Dissociation constant

Last updated

In chemistry, biochemistry, and pharmacology, a dissociation constant () is a specific type of equilibrium constant that measures the propensity of a larger object to separate (dissociate) reversibly into smaller components, as when a complex falls apart into its component molecules, or when a salt splits up into its component ions. The dissociation constant is the inverse of the association constant. In the special case of salts, the dissociation constant can also be called an ionization constant. [1] [2] For a general reaction:

Contents

in which a complex breaks down into xA subunits and yB subunits, the dissociation constant is defined as

where [A], [B], and [AxBy] are the equilibrium concentrations of A, B, and the complex AxBy, respectively.

One reason for the popularity of the dissociation constant in biochemistry and pharmacology is that in the frequently encountered case where x = y = 1, KD has a simple physical interpretation: when , then or equivalently . That is, KD, which has the dimensions of concentration, equals the concentration of free A at which half of the total molecules of B are associated with A. This simple interpretation does not apply for higher values of x or y. It also presumes the absence of competing reactions, though the derivation can be extended to explicitly allow for and describe competitive binding.[ citation needed ] It is useful as a quick description of the binding of a substance, in the same way that EC50 and IC50 describe the biological activities of substances.

Concentration of bound molecules

Molecules with one binding site

Experimentally, the concentration of the molecule complex [AB] is obtained indirectly from the measurement of the concentration of a free molecules, either [A] or [B]. [3] In principle, the total amounts of molecule [A]0 and [B]0 added to the reaction are known. They separate into free and bound components according to the mass conservation principle:

To track the concentration of the complex [AB], one substitutes the concentration of the free molecules ([A] or [B]), of the respective conservation equations, by the definition of the dissociation constant,

This yields the concentration of the complex related to the concentration of either one of the free molecules

Macromolecules with identical independent binding sites

Many biological proteins and enzymes can possess more than one binding site. [3] Usually, when a ligand L binds with a macromolecule M, it can influence binding kinetics of other ligands L binding to the macromolecule. A simplified mechanism can be formulated if the affinity of all binding sites can be considered independent of the number of ligands bound to the macromolecule. This is valid for macromolecules composed of more than one, mostly identical, subunits. It can be then assumed that each of these n subunits are identical, symmetric and that they possess only one single binding site. Then, the concentration of bound ligands becomes

In this case, , but comprises all partially saturated forms of the macromolecule:

where the saturation occurs stepwise

For the derivation of the general binding equation a saturation function is defined as the quotient from the portion of bound ligand to the total amount of the macromolecule:

K’n are so-called macroscopic or apparent dissociation constants and can result from multiple individual reactions. For example, if a macromolecule M has 3 binding sites, K’1 describes a ligand being bound to any of the 3 binding sites. In this example, K’2 describes 2 molecules being bound and K’3 3 molecules being bound to the macromolecule. The microscopic or individual dissociation constant describes the equilibrium of ligands binding to specific binding sites. Because we assume identical binding sites with no cooperativity, the microscopic dissociation constant must be equal for every binding site and can be abbreviated simply as KD. In our example, K’1 is the amalgamation of a ligand binding to either of the 3 possible binding sites (I, II and III), hence 3 microscopic dissociation constants and 3 distinct states of the ligand-macromolecule complex. For K’2 there are 6 different microscopic dissociation constants (I-II, I-III, II-I, II-III, III-I, III-II) but only 3 distinct states (it does not matter whether you bind pocket I and II or first II and then I). For K’3 there are 3 different dissociation constants - there are only three possibilities to which pocket is filled last (I, II or III) - and 1 state (I-II-III).

Even when the microscopic dissociation constant is the same for each individual binding event, the macroscopic outcome (K’1, K’2 and K’3) is not equal. This can be understood intuitively for our example of 3 possible binding sites. K’1 describes the reaction from one state (no ligand bound) to 3 states (1 ligand bound to either of the 3 binding sides). The apparent K’1 would therefore be 3 times smaller than the individual KD. K’2 describes the reaction from 3 states (1 ligand bound) to 3 states (2 ligands bound), therefore K’2 would be equal to KD. K’3 describes the reaction from 3 states (2 ligands bound) to 1 state (3 ligands bound), hence the apparent dissociation constant K’3 being 3 times bigger than the microscopic dissociation constant KD. The general relationship between both types of dissociation constants for n binding sites is

Hence, the ratio of bound ligand to macromolecules becomes

where is the binomial coefficient. Then, the first equation is proved by applying the binomial rule

Protein-ligand binding

The dissociation constant is commonly used to describe the affinity between a ligand (such as a drug) and a protein ; i.e., how tightly a ligand binds to a particular protein. Ligand-protein affinities are influenced by non-covalent intermolecular interactions between the two molecules such as hydrogen bonding, electrostatic interactions, hydrophobic and van der Waals forces. Affinities can also be affected by high concentrations of other macromolecules, which causes macromolecular crowding. [4] [5]

The formation of a ligand-protein complex can be described by a two-state process

the corresponding dissociation constant is defined

where , and represent molar concentrations of the protein, ligand, and protein-ligand complex, respectively.

The dissociation constant has molar units (M) and corresponds to the ligand concentration at which half of the proteins are occupied at equilibrium, [6] i.e., the concentration of ligand at which the concentration of protein with ligand bound equals the concentration of protein with no ligand bound . The smaller the dissociation constant, the more tightly bound the ligand is, or the higher the affinity between ligand and protein. For example, a ligand with a nanomolar (nM) dissociation constant binds more tightly to a particular protein than a ligand with a micromolar (μM) dissociation constant.

Sub-picomolar dissociation constants as a result of non-covalent binding interactions between two molecules are rare. Nevertheless, there are some important exceptions. Biotin and avidin bind with a dissociation constant of roughly 10−15 M = 1 fM = 0.000001 nM. [7] Ribonuclease inhibitor proteins may also bind to ribonuclease with a similar 10−15 M affinity. [8] The dissociation constant for a particular ligand-protein interaction can change significantly with solution conditions (e.g., temperature, pH and salt concentration). The effect of different solution conditions is to effectively modify the strength of any intermolecular interactions holding a particular ligand-protein complex together.

Drugs can produce harmful side effects through interactions with proteins for which they were not meant to or designed to interact. Therefore, much pharmaceutical research is aimed at designing drugs that bind to only their target proteins (Negative Design) with high affinity (typically 0.1-10 nM) or at improving the affinity between a particular drug and its in-vivo protein target (Positive Design).

Antibodies

In the specific case of antibodies (Ab) binding to antigen (Ag), usually the term affinity constant refers to the association constant.

This chemical equilibrium is also the ratio of the on-rate (kforward or ka) and off-rate (kback or kd) constants. Two antibodies can have the same affinity, but one may have both a high on- and off-rate constant, while the other may have both a low on- and off-rate constant.

Acid–base reactions

For the deprotonation of acids, K is known as Ka, the acid dissociation constant. Stronger acids, for example sulfuric or phosphoric acid, have larger dissociation constants; weaker acids, like acetic acid, have smaller dissociation constants.

(The symbol , used for the acid dissociation constant, can lead to confusion with the association constant and it may be necessary to see the reaction or the equilibrium expression to know which is meant.)

Acid dissociation constants are sometimes expressed by , which is defined as:

This notation is seen in other contexts as well; it is mainly used for covalent dissociations (i.e., reactions in which chemical bonds are made or broken) since such dissociation constants can vary greatly.

A molecule can have several acid dissociation constants. In this regard, that is depending on the number of the protons they can give up, we define monoprotic, diprotic and triprotic acids. The first (e.g., acetic acid or ammonium) have only one dissociable group, the second (carbonic acid, bicarbonate, glycine) have two dissociable groups and the third (e.g., phosphoric acid) have three dissociable groups. In the case of multiple pK values they are designated by indices: pK1, pK2, pK3 and so on. For amino acids, the pK1 constant refers to its carboxyl (-COOH) group, pK2 refers to its amino (-NH2) group and the pK3 is the pK value of its side chain.

Dissociation constant of water

The dissociation constant of water is denoted Kw:

The concentration of water, [H2O], is omitted by convention, which means that the value of Kw differs from the value of Keq that would be computed using that concentration.

The value of Kw varies with temperature, as shown in the table below. This variation must be taken into account when making precise measurements of quantities such as pH.

Water temperatureKwpKw [9]
0 °C0.112×10−1414.95
25 °C1.023×10−1413.99
50 °C5.495×10−1413.26
75 °C19.95×10−1412.70
100 °C56.23×10−1412.25

See also

Related Research Articles

In chemistry, an acid dissociation constant is a quantitative measure of the strength of an acid in solution. It is the equilibrium constant for a chemical reaction

Cooperative binding occurs in molecular binding systems containing more than one type, or species, of molecule and in which one of the partners is not mono-valent and can bind more than one molecule of the other species. In general, molecular binding is an interaction between molecules that results in a stable physical association between those molecules.

The equilibrium constant of a chemical reaction is the value of its reaction quotient at chemical equilibrium, a state approached by a dynamic chemical system after sufficient time has elapsed at which its composition has no measurable tendency towards further change. For a given set of reaction conditions, the equilibrium constant is independent of the initial analytical concentrations of the reactant and product species in the mixture. Thus, given the initial composition of a system, known equilibrium constant values can be used to determine the composition of the system at equilibrium. However, reaction parameters like temperature, solvent, and ionic strength may all influence the value of the equilibrium constant.

<span class="mw-page-title-main">Dissociation (chemistry)</span> Separation of molecules or ionic compounds into smaller constituent entities

Dissociation in chemistry is a general process in which molecules (or ionic compounds such as salts, or complexes) separate or split into other things such as atoms, ions, or radicals, usually in a reversible manner. For instance, when an acid dissolves in water, a covalent bond between an electronegative atom and a hydrogen atom is broken by heterolytic fission, which gives a proton (H+) and a negative ion. Dissociation is the opposite of association or recombination.

<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.

In particle physics, the quantum yield of a radiation-induced process is the number of times a specific event occurs per photon absorbed by the system.

<span class="mw-page-title-main">Atom transfer radical polymerization</span>

Atom transfer radical polymerization (ATRP) is an example of a reversible-deactivation radical polymerization. Like its counterpart, ATRA, or atom transfer radical addition, ATRP is a means of forming a carbon-carbon bond with a transition metal catalyst. Polymerization from this method is called atom transfer radical addition polymerization (ATRAP). As the name implies, the atom transfer step is crucial in the reaction responsible for uniform polymer chain growth. ATRP was independently discovered by Mitsuo Sawamoto and by Krzysztof Matyjaszewski and Jin-Shan Wang in 1995.

The Scatchard equation is an equation used in molecular biology to calculate the affinity and number of binding sites of a receptor for a ligand. It is named after the American chemist George Scatchard.

In biochemistry, avidity refers to the accumulated strength of multiple affinities of individual non-covalent binding interactions, such as between a protein receptor and its ligand, and is commonly referred to as functional affinity. Avidity differs from affinity, which describes the strength of a single interaction. However, because individual binding events increase the likelihood of occurrence of other interactions, avidity should not be thought of as the mere sum of its constituent affinities but as the combined effect of all affinities participating in the biomolecular interaction. A particular important aspect relates to the phenomenon of 'avidity entropy'. Biomolecules often form heterogenous complexes or homogeneous oligomers and multimers or polymers. If clustered proteins form an organized matrix, such as the clathrin-coat, the interaction is described as a matricity.

In biochemistry, receptor–ligand kinetics is a branch of chemical kinetics in which the kinetic species are defined by different non-covalent bindings and/or conformations of the molecules involved, which are denoted as receptor(s) and ligand(s). Receptor–ligand binding kinetics also involves the on- and off-rates of binding.

Equilibrium constants are determined in order to quantify chemical equilibria. When an equilibrium constant K is expressed as a concentration quotient,

Molecular binding is an attractive interaction between two molecules that results in a stable association in which the molecules are in close proximity to each other. It is formed when atoms or molecules bind together by sharing of electrons. It often, but not always, involves some chemical bonding.

In coordination chemistry, a stability constant is an equilibrium constant for the formation of a complex in solution. It is a measure of the strength of the interaction between the reagents that come together to form the complex. There are two main kinds of complex: compounds formed by the interaction of a metal ion with a ligand and supramolecular complexes, such as host–guest complexes and complexes of anions. The stability constant(s) provide(s) the information required to calculate the concentration(s) of the complex(es) in solution. There are many areas of application in chemistry, biology and medicine.

In chemistry, dissociative substitution describes a reaction pathway by which compounds interchange ligands. The term is typically applied to coordination and organometallic complexes, but resembles the SN1 mechanism in organic chemistry. This pathway can be well described by the cis effect, or the labilization of CO ligands in the cis position. The opposite pathway is associative substitution, being analogous to SN2 pathway. Pathways that are intermediate between the pure dissociative and pure associative pathways are called interchange mechanisms.

In chemistry, binding selectivity is defined with respect to the binding of ligands to a substrate forming a complex. Binding selectivity describes how a ligand may bind more preferentially to one receptor than another. A selectivity coefficient is the equilibrium constant for the reaction of displacement by one ligand of another ligand in a complex with the substrate. Binding selectivity is of major importance in biochemistry and in chemical separation processes.

Equilibrium chemistry is concerned with systems in chemical equilibrium. The unifying principle is that the free energy of a system at equilibrium is the minimum possible, so that the slope of the free energy with respect to the reaction coordinate is zero. This principle, applied to mixtures at equilibrium provides a definition of an equilibrium constant. Applications include acid–base, host–guest, metal–complex, solubility, partition, chromatography and redox equilibria.

Antigen-antibody interaction, or antigen-antibody reaction, is a specific chemical interaction between antibodies produced by B cells of the white blood cells and antigens during immune reaction. The antigens and antibodies combine by a process called agglutination. It is the fundamental reaction in the body by which the body is protected from complex foreign molecules, such as pathogens and their chemical toxins. In the blood, the antigens are specifically and with high affinity bound by antibodies to form an antigen-antibody complex. The immune complex is then transported to cellular systems where it can be destroyed or deactivated.

<span class="mw-page-title-main">TNP-ATP</span>

TNP-ATP is a fluorescent molecule that is able to determine whether a protein binds to ATP, and the constants associated with that binding. It is primarily used in fluorescence spectroscopy, but is also very useful as an acceptor molecule in FRET, and as a fluorescent probe in fluorescence microscopy and X-ray crystallography.

<span class="mw-page-title-main">Protein–ligand complex</span>

A protein–ligand complex is a complex of a protein bound with a ligand that is formed following molecular recognition between proteins that interact with each other or with other molecules. Formation of a protein-ligand complex is based on molecular recognition between biological macromolecules and ligands, where ligand means any molecule that binds the protein with high affinity and specificity. Molecular recognition is not a process by itself since it is part of a functionally important mechanism involving the essential elements of life like in self-replication, metabolism, and information processing. For example DNA-replication depends on recognition and binding of DNA double helix by helicase, DNA single strand by DNA-polymerase and DNA segments by ligase. Molecular recognition depends on affinity and specificity. Specificity means that proteins distinguish the highly specific binding partner from less specific partners and affinity allows the specific partner with high affinity to remain bound even if there are high concentrations of less specific partners with lower affinity.

<span class="mw-page-title-main">Competitive inhibition</span> Interruption of a chemical pathway

Competitive inhibition is interruption of a chemical pathway owing to one chemical substance inhibiting the effect of another by competing with it for binding or bonding. Any metabolic or chemical messenger system can potentially be affected by this principle, but several classes of competitive inhibition are especially important in biochemistry and medicine, including the competitive form of enzyme inhibition, the competitive form of receptor antagonism, the competitive form of antimetabolite activity, and the competitive form of poisoning.

References

  1. "Dissociation Constant". Chemistry LibreTexts. 2015-08-09. Retrieved 2020-10-26.
  2. Bioanalytical Chemistry Textbook De Gruyter 2021 https://doi.org/10.1515/9783110589160-206
  3. 1 2 Bisswanger, Hans (2008). Enzyme Kinetics: Principles and Methods (PDF). Weinheim: Wiley-VCH. p. 302. ISBN   978-3-527-31957-2.
  4. Zhou, H.; Rivas, G.; Minton, A. (2008). "Macromolecular crowding and confinement: biochemical, biophysical, and potential physiological consequences". Annual Review of Biophysics. 37: 375–397. doi:10.1146/annurev.biophys.37.032807.125817. PMC   2826134 . PMID   18573087.
  5. Minton, A. P. (2001). "The influence of macromolecular crowding and macromolecular confinement on biochemical reactions in physiological media" (PDF). The Journal of Biological Chemistry. 276 (14): 10577–10580. doi: 10.1074/jbc.R100005200 . PMID   11279227.
  6. Björkelund, Hanna; Gedda, Lars; Andersson, Karl (2011-01-31). "Comparing the Epidermal Growth Factor Interaction with Four Different Cell Lines: Intriguing Effects Imply Strong Dependency of Cellular Context". PLOS ONE. 6 (1): e16536. Bibcode:2011PLoSO...616536B. doi: 10.1371/journal.pone.0016536 . ISSN   1932-6203. PMC   3031572 . PMID   21304974.
  7. Livnah, O.; Bayer, E.; Wilchek, M.; Sussman, J. (1993). "Three-dimensional structures of avidin and the avidin-biotin complex". Proceedings of the National Academy of Sciences of the United States of America. 90 (11): 5076–5080. Bibcode:1993PNAS...90.5076L. doi: 10.1073/pnas.90.11.5076 . PMC   46657 . PMID   8506353.
  8. Johnson, R.; Mccoy, J.; Bingman, C.; Phillips Gn, J.; Raines, R. (2007). "Inhibition of human pancreatic ribonuclease by the human ribonuclease inhibitor protein". Journal of Molecular Biology. 368 (2): 434–449. doi:10.1016/j.jmb.2007.02.005. PMC   1993901 . PMID   17350650.
  9. Bandura, Andrei V.; Lvov, Serguei N. (2006). "The Ionization Constant of Water over Wide Ranges of Temperature and Density" (PDF). Journal of Physical and Chemical Reference Data. 35 (1): 15–30. Bibcode:2006JPCRD..35...15B. doi:10.1063/1.1928231.