Enzyme Inhibition
Updated:
Irreversible inhibition is a type of enzyme inhibition that permanently inactivates the enzyme by forming a strong covalent bond with an amino acid residue at or near the active site. This prevents the enzyme from catalyzing its normal reaction and reduces the overall enzyme activity in the system.
Irreversible inhibitors are often highly reactive compounds that target specific amino acid residues that are essential for the enzyme function. For example, some irreversible inhibitors react with serine or cysteine residues that have nucleophilic hydroxyl or sulfhydryl groups, respectively. These groups can attack the electrophilic center of the inhibitor and form a stable covalent bond.
Irreversible inhibition can be used as a tool to study the structure and function of enzymes, as well as to design drugs that selectively block the activity of certain enzymes involved in diseases. Some examples of irreversible inhibitors are:
- Diisopropyl fluorophosphate (DIPF), a nerve agent that inhibits acetylcholinesterase, an enzyme that breaks down the neurotransmitter acetylcholine. By blocking this enzyme, DIPF causes an accumulation of acetylcholine in the synapses and leads to muscle paralysis and respiratory failure.
- Iodoacetamide, a compound that modifies cysteine residues and can be used to identify whether one or more cysteine residues are required for enzyme activity. For example, iodoacetamide can inhibit papain, a cysteine protease that cleaves peptide bonds.
- Penicillin, an antibiotic that inhibits the transpeptidase enzyme that forms the cross-links in the bacterial cell wall. By binding covalently to a serine residue in the active site of the enzyme, penicillin prevents the formation of a stable peptidoglycan layer and makes the bacteria susceptible to osmotic lysis.
Irreversible inhibitors are molecules that bind covalently to an enzyme and permanently inactivate it. They often target a specific amino acid residue at or near the active site of the enzyme, and modify its chemical properties or structure. Some examples of irreversible inhibitors are:
- DIPF (di-isopropylphosphofluoridate): This compound is a component of nerve gases, and it reacts with a serine residue in the active site of the enzyme acetylcholinesterase. Acetylcholinesterase is responsible for breaking down the neurotransmitter acetylcholine in the synaptic cleft, and thus terminating the nerve impulse. By inhibiting acetylcholinesterase, DIPF causes a buildup of acetylcholine in the synapse, leading to overstimulation of the nerve cells and paralysis of the muscles.
- Iodoacetamide: This compound modifies cysteine residues by adding an iodoacetyl group to their sulfhydryl groups. Cysteine residues are often involved in catalysis or substrate binding in enzymes, and their modification can impair the enzyme function. Iodoacetamide can be used as a diagnostic tool to determine whether one or more cysteine residues are essential for enzyme activity.
- Penicillin: This antibiotic inhibits the enzyme glycopeptide transpeptidase, which is involved in the synthesis of the bacterial cell wall. Penicillin covalently attaches to a serine residue in the active site of the enzyme, and prevents it from forming cross-links between the peptidoglycan chains that make up the cell wall. As a result, the bacteria become susceptible to osmotic lysis and die.
Reversible inhibition is a type of enzyme inhibition that can be reversed by removing the inhibitor from the enzyme. Reversible inhibitors bind to the enzyme through non-covalent interactions, such as hydrogen bonds, ionic bonds, or hydrophobic interactions, and can dissociate from the enzyme easily. Reversible inhibitors do not permanently alter or damage the enzyme, but they reduce its catalytic activity by interfering with its substrate binding or catalytic mechanism.
Reversible inhibitors can be classified into different types based on their mode of action and their effect on the enzyme kinetics. The main types of reversible inhibitors are competitive, non-competitive, mixed, and uncompetitive inhibitors. Each type of inhibitor has a different effect on the parameters of the Michaelis-Menten equation, such as Vmax (the maximum rate of reaction) and Km (the substrate concentration at half of Vmax). These effects can be visualized by plotting the reaction rate versus substrate concentration (Michaelis-Menten plot) or the inverse of these values (Lineweaver-Burk plot).
Reversible inhibition can be overcome by increasing the substrate concentration, changing the pH or temperature, or adding an activator that enhances the enzyme activity. Reversible inhibitors are often used as drugs to modulate the activity of specific enzymes involved in various metabolic pathways or diseases. For example, some antiviral drugs inhibit the enzymes that synthesize viral nucleic acids, some antibiotics inhibit the enzymes that synthesize bacterial cell walls, and some antihypertensive drugs inhibit the enzymes that regulate blood pressure. Reversible inhibitors are also important in regulating enzyme activity in normal physiological conditions, such as feedback inhibition and allosteric regulation.
Competitive inhibitors are molecules that bind to the active site of an enzyme and prevent the substrate from binding. They compete with the substrate for the same binding site and reduce the rate of enzyme-catalyzed reactions.
Competitive inhibitors usually have a similar structure or shape to the normal substrate of the enzyme. This allows them to fit into the active site and block the access of the substrate. For example, malonate is a competitive inhibitor of succinate dehydrogenase because it has a similar structure to succinate, the natural substrate of the enzyme.
Competitive inhibition is reversible and can be overcome by increasing the concentration of the substrate. This increases the likelihood of the substrate binding to the enzyme and displacing the inhibitor. The higher the affinity of the inhibitor for the enzyme, the more substrate is needed to overcome the inhibition.
The effect of competitive inhibition on enzyme kinetics can be analyzed using a Lineweaver-Burk plot, which is a double reciprocal plot of 1/V (the inverse of reaction velocity) versus 1/[S] (the inverse of substrate concentration). In the presence of a competitive inhibitor, the slope of the plot increases, indicating a lower affinity of the enzyme for the substrate. The y-intercept, which represents 1/Vmax (the inverse of maximum reaction velocity), remains unchanged, indicating that Vmax is not affected by competitive inhibition. The x-intercept, which represents -1/Km (the negative inverse of Michaelis constant), shifts to the left, indicating that Km increases in the presence of a competitive inhibitor.
The following diagram illustrates how competitive inhibition affects enzyme kinetics:
[Insert diagram here]
Competitive inhibitors are often used as drugs to inhibit enzymes that are involved in disease processes. For example, methotrexate is a competitive inhibitor of dihydrofolate reductase, an enzyme that is essential for DNA synthesis in bacteria and cancer cells. By inhibiting this enzyme, methotrexate prevents the growth and proliferation of these cells. Another example is statins, which are competitive inhibitors of HMG-CoA reductase, an enzyme that is involved in cholesterol synthesis in the liver. By inhibiting this enzyme, statins lower blood cholesterol levels and reduce the risk of cardiovascular diseases.
Succinate dehydrogenase (SDH) is an enzyme that catalyzes the oxidation of succinate to fumarate in the citric acid cycle. It is also part of the electron transport chain, where it transfers electrons from succinate to ubiquinone.
A competitive inhibitor of SDH is malonate, which has a similar structure to succinate but with one less carbon atom. Malonate can bind to the active site of SDH and prevent succinate from binding and being oxidized. This reduces the activity of SDH and the production of fumarate and ubiquinone.
The inhibition by malonate is reversible, meaning that it can be overcome by increasing the concentration of succinate. As the substrate concentration increases, the chances of succinate binding to the active site become higher than those of malonate. This restores the normal function of SDH and the citric acid cycle.
The effect of malonate on SDH can be demonstrated experimentally by measuring the rate of oxygen consumption by isolated mitochondria in the presence of different concentrations of succinate and malonate. The rate of oxygen consumption reflects the rate of electron transport chain activity, which depends on the availability of ubiquinone produced by SDH.
As shown in the figure below, when malonate is absent, the rate of oxygen consumption increases linearly with increasing succinate concentration. This indicates that SDH is not inhibited and can oxidize succinate efficiently. However, when malonate is present, the rate of oxygen consumption is lower at any given succinate concentration. This indicates that SDH is inhibited by malonate and cannot oxidize succinate as fast. The inhibition becomes less pronounced as the succinate concentration increases, showing that malonate is a competitive inhibitor.
[Insert figure here]
Malonate is an example of a competitive inhibitor that mimics the structure of the substrate and competes with it for binding to the active site. By understanding how competitive inhibitors work, we can design drugs that target specific enzymes and modulate their activity. For example, some antiviral drugs work by competitively inhibiting viral enzymes that are essential for viral replication.
Non-competitive inhibitors are a type of reversible inhibitors that bind to the enzyme or the enzyme–substrate complex at a site different from the active site, decreasing the activity of the enzyme. Unlike competitive inhibitors, non-competitive inhibitors do not have structural similarities to the normal substrate for the enzyme. They can bind to the enzyme regardless of whether the substrate is present or not.
Non-competitive inhibitors affect the enzyme activity by changing the overall three-dimensional shape of the enzyme. This alters the active site and reduces its affinity for the substrate, meaning that the substrate cannot bind as effectively or as frequently. As a result, the rate of enzyme-catalyzed reaction decreases.
The effect of non-competitive inhibitors cannot be overcome by increasing the substrate concentration, because the inhibitor can still bind to the enzyme–substrate complex and prevent the formation of products. Therefore, non-competitive inhibitors lower the maximum rate of reaction (Vmax) of the enzyme, but do not affect the Michaelis constant (Km), which is a measure of how well the substrate binds to the enzyme.
Non-competitive inhibitors are often used as drugs to treat various diseases or disorders by inhibiting specific enzymes involved in metabolic pathways. For example, aspirin is a non-competitive inhibitor of cyclooxygenase, an enzyme that produces prostaglandins, which are involved in inflammation and pain. By inhibiting cyclooxygenase, aspirin reduces inflammation and pain.
Renin is an enzyme that catalyzes the conversion of angiotensinogen to angiotensin I, the first step in the renin–angiotensin–aldosterone system that regulates blood pressure and fluid balance. Renin is produced by the kidneys and released into the bloodstream in response to low blood pressure or sodium levels.
Pepstatin is a synthetic peptide that inhibits renin by binding to a site different from the active site where angiotensinogen binds. Pepstatin has a very high affinity for renin and forms a reversible complex with it. This complex prevents renin from interacting with angiotensinogen and reduces the formation of angiotensin I and II.
Pepstatin is a classical non-competitive inhibitor of renin, meaning that it does not compete with the substrate for the active site, but rather reduces the catalytic activity of the enzyme. The inhibition by pepstatin is not affected by the concentration of angiotensinogen, so increasing the substrate concentration does not overcome the inhibition. Pepstatin also lowers the maximum velocity (Vmax) of renin, but does not change its affinity (Km) for angiotensinogen.
Pepstatin was the first synthetic renin inhibitor discovered in 1972, but it was not suitable for clinical use because of its poor pharmacokinetic properties. It was mainly used as a research tool to study the mechanism and structure of renin and its inhibitors. Later, more potent and orally bioavailable non-peptide renin inhibitors were developed, such as aliskiren, which is currently used as an antihypertensive drug.
We are Compiling this Section. Thanks for your understanding.