Enzyme inhibitor

Source: Wikipedia, the free encyclopedia.
(Redirected from
Enzyme inhibitors
)

substrates (S) to products
(P). Bottom: by binding to the enzyme, inhibitor (I) blocks binding of substrate. Binding site shown in blue checkerboard, substrate as black rectangle, and inhibitor as green rounded rectangle.

An enzyme inhibitor is a

substrate molecules are converted into products.[1] An enzyme facilitates a specific chemical reaction by binding the substrate to its active site, a specialized area on the enzyme that accelerates the most difficult step of the reaction
.

An enzyme inhibitor stops ("inhibits") this process, either by binding to the enzyme's active site (thus preventing the substrate itself from binding) or by binding to another site on the enzyme such that the enzyme's

non-covalently
and may spontaneously leave the enzyme, allowing the enzyme to resume its function. Reversible inhibitors produce different types of inhibition depending on whether they bind to the enzyme, the enzyme-substrate complex, or both.

Enzyme inhibitors play an important role in all cells, since they are generally specific to one enzyme each and serve to control that enzyme's activity. For example, enzymes in a

predators
.

Many

specific and generally produce few side effects in humans, provided that no analogous enzyme is found in humans. (This is often the case, since such pathogens and humans are genetically distant.) Medicinal enzyme inhibitors often have low dissociation constants, meaning that only a minute amount of the inhibitor is required to inhibit the enzyme. A low concentration of the enzyme inhibitor reduces the risk for liver and kidney damage and other adverse drug reactions in humans. Hence the discovery and refinement of enzyme inhibitors is an active area of research in biochemistry and pharmacology
.

Structural classes

Enzyme inhibitors are a chemically diverse set of substances that range in size from organic small molecules to macromolecular proteins.

Small molecule inhibitors include essential primary metabolites that inhibit upstream enzymes that produce those metabolites. This provides a negative feedback loop that prevents over production of metabolites and thus maintains cellular homeostasis (steady internal conditions).[3][2] Small molecule enzyme inhibitors also include secondary metabolites, which are not essential to the organism that produces them, but provide the organism with an evolutionary advantage, in that they can be used to repel predators or competing organisms or immobilize prey.[4] In addition, many drugs are small molecule enzyme inhibitors that target either disease-modifying enzymes in the patient[1]: 5  or enzymes in pathogens which are required for the growth and reproduction of the pathogen.[5]

In addition to small molecules, some proteins act as enzyme inhibitors. The most prominent example are

N-terminal peptide that binds to the active site of enzyme that intramolecularly blocks its activity as a protective mechanism against uncontrolled catalysis. The N‑terminal peptide is cleaved (split) from the zymogen enzyme precursor by another enzyme to release an active enzyme.[8]

The

allosteric ("alternative" orientation) inhibitors.[9] The mechanisms of allosteric inhibition are varied and include changing the conformation (shape) of the enzyme such that it can no longer bind substrate (kinetically indistinguishable from competitive orthosteric inhibition)[10] or alternatively stabilise binding of substrate to the enzyme but lock the enzyme in a conformation which is no longer catalytically active.[11]

Reversible inhibitors

Inhibition mechanism schematic
chemical equilibrium reaction formula for competitive, uncompetitive, non-competitive, and mixed inhibition
Kinetic mechanisms for reversible inhibition. Substrate (S) binding to enzyme (E) in blue, catalysis releasing product (P) in red, inhibitor (I) binding to enzyme in green.
schematic diagram of the three types of reversible inhibitors
Schematics for reversible inhibition. Binding site in blue, substrate in black, inhibitor in green, and allosteric site in light green.
Competitive inhibitors usually bind to the active site. Non-competitive bind to a remote (allosteric) site. Uncompetitive inhibitors only bind once the substrate is bound, fully disrupting catalysis, and mixed inhibition is similar but with only partial disruption of catalysis.

Reversible inhibitors attach to enzymes with

ionic bonds.[12] Multiple weak bonds between the inhibitor and the enzyme active site
combine to produce strong and specific binding.

In contrast to irreversible inhibitors, reversible inhibitors generally do not undergo chemical reactions when bound to the enzyme and can be easily removed by dilution or dialysis. A special case is covalent reversible inhibitors that form a chemical bond with the enzyme, but the bond can be cleaved so the inhibition is fully reversible.[13]

Reversible inhibitors are generally categorized into four types, as introduced by Cleland in 1963.[14] They are classified according to the effect of the inhibitor on the Vmax (maximum reaction rate catalysed by the enzyme) and Km (the concentration of substrate resulting in half maximal enzyme activity) as the concentration of the enzyme's substrate is varied.[15][16]

Competitive

In competitive inhibition the substrate and inhibitor cannot bind to the enzyme at the same time.[17]: 134  This usually results from the inhibitor having an affinity for the active site of an enzyme where the substrate also binds; the substrate and inhibitor compete for access to the enzyme's active site. This type of inhibition can be overcome by sufficiently high concentrations of substrate (Vmax remains constant), i.e., by out-competing the inhibitor.[17]: 134–135  However, the apparent Km will increase as it takes a higher concentration of the substrate to reach the Km point, or half the Vmax. Competitive inhibitors are often similar in structure to the real substrate (see for example the "methotrexate versus folate" figure in the "Drugs" section).[17]: 134 

Uncompetitive

In uncompetitive inhibition the inhibitor binds only to the enzyme-substrate complex.[17]: 139  This type of inhibition causes Vmax to decrease (maximum velocity decreases as a result of removing activated complex) and Km to decrease (due to better binding efficiency as a result of Le Chatelier's principle and the effective elimination of the ES complex thus decreasing the Km which indicates a higher binding affinity).[18] Uncompetitive inhibition is rare.[17]: 139 [19]

Non-competitive

In

activity but does not affect the binding of substrate.[16] This type of inhibitor binds with equal affinity to the free enzyme as to the enzyme-substrate complex. It can be thought of as having the ability of competitive and uncompetitive inhibitors, but with no preference to either type. As a result, the extent of inhibition depends only on the concentration of the inhibitor. Vmax will decrease due to the inability for the reaction to proceed as efficiently, but Km will remain the same as the actual binding of the substrate, by definition, will still function properly.[20]

Mixed

In

tertiary structure or three-dimensional shape) of the enzyme so that the affinity of the substrate for the active site is reduced.[22]

These four types of inhibition can also be distinguished by the effect of increasing the substrate concentration [S] on the degree of inhibition caused by a given amount of inhibitor. For competitive inhibition the degree of inhibition is reduced by increasing [S], for noncompetitive inhibition the degree of inhibition is unchanged, and for uncompetitive (also called anticompetitive) inhibition the degree of inhibition increases with [S].[23]

Quantitative description

Reversible inhibition can be described quantitatively in terms of the inhibitor's

Michaelis-Menten scheme (shown in the "inhibition mechanism schematic" diagram), an enzyme (E) binds to its substrate (S) to form the enzyme–substrate complex ES. Upon catalysis, this complex breaks down to release product P and free enzyme.[24]: 55  The inhibitor (I) can bind to either E or ES with the dissociation constants Ki or Ki', respectively.[24]
: 87 

When an enzyme has multiple substrates, inhibitors can show different types of inhibition depending on which substrate is considered. This results from the active site containing two different binding sites within the active site, one for each substrate. For example, an inhibitor might compete with substrate A for the first binding site, but be a non-competitive inhibitor with respect to substrate B in the second binding site.[26]

Traditionally reversible enzyme inhibitors have been classified as competitive, uncompetitive, or non-competitive, according to their effects on Km and Vmax.[14] These three types of inhibition result respectively from the inhibitor binding only to the enzyme E in the absence of substrate S, to the enzyme–substrate complex ES, or to both. The division of these classes arises from a problem in their derivation and results in the need to use two different binding constants for one binding event.[27] It is further assumed that binding of the inhibitor to the enzyme results in 100% inhibition and fails to consider the possibility of partial inhibition.[27] The common form of the inhibitory term also obscures the relationship between the inhibitor binding to the enzyme and its relationship to any other binding term be it the Michaelis–Menten equation or a dose response curve associated with ligand receptor binding. To demonstrate the relationship the following rearrangement can be made:[28]

This rearrangement demonstrates that similar to the Michaelis–Menten equation, the maximal rate of reaction depends on the proportion of the enzyme population interacting with its substrate.

fraction of the enzyme population bound by substrate

fraction of the enzyme population bound by inhibitor

the effect of the inhibitor is a result of the percent of the enzyme population interacting with inhibitor. The only problem with this equation in its present form is that it assumes absolute inhibition of the enzyme with inhibitor binding, when in fact there can be a wide range of effects anywhere from 100% inhibition of substrate turn over to no inhibition. To account for this the equation can be easily modified to allow for different degrees of inhibition by including a delta Vmax term.[29]: 361 

or

This term can then define the residual enzymatic activity present when the inhibitor is interacting with individual enzymes in the population. However the inclusion of this term has the added value of allowing for the possibility of activation if the secondary Vmax term turns out to be higher than the initial term. To account for the possibly of activation as well the notation can then be rewritten replacing the inhibitor "I" with a modifier term (stimulator or inhibitor) denoted here as "X".[28]: eq 13 

While this terminology results in a simplified way of dealing with kinetic effects relating to the maximum velocity of the Michaelis–Menten equation, it highlights potential problems with the term used to describe effects relating to the Km. The Km relating to the affinity of the enzyme for the substrate should in most cases relate to potential changes in the binding site of the enzyme which would directly result from enzyme inhibitor interactions. As such a term similar to the delta Vmax term proposed above to modulate Vmax should be appropriate in most situations:[28]: eq 14 

Dissociation constants

2D plots of 1/[S] concentration (x-axis) and 1/V (y-axis) demonstrating that as inhibitor concentration is changed, competitive inhibitor lines intersect at a single point on the y-axis, non-competitive inhibitors intersect at the x-axis, and mixed inhibitors intersect a point that is on neither axis
Lineweaver–Burk diagrams of different types of reversible enzyme inhibitors. The arrow shows the effect of increasing concentrations of inhibitor.

An enzyme inhibitor is characterised by its

enzyme activity under various substrate and inhibitor concentrations, and fitting the data via nonlinear regression[32] to a modified Michaelis–Menten equation.[21]

where the modifying factors α and α' are defined by the inhibitor concentration and its two dissociation constants

Thus, in the presence of the inhibitor, the enzyme's effective Km and Vmax become (α/α')Km and (1/α')Vmax, respectively. However, the modified Michaelis-Menten equation assumes that binding of the inhibitor to the enzyme has reached equilibrium, which may be a very slow process for inhibitors with sub-nanomolar dissociation constants. In these cases the inhibition becomes effectively irreversible, hence it is more practical to treat such tight-binding inhibitors as irreversible (see below).

The effects of different types of reversible enzyme inhibitors on enzymatic activity can be visualised using graphical representations of the Michaelis–Menten equation, such as

Hanes-Woolf plots.[17]: 140–144  An illustration is provided by the three Lineweaver–Burk plots depicted in the Lineweaver–Burk diagrams figure. In the top diagram the competitive inhibition lines intersect on the y-axis, illustrating that such inhibitors do not affect Vmax. In the bottom diagram the non-competitive inhibition lines intersect on the x-axis, showing these inhibitors do not affect Km. However, since it can be difficult to estimate Ki and Ki' accurately from such plots,[33] it is advisable to estimate these constants using more reliable nonlinear regression methods.[33]

Special cases

Partially competitive

The mechanism of partially competitive inhibition is similar to that of non-competitive, except that the EIS complex has catalytic activity, which may be lower or even higher (partially competitive activation) than that of the enzyme–substrate (ES) complex. This inhibition typically displays a lower Vmax, but an unaffected Km value.[18]

Substrate or product

Substrate or product inhibition is where either an enzymes substrate or product also act as an inhibitor. This inhibition may follow the competitive, uncompetitive or mixed patterns. In substrate inhibition there is a progressive decrease in activity at high substrate concentrations, potentially from an enzyme having two competing substrate-binding sites. At low substrate, the high-affinity site is occupied and normal kinetics are followed. However, at higher concentrations, the second inhibitory site becomes occupied, inhibiting the enzyme.[34] Product inhibition (either the enzyme's own product, or a product to an enzyme downstream in its metabolic pathway) is often a regulatory feature in metabolism and can be a form of negative feedback.[2]

Slow-tight

Slow-tight inhibition occurs when the initial enzyme–inhibitor complex EI undergoes

irreversible inhibition.[17]
: 168 

Multi-substrate analogues

TGDDF / GDDF MAIs where blue depicts the tetrahydrofolate cofactor analogue, black GAR or thioGAR and red, the connecting atoms.
TGDDF/GDDF multi-substrate adduct inhibitor. Substrate analogue in black, cofactor analogue in blue, non-cleavable linker in red.
Ritonavir is similar to the natural substrate.
Peptide-based HIV-1 protease inhibitor ritonavir with substrate binding sites located in enzyme labelled as S2, S1, S1', and S2'.
Tipranavir is not similar to the natural substrate.
Nonpeptidic HIV-1 protease inhibitor tipranavir

Multi-substrate analogue inhibitors are high affinity selective inhibitors that can be prepared for enzymes that catalyse reactions with more than one substrate by capturing the binding energy of each of those substrate into one molecule.

PTC124)[40] with cellular cofactors such as nicotinamide adenine dinucleotide (NADH) and adenosine triphosphate (ATP) respectively.[41]

Examples

As enzymes have evolved to bind their substrates tightly, and most reversible inhibitors bind in the active site of enzymes, it is unsurprising that some of these inhibitors are strikingly similar in structure to the substrates of their targets. Inhibitors of

antiretroviral drugs used to treat HIV/AIDS.[43][44] The structure of ritonavir, a peptidomimetic (peptide mimic) protease inhibitor containing three peptide bonds, as shown in the "competitive inhibition" figure above. As this drug resembles the peptide that is the substrate of the HIV protease, it competes with the substrate in the enzyme's active site.[45]

Enzyme inhibitors are often designed to mimic the transition state or intermediate of an enzyme-catalysed reaction.[46] This ensures that the inhibitor exploits the transition state stabilising effect of the enzyme, resulting in a better binding affinity (lower Ki) than substrate-based designs. An example of such a transition state inhibitor is the antiviral drug oseltamivir; this drug mimics the planar nature of the ring oxonium ion in the reaction of the viral enzyme neuraminidase.[47]

However, not all inhibitors are based on the structures of substrates. For example, the structure of another HIV protease inhibitor

peptidases and are less likely to be degraded.[48]

In drug design it is important to consider the concentrations of substrates to which the target enzymes are exposed. For example, some protein kinase inhibitors have chemical structures that are similar to ATP, one of the substrates of these enzymes.[49] However, drugs that are simple competitive inhibitors will have to compete with the high concentrations of ATP in the cell. Protein kinases can also be inhibited by competition at the binding sites where the kinases interact with their substrate proteins, and most proteins are present inside cells at concentrations much lower than the concentration of ATP. As a consequence, if two protein kinase inhibitors both bind in the active site with similar affinity, but only one has to compete with ATP, then the competitive inhibitor at the protein-binding site will inhibit the enzyme more effectively.[50]

Irreversible inhibitors

Types

Irreversible inhibitors

DFP, see the "DFP reaction" diagram), and also cysteine, threonine, or tyrosine.[53]

Irreversible inhibition is different from irreversible enzyme inactivation.[54] Irreversible inhibitors are generally specific for one class of enzyme and do not inactivate all proteins; they do not function by destroying protein structure but by specifically altering the active site of their target. For example, extremes of pH or temperature usually cause denaturation of all protein structure, but this is a non-specific effect. Similarly, some non-specific chemical treatments destroy protein structure: for example, heating in concentrated hydrochloric acid will hydrolyse the peptide bonds holding proteins together, releasing free amino acids.[55]

Irreversible inhibitors display time-dependent inhibition and their potency therefore cannot be characterised by an IC50 value. This is because the amount of active enzyme at a given concentration of irreversible inhibitor will be different depending on how long the inhibitor is pre-incubated with the enzyme. Instead, kobs/[I] values are used,[56] where kobs is the observed pseudo-first order rate of inactivation (obtained by plotting the log of % activity versus time) and [I] is the concentration of inhibitor. The kobs/[I] parameter is valid as long as the inhibitor does not saturate binding with the enzyme (in which case kobs = kinact) where kinact is the rate of inactivation.

Measuring

Irreversible inhibition mechanism
Depiction of the reversible chemical equilibria between enzyme + substrate, enzyme/substrate complex, and enzyme + product, and two competing equilibria. The first is between enzyme + inhibitor, enzyme/inhibitor non-covalent complex, followed by irreversible formation of the covalent complex. The second is between enzyme/substrate complex + inhibitor, noncovalent enzyme/substrate, followed by irreversible formation of the covalent complex
Kinetic mechanism for irreversible inhibition. Substrate binding in blue, catalysis in red, inhibitor binding in green, inactivation reaction in dark green.

Irreversible inhibitors first form a reversible non-covalent complex with the enzyme (EI or ESI). Subsequently, a chemical reaction occurs between the enzyme and inhibitor to produce the covalently modified "dead-end complex" EI* (an irreversible covalent complex). The rate at which EI* is formed is called the inactivation rate or kinact.[13] Since formation of EI may compete with ES, binding of irreversible inhibitors can be prevented by competition either with substrate or with a second, reversible inhibitor. This protection effect is good evidence of a specific reaction of the irreversible inhibitor with the active site.

The binding and inactivation steps of this reaction are investigated by incubating the enzyme with inhibitor and assaying the amount of activity remaining over time. The activity will be decreased in a time-dependent manner, usually following exponential decay. Fitting these data to a rate equation gives the rate of inactivation at this concentration of inhibitor. This is done at several different concentrations of inhibitor. If a reversible EI complex is involved the inactivation rate will be saturable and fitting this curve will give kinact and Ki.[57]

Another method that is widely used in these analyses is

MALDI-TOF mass spectrometer.[59] In a complementary technique, peptide mass fingerprinting involves digestion of the native and modified protein with a protease such as trypsin. This will produce a set of peptides that can be analysed using a mass spectrometer. The peptide that changes in mass after reaction with the inhibitor will be the one that contains the site of modification.[60]

Slow binding

2D chemical structure diagram depicting a lysine residue from the enzyme first reacting with DFMO, elimination of fluoride and carbon dioxide, followed by cysteine attacking the covalent lysine-DFMO adduct freeing the lysine residue to form an irreversible cysteine adduct
Chemical mechanism for irreversible inhibition of ornithine decarboxylase by DFMO. Pyridoxal 5'-phosphate (Py) and enzyme (E) are not shown. Adapted from Poulin et al, 1992.[61]

Not all irreversible inhibitors form covalent adducts with their enzyme targets. Some reversible inhibitors bind so tightly to their target enzyme that they are essentially irreversible. These tight-binding inhibitors may show kinetics similar to covalent irreversible inhibitors. In these cases some of these inhibitors rapidly bind to the enzyme in a low-affinity EI complex and this then undergoes a slower rearrangement to a very tightly bound EI* complex (see the "irreversible inhibition mechanism" diagram). This kinetic behaviour is called slow-binding.

acyclovir.[65]

Some examples

3D cartoon diagram of the trypanothione reductase protein bound to two molecules of inhibitors depicted as a stick diagrams.
Trypanothione reductase with the lower molecule of an inhibitor bound irreversibly and the upper one reversibly. Created from Bond et al, 2004.[66] (PDB: 1GXF​)

synapses of neurons, and consequently is a potent neurotoxin, with a lethal dose of less than 100 mg.[68]

Suicide inhibition is an unusual type of irreversible inhibition where the enzyme converts the inhibitor into a reactive form in its active site.[69] An example is the inhibitor of polyamine biosynthesis, α-difluoromethylornithine (DFMO), which is an analogue of the amino acid ornithine, and is used to treat African trypanosomiasis (sleeping sickness). Ornithine decarboxylase can catalyse the decarboxylation of DFMO instead of ornithine (see the "DFMO inhibitor mechanism" diagram). However, this decarboxylation reaction is followed by the elimination of a fluorine atom, which converts this catalytic intermediate into a conjugated imine, a highly electrophilic species. This reactive form of DFMO then reacts with either a cysteine or lysine residue in the active site to irreversibly inactivate the enzyme.[61]

Since irreversible inhibition often involves the initial formation of a non-covalent enzyme inhibitor (EI) complex,[13] it is sometimes possible for an inhibitor to bind to an enzyme in more than one way. For example, in the figure showing trypanothione reductase from the human protozoan parasite Trypanosoma cruzi, two molecules of an inhibitor called quinacrine mustard are bound in its active site. The top molecule is bound reversibly, but the lower one is bound covalently as it has reacted with an amino acid residue through its nitrogen mustard group.[70]

Applications

Enzyme inhibitors are found in nature

metabolic processes and are essential for life.[3][1] In addition, naturally produced poisons are often enzyme inhibitors that have evolved for use as toxic agents against predators, prey, and competing organisms.[4] These natural toxins include some of the most poisonous substances known.[73] Artificial inhibitors are often used as drugs, but can also be insecticides such as malathion, herbicides such as glyphosate,[74] or disinfectants such as triclosan. Other artificial enzyme inhibitors block acetylcholinesterase, an enzyme which breaks down acetylcholine, and are used as nerve agents in chemical warfare.[75]

Metabolic regulation

Enzyme inhibition is a common feature of

pyruvate. A key step for the regulation of glycolysis is an early reaction in the pathway catalysed by phosphofructokinase‑1 (PFK1). When ATP levels rise, ATP binds an allosteric site in PFK1 to decrease the rate of the enzyme reaction; glycolysis is inhibited and ATP production falls. This negative feedback control helps maintain a steady concentration of ATP in the cell. However, metabolic pathways are not just regulated through inhibition since enzyme activation is equally important. With respect to PFK1, fructose 2,6-bisphosphate and ADP are examples of metabolites that are allosteric activators.[77]

Physiological enzyme inhibition can also be produced by specific protein inhibitors. This mechanism occurs in the pancreas, which synthesises many digestive precursor enzymes known as zymogens. Many of these are activated by the trypsin protease, so it is important to inhibit the activity of trypsin in the pancreas to prevent the organ from digesting itself. One way in which the activity of trypsin is controlled is the production of a specific and potent trypsin inhibitor protein in the pancreas. This inhibitor binds tightly to trypsin, preventing the trypsin activity that would otherwise be detrimental to the organ.[78] Although the trypsin inhibitor is a protein, it avoids being hydrolysed as a substrate by the protease by excluding water from trypsin's active site and destabilising the transition state.[79] Other examples of physiological enzyme inhibitor proteins include the barstar inhibitor of the bacterial ribonuclease barnase.[80]

Natural poisons

photo of three piles of legume seeds coloured brown, pea green, and brown/orange
To discourage seed predation, legumes contain trypsin inhibitors that interfere with digestion.

Animals and plants have evolved to synthesise a vast array of poisonous products including secondary metabolites,[81] peptides and proteins[82] that can act as inhibitors. Natural toxins are usually small organic molecules and are so diverse that there are probably natural inhibitors for most metabolic processes.[83] The metabolic processes targeted by natural poisons encompass more than enzymes in metabolic pathways and can also include the inhibition of receptor, channel and structural protein functions in a cell. For example, paclitaxel (taxol), an organic molecule found in the Pacific yew tree, binds tightly to tubulin dimers and inhibits their assembly into microtubules in the cytoskeleton.[84]

Many natural poisons act as

competitive antagonist of the muscarinic acetylcholine receptors.[87]

Although many natural toxins are secondary metabolites, these poisons also include peptides and proteins. An example of a toxic peptide is

death cap mushroom. This is a potent enzyme inhibitor, in this case preventing the RNA polymerase II enzyme from transcribing DNA.[88] The algal toxin microcystin is also a peptide and is an inhibitor of protein phosphatases.[89] This toxin can contaminate water supplies after algal blooms and is a known carcinogen that can also cause acute liver haemorrhage and death at higher doses.[90]

Proteins can also be natural poisons or

glycosidase that inactivates ribosomes.[93] Since ricin is a catalytic irreversible inhibitor, this allows just a single molecule of ricin to kill a cell.[94]

Drugs

2D chemical structural diagrams comparing folic acid and methotrexate
The coenzyme folic acid (top) compared to the anti-cancer drug methotrexate (bottom)
2D structural diagram of sildenafil
The structure of sildenafil (Viagra)

The most common uses for enzyme inhibitors are as drugs to treat disease. Many of these inhibitors target a human enzyme and aim to correct a pathological condition. For instance,

suicide inhibitor of the cyclooxygenase enzyme.[95] This inhibition in turn suppresses the production of proinflammatory prostaglandins and thus aspirin may be used to reduce pain, fever, and inflammation.[95]

As of 2017,[update] an estimated 29% of approved drugs are enzyme inhibitors

inflammatory diseases in including arthritis, asthma, and Crohn's disease.[98]

An example of the structural similarity of some inhibitors to the substrates of the enzymes they target is seen in the figure comparing the drug

folic acid. Folic acid is the oxidised form of the substrate of dihydrofolate reductase, an enzyme that is potently inhibited by methotrexate. Methotrexate blocks the action of dihydrofolate reductase and thereby halts thymidine biosynthesis.[42] This block of nucleotide biosynthesis is selectively toxic to rapidly growing cells, therefore methotrexate is often used in cancer chemotherapy.[99]

A common treatment for

cGMP specific phosphodiesterase type 5, the enzyme that degrades the signalling molecule cyclic guanosine monophosphate.[101] This signalling molecule triggers smooth muscle relaxation and allows blood flow into the corpus cavernosum
, which causes an erection. Since the drug decreases the activity of the enzyme that halts the signal, it makes this signal last for a longer period of time.

Antibiotics

3D cartoon diagram of transpeptidase bound to penicillin G depicted as sticks
The structure of a complex between penicillin G and the Streptomyces transpeptidase (PDB: 1PWC​)

Drugs are also used to inhibit enzymes needed for the survival of pathogens. For example, bacteria are surrounded by a thick cell wall made of a net-like polymer called peptidoglycan. Many antibiotics such as penicillin and vancomycin inhibit the enzymes that produce and then cross-link the strands of this polymer together.[102][103] This causes the cell wall to lose strength and the bacteria to burst. In the figure, a molecule of penicillin (shown in a ball-and-stick form) is shown bound to its target, the transpeptidase from the bacteria Streptomyces R61 (the protein is shown as a ribbon diagram).

Antibiotic drug design is facilitated when an enzyme that is essential to the pathogen's survival is absent or very different in humans.[104] Humans do not make peptidoglycan, therefore antibiotics that inhibit this process are selectively toxic to bacteria.[105] Selective toxicity is also produced in antibiotics by exploiting differences in the structure of the ribosomes in bacteria,[106] or how they make fatty acids.[107]

Antivirals

Drugs that inhibit enzymes needed for the

human cytomegalovirus.[113]

Pesticides

Many

synaptic cleft rather than cleaved. A large number of AChE inhibitors are used in both medicine and agriculture.[116] Reversible competitive inhibitors, such as edrophonium, physostigmine, and neostigmine, are used in the treatment of myasthenia gravis[117] and in anaesthesia to reverse muscle blockade.[118] The carbamate pesticides are also examples of reversible AChE inhibitors. The organophosphate pesticides such as malathion, parathion, and chlorpyrifos irreversibly inhibit acetylcholinesterase.[119]

Herbicides

The herbicide

3-phosphoshikimate 1-carboxyvinyltransferase,[120] other herbicides, such as the sulfonylureas inhibit the enzyme acetolactate synthase.[121] Both enzymes are needed for plants to make branched-chain amino acids. Many other enzymes are inhibited by herbicides, including enzymes needed for the biosynthesis of lipids and carotenoids and the processes of photosynthesis and oxidative phosphorylation.[122]

Discovery and design

photo of robots at work
Robots are used for the high-throughput screening of chemical libraries to discover new enzyme inhibitors.

New drugs are the products of a long drug development process, the first step of which is often the discovery of a new enzyme inhibitor.[123] There are two principle approaches of discovering these inhibitors.[124]

The first general method is

transition state analogues generally possess higher affinity for the enzyme compared to the substrate, and therefore are effective inhibitors.[46]

The second way of discovering new enzyme inhibitors is

Hits from any of the above approaches can be

molecular docking[132] and molecular mechanics can be used to assist in the optimisation process.[133] New inhibitors are used to obtain crystallographic structures of the enzyme in an inhibitor/enzyme complex to show how the molecule is binding to the active site, allowing changes to be made to the inhibitor to optimise binding in a process known as structure-based drug design.[1]
: 66  This test and improve cycle is repeated until a sufficiently potent inhibitor is produced.

See also

  • Activity-based proteomics – a branch of proteomics that uses covalent enzyme inhibitors as reporters to monitor enzyme activity.
  • Antimetabolite – an enzyme inhibitor that is used to interfere with cell growth and division
  • Transition state analogue
    – a type of enzyme inhibitor that mimics the transition state of the chemical reaction catalysed by the enzyme

References

  1. ^ .
  2. ^ .
  3. ^ from the original on 28 March 2023. Retrieved 14 April 2022.
  4. ^ .
  5. .
  6. .
  7. .
  8. .
  9. ^ from the original on 28 March 2023. Retrieved 20 July 2022.
  10. .
  11. from the original on 20 July 2022. Retrieved 20 July 2022.
  12. from the original on 7 June 2022. Retrieved 7 June 2022.
  13. ^ .
  14. ^ .
  15. from the original on 26 September 2009. Retrieved 31 August 2017.
  16. ^ a b c "Types of Inhibition". NIH Center for Translational Therapeutics. Archived from the original on 8 September 2011. Retrieved 2 April 2012.
  17. ^ .
  18. ^ .
  19. .
  20. from the original on 28 November 2021. Retrieved 3 April 2022.
  21. ^ .
  22. . In some cases, the inhibitor may bind to a distinct site on the enzyme that is in allosteric communication with the substrate binding pocket. In many cases, allosteric, substrate competitive compounds result in conformational changes to the enzyme that change the ability of the enzyme to bind substrate.
  23. .
  24. ^ .
  25. .
  26. .
  27. ^ .
  28. ^ .
  29. ^ Walsh R (May 2012). "Alternative perspectives of enzyme kinetic modeling". Medicinal Chemistry and Drug Design. (16). InTech: 357–372.
  30. from the original on 15 June 2022. Retrieved 9 April 2022.
  31. .
  32. .
  33. ^ .
  34. .
  35. .
  36. .
  37. .
  38. .
  39. .
  40. .
  41. .
  42. ^ .
  43. .
  44. .
  45. .
  46. ^ .
  47. .
  48. .
  49. .
  50. .
  51. ^ from the original on 28 March 2023. Retrieved 3 June 2022.
  52. .
  53. .
  54. . Enzyme inactivation is generally explained as a chemical process involving several phenomena like aggregation, dissociation into subunits, or denaturation (conformational changes), which occur simultaneously during the inactivation of a specific enzyme.
  55. .
  56. .
  57. .
  58. .
  59. from the original on 20 July 2022. Retrieved 20 July 2022.
  60. .
  61. ^ .
  62. .
  63. .
  64. .
  65. .
  66. .
  67. on 28 February 2018.
  68. .
  69. .
  70. .
  71. (PDF) from the original on 5 December 2022. Retrieved 3 April 2022.
  72. .
  73. from the original on 25 April 2013. Retrieved 3 June 2022.
  74. from the original on 28 March 2023. Retrieved 14 April 2022.
  75. from the original on 17 January 2023. Retrieved 14 April 2022.
  76. .
  77. .
  78. .
  79. from the original on 28 March 2023. Retrieved 14 April 2022.
  80. .
  81. .
  82. .
  83. .
  84. .
  85. .
  86. .
  87. .
  88. .
  89. .
  90. .
  91. .
  92. .
  93. .
  94. .
  95. ^ from the original on 28 March 2023. Retrieved 14 April 2022.
  96. ^ . Figure 1C: Clinical success of privileged protein family classes (% approved drugs targeting each target class): Reductase 7.62, Kinase 5.94, Protease 3.35, Hydrolase 2.76, NPTase 2.09, Transferase 1.92, Lyase 1.59, Isomerase 1.51, Phosphodiesterase 1.50, Cytochrome p450 0.84, Epigenetic eraser 0.33, Total enzyme targets of approved drugs = 29.45%
  97. .
  98. .
  99. .
  100. .
  101. .
  102. from the original on 28 March 2023. Retrieved 20 July 2022.
  103. .
  104. .
  105. from the original on 9 April 2022. Retrieved 9 April 2022.
  106. .
  107. .
  108. .
  109. .
  110. .
  111. .
  112. .
  113. .
  114. .
  115. .
  116. from the original on 21 July 2022. Retrieved 21 July 2022.
  117. .
  118. ^ Butterworth JF, IV, Mackey DC, Wasnick JD, eds. (2013). "Chapter 12. Cholinesterase Inhibitors & Other Pharmacologic Antagonists to Neuromuscular Blocking Agents.". Morgan & Mikhail's Clinical Anesthesiology (5th ed.). McGraw Hill. Archived from the original on 30 June 2022. Retrieved 9 April 2022.
  119. PMID 28117022
    .
  120. .
  121. from the original on 27 June 2022. Retrieved 27 June 2022. Chapter 10.2.1: Sulfonylurea acetolactate synthase inhibitors
  122. .
  123. .
  124. .
  125. ^ Lindquist RN (October 2013). "The design of enzyme inhibitors: Transition state analogues.". In Ariëns EJ (ed.). Drug Design: Medicinal Chemistry: A Series of Monographs. Vol. 5. Elsevier. pp. 24–80.
  126. PMID 16796557
    .
  127. .
  128. .
  129. .
  130. .
  131. .
  132. .
  133. .

External links

  • "BRENDA". Archived from the original on 1 April 2022., Database of enzymes giving lists of known inhibitors for each entry
  • "PubChem". National Center for Biotechnology Information. National Library of Medicine. Database of drugs and enzyme inhibitors
  • "Symbolism and Terminology in Enzyme Kinetics". Archived from the original on 20 June 2006. Recommendations of the Nomenclature Committee of the International Union of Biochemistry (NC-IUB) on enzyme inhibition terminology