You are on page 1of 17

Adsorption

Adsorption is the adhesion of atoms, ions or molecules from a


gas, liquid or dissolved solid to a surface.[1] This process creates a
film of the adsorbate on the surface of the adsorbent. This process
differs from absorption, in which a fluid (the absorbate) is
dissolved by or permeates a liquid or solid (the absorbent).[2]
Adsorption is a surface phenomenon, while absorption involves
Brunauer, Emmett and Teller's
the whole volume of the material, although adsorption does often
model of multilayer adsorption is a
precede absorption.[3] The term sorption encompasses both random distribution of molecules on
processes, while desorption is the reverse of it. the material surface.

Like surface tension, adsorption is a


consequence of surface energy. In a bulk
material, all the bonding requirements (be they IUPAC definition
ionic, covalent or metallic) of the constituent Increase in the concentration of a substance at
atoms of the material are fulfilled by other the interface of a condensed and a liquid or
atoms in the material. However, atoms on the gaseous layer owing to the operation of surface
surface of the adsorbent are not wholly forces.
surrounded by other adsorbent atoms and
therefore can attract adsorbates. The exact Note 1: Adsorption of proteins is of great
nature of the bonding depends on the details of importance when a material is in contact with
the species involved, but the adsorption process blood or body fluids. In the case of blood,
is generally classified as physisorption albumin, which is largely predominant, is
(characteristic of weak van der Waals forces) or generally adsorbed first, and then
chemisorption (characteristic of covalent rearrangements occur in favor of other minor
bonding). It may also occur due to electrostatic proteins according to surface affinity against
mass law selection (Vroman effect).
attraction.[5][6]
Note 2: Adsorbed molecules are those that are
Adsorption is present in many natural, physical,
biological and chemical systems and is widely resistant to washing with the same solvent
medium in the case of adsorption from
used in industrial applications such as
solutions. The washing conditions can thus
heterogeneous catalysts,[7][8] activated
modify the measurement results, particularly
charcoal, capturing and using waste heat to
provide cold water for air conditioning and when the interaction energy is low.[4]
other process requirements (adsorption
chillers), synthetic resins, increasing storage capacity of carbide-derived carbons and water
purification. Adsorption, ion exchange and chromatography are sorption processes in which certain
adsorbates are selectively transferred from the fluid phase to the surface of insoluble, rigid particles
suspended in a vessel or packed in a column. Pharmaceutical industry applications, which use
adsorption as a means to prolong neurological exposure to specific drugs or parts thereof, are lesser
known.

The word "adsorption" was coined in 1881 by German physicist Heinrich Kayser (1853–1940).[9]
Contents
Isotherms
Freundlich
Langmuir
BET
Kisliuk
Adsorption enthalpy
Single-molecule explanation
Quantum mechanical – thermodynamic modelling for surface area and porosity
Adsorbents
Characteristics and general requirements
Silica gel
Zeolites
Activated carbon
Water adsorption
Adsorption solar heating and storage
Carbon capture and storage
Protein and surfactant adsorption
Adsorption chillers
Portal site mediated adsorption
Adsorption spillover
Polymer adsorption
Adsorption in viruses
In popular culture
See also
References
Further reading
External links

Isotherms
The adsorption of gases and solutes is usually described through isotherms, that is, the amount of
adsorbate on the adsorbent as a function of its pressure (if gas) or concentration (for liquid phase
solutes) at constant temperature. The quantity adsorbed is nearly always normalized by the mass of
the adsorbent to allow comparison of different materials. To date, 15 different isotherm models have
been developed.[10]

Freundlich
The first mathematical fit to an isotherm was published by Freundlich and Kuster (1906) and is a
purely empirical formula for gaseous adsorbates:

where is the mass of adsorbate adsorbed, is the mass of the adsorbent, is the pressure of
adsorbate (this can be changed to concentration if investigating solution rather than gas), and and
are empirical constants for each adsorbent–adsorbate pair at a given temperature. The function is not
adequate at very high pressure because in reality has an asymptotic maximum as pressure
increases without bound. As the temperature increases, the constants and change to reflect the
empirical observation that the quantity adsorbed rises more slowly and higher pressures are required
to saturate the surface.

Langmuir

Irving Langmuir was the first to derive a scientifically based adsorption isotherm in 1918.[11] The
model applies to gases adsorbed on solid surfaces. It is a semi-empirical isotherm with a kinetic basis
and was derived based on statistical thermodynamics. It is the most common isotherm equation to
use due to its simplicity and its ability to fit a variety of adsorption data. It is based on four
assumptions:

1. All of the adsorption sites are equivalent, and each site can only accommodate one molecule.
2. The surface is energetically homogeneous, and adsorbed molecules do not interact.
3. There are no phase transitions.
4. At the maximum adsorption, only a monolayer is formed. Adsorption only occurs on localized sites
on the surface, not with other adsorbates.

These four assumptions are seldom all true: there are always imperfections on the surface, adsorbed
molecules are not necessarily inert, and the mechanism is clearly not the same for the very first
molecules to adsorb to a surface as for the last. The fourth condition is the most troublesome, as
frequently more molecules will adsorb to the monolayer; this problem is addressed by the BET
isotherm for relatively flat (non-microporous) surfaces. The Langmuir isotherm is nonetheless the
first choice for most models of adsorption and has many applications in surface kinetics (usually
called Langmuir–Hinshelwood kinetics) and thermodynamics.

Langmuir suggested that adsorption takes place through this mechanism: , where A is
a gas molecule, and S is an adsorption site. The direct and inverse rate constants are k and k−1. If we
define surface coverage, , as the fraction of the adsorption sites occupied, in the equilibrium we have:

or

where is the partial pressure of the gas or the molar concentration of the solution.
For very low
pressures , and for high pressures .
The value of is difficult to measure experimentally; usually, the adsorbate is a gas and the quantity
adsorbed is given in moles, grams, or gas volumes at standard temperature and pressure (STP) per
gram of adsorbent. If we call vmon the STP volume of adsorbate required to form a monolayer on the
adsorbent (per gram of adsorbent), then , and we obtain an expression for a straight line:

Through its slope and y intercept we can obtain vmon and K, which are constants for each adsorbent–
adsorbate pair at a given temperature. vmon is related to the number of adsorption sites through the
ideal gas law. If we assume that the number of sites is just the whole area of the solid divided into the
cross section of the adsorbate molecules, we can easily calculate the surface area of the adsorbent.
The
surface area of an adsorbent depends on its structure: the more pores it has, the greater the area,
which has a big influence on reactions on surfaces.

If more than one gas adsorbs on the surface, we define as the fraction of empty sites, and we have:

Also, we can define as the fraction of the sites occupied by the j-th gas:

where i is each one of the gases that adsorb.

Note:

1) To choose between the langmuir and freundlich equations, the enthalpies of adsorption must be
investigated.[12] While the langmuir model assumes that the energy of adsorption remains constant
with surface occupancy, the Freundlich equation is derived with the assumption that the heat of
adsorption continually decrease as the binding sites are occupied.[13] The choice of the model based
on best fitting of the data is a common misconception.[12]

2) The use of the linearized form of the langmuir model is no longer common practice. Advances in
computational power allowed for nonlinear regression to be performed quickly and with higher
confidence since no data transformation is required.

BET

Often molecules do form multilayers, that is, some are adsorbed on already adsorbed molecules, and
the Langmuir isotherm is not valid. In 1938 Stephen Brunauer, Paul Emmett, and Edward Teller
developed a model isotherm that takes that possibility into account. Their theory is called BET theory,
after the initials in their last names. They modified Langmuir's mechanism as follows:

A(g) + S ⇌ AS,
A(g) + AS ⇌ A2S,

A(g) + A2S ⇌ A3S and so on.

The derivation of the formula is more complicated than


Langmuir's (see links for complete derivation). We obtain:

where x is the pressure divided by the vapor pressure for the


adsorbate at that temperature (usually denoted ), v is the
STP volume of adsorbed adsorbate, vmon is the STP volume of the Langmuir (blue) and BET (red)
amount of adsorbate required to form a monolayer, and c is the isotherms
equilibrium constant K we used in Langmuir isotherm multiplied
by the vapor pressure of the adsorbate. The key assumption used
in deriving the BET equation that the successive heats of adsorption for all layers except the first are
equal to the heat of condensation of the adsorbate.

The Langmuir isotherm is usually better for chemisorption, and the BET isotherm works better for
physisorption for non-microporous surfaces.

Kisliuk

In other instances, molecular interactions between gas molecules


previously adsorbed on a solid surface form significant
interactions with gas molecules in the gaseous phases. Hence,
adsorption of gas molecules to the surface is more likely to occur
around gas molecules that are already present on the solid surface,
rendering the Langmuir adsorption isotherm ineffective for the Two adsorbate nitrogen molecules
purposes of modelling. This effect was studied in a system where adsorbing onto a tungsten
nitrogen was the adsorbate and tungsten was the adsorbent by adsorbent from the precursor state
Paul Kisliuk (1922–2008) in 1957. [14] To compensate for the around an island of previously
increased probability of adsorption occurring around molecules adsorbed adsorbate (left) and via
random adsorption (right)
present on the substrate surface, Kisliuk developed the precursor
state theory, whereby molecules would enter a precursor state at
the interface between the solid adsorbent and adsorbate in the
gaseous phase. From here, adsorbate molecules would either adsorb to the adsorbent or desorb into
the gaseous phase. The probability of adsorption occurring from the precursor state is dependent on
the adsorbate's proximity to other adsorbate molecules that have already been adsorbed. If the
adsorbate molecule in the precursor state is in close proximity to an adsorbate molecule that has
already formed on the surface, it has a sticking probability reflected by the size of the SE constant and
will either be adsorbed from the precursor state at a rate of kEC or will desorb into the gaseous phase
at a rate of kES. If an adsorbate molecule enters the precursor state at a location that is remote from
any other previously adsorbed adsorbate molecules, the sticking probability is reflected by the size of
the SD constant.

These factors were included as part of a single constant termed a "sticking coefficient", kE, described
below:
As SD is dictated by factors that are taken into account by the Langmuir model, SD can be assumed to
be the adsorption rate constant. However, the rate constant for the Kisliuk model (R’) is different
from that of the Langmuir model, as R’ is used to represent the impact of diffusion on monolayer
formation and is proportional to the square root of the system's diffusion coefficient. The Kisliuk
adsorption isotherm is written as follows, where θ(t) is fractional coverage of the adsorbent with
adsorbate, and t is immersion time:

Solving for θ(t) yields:

Adsorption enthalpy

Adsorption constants are equilibrium constants, therefore they obey the van 't Hoff equation:

As can be seen in the formula, the variation of K must be isosteric, that is, at constant coverage.
If we
start from the BET isotherm and assume that the entropy change is the same for liquefaction and
adsorption, we obtain

that is to say, adsorption is more exothermic than liquefaction.

Single-molecule explanation

The adsorption of ensemble molecules on a surface or interface can be divided into two processes:
adsorption and desorption. If the adsorption rate wins the desorption rate, the molecules will
accumulate over time giving the adsorption curve over time. If the desorption rate is larger, the
number of molecules on the surface will decrease over time. The adsorption rate is dependent on the
temperature, the diffusion rate of the solute (related to mean free path for pure gas), and the energy
barrier between the molecule and the surface. The diffusion and key elements of the adsorption rate
can be calculated using Fick's laws of diffusion and Einstein relation (kinetic theory). Under ideal
conditions, when there is no energy barrier and all molecules that diffuse and collide with the surface
get adsorbed, the number of molecules adsorbed at a surface of area on an infinite area surface
can be directly integrated from Fick's second law differential equation to be:[15]
where is the surface area (unit m2), is the number concentration of the molecule in the bulk
solution (unit #/m3), is the diffusion constant (unit m2/s), and is time (unit s). Further
simulations and analysis of this equation [16] show that the square root dependence on the time is
originated from the decrease of the concentrations near the surface under ideal adsorption conditions.
Also, this equation only works for the beginning of the adsorption when a well-behaved concentration
gradient forms near the surface. Correction on the reduction of the adsorption area and slowing down
of the concentration gradient evolution have to be considered over a longer time.[17] Under real
experimental conditions, the flow and the small adsorption area always make the adsorption rate
faster than what this equation predicted, and the energy barrier will either accelerate this rate by
surface attraction or slow it down by surface repulsion. Thus, the prediction from this equation is
often a few to several orders of magnitude away from the experimental results. Under special cases,
such as a very small adsorption area on a large surface, and under chemical equilibrium when there is
no concentration gradience near the surface, this equation becomes useful to predict the adsorption
rate with debatable special care to determine a specific value of in a particular measurement.[16]

The desorption of a molecule from the surface depends on the binding energy of the molecule to the
surface and the temperature. The typical overall adsorption rate is thus often a combined result of the
adsorption and desorption.

Quantum mechanical – thermodynamic modelling for


surface area and porosity
Since 1980 two theories were worked on to explain adsorption and obtain equations that work. These
two are referred to as the chi hypothesis, the quantum mechanical derivation, and excess surface work
(ESW).[18] Both these theories yield the same equation for flat surfaces:

where U is the unit step function. The definitions of the other symbols is as follows:

where "ads" stands for "adsorbed", "m" stands for "monolayer equivalence" and "vap" is reference to
the vapor pressure of the liquid adsorptive at the same temperature as the solid sample. The unit
function creates the definition of the molar energy of adsorption for the first adsorbed molecule by:

The plot of adsorbed versus is referred to as the chi plot. For flat surfaces, the slope of the chi
plot yields the surface area. Empirically, this plot was noticed as being a very good fit to the isotherm
by Polanyi[19][20][21] and also by deBoer and Zwikker[22] but not pursued. This was due to criticism in
the former case by Einstein and in the latter case by Brunauer. This flat surface equation may be used
as a "standard curve" in the normal tradition of comparison curves, with the exception that the porous
sample's early portion of the plot of versus acts as a self-standard. Ultramicroporous,
microporous and mesoporous conditions may be analyzed using this technique. Typical standard
deviations for full isotherm fits including porous samples are less than 2%.
Notice that in this description of physical adsorption, the entropy of adsorption is consistent with the
Dubinin thermodynamic criterion, that is the entropy of adsorption from the liquid state to the
adsorbed state is approximately zero.

Adsorbents

Characteristics and general requirements

Adsorbents are used usually in the form of spherical pellets, rods,


moldings, or monoliths with a hydrodynamic radius between 0.25
and 5 mm. They must have high abrasion resistance, high thermal
stability and small pore diameters, which results in higher
exposed surface area and hence high capacity for adsorption. The
adsorbents must also have a distinct pore structure that enables
fast transport of the gaseous vapors.

Most industrial adsorbents fall into one of three classes: Activated carbon is used as an
adsorbent
Oxygen-containing compounds – Are typically hydrophilic and
polar, including materials such as silica gel, limestone (calcium
carbonate)[23] and zeolites.
Carbon-based compounds – Are typically hydrophobic and non-polar, including materials such as
activated carbon and graphite.
Polymer-based compounds – Are polar or non-polar, depending on the functional groups in the
polymer matrix.

Silica gel

Silica gel is a chemically inert, non-toxic, polar and dimensionally


stable (< 400  °C or 750  °F) amorphous form of SiO2. It is
prepared by the reaction between sodium silicate and acetic acid,
which is followed by a series of after-treatment processes such as
aging, pickling, etc. These after-treatment methods results in
various pore size distributions.

Silica is used for drying of process air (e.g. oxygen, natural gas)
and adsorption of heavy (polar) hydrocarbons from natural gas.
Silica gel adsorber for NO2, Fixed
Nitrogen Research Laboratory,
Zeolites ca.1930s

Zeolites are natural or synthetic crystalline aluminosilicates,


which have a repeating pore network and release water at high temperature. Zeolites are polar in
nature.

They are manufactured by hydrothermal synthesis of sodium aluminosilicate or another silica source
in an autoclave followed by ion exchange with certain cations (Na+, Li+, Ca2+, K+, NH4+). The channel
diameter of zeolite cages usually ranges from 2 to 9 Å. The ion exchange process is followed by drying
of the crystals, which can be pelletized with a binder to form macroporous pellets.

Zeolites are applied in drying of process air, CO2 removal from natural gas, CO removal from
reforming gas, air separation, catalytic cracking, and catalytic synthesis and reforming.

Non-polar (siliceous) zeolites are synthesized from aluminum-free silica sources or by dealumination
of aluminum-containing zeolites. The dealumination process is done by treating the zeolite with
steam at elevated temperatures, typically greater than 500  °C (930  °F). This high temperature heat
treatment breaks the aluminum-oxygen bonds and the aluminum atom is expelled from the zeolite
framework.

Activated carbon

Activated carbon is a highly porous, amorphous solid consisting of microcrystallites with a graphite
lattice, usually prepared in small pellets or a powder. It is non-polar and cheap. One of its main
drawbacks is that it reacts with oxygen at moderate temperatures (over 300 °C).

Activated carbon can be manufactured from carbonaceous


material, including coal (bituminous, subbituminous, and
lignite), peat, wood, or nutshells (e.g., coconut). The
manufacturing process consists of two phases, carbonization and
activation.[24][25] The carbonization process includes drying and
then heating to separate by-products, including tars and other
hydrocarbons from the raw material, as well as to drive off any
gases generated. The process is completed by heating the
material over 400 °C (750 °F) in an oxygen-free atmosphere that
cannot support combustion. The carbonized particles are then
"activated" by exposing them to an oxidizing agent, usually Activated carbon nitrogen isotherm
steam or carbon dioxide at high temperature. This agent burns showing a marked microporous type I
off the pore blocking structures created during the carbonization behavior
phase and so, they develop a porous, three-dimensional graphite
lattice structure. The size of the pores developed during
activation is a function of the time that they spend in this stage. Longer exposure times result in larger
pore sizes. The most popular aqueous phase carbons are bituminous based because of their hardness,
abrasion resistance, pore size distribution, and low cost, but their effectiveness needs to be tested in
each application to determine the optimal product.

Activated carbon is used for adsorption of organic substances[26] and non-polar adsorbates and it is
also usually used for waste gas (and waste water) treatment. It is the most widely used adsorbent
since most of its chemical (e.g. surface groups) and physical properties (e.g. pore size distribution and
surface area) can be tuned according to what is needed. Its usefulness also derives from its large
micropore (and sometimes mesopore) volume and the resulting high surface area. Recent research
works reported activated carbon as an effective agent to adsorb cationic species of toxic metals from
multi-pollutant systems and also proposed possible adsorption mechanisms with supporting
evidences.[27]

Water adsorption
The adsorption of water at surfaces is of broad importance in chemical engineering, materials science
and catalysis. Also termed surface hydration, the presence of physically or chemically adsorbed water
at the surfaces of solids plays an important role in governing interface properties, chemical reaction
pathways and catalytic performance in a wide range of systems. In the case of physically adsorbed
water, surface hydration can be eliminated simply through drying at conditions of temperature and
pressure allowing full vaporization of water. For chemically adsorbed water, hydration may be in the
form of either dissociative adsorption, where H2O molecules are dissociated into surface adsorbed -H
and -OH, or molecular adsorption (associative adsorption) where individual water molecules remain
intact [28]

Adsorption solar heating and storage


The low cost ($200/ton) and high cycle rate (2,000 ×) of synthetic zeolites such as Linde 13X with
water adsorbate has garnered much academic and commercial interest recently for use for thermal
energy storage (TES), specifically of low-grade solar and waste heat. Several pilot projects have been
funded in the EU from 2000 to the present (2020). The basic concept is to store solar thermal energy
as chemical latent energy in the zeolite. Typically, hot dry air from flat plate solar collectors is made to
flow through a bed of zeolite such that any water adsorbate present is driven off. Storage can be
diurnal, weekly, monthly, or even seasonal depending on the volume of the zeolite and the area of the
solar thermal panels. When heat is called for during the night, or sunless hours, or winter, humidified
air flows through the zeolite. As the humidity is adsorbed by the zeolite, heat is released to the air and
subsequently to the building space. This form of TES, with specific use of zeolites, was first taught by
Guerra in 1978.[29]

Carbon capture and storage


Typical adsorbents proposed for carbon capture and storage are zeolites and MOFs.[30] The
customization of adsorbents makes them a potentially attractive alternative to absorption. Because
adsorbents can be regenerated by temperature or pressure swing, this step can be less energy
intensive than absorption regeneration methods.[31] Major problems that are present with adsorption
cost in carbon capture are: regenerating the adsorbent, mass ratio, solvent/MOF, cost of adsorbent,
production of the adsorbent, lifetime of adsorbent.[32]

In sorption enhanced water gas shift (SEWGS) technology a pre-combustion carbon capture process,
based on solid adsorption, is combined with the water gas shift reaction (WGS) in order to produce a
high pressure hydrogen stream.[33] The CO2 stream produced can be stored or used for other
industrial processes.[34]

Protein and surfactant adsorption


Protein adsorption is a process that has a fundamental role in the field of biomaterials. Indeed,
biomaterial surfaces in contact with biological media, such as blood or serum, are immediately coated
by proteins. Therefore, living cells do not interact directly with the biomaterial surface, but with the
adsorbed proteins layer. This protein layer mediates the interaction between biomaterials and cells,
translating biomaterial physical and chemical properties into a "biological language".[35] In fact, cell
membrane receptors bind to protein layer bioactive sites and these receptor-protein binding events
are transduced, through the cell membrane, in a manner that stimulates specific intracellular
processes that then determine cell adhesion, shape, growth and differentiation. Protein adsorption is
influenced by many surface properties such as surface wettability, surface chemical composition [36]
and surface nanometre-scale morphology.[37]
Surfactant adsorption is a similar phenomenon, but
utilising surfactant molecules in the place of proteins.[38]

Adsorption chillers
Combining an adsorbent with a refrigerant, adsorption chillers
use heat to provide a cooling effect. This heat, in the form of hot
water, may come from any number of industrial sources including
waste heat from industrial processes, prime heat from solar
thermal installations or from the exhaust or water jacket heat of a
piston engine or turbine.

Although there are similarities between adsorption chillers and A schematic diagram of an
absorption refrigeration, the former is based on the interaction adsorption chiller: (1) heat is lost
between gases and solids. The adsorption chamber of the chiller is through evaporation of refrigerant,
filled with a solid material (for example zeolite, silica gel, alumina, (2) refrigerant vapour is adsorbed
active carbon or certain types of metal salts), which in its neutral onto the solid medium, (3)
state has adsorbed the refrigerant. When heated, the solid desorbs refrigerant is desorbed from the
(releases) refrigerant vapour, which subsequently is cooled and solid medium section not in use, (4)
liquefied. This liquid refrigerant then provides a cooling effect at refrigerant is condensed and
the evaporator from its enthalpy of vaporization. In the final stage returned to the start, (5) & (6) solid
the refrigerant vapour is (re)adsorbed into the solid.[39] As an medium is cycled between
adsorption chiller requires no compressor, it is relatively quiet. adsorption and desorption to
regenerate it.

Portal site mediated adsorption


Portal site mediated adsorption is a model for site-selective activated gas adsorption in metallic
catalytic systems that contain a variety of different adsorption sites. In such systems, low-
coordination "edge and corner" defect-like sites can exhibit significantly lower adsorption enthalpies
than high-coordination (basal plane) sites. As a result, these sites can serve as "portals" for very rapid
adsorption to the rest of the surface. The phenomenon relies on the common "spillover" effect
(described below), where certain adsorbed species exhibit high mobility on some surfaces. The model
explains seemingly inconsistent observations of gas adsorption thermodynamics and kinetics in
catalytic systems where surfaces can exist in a range of coordination structures, and it has been
successfully applied to bimetallic catalytic systems where synergistic activity is observed.

In contrast to pure spillover, portal site adsorption refers to surface diffusion to adjacent adsorption
sites, not to non-adsorptive support surfaces.

The model appears to have been first proposed for carbon monoxide on silica-supported platinum by
Brandt et al. (1993).[40] A similar, but independent model was developed by King and co-
workers[41][42][43] to describe hydrogen adsorption on silica-supported alkali promoted ruthenium,
silver-ruthenium and copper-ruthenium bimetallic catalysts. The same group applied the model to CO
hydrogenation (Fischer–Tropsch synthesis).[44] Zupanc et al. (2002) subsequently confirmed the
same model for hydrogen adsorption on magnesia-supported caesium-ruthenium bimetallic
catalysts.[45] Trens et al. (2009) have similarly described CO surface diffusion on carbon-supported Pt
particles of varying morphology.[46]
Adsorption spillover
In the case catalytic or adsorbent systems where a metal species is dispersed upon a support (or
carrier) material (often quasi-inert oxides, such as alumina or silica), it is possible for an adsorptive
species to indirectly adsorb to the support surface under conditions where such adsorption is
thermodynamically unfavorable. The presence of the metal serves as a lower-energy pathway for
gaseous species to first adsorb to the metal and then diffuse on the support surface. This is possible
because the adsorbed species attains a lower energy state once it has adsorbed to the metal, thus
lowering the activation barrier between the gas phase species and the support-adsorbed species.

Hydrogen spillover is the most common example of an adsorptive spillover. In the case of hydrogen,
adsorption is most often accompanied with dissociation of molecular hydrogen (H2) to atomic
hydrogen (H), followed by spillover of the hydrogen atoms present.

The spillover effect has been used to explain many observations in heterogeneous catalysis and
adsorption.[47]

Polymer adsorption
Adsorption of molecules onto polymer surfaces is central to a number of applications, including
development of non-stick coatings and in various biomedical devices. Polymers may also be adsorbed
to surfaces through polyelectrolyte adsorption.

Adsorption in viruses
Adsorption is the first step in the viral life cycle. The next steps are penetration, uncoating, synthesis
(transcription if needed, and translation), and release. The virus replication cycle, in this respect, is
similar for all types of viruses. Factors such as transcription may or may not be needed if the virus is
able to integrate its genomic information in the cell's nucleus, or if the virus can replicate itself
directly within the cell's cytoplasm.

In popular culture
The game of Tetris is a puzzle game in which blocks of 4 are adsorbed onto a surface during game
play. Scientists have used Tetris blocks "as a proxy for molecules with a complex shape" and their
"adsorption on a flat surface" for studying the thermodynamics of nanoparticles.[48][49]

See also
Adatom
Cryo-adsorption
Dual-polarization interferometry
Fluidized bed concentrator
Kelvin probe force microscope
Micromeritics
Molecular sieve
Polanyi adsorption
Pressure swing adsorption
Random sequential adsorption

References
1. "Glossary" (https://web.archive.org/web/20080218094403/http://www.brownfieldstsc.org/glossary.
cfm?q=1). The Brownfields and Land Revitalization Technology Support Center. Archived from the
original (http://www.brownfieldstsc.org/glossary.cfm?q=1) on 2008-02-18. Retrieved 2009-12-21.
2. "absorption (chemistry)" (https://web.archive.org/web/20181005080455/http://www.memidex.com/
absorption+chemistry). Memidex (WordNet) Dictionary/Thesaurus. Archived from the original (htt
p://www.memidex.com/absorption+chemistry) on 2018-10-05. Retrieved 2010-11-02.
3. Atkins, P. W.; De Paula, Julio; Keeler, James (2018). Atkins' Physical chemistry (Eleventh ed.).
Oxford, United Kingdom. ISBN 978-0-19-876986-6. OCLC 1020028162 (https://www.worldcat.org/
oclc/1020028162).
4. Glossary of atmospheric chemistry terms (Recommendations 1990) (http://goldbook.iupac.org/A0
0155.html). Pure and Applied Chemistry. Vol. 62. 1990. p. 2167. doi:10.1351/goldbook.A00155 (ht
tps://doi.org/10.1351%2Fgoldbook.A00155). ISBN 978-0-9678550-9-7.
5. Ferrari, L.; Kaufmann, J.; Winnefeld, F.; Plank, J. (2010). "Interaction of cement model systems
with superplasticizers investigated by atomic force microscopy, zeta potential, and adsorption
measurements". J. Colloid Interface Sci. 347 (1): 15–24. Bibcode:2010JCIS..347...15F (https://ui.a
dsabs.harvard.edu/abs/2010JCIS..347...15F). doi:10.1016/j.jcis.2010.03.005 (https://doi.org/10.10
16%2Fj.jcis.2010.03.005). PMID 20356605 (https://pubmed.ncbi.nlm.nih.gov/20356605).
6. Khosrowshahi, Mobin Safarzadeh; Abdol, Mohammad Ali; Mashhadimoslem, Hossein; Khakpour,
Elnaz; Emrooz, Hosein Banna Motejadded; Sadeghzadeh, Sadegh; Ghaemi, Ahad (26 May
2022). "The role of surface chemistry on CO2 adsorption in biomass-derived porous carbons by
experimental results and molecular dynamics simulations". Scientific Reports. 12 (1): 8917.
doi:10.1038/s41598-022-12596-5 (https://doi.org/10.1038%2Fs41598-022-12596-5).
PMID 35618757 (https://pubmed.ncbi.nlm.nih.gov/35618757). S2CID 249096513 (https://api.sem
anticscholar.org/CorpusID:249096513).
7. Czelej, K.; Cwieka, K.; Kurzydlowski, K.J. (May 2016). "CO2 stability on the Ni low-index surfaces:
Van der Waals corrected DFT analysis". Catalysis Communications. 80 (5): 33–38.
doi:10.1016/j.catcom.2016.03.017 (https://doi.org/10.1016%2Fj.catcom.2016.03.017).
8. Czelej, K.; Cwieka, K.; Colmenares, J.C.; Kurzydlowski, K.J. (2016). "Insight on the Interaction of
Methanol-Selective Oxidation Intermediates with Au- or/and Pd-Containing Monometallic and
Bimetallic Core@Shell Catalysts". Langmuir. 32 (30): 7493–7502.
doi:10.1021/acs.langmuir.6b01906 (https://doi.org/10.1021%2Facs.langmuir.6b01906).
PMID 27373791 (https://pubmed.ncbi.nlm.nih.gov/27373791).
9. Kayser, Heinrich (1881). "Über die Verdichtung von Gasen an Oberflächen in ihrer Abhängigkeit
von Druck und Temperatur" (https://books.google.com/books?id=ZxVbAAAAYAAJ&pg=PA526).
Annalen der Physik und Chemie. 248 (4): 526–537. Bibcode:1881AnP...248..526K (https://ui.adsa
bs.harvard.edu/abs/1881AnP...248..526K). doi:10.1002/andp.18812480404 (https://doi.org/10.100
2%2Fandp.18812480404).. In this study of the adsorption of gases by charcoal, the first use of
the word "adsorption" appears on page 527: "Schon Saussure kannte die beiden für die Grösse
der Adsorption massgebenden Factoren, den Druck und die Temperatur, da er Erniedrigung des
Druckes oder Erhöhung der Temperatur zur Befreiung der porösen Körper von Gasen benutzte."
("Saussaure already knew the two factors that determine the quantity of adsorption – [namely,] the
pressure and temperature – since he used the lowering of the pressure or the raising of the
temperature to free the porous substances of gases.")
10. Foo, K. Y.; Hameed, B. H. (2010). "Insights into the modeling of adsorption isotherm systems".
Chemical Engineering Journal. 156 (1): 2–10. doi:10.1016/j.cej.2009.09.013 (https://doi.org/10.10
16%2Fj.cej.2009.09.013). ISSN 1385-8947 (https://www.worldcat.org/issn/1385-8947).
S2CID 11760738 (https://api.semanticscholar.org/CorpusID:11760738).
11. Czepirski, L.; Balys, M. R.; Komorowska-Czepirska, E. (2000). "Some generalization of Langmuir
adsorption isotherm" (http://www.yumpu.com/en/document/view/5463272/article-14-some-generali
zation-of-langmuir-adsorption-isotherm-agh). Internet Journal of Chemistry. 3 (14). ISSN 1099-
8292 (https://www.worldcat.org/issn/1099-8292).
12. Burke GM, Wurster DE, Buraphacheep V, Berg MJ, Veng-Pedersen P, Schottelius DD. Model
selection for the adsorption of phenobarbital by activated charcoal. Pharm Res. 1991;8(2):228‐
231. doi:10.1023/a:1015800322286
13. Physical Chemistry of Surfaces. Arthur W. Adamson. Interscience (Wiley), New York 6th ed
14. Kisliuk, P. (January 1957). "The sticking probabilities of gases chemisorbed on the surfaces of
solids". Journal of Physics and Chemistry of Solids. 3 (1–2): 95–101.
Bibcode:1957JPCS....3...95K (https://ui.adsabs.harvard.edu/abs/1957JPCS....3...95K).
doi:10.1016/0022-3697(57)90054-9 (https://doi.org/10.1016%2F0022-3697%2857%2990054-9).
15. Langmuir, I.; Schaefer, V.J. (1937). "The Effect of Dissolved Salts on Insoluble Monolayers".
Journal of the American Chemical Society. 29 (11): 2400–2414. doi:10.1021/ja01290a091 (https://
doi.org/10.1021%2Fja01290a091).
16. Chen, Jixin (2020). "Stochastic Adsorption of Diluted Solute Molecules at Interfaces". ChemRxiv.
doi:10.26434/chemrxiv.12402404 (https://doi.org/10.26434%2Fchemrxiv.12402404).
S2CID 242860958 (https://api.semanticscholar.org/CorpusID:242860958).
17. Ward, A.F.H.; Tordai, L. (1946). "Time-dependence of Boundary Tensions of Solutions I. The Role
of Diffusion in Time-effects". Journal of Chemical Physics. 14 (7): 453–461.
Bibcode:1946JChPh..14..453W (https://ui.adsabs.harvard.edu/abs/1946JChPh..14..453W).
doi:10.1063/1.1724167 (https://doi.org/10.1063%2F1.1724167).
18. Condon, James (2020). Surface Area and Porosity Determinations by Physisorption,
Measurement, Classical Theory and Quantum Theory, 2nd edition. Amsterdam.NL: Elsevier.
pp. Chapters 3, 4 and 5. ISBN 978-0-12-818785-2.
19. Polanyi, M. (1914). "Über die Adsorption vom Standpunkt des dritten Wärmesatzes".
Verhandlungen der Deutschen Physikalischen Gesellschaft (in German). 16: 1012.
20. Polanyi, M. (1920). "Neueres über Adsorption und Ursache der Adsorptionskräfte". Zeitschrift für
Elektrochemie. 26: 370–374.
21. Polanyi, M. (1929). "Grundlagen der Potentialtheorie der Adsorption". Zeitschrift für Elektrochemie
(in German). 35: 431–432.
22. deBoer, J.H.; Zwikker, C. (1929). "Adsorption als Folge von Polarisation". Zeitschrift für
Physikalische Chemie (in German). B3: 407–420.
23. Viswambari Devi, R; Nair, Vijay V; Sathyamoorthy, P; Doble, Mukesh (2022). "Mixture of CaCO3
Polymorphs Serves as Best Adsorbent of Heavy Metals in Quadruple System". Journal of
Hazardous, Toxic, and Radioactive Waste. 26 (1). doi:10.1061/(ASCE)HZ.2153-5515.0000651 (ht
tps://doi.org/10.1061%2F%28ASCE%29HZ.2153-5515.0000651). S2CID 240454883 (https://api.s
emanticscholar.org/CorpusID:240454883).
24. Spessato, L. et al. KOH-super activated carbon from biomass waste: Insights into the
paracetamol adsorption mechanism and thermal regeneration cycles. Journal of Hazardous
Materials, Vol. 371, Pages 499-505, 2019.
25. Spessato, L. et al. Optimization of Sibipiruna activated carbon preparation by simplex-centroid
mixture design for simultaneous adsorption of rhodamine B and metformin. Journal of Hazardous
Materials, Vol. 411, Page 125166, 2021.
26. Malhotra, Milan; Suresh, Sumathi; Garg, Anurag (2018). "Tea waste derived activated carbon for
the adsorption of sodium diclofenac from wastewater: adsorbent characteristics, adsorption
isotherms, kinetics, and thermodynamics". Environmental Science and Pollution Research. 25
(32): 32210–32220. doi:10.1007/s11356-018-3148-y (https://doi.org/10.1007%2Fs11356-018-314
8-y). PMID 30221322 (https://pubmed.ncbi.nlm.nih.gov/30221322). S2CID 52280860 (https://api.s
emanticscholar.org/CorpusID:52280860).
27. Mohan, S; Nair, Vijay V (2020). "Comparative study of separation of heavy metals from leachate
using activated carbon and fuel ash". Journal of Hazardous Toxic and Radioactive Waste. 24 (4):
473–491. doi:10.1061/(ASCE)HZ.2153-5515.0000520 (https://doi.org/10.1061%2F%28ASCE%29
HZ.2153-5515.0000520). PMID 04020031 (https://pubmed.ncbi.nlm.nih.gov/04020031).
S2CID 219747988 (https://api.semanticscholar.org/CorpusID:219747988).
28. Assadi, M. Hussein N.; Hanaor, Dorian A H (June 2016). "The effects of copper doping on
photocatalytic activity at (101) planes of anatase TiO2: A theoretical study". Applied Surface
Science. 387 (387): 682–689. arXiv:1811.09157 (https://arxiv.org/abs/1811.09157).
Bibcode:2016ApSS..387..682A (https://ui.adsabs.harvard.edu/abs/2016ApSS..387..682A).
doi:10.1016/j.apsusc.2016.06.178 (https://doi.org/10.1016%2Fj.apsusc.2016.06.178).
S2CID 99834042 (https://api.semanticscholar.org/CorpusID:99834042).
29. U.S. Pat. No. 4,269,170, "Adsorption solar heating and storage"; Inventor: John M. Guerra;
Granted May 26, 1981
30. Berend, Smit; Reimer, Jeffery A; Oldenburg, Curtis M; Bourg, Ian C (2014). Introduction to carbon
capture and sequestration. Imperial College Press. ISBN 9781306496834.
31. D'Alessandro, Deanna M.; Smit, Berend; Long, Jeffrey R. (16 August 2010). "Carbon Dioxide
Capture: Prospects for New Materials" (http://infoscience.epfl.ch/record/200571). Angewandte
Chemie International Edition. 49 (35): 6058–6082. doi:10.1002/anie.201000431 (https://doi.org/1
0.1002%2Fanie.201000431). PMID 20652916 (https://pubmed.ncbi.nlm.nih.gov/20652916).
32. Sathre, Roger; Masanet, Eric (2013-03-18). "Prospective life-cycle modeling of a carbon capture
and storage system using metal–organic frameworks for CO2 capture". RSC Advances. 3 (15):
4964. Bibcode:2013RSCAd...3.4964S (https://ui.adsabs.harvard.edu/abs/2013RSCAd...3.4964S).
doi:10.1039/C3RA40265G (https://doi.org/10.1039%2FC3RA40265G). ISSN 2046-2069 (https://w
ww.worldcat.org/issn/2046-2069).
33. Jansen, Daniel; van Selow, Edward; Cobden, Paul; Manzolini, Giampaolo; Macchi, Ennio;
Gazzani, Matteo; Blom, Richard; Henriksen, Partow Pakdel; Beavis, Rich; Wright, Andrew (2013).
"SEWGS Technology is Now Ready for Scale-up!". Energy Procedia. 37: 2265–2273.
doi:10.1016/j.egypro.2013.06.107 (https://doi.org/10.1016%2Fj.egypro.2013.06.107).
34. (Eric) van Dijk, H. A. J.; Cobden, Paul D.; Lukashuk, Liliana; de Water, Leon van; Lundqvist,
Magnus; Manzolini, Giampaolo; Cormos, Calin-Cristian; van Dijk, Camiel; Mancuso, Luca; Johns,
Jeremy; Bellqvist, David (1 October 2018). "STEPWISE Project: Sorption-Enhanced Water-Gas
Shift Technology to Reduce Carbon Footprint in the Iron and Steel Industry". Johnson Matthey
Technology Review. 62 (4): 395–402. doi:10.1595/205651318X15268923666410 (https://doi.org/1
0.1595%2F205651318X15268923666410). hdl:11311/1079169 (https://hdl.handle.net/11311%2F1
079169). S2CID 139928989 (https://api.semanticscholar.org/CorpusID:139928989).
35. Wilson, CJ; Clegg, RE; Leavesley, DI; Pearcy, MJ (2005). "Mediation of Biomaterial-Cell
Interactions by Adsorbed Proteins: A Review". Tissue Engineering. 11 (1): 1–18.
doi:10.1089/ten.2005.11.1 (https://doi.org/10.1089%2Ften.2005.11.1). PMID 15738657 (https://pu
bmed.ncbi.nlm.nih.gov/15738657).
36. Sivaraman B.; Fears K.P.; Latour R.A. (2009). "Investigation of the effects of surface chemistry
and solution concentration on the conformation of adsorbed proteins using an improved circular
dichroism method" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC2891683). Langmuir. 25 (5):
3050–6. doi:10.1021/la8036814 (https://doi.org/10.1021%2Fla8036814). PMC 2891683 (https://w
ww.ncbi.nlm.nih.gov/pmc/articles/PMC2891683). PMID 19437712 (https://pubmed.ncbi.nlm.nih.go
v/19437712).
37. Scopelliti, Pasquale Emanuele; Borgonovo, Antonio; Indrieri, Marco; Giorgetti, Luca; Bongiorno,
Gero; Carbone, Roberta; Podestà, Alessandro; Milani, Paolo (2010). Zhang, Shuguang (ed.).
"The effect of surface nanometre-scale morphology on protein adsorption" (https://www.ncbi.nlm.n
ih.gov/pmc/articles/PMC2912332). PLoS ONE. 5 (7): e11862. Bibcode:2010PLoSO...511862S (htt
ps://ui.adsabs.harvard.edu/abs/2010PLoSO...511862S). doi:10.1371/journal.pone.0011862 (http
s://doi.org/10.1371%2Fjournal.pone.0011862). PMC 2912332 (https://www.ncbi.nlm.nih.gov/pmc/
articles/PMC2912332). PMID 20686681 (https://pubmed.ncbi.nlm.nih.gov/20686681).
38. Cheraghian, Goshtasp (2017). "Evaluation of Clay and Fumed Silica Nanoparticles on Adsorption
of Surfactant Polymer during Enhanced Oil Recovery" (https://doi.org/10.1627%2Fjpi.60.85).
Journal of the Japan Petroleum Institute. 60 (2): 85–94. doi:10.1627/jpi.60.85 (https://doi.org/10.1
627%2Fjpi.60.85).
39. Pilatowsky, I.; Romero, R.J.; Isaza, C.A.; Gamboa, S.A.; Sebastian, P.J.; Rivera, W. (2011).
"Chapter 5: Sorption Refrigeration Systems". Cogeneration Fuel Cell-Sorption Air Conditioning
Systems. Green Energy and Technology. Springer. pp. 99, 100. doi:10.1007/978-1-84996-028-1_5
(https://doi.org/10.1007%2F978-1-84996-028-1_5). ISBN 978-1-84996-027-4.
40. Brandt, Robert K.; Hughes, M.R.; Bourget, L.P.; Truszkowska, K.; Greenler, Robert G. (April
1993). "The interpretation of CO adsorbed on Pt/SiO2 of two different particle-size distributions".
Surface Science. 286 (1–2): 15–25. Bibcode:1993SurSc.286...15B (https://ui.adsabs.harvard.edu/
abs/1993SurSc.286...15B). doi:10.1016/0039-6028(93)90552-U (https://doi.org/10.1016%2F0039-
6028%2893%2990552-U).
41. Uner, D.O.; Savargoankar, N.; Pruski, M.; King, T.S. (1997). "The effects of alkali promoters on the
dynamics of hydrogen chemisorption and syngas reaction kinetics on Ru/SiO2 surfaces".
Dynamics of Surfaces and Reaction Kinetics in Heterogeneous Catalysis, Proceedings of the
International Symposium. Studies in Surface Science and Catalysis. Vol. 109. pp. 315–324.
doi:10.1016/s0167-2991(97)80418-1 (https://doi.org/10.1016%2Fs0167-2991%2897%2980418-
1). ISBN 9780444826091.
42. Narayan, R.L; King, T.S (March 1998). "Hydrogen adsorption states on silica-supported Ru–Ag
and Ru–Cu bimetallic catalysts investigated via microcalorimetry". Thermochimica Acta. 312 (1–
2): 105–114. doi:10.1016/S0040-6031(97)00444-9 (https://doi.org/10.1016%2FS0040-6031%289
7%2900444-9).
43. VanderWiel, David P.; Pruski, Marek; King, Terry S. (November 1999). "A Kinetic Study on the
Adsorption and Reaction of Hydrogen over Silica-Supported Ruthenium and Silver–Ruthenium
Catalysts during the Hydrogenation of Carbon Monoxide" (https://lib.dr.iastate.edu/rtd/12532).
Journal of Catalysis. 188 (1): 186–202. doi:10.1006/jcat.1999.2646 (https://doi.org/10.1006%2Fjc
at.1999.2646).
44. Uner, D. O. (1 June 1998). "A Sensible Mechanism of Alkali Promotion in Fischer−Tropsch
Synthesis: Adsorbate Mobilities". Industrial & Engineering Chemistry Research. 37 (6): 2239–
2245. doi:10.1021/ie970696d (https://doi.org/10.1021%2Fie970696d).
45. Zupanc, C.; Hornung, A.; Hinrichsen, O.; Muhler, M. (July 2002). "The Interaction of Hydrogen
with Ru/MgO Catalysts". Journal of Catalysis. 209 (2): 501–514. doi:10.1006/jcat.2002.3647 (http
s://doi.org/10.1006%2Fjcat.2002.3647).
46. Trens, Philippe; Durand, Robert; Coq, Bernard; Coutanceau, Christophe; Rousseau, Séverine;
Lamy, Claude (November 2009). "Poisoning of Pt/C catalysts by CO and its consequences over
the kinetics of hydrogen chemisorption". Applied Catalysis B: Environmental. 92 (3–4): 280–284.
doi:10.1016/j.apcatb.2009.08.004 (https://doi.org/10.1016%2Fj.apcatb.2009.08.004).
47. Rozanov, Valerii V; Krylov, Oleg V (28 February 1997). "Hydrogen spillover in heterogeneous
catalysis". Russian Chemical Reviews. 66 (2): 107–119. Bibcode:1997RuCRv..66..107R (https://u
i.adsabs.harvard.edu/abs/1997RuCRv..66..107R). doi:10.1070/rc1997v066n02abeh000308 (http
s://doi.org/10.1070%2Frc1997v066n02abeh000308).
48. Ford, Matt (6 May 2009). "The thermodynamics of Tetris" (https://arstechnica.com/science/2009/0
5/the-thermodynamics-of-tetris/). Ars Technica.
49. Barnes, Brian C.; Siderius, Daniel W.; Gelb, Lev D. (2009). "Structure, Thermodynamics, and
Solubility in Tetromino Fluids" (https://doi.org/10.1021%2Fla900196b). Langmuir. 25 (12): 6702–
16. doi:10.1021/la900196b (https://doi.org/10.1021%2Fla900196b). PMID 19397254 (https://pubm
ed.ncbi.nlm.nih.gov/19397254).

Further reading
Cussler, E. L. (1997). Diffusion: Mass Transfer in Fluid Systems (2nd ed.). New York: Cambridge
University Press. pp. 308–330. ISBN 978-0-521-45078-2.

External links
Derivation of Langmuir and BET isotherms (http://www.jhu.edu/%7Echem/fairbr/OLDS/derive.htm
l), at JHU.edu
Carbon Adsorption (https://web.archive.org/web/20090105233739/http://www.megtec.com/solvent
-recovery-carbon-adsorption-p-685-l-en.html), at MEGTEC.com

Retrieved from "https://en.wikipedia.org/w/index.php?title=Adsorption&oldid=1096359128"

This page was last edited on 4 July 2022, at 00:14 (UTC).

Text is available under the Creative Commons Attribution-ShareAlike License 3.0;


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