Jump to content
 







Main menu
   


Navigation  



Main page
Contents
Current events
Random article
About Wikipedia
Contact us
Donate
 




Contribute  



Help
Learn to edit
Community portal
Recent changes
Upload file
 








Search  

































Create account

Log in
 









Create account
 Log in
 




Pages for logged out editors learn more  



Contributions
Talk
 



















Contents

   



(Top)
 


1 Concentration of bound molecules  



1.1  Molecules with one binding site  





1.2  Macromolecules with identical independent binding sites  







2 Proteinligand binding  



2.1  Antibodies  







3 Acidbase reactions  





4 Dissociation constant of water  





5 See also  





6 References  














Dissociation constant






العربية
Беларуская
Català
Čeština
Deutsch
Eesti
Español
فارسی
Français
Galego

Հայերեն
ि
Bahasa Indonesia
Italiano
עברית
Lietuvių

Norsk bokmål
Polski
Português
Română
Русский
Саха тыла
Српски / srpski
Srpskohrvatski / српскохрватски
Svenska
Українська

 

Edit links
 









Article
Talk
 

















Read
Edit
View history
 








Tools
   


Actions  



Read
Edit
View history
 




General  



What links here
Related changes
Upload file
Special pages
Permanent link
Page information
Cite this page
Get shortened URL
Download QR code
Wikidata item
 




Print/export  



Download as PDF
Printable version
 
















Appearance
   

 






From Wikipedia, the free encyclopedia
 


Inchemistry, biochemistry, and pharmacology, a dissociation constant (KD) 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:

in which a complex breaks down into x A subunits and y B subunits, the dissociation constant is defined as

where [A], [B], and [Ax By] are the equilibrium concentrations of A, B, and the complex Ax By, 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 [A] = KD, then [B] = [AB] 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 xory. 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[edit]

Molecules with one binding site[edit]

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[edit]

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 a 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:

Kn are so-called macroscopic or apparent dissociation constants and can result from multiple individual reactions. For example, if a macromolecule M has three binding sites, K1 describes a ligand being bound to any of the three binding sites. In this example, K2 describes two molecules being bound and K3 three 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, K1 is the amalgamation of a ligand binding to either of the three possible binding sites (I, II and III), hence three microscopic dissociation constants and three distinct states of the ligand–macromolecule complex. For K2 there are six different microscopic dissociation constants (I–II, I–III, II–I, II–III, III–I, III–II) but only three distinct states (it does not matter whether you bind pocket I first and then II or II first and then I). For K3 there are three different dissociation constants — there are only three possibilities for which pocket is filled last (I, II or III) — and one state (I–II–III).

Even when the microscopic dissociation constant is the same for each individual binding event, the macroscopic outcome (K1, K2 and K3) is not equal. This can be understood intuitively for our example of three possible binding sites. K1 describes the reaction from one state (no ligand bound) to three states (one ligand bound to either of the three binding sides). The apparent K1 would therefore be three times smaller than the individual KD. K2 describes the reaction from three states (one ligand bound) to three states (two ligands bound); therefore, K2 would be equal to KD. K3 describes the reaction from three states (two ligands bound) to one state (three ligands bound); hence, the apparent dissociation constant K3 is three 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[edit]

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 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[edit]

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 (kforwardorka) and off-rate (kbackorkd) 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[edit]

For the deprotonationofacids, K is known as Ka, the acid dissociation constant. Strong acids, such as sulfuricorphosphoric acid, have large dissociation constants; weak acids, such as acetic acid, have small dissociation constants.

The symbol Ka, 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 pKa, which is defined by

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 (e.g., 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[edit]

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 temperature Kw pKw[9]
000 °C 00.112×10−14 14.95
025 °C 01.023×10−14 13.99
050 °C 05.495×10−14 13.26
075 °C 19.950×10−14 12.70
100 °C 56.230×10−14 12.25

See also[edit]

References[edit]

  1. ^ "Dissociation Constant". Chemistry LibreTexts. 2015-08-09. Retrieved 2020-10-26.
  • ^ Bioanalytical Chemistry Textbook De Gruyter 2021 https://doi.org/10.1515/9783110589160-206
  • ^ a b Bisswanger, Hans (2008). Enzyme Kinetics: Principles and Methods (PDF). Weinheim: Wiley-VCH. p. 302. ISBN 978-3-527-31957-2.
  • ^ 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.
  • ^ 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.
  • ^ 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.
  • ^ 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.
  • ^ 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.
  • ^ 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. Archived from the original (PDF) on 2013-05-12. Retrieved 2017-07-13.

  • Retrieved from "https://en.wikipedia.org/w/index.php?title=Dissociation_constant&oldid=1210956376"

    Categories: 
    Equilibrium chemistry
    Enzyme kinetics
    Hidden categories: 
    Articles with short description
    Short description matches Wikidata
    All articles with unsourced statements
    Articles with unsourced statements from June 2016
    Articles with GND identifiers
    Pages that use a deprecated format of the chem tags
     



    This page was last edited on 29 February 2024, at 03:51 (UTC).

    Text is available under the Creative Commons Attribution-ShareAlike License 4.0; 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.



    Privacy policy

    About Wikipedia

    Disclaimers

    Contact Wikipedia

    Code of Conduct

    Developers

    Statistics

    Cookie statement

    Mobile view



    Wikimedia Foundation
    Powered by MediaWiki