defects from solid solutions, whereas adsorption is generally used to describe such partitioning from liquids and gases to surfaces. The molecular-level segregation discussed in this article is distinct from other types of materials phenomena that are often called segregation, such as particle segregation in granular materials, and phase separation or precipitation, wherein molecules are segregated in to macroscopic regions of different compositions. Segregation has many practical consequences, ranging from the formation of soap bubbles, to microstructural engineering in materials science, to the stabilization of colloidal suspensions.

Supercritical adsorption also referred to as the adsorption of supercritical fluids, is the adsorption at above-critical temperatures. There are different tacit understandings of supercritical fluids. For example, “a fluid is considered to be ‘supercritical’ when its temperature and pressure exceed the temperature and pressure at the critical point”. In the studies of supercritical extraction, however, “supercritical fluid” is applied for a narrow temperature region of 1-1.2 or to +10 K, which is called the supercritical region.

Adsorption is the adhesion of ions or molecules onto the surface of another phase. Adsorption may occur via physisorption and chemisorption. Ions and molecules can adsorb to many types of surfaces including polymer surfaces. A polymer is a large molecule composed of repeating subunits bound together by covalent bonds. In dilute solution, polymers form globule structures. When a polymer adsorbs to a surface that it interacts favorably with, the globule is essentially squashed, and the polymer has a pancake structure.

The potential theory of Polanyi, also called Polanyi adsorption potential theory, is a model of adsorption proposed by Michael Polanyi where adsorption can be measured through the equilibrium between the chemical potential of a gas near the surface and the chemical potential of the gas from a large distance away. In this model, he assumed that the attraction largely due to Van Der Waals forces of the gas to the surface is determined by the position of the gas particle from the surface, and that the gas behaves as an ideal gas until condensation where the gas exceeds its equilibrium vapor pressure. While the adsorption theory of Henry is more applicable in low pressure and BET adsorption isotherm equation is more useful at from 0.05 to 0.35 P/Po, the Polanyi potential theory has much more application at higher P/Po (~0.1–0.8).

References

  1. 1 2 Hanaor, D. A. H.; Ghadiri, M.; Chrzanowski, W.; Gan, Y. (2014). "Scalable Surface Area Characterization by Electrokinetic Analysis of Complex Anion Adsorption" (PDF). Langmuir. 30 (50): 15143–15152. arXiv: 2106.03411 . doi:10.1021/la503581e. PMID   25495551. S2CID   4697498.
  2. Langmuir, Irving (June 1918). "The Adsorption of Gases on Plane Surface of Glass, Mica and Platinum". Journal of the American Chemical Society. 40 (9): 1361–1402. doi:10.1021/ja02242a004.
  3. Langmuir, Irving (1916). "Part I". The Research Laboratory of the General Electric Company: 2221.
  4. Langmuir, Irving (1918). "Part II". The Research Laboratory of the General Electric Company: 1848.
  5. 1 2 Masel, Richard (1996). Principles of Adsorption and Reaction on Solid Surfaces. Wiley Interscience. p. 240. ISBN   978-0-471-30392-3.
  6. Arthur W. Adamson. "Physical Chemistry of Surfaces". Sixth edition.
  7. Kleman, Maurice; Lavrentovich, Oleg D.; Laverntovich, Oleg D. (2004). "Surface Phenomena". Soft Matter Physics: An Introduction. Partially Ordered Systems. pp. 472–518. doi:10.1007/978-0-387-21759-8_13. ISBN   978-0-387-95267-3.
  8. Masel, Richard (1996). Principles of Adsorption and Reaction on Solid Surfaces. Wiley Interscience. p. 242. ISBN   978-0-471-30392-3.
  9. Cahill, David (2008). "Lecture Notes 5 Page 2" (pdf). University of Illinois, Urbana Champaign. Retrieved 2008-11-09.
  10. Volmer, M. A.; Mahnert, P. (1925). "Solution of Solid Substances in Liquid Surfaces and the Characteristics of Layers Thus Produced". Z. Phys. Chem. 115: 253. doi:10.1515/zpch-1925-11519. S2CID   100362186.
  11. 1 2 Masel, Richard (1996). Principles of Adsorption and Reaction on Solid Surfaces. Wiley Interscience. p. 244. ISBN   978-0-471-30392-3.
  12. Cahill, David (2008). "Lecture Notes 5 Page 13" (pdf). University of Illinois, Urbana Champaign. Retrieved 2008-11-09.
  13. Kirk A. VanDer Kamp, Dongmei Qiang, Aktham Aburub, and Dale Eric Wurster, "Modified Langmuir-like Model for Modeling the Adsorption from Aqueous Solutions by Activated Carbons", Langmuir 2005 21 (1), 217–224, doi : 10.1021/la040093o.
  14. Freundlich, H. (1909). "Eine darstellung der chemie der kolloide und verwanter gebiete". Kapillarchemie (in German).
  15. Adamson, A. W. (1997). Physical chemistry of surfaces. p. 699.
  16. Toth, J. (1971). "State equations of the solid gas interface layer". Acta Chim. Acad. Sci. Hung. 69: 311.
  17. Temkin, M. I.; Pyzhev, V. (1940). "Kinetics of ammonia synthesis on promoted iron catalyst". Acta Phys. Chim. USSR. 12: 327.
  18. Brunauer, Stephen; Emmett, P. H.; Teller, Edward (1938). "Adsorption of gases in multimolecular layers". Journal of the American Chemical Society. 60 (2): 309–319. Bibcode:1938JAChS..60..309B. doi:10.1021/ja01269a023. ISSN   0002-7863.
  19. Marczewski, A. W. (2002). "Basics of liquid adsorption". www.adsorption.org. Retrieved 2008-11-24.