What allows drugs to gather in high concentration in a cell or body compartment?

The development of irreversible antagonists and subsequent crystallization of hCB1R is described in Chapter 10 of this volume (“Ligand-Assisted Protein Structure (LAPS): An Experimental Paradigm for Characterizing Cannabinoid-Receptor Ligand-Binding Domains” by Janero et al.).

From: Methods in Enzymology, 2017

How drugs act : General principles

James M. Ritter DPhil FRCP HonFBPhS FMedSci, in Rang & Dale's Pharmacology, 2020

Irreversible Competitive Antagonism

▼ Irreversible competitive (ornon-equilibrium)antagonism occurs when the antagonist binds to the same site on the receptor as the agonist but dissociates very slowly, or not at all, from the receptors, with the result that no change in the antagonist occupancy takes place when the agonist is applied.4

The predicted effects of reversible and irreversible antagonists are compared inFig. 2.4.

In some cases (Fig. 2.6A), the theoretical effect is accurately reproduced with the antagonist reducing the maximum response. However, the distinction between reversible and irreversible competitive antagonism (or even non-competitive antagonism) is not always so clear. This is because of the phenomenon of spare receptors (seep. 10); if the agonist occupancy required to produce a maximal biological response is very small (say 1% of the total receptor pool), then it is possible to block irreversibly nearly 99% of the receptors without reducing the maximal response. The effect of a lesser degree of antagonist occupancy will be to produce a parallel shift of the log concentration–effect curve that is indistinguishable from reversible competitive antagonism (Fig. 2.6B). Only when the antagonist occupancy exceeds 99% will the maximum response will be reduced.

Irreversible competitive antagonism occurs with drugs that possess reactive groups that form covalent bonds with the receptor. These are mainly used as experimental tools for investigating receptor function, and few are used clinically. Irreversible enzyme inhibitors that act similarly are clinically used, however, and include drugs such asaspirin (Ch. 27),omeprazole (Ch. 31) and monoamine oxidase inhibitors (Ch. 48).

Renal and Genitourinary Systems

Stan K. Bardal BSc (Pharm), MBA, PhD, ... Douglas S. Martin PhD, in Applied Pharmacology, 2011

FYI

Phenoxybenzamine is an irreversible antagonist of α1 receptors, whereas other α1 antagonists are reversible.

Phentolamine antagonizes both α1 and α2 receptors. Through antagonism of presynaptic α2 receptors, it increases presynaptic noradrenaline release, which in turn stimulates the heart.

Yohimbine is an α2 antagonist with limited antagonism of α1 receptors. It is not widely used, but as with phentolamine, it can promote norepinephrine release and can be used in patients with orthostatic hypotension. It was also once used to treat erectile dysfunction, although it has been effectively replaced by the phosphodiesterase inhibitors (e.g., sildenafil).

Read full chapter

URL: //www.sciencedirect.com/science/article/pii/B9781437703108000257

Basic Principles of Pharmacology

Robert B. Raffa PhD, in Netter's Illustrated Pharmacology, 2014

Overview

Pharmacology is the study of drug action at both the molecular and the whole-organism levels. At the molecular level,drug action refers to the mechanism by which a drug or other molecule produces a biologic effect. At the whole-organism level,drug action refers to the therapeutic effects of a drug and its unwanted (ie, adverse, or side) effects. Drugs can produce biologic effects in several ways, eg, killing harmful invading organisms such as bacteria and viruses; killing the body's own cells that have gone awry (eg, cancer cells); neutralizing acid (mechanism of action of antacids); modifying ongoing underactive or overactive physiologic processes. In the last case, direct replacement of chemicals (eg, insulin) or indirect or more subtle modulation of biochemical processes (eg, inhibition of enzyme action) may be required.

Drugs can be said to modify the communication system within an organism. The modification should not interfere with the fidelity of the signal and should not activate unwanted compensatory responses. Drugs should selectively target specific cellular components that function in the normal signaling process. The study of molecular, biochemical, and physiologic effects of drugs on cellular systems and drug mechanisms of action is termedpharmacodynamics.

Equally important to drug action are the absorption, distribution, metabolism, and elimination (ADME) of drugs. The study of these processes (which involves the movement of the drug molecules through various physiologic compartments) and how they affect drug use and usefulness is termedpharmacokinetics. Complete understanding of the action of a drug involves knowledge of both pharmacodynamic (PD) and pharmacokinetic (PK) properties. In addition, the physical characteristics of an individual patient (eg, age, sex, weight, liver function, kidney function) dictate how the PD and PK characteristics of the drug are manifested.

Pharmacognosy is the study of drugs from natural sources.Pharmacy is the clinical practice devoted to the formulation and proper and safe distribution and use of therapeutic agents.

Therapeutic drug action involves interaction between an exogenous chemical and the endogenous biochemical target. The study of chemical structures of drugs and the study of normal and abnormal physiology are thus interrelated. Only by a clear understanding of the anatomy, physiology, and pathology of the organism can the proper drugs be designed and administered. The study of pharmacology therefore involves broad-based knowledge of the drug molecule, the organism, and the interaction between them.

Figure 1-1. Externaland Internal Threats

Invading organisms such as bacteria, viruses, fungi, and helminths can threaten the health of the host. Cancer cells are abnormal and differ from normal cells in terms of chromosome alterations, uncontrolled proliferation, dedifferentiation and loss of function, and invasiveness. Drug therapy (chemotherapy) aims to kill invading organisms or aberrant cells directly or to reduce their numbers to a level that can be managed by a host-mounted defense. Typical drug targets for invading organisms include biochemical processes needed for cell wall synthesis or integrity. Drug targets for abnormal cells include cell-cycle regulation and enzymes involved in protein synthesis, so as to inhibit cancer cell replication. In both cases, optimal treatment occurs when a drug or combination of drugs displays selectivity against invaders or cancer cells. Such therapy—with separation between a desired therapeutic effect and unwanted (adverse or side) effects—minimizes harmful drug effects.

Figure 1-2. Endogenous Chemical Balance

When the amount of an endogenous substance is insufficient for normal functions, it may be possible to supply it from sources outside of the body (exogenous supply). Examples include insulin used for diabetes and dopamine used for parkinsonism. The exogenous material may originate from humans, animals, microorganisms, or minerals or it may be synthesized—a product of technology. It can be the substance itself or a precursor metabolized to the substance (eg, levodopa is metabolized to dopamine). Excess amounts can also be harmful, eg, excess stomach acid can cause or exacerbate ulcer formation. Gastric acid levels can be reduced directly by using an antacid (a base such as calcium carbonate or magnesium hydroxide). An alternative approach—–inhibiting acid secretion—can be achieved by antagonizing the action of histamine on H2 receptors of parietal cells (eg, with cimetidine) or by interfering with the proton pump that transports acid across parietal cells (eg, with omeprazole).

Figure 1-3. Modulate Physiologic Processes

Drugs use different mechanisms to modify normal homeostatic and biochemical communication in cellular and physiologic processes. They mimic (eg, carbachol) or block neurotransmitters that transmit information across synapses. Chemical substances such as hormones also act over long distances in the body. Drugs that mimic hormones include oxandrolone; mifepristone blocks hormone action. Drugs selectively modify physiologic processes by targeting enzymes, DNA, neurotransmitters, or other chemical mediators or components of signaling processes such as receptors. The total effect depends on whether a drug promotes or reduces endogenous activity. Drugs with other mechanisms of action are chelating agents (contain metal atoms that form chemical bonds with toxins or drugs), antimetabolites (masquerade as endogenous substances but are inactive or less active than these substrates), irritants (stimulate physiologic processes), and nutritional or replacement agents (eg, vitamins, minerals).

Figure 1-4. Chemical Transmissionatthe Synapse

Communication (transmission of information) across synapses occurs via chemical messengers—neurotransmitters—stored in vesicles in presynaptic neurons. Action potentials at presynaptic axon terminals initiate steps that release neurotransmitter molecules into a synapse, which cross the synaptic cleft and bind reversibly to postsynaptic receptors. Receptor activation leads to cellular response. Receptor activators (eg, drugs) areagonists;antagonists are drugs that combine with but do not activate receptors. Transmitters are removed from synapses by enzymatic destruction, diffusion, and active reuptake into presynaptic neurons. Major peripheral neurotransmitters are acetylcholine and catecholamines (eg, epinephrine, dopamine). In the brain and spinal cord, major excitatory neurotransmitters are glutamate and aspartate; major inhibitory neurotransmitters are GABA and glycine. 5-HT, or serotonin, and neuropeptides are other neurotransmitters.

Figure 1-5. Synapse Morphology

Asynapse is a region including the axon terminal of a presynaptic neuron, the plasma membrane of the postsynaptic (receiving) cell, and the physical space between the cells (synaptic cleft). Postsynaptic cells can be neurons or other cells (eg, effector cells in muscle). At synapses, electrical transmissions—action potentials along presynaptic neurons—are translated into chemical signals, which lead to postsynaptic cell responses: increase (excitation), decrease (inhibition), or modulation of neuron activity or biochemistry. Synaptic transmission involves many steps, all possible drug targets. Steps occur in presynaptic neurons (eg, neurotransmitter synthesis and storage in vesicles), at presynaptic membranes (eg, vesicle docking with membranes, neurotransmitter exocytosis), in synaptic clefts (eg, enzymatic reuptake), on postsynaptic membranes (eg, binding to receptors, change in ion channel function), and in postsynaptic neurons (eg, effects on second-messenger transduction).

Figure 1-6. Receptorsand Signaling

Receptors are the first molecules in or on a cell that respond to a neurotransmitter, a hormone, or another endogenous or exogenous signaling molecule (ligand) and transmit messages (via transduction) from the molecule to the cell machinery. Receptors ensure fidelity of the intended communication by responding only to the intended signaling molecule or to molecules with closely related chemical structures (such as drugs with the required shape). Receptors are composed primarily of long sequences (typically hundreds) of amino acids. The body has dozens of receptor types to maintain communication pathways that must be differentiated from each other and serve different purposes. An individual cell may express one or many types of receptors, with the number depending on age, health, or other factors.

Figure 1-7. Receptor Subtypes

Receptors can be classified into subtypes, as first noted for receptors for the structurally related catecholamines epinephrine, isoproterenol, and norepinephrine. The order of potency (structure-activity relation, or SAR) of these drugs in some tissues is norepinephrine > epinephrine > isoproterenol; in other tissues, it is the reverse. Catecholamine receptors (adrenoceptors) exist in pharmacologically distinct types (α and β) and subtypes (eg, α1, α2, and so on). Subtypes are differentiated by amino acid sequence and posttranslational processing, as shown for dopamine receptor subtypes. A clinical example of receptor subtype targeting involves asthma treatment. Activation of adrenoceptors in the lung relaxes smooth muscles and dilates bronchioles to ease breathing. To avoid stimulation of heart adrenoceptors, β2-selective drugs (eg, albuterol, metaproterenol, ritodrine, terbutaline) were developed to activate only lung adrenoceptors; β1-selective drugs would affect the heart.

Figure 1-8. Agonists

Certain molecules have physiochemical and stereochemical (3-dimensional) characteristics that impartaffinity for a receptor, affinity being the quantifiable tendency of a drug molecule to form a complex with (bind to) a receptor. Binding involves interaction between a ligand molecule (L) and a receptor molecule (R) to form a ligand-receptor complex (LR): L + R ↔ LR. Affinity is quantified by the reciprocal of the equilibrium constant of this interaction and is commonly reported (often designatedKd orKi); the greater the affinity is, the smaller theK value is. Drugs can activate receptors and thus elicit a biologic effect (ie, have intrinsic activity, or efficacy). Such molecules have shapes complementary to receptor shapes and somehow alter the activity of a receptor. Full agonists possess high efficacy and can elicit a maximal tissue response, whereas partial agonists have intermediate levels of efficacy (the tissue response is submaximal even when all receptors are occupied).

Figure 1-9. Antagonists

Some molecules have physiochemical and stereochemical traits that impart affinity for a receptor but cannot activate it. Such molecules bind to (occupy) receptors and block access of agonists, thereby reducing the effects of agonists. Such pharmacologic antagonists do not elicit biologic effects directly; they modify the physiologic process that is maintained by agonist action (eg, by neurotransmitters). Examples of drugs that are receptor antagonists are atropine (muscarinic cholinergic),d-tubocurarine (nicotinic cholinergic), atenolol (adrenoceptor), spironolactone (mineralocorticoid), diphenhydramine (histamine H1), ondansetron (5-HT), flumazenil (benzodiazepine), haloperidol (dopamine), and naloxone (opioid). Chemical antagonism (eg, neutralization of gastric acid by chemical bases) or physiologic antagonism, in which an effect of one drug opposes an effect of another agent (eg, epinephrine used to counteract the histamine response to a bee sting), of drug effects can also occur.

Figure 1-10. Stereochemistryand 3-Dimensional Fit

One enantiomer of a racemic pair is often observed to bind more avidly to (has greater affinity for) a receptor than does the other enantiomer of the pair. Because the only difference between them is the stereochemistry, the 3-dimensional shape of a molecule must be a crucial characteristic for binding affinity. The relation between chemical structure and biologic response is known as theSAR and is a common focus of drug discovery efforts. Computer modeling of the ligand-receptor fit provides a visual representation of the fit of a ligand into the receptor pocket. It can also be used for virtual screening for goodness of fit of potential drug candidates before they are synthesized.

Figure 1-11. Receptor-Effector Coupling

In most cases, a drug activates or inhibits only 1 molecule in a long series of biochemical reactions. When a drug binds to a receptor on a cell membrane, the extracellular drug signal must be passed to the intracellular physiologic processes, ie, it must be converted (transduced) to an intracellular message, the process termedsignal transduction, which occurs via many mechanisms. The effect of a drug depends on its receptors, the transduction pathways to which it is coupled, its level of receptor expression in cells, and its cellular response capacity. In the simplest case (A), a drug binds to 1 receptor coupled to 1 effector (transduction pathway) and produces 1 effect. A drug can bind to 1 receptor coupled to more than 1 effector (B) so it produces more than 1 effect in the same or different cells. A drug can also have affinity for more than 1 receptor (C), with each receptor coupled to a different effector. Effect 2 can be a therapeutic end point or an adverse effect.

Figure 1-12. Signal Transductionand Cross Talk

Receptors provide specificity for cell responses to only certain extracellular chemical signals. Different receptor types can have 1 or more intracellular second-messenger transduction mechanisms without loss of ligand specificity. Different ligands acting through different receptors can thus have the same or different effects via 1 messenger system. In some invertebrate organisms all ionotropic (ion channel) receptors shown here regulate Cl− influx and have the effects shown. In mamals only the GABA receptor regulates Cl− influx. The others have other transduction mechanisms and produce different effects. The effect depends on ligand concentration, cell type, and expression of receptor and second messenger system components. Integrated communication between and within cells thus occurs. A cell with multiple receptor types can be regulated by various ligands and by interaction among receptor types. Interaction among receptor types constitutes receptor cross talk, which allows cells diverse and sophisticated response possibilities.

Figure 1-13. Second-Messenger Pathways

Signal transduction commonly occurs by means of several general mechanisms: (1) ligand-gated ion channels modulate the influx or outflow of ions that alter transmembrane potential or modulate intracellular biochemical reactions (eg, the calcium-calmodulin system); (2) ligand binding to GPCRs modulates enzyme activity (eg, adenylyl cyclase or phospholipase C); (3) ligand binding activates a catalytic portion of the receptor (eg, tyrosine kinase activity); (4) a ligand enters the cell nucleus and alters protein (receptor) synthesis; and (5) a ligand amplifies or attenuates nitric oxide synthesis and the subsequent production of cGMP.

Figure 1-14. Ligand-Gated Ion Channels

Some drugs bind to molecules (ion channels) that form transmembrane pores for ions (usually Na+, K+, Ca2+, Cl−), the channels being composed of many subunits. A drug's binding to 1 or more subunits modifies the receptor function (ion passage), ie, the channels are ligand gated. A single ion channel can accommodate multiple drugs, with each drug binding to a different subunit or site on or within (extracellular, transmembrane, or intracellular) the channel. Membrane-bound channels include nicotinic cholinergic, ionotropic glutamate, GABAA, 5-HT3 (serotonin), and glycine receptors. Intracellular channels include those for Ca2+ on the sarcoplasmic reticulum, endoplasmic reticulum, and mitochondria. Barbiturates, for example, bind to sites on the GABAA receptor complex, which increases Cl− influx and produces increased resting transmembrane potential difference and decreased cell excitability. One drug that modifies activity of an intracellular ligand-gated ion (Ca2+) channel is caffeine.

Figure 1-15. G Protein–Coupled Receptors

Some drugs bind to receptors whose transduction involves a physical association of a receptor with G proteins—the GPCRs. GPCRs, a large family of receptors, mediate effects of neurotransmitters, hormones, and drugs. GPCRs are large proteins that span a cell membrane many times; many drug-related GPCRs, the 7-TM GPCRs, do this 7 times (amino terminus is outside the cell; carboxy terminus is inside). Examples are receptors for epinephrine, norepinephrine, dopamine, 5-HT, ACh (muscarinic), histamine, adenosine, purines, GABA, glutamate, opioids, and vasopressin. Binding of an agonist (drug or endogenous ligand) to a GPCR activates associated G proteins by GTP-GDP exchange, which stimulates dissociation of α from βγ subunits. Inherent GTPase activity within the α subunit restores the initial conditions. One receptor can be coupled to more than 1 type of G protein. Some G proteins activate and others inhibit biochemical steps in signal transduction.

Figure 1-16. Trk Receptors

Some drugs bind to receptors that are composed of an extracellular ligand-binding domain, a transmembrane region, and an intracellular domain that has tyrosine kinase (trk) activity. When activated, these receptors catalyze the intracellular phosphorylation of tyrosine residues in target proteins that are important for cellular growth and differentiation and responses to metabolic stimuli. Examples of ligands (and drug mimetics) that bind totrk receptors include insulin, nerve growth factor, platelet-derived growth factor, cytokines, and other growth factors. It is hypothesized that agonists cause a change in the conformation of the receptor, thereby promoting its action as a tyrosine kinase.

Figure 1-17. Nuclear Receptors

Some drugs produce their effects by binding to receptors located in the cytoplasm or the nucleus of the cell. For example, steroid hormones, thyroid hormone, corticosteroids, vitamin D, and retinoids diffuse through the plasma membrane of the cell and bind to their respective receptors in the cytoplasm. The complex or activated receptors then act as transcription factors by entering the nucleus and binding to DNA hormone-response elements within the nucleus. The DNA-binding domain recognizes certain base sequences, which leads to promotion or repression of particular genes. Regulation of gene transcription by this mechanism can lead to long-term effects. One class of nuclear receptors functions in increased expression of drug-metabolizing enzymes induced by many drugs.

Figure 1-18. Up-regulationand Down-regulationof Receptors

The type and number of receptors that a cell expresses are the net effect of simultaneous receptor synthesis and destruction. In addition to other factors, the number of receptors is modified by long-term exposure to drugs. Chronic stimulation by agonists tends to decrease receptor number (down-regulation), whereas chronic inhibition by antagonists tends to increase the number of receptors (up-regulation). The cellular response opposes the drug-induced effect and may be a defense mechanism. Also, the effect of subsequent administration of drug is greater (or less) than that of initial exposure, and abrupt withdrawal of drug leaves the cell overresponsive or underresponsive to the endogenous ligand. Down-regulation is one mechanism by which pharmacologic tolerance can occur, in which increasing doses of a drug must be used to achieve the same effect.

Figure 1-19. Dose-Response Curves

A direct relation exists between the concentration or dose of a drug and the magnitude of its biologic effect. As a graph, this relation is commonly referred to as aDRC. A DRC can be plotted by using a continuous (graded) or binary (quantal) measure of effect and a linear or logarithmic representation of dose (the latter producing the familiar S-shaped DRC). Each of a drug's usually multiple effects can be represented by a DRC. When the effect is mediated by receptors, the shape of the DRC is consistent with a reversible interaction between ligand (L) and receptor (R):nL +mR ↔ LnRm, wherem andn usually equal 1. The general relation between ligand [L] (drug concentration) and effectE is given by

E=Emax⋅[L][L]+[L]50%effect.

Figure 1-20. Potency

Potency is the drug quantity required for a specified level of a specified effect. For the drug with a DRC given by line A (A), potency is 1 mg/kg for the 50% level of effect A. The 50% level is usually used, with potency shown as an ED50 value. Potency represents ADME and PD properties. Potency for desirable and adverse effects can be established: the potency of one drug for effects A, B, and C (A) is 1, 10, and 100 mg/kg. Potency is thus related to the relative position of a DRC along the horizontal axis. Potency is also used to compare drugs with similar effects (B): 1 mg/kg of drug A is needed for 50% of the effect. Ten times the amount of drug B (10 mg/kg) is required for this level, so drug A is more potent than drug B; both are more potent than drug C. Potency is clinically important only if a drug is expensive or the amount needed is too large. The ED50/LD50 ratio (therapeutic index) is used to compare potency (ED50) with lethality (LD50).

Figure 1-21. Efficacy

At a molecular level,efficacy is the ability of a drug to produce an effect (agonists have positive efficacy, and antagonists have zero efficacy) and the degree of effect per drug molecule bound. At an organism level, it refers to the maximum effect of a drug. Maximum effects of drugs whose DRCs are given by lines A, B, and C is 100%, 50%, and 25%, with the order of efficacy being A > B > C. Efficacy is thus associated with the position of a DRC along the vertical axis. Drugs with a maximal possible effect arefull agonists;partial agonists are drugs whose effect is less than maximal. Some agonists elicit this effect by occupying less than 100% of available receptors, and the other receptors are calledspare receptors. Efficacy is associated with the molecular actions of a drug, not its PK properties. Efficacy can be determined for each of a drug's effects. Unlike potency, efficacy is relatively important clinically because it indicates the maximum attainable effect of a drug.

Figure 1-22. Inverse Agonists

Drug receptors were first thought to be binary switches—either on (activated) or off (resting). Agonists turned the switch on; antagonists blocked agonists' access to receptors. Today, a receptor is viewed as a continuous switch, with the resting state between on or off. Two types of agonists can exist at these receptors: those that move the receptor from resting toward on and those that move it toward off. Both types are agonists, because both have affinity and intrinsic activity. For example, the channel pore of a ligand-gated ion-channel receptor may have a certain resting diameter; some agonists bind to the receptor and increase pore size (increase ion flux), whereas others decrease pore size (decrease ion flux). Which agonist is said to be the inverse of another is arbitrary and depends on which was discovered first. Classic examples of inverse agonists reduce Cl− flow through a GABAA receptor and cause rather than inhibit anxiety. The same antagonist should block both types of agonist.

Figure 1-23. Antagonists: Surmountable (Reversible)and Nonsurmountable (Irreversible)

The ability of an antagonist to alter an agonist effect depends on the affinity of the antagonist for the shared receptor. With weak, reversible antagonist binding (eg, hydrogen bonds), thermal agitation causes some antagonist molecules to uncouple from receptor and agonists successfully compete for receptor sites. If the agonist DRC with surmountable antagonists shifts to the right along the horizontal (dose) axis, the same maximal effect can occur. If antagonist molecules bind to a receptor irreversibly (eg, covalent chemical bonds) or irreversibly alter receptor sites, those sites are unavailable for agonist molecules. Antagonist molecules do not uncouple from a receptor; agonist molecules cannot compete for unoccupied sites. Fewer drug-receptor complexes mean diminished drug effect. The agonist DRC with irreversible antagonists shifts to the right along the dose axis and downward. The same maximal effect cannot be achieved by the agonist at any dose (nonsurmountable antagonism).

Figure 1-24. Routesof Administration

The oral route is generally the most convenient, economic, and safe. Most drugs are rapidly and well absorbed along the GI tract, although some (eg, insulin) are not because of inactivation by enzymes. Drugs given intravenously enter the systemic circulation rapidly; drugs given intra-arterially reach a target site in high concentration. Subcutaneous and intramuscular routes rely on diffusion of the drug into the bloodstream, which can be influenced by warming or cooling the area or by other drugs. Inhalation produces a rapid response to a drug because of the large surface area of the lungs and their extensive blood supply. Transdermal application is becoming an increasing popular mode of administration. Other routes or sites of drug administration include dermal (for local action), mucous membranes (for systemic action), insufflation (lungs), intraneural (nerves), optic (eyes), otic (ears), intraperitoneal (abdomen), and epidural (spinal cord).

Figure 1-25. First-Pass Effect

Drugs that are administered into the GI tract (orally or rectally) are subject to a first-pass effect. Venous drainage of blood from most portions of the GI tract enters the portal circulation, which delivers blood to the liver. In the liver (sometimes the gut wall), drug molecules can be biotransformed (term preferred tometabolized) to less active substances (usually). The amount of active drug that enters the systemic circulation after GI administration is thus less—by the amount of the first-pass effect—than that after another route of administration. The magnitude of this effect on a drug's systemic bioavailability (F) is expressed as the extraction ratio (ER):

F=f×(1−ER)=f×(1−Clliver/Q),

wheref is the extent of absorption,Clliver is the hepatic clearance, andQ is the hepatic blood flow (normally approximately 90 L/h in a 70-kg person). Two related drugs that have comparable bioavailability and similartmax (time to peak concentration) are said to bebioequivalent.

Figure 1-26. Membrane Transport

The biologic membrane is a phospholipid bilayer, a hydrophobic core (lipid layer) between 2 hydrophilic portions (phospho groups). Small molecules can pass through membrane pores. Drugs can pass across membranes by passive diffusion (through lipid or aqueous channels), by active transport (combining with carriers), or by pinocytosis. To cross membranes, most drugs must be both water soluble (hydrophilic or lipophobic) and fat soluble (lipophilic or hydrophobic), which is achieved by weak acids (HA ↔ H+ + A−) and weak bases (BH+ ↔ B + H+), whose charged (hydrophilic) and uncharged (lipophilic) forms are in equilibrium. The extent of drug absorption is a function of pKa of the drug and pH of the local environment. Equations for determining distribution of protonated and nonprotonated forms of a drug across a membrane are

Acids:pKa=pH+log(HA/A−)

Bases:pKa= pH+log(BH+/B).

For reference, pH values in the stomach are 1.0 to 1.5; that in blood plasma is approximately 7.4.

Figure 1-27. Distribution

After absorption, drugs enter the systemic circulation and are distributed widely in the body; they leave the bloodstream and enter cells, with the amount entering depending on local blood flow, capillary permeability, and relative drug lipophilicity. Drugs in the blood are either unbound or bound reversibly to plasma proteins (eg, albumin) in equilibrium. The unbound portion is bioactive. Binding of drugs to these proteins is determined by affinity between drug and protein and protein binding capacity. Only a few binding sites are available, so a high dose can saturate binding sites, and additional drug circulates unbound in the bloodstream. If 2 or more drugs have affinity for the same binding sites, the one with highest affinity will bind, which increases plasma concentration of displaced drug. These effects, which may have clinical consequences, must be considered for the dosing regimen. Drugs with high plasma protein binding (≥95%) include lithium, midazolam, and warfarin (99%).

Figure 1-28. Barriers

Because of various anatomical and physiologic features, endothelial cells of the capillaries can limit passage of drugs from the bloodstream to tissues. For example, endothelial cells of brain capillaries, whose tight junctions merge into a continuous wall, are highly impermeable to many substances. Thus, a blood-brain barrier is established that generally limits accessibility of a good number of drugs, many of which are ionized in the blood at pH 7.4, to the brain. Water-soluble drugs, polar drugs, and ionized forms of drugs cannot cross this blood-brain barrier because they cannot pass through slit junctions and have difficulty traversing the lipid cell membrane. Lipid-soluble drugs pass more readily through cell membranes. In the liver, large fenestrations allow most drugs free access to the hepatic interstitium (with subsequent metabolism of the drugs). The placenta limits but does not prevent entry of drugs into the fetal circulation.

Figure 1-29. Metabolism (Biotransformation)of Drugs

Drugs undergo biotransformation by many of the same reactions as endogenous compounds. Drugs are usually metabolized to less active and more ionized (water-soluble) forms, but equally or more active metabolites can also be created. An inactive parent drug that forms active metabolites is called aprodrug. Although drug metabolism occurs in almost all tissues, including the GI tract, the liver is the major site because of its strategic place in the portal circulation and its many metabolic enzymes. Two general types of drug metabolic reactions occur: phase 1, involving chemical modification, typically by oxidation, reduction, or hydrolysis; and phase 2, in which an endogenous chemical is covalently attached (conjugated) (glucose conjugation, or glucuronidation, the most common). Drugs often undergo multiple phase 1 and 2 reactions, which produces many metabolites, each with its own pharmacologic profile. Liver disease alters drug metabolism, so appropriate dosage adjustment is required.

Figure 1-30. Cytochrome P-450 (CYP450) Enzymes

A major enzyme system that catalyzes phase 1–type drug metabolism reactions is the microsomal CYP450 mixed-function oxidase (monooxygenase) system located in the endoplasmic reticulum in liver, GI tract, lungs, kidney, and other tissues. These enzymes catalyze an oxidation-reduction process that requires CYP450, CYP450 reductase, NADPH (reducing agent), and O2. The only common feature of the many drugs metabolized by this pathway is lipid solubility. The pie chart shown above indicates the approximate percent of current drugs that are metabolized by the indicated CYP isozymes. Known polymorphisms in these enzymes require a drug dosage adjustment. If 2 drugs are metabolized by the same CYP isozyme, they can interfere with each other's normal route or rate of metabolism, and a drug interaction may decrease or increase plasma drug concentrations. An example is interaction between fluoxetine (a selective serotonin reuptake inhibitor) and St John's wort.

Figure 1-31. Metabolic Enzyme Inductionand Inhibition

Multiple factors, including drugs, can either increase or decrease metabolic enzyme activity. Long-term administration of drugs often induces CYP450 activity dramatically by enhancing the rate of synthesis or reducing the rate of degradation of these hepatic microsomal enzymes. Enzyme induction results in more rapid metabolism of the drug and all other drugs metabolized by the same enzymes. As a result, plasma levels and biologic effects of the drugs decrease (except for prodrugs, whose biologic effects increase). Barbiturates are well-known strong inducers of CYP450 enzymes. Other substances can inhibit CYP450 enzymatic activity. In this case, the metabolism of other drugs through this pathway is reduced, which results in increased blood levels of these other drugs. The clinical consequences of the altered blood levels can be greater biologic effects (except for prodrugs) or increased toxicity.

Figure 1-32. Elimination

The major route of drug elimination is through the kidneys, which receive one fifth to one fourth of the cardiac output. Other routes are feces and lungs (especially for anesthetic gases). The rate of elimination of most drugs follows first-order kinetics (exponential decline). The time for the plasma levels of a drug to reach half the initial value is thehalf-life (t1/2). A notable exception is ethanol, which follows zero-order (linear) kinetics at subintoxicating concentrations. Theclearance of a drug from the body is the sum of clearances from all elimination routes, eg, clearance from the kidney is given by the volume of plasma that is completely cleared of the drug per unit time (usually 1 minute). In this case, the amount of drug in urine is measured. Kidney clearance of drug X (CX) is calculated from drug concentrations in urine (UX) and plasma (PX), and urine volume (V):CLX = (UX ×V)/PX. A kidney disorder alters the rate of drug elimination, so the dosage must be adjusted.

Neurological disease

Elaine M Aldred BSc (Hons), DC, Lic Ac, Dip Herb Med, Dip CHM, ... Kenneth Vall, in Pharmacology, 2009

Monoamine Oxidase Inhibitors (MAOIs)

Non-competitive irreversible antagonists (see Chapter 19 ‘Pharmacodynamics: how drugs elicit a physiological effect’, p. 138) of monoamine oxidase type. They break down noradrenaline (norepinephrine) and serotonin, leading to an increase in transmitter activity.

The enzyme (monoamine oxidase) is inhibited not only in the brain, but also in the peripheral neurons, enterocytes and platelets.

Inhibition of the enzyme leads to increases in serotonin, noradrenaline (norepinephrine) and dopamine (see Chapter 31 ‘The nervous system’, p. 242) in the brain.

Because of this widespread inhibition, MAOIs are now used only under tight supervision, for example in patients who are institutionalized and whose diets can be controlled. Other antidepressants tend to be favoured because they have fewer severe side effects.

It can take 2–3 weeks before enough of the drug accumulates for the effects to be felt, and the same amount again to recover because these are irreversible antagonists and it takes time for the enzyme to be remanufactured.

• Adverse Effects

Postural hypotension.

Headache.

Anticholinergic side effects: dry mouth, constipation, sexual dysfunction.

Drug-induced liver damage.

Hypertensive crisis: this is noteworthy because it occurs following foods, drinks or drugs containing the amine tyramine (cheese, meat, yeast extract, soy, beer, some wines and many processed foods). Inhibition of the monoamine oxidase enzyme in the gut wall makes it easier to absorb tyramine and other sympathomimetic substances (i.e. simulating an apparent release of noradrenaline-like substances; see Chapter 31 ‘The nervous system’, p. 244) from food.

Herbal remedies, particularly those containing alkaloids, can be detrimental to a patient’s health. Amines can displace the normal noradrenaline from the storage sites, which leads to hypertension, tachycardia and headaches. If the rise in blood pressure is severe enough then there is the possibility of a cerebrovascular accident.

Read full chapter

URL: //www.sciencedirect.com/science/article/pii/B9780443068980000335

Complex Regional Pain Syndrome

Anton N. Sidawy MD, MPH, in Rutherford's Vascular Surgery and Endovascular Therapy, 2019

Pharmacologic Therapy

Pharmacologic therapy, particularly as it applies to early stages of CRPS, can be combined with intermittent sympathetic blockades and physical therapy. Drug therapy may require nonspecific analgesics, but these should be superimposed, only when necessary, on a “background” of medication designed to attenuate the symptoms by direct effect. Of these, phenytoin, amitriptyline, carbamazepine, and baclofen can be effective, and they are usually used in that order owing to increasing side effects. Tricyclic antidepressant amitriptyline hydrochloride (Elavil) is used in doses of 50 to 75 mg nightly.

Opioids: Opioids inhibit central nociceptive neurons through interaction with µ-receptors. There are no long-term studies of oral opioids in the treatment of neuropathic pain, including CRPS; however, the expert opinion of pain clinicians is that opioids should be part of a comprehensive pain regimen protocol. Opioids should be tried early in the course of CRPS, not as a “last resort.”

Tricyclic Antidepressants: Tricyclic antidepressants are some of the best-studied drugs in the treatment of neuropathic pain, and they have shown an analgesic effect. They inhibit the reuptake of monoaminergic transmitters. There is solid evidence that the reuptake and noradrenaline blocker amitriptyline and the selective noradrenaline blocker desipramine produce pain relief in several neuropathies.

GABA-Agonists: GABA-agonists such as gabapentin (Neurontin) have also been used in the treatment of CRPS. The action of gabapentin probably includes the inhibition of calcium channels.

α-Adrenergic Blocking Agents: The use of α-adrenergic blockers is based primarily on the fact that patients with CRPS have altered blood flow as a result of increased local secretion of norepinephrine and vascular endothelial hypersensitivity. Inhibition of the receptors leading to vascular dilation and increased blood flow may be helpful. Pain relief following intravenous phentolamine administration has been suggested as a diagnostic tool, as well as a prognostic guide for favorable response to sympathetic blockade.44,53 Its 15-minute plasma half-life precludes its use as a therapeutic modality. Other medications include phenoxybenzamine and prazosin.44,45 Phenoxybenzamine has antiadrenergic effects and is an irreversible α-antagonist and has been shown to be beneficial in treating CRPS. It is considered a third-line agent and works best in cases lasting less than 3 months.53

Oral β-Blocking Agents: Propranolol, which antagonizes serotonin, has been partially successful.43,44

Oral Calcium Channel Blocking Agents: These agents produce smooth muscle relaxation of the arteriole walls, leading to increased blood flow.

Bisphosphonate Therapy: This therapy is based on the concept that pain results from osteopenia as part of CRPS. These are potent inhibitors of osteoclastic activity, which are used for management of osteoporosis and other metabolic bone diseases. Their primary action is reduction of bone turnover.

Evaluation of the Biological Activity of Compounds

Iain G. Dougall, John Unitt, in The Practice of Medicinal Chemistry (Fourth Edition), 2015

Irreversible, Noncompetitive, and Allosteric Antagonists

Several other forms of antagonists have been identified and will be discussed briefly. Irreversible antagonists form covalent bonds with the receptor protein and thus prevent binding of agonists. They therefore effectively decrease the receptor pool and, by so doing, decrease the response eliciting capacity (efficacy) of the agonist, as this is dependent on [Rtot]. As equilibrium is not attained, the antagonist affinity cannot be measured, but such agents (e.g., Phenoxybenzamine) form the basis of the receptor inactivation method developed by Furchgott [31] for estimating agonist affinities and efficacies. Practically, however, the reliance of this method on the availability of suitable alkylating agents excludes its use in most receptor systems.

Noncompetitive antagonists bind to receptors and make them functionally inoperative either by preclusion of agonist binding (due to negligible dissociation of the antagonist during the response-gathering phase of the experiment) or through some other biochemical mechanism that obviates agonist effect (e.g., pore blockade of ion channels). Under these circumstances, increasing the agonist concentration cannot overcome the effect of the antagonist, and a distinctive feature of noncompetitive antagonists is the depressive effect they have on the maximal agonist response (α). The magnitude of the depression will however depend on the agonist under study and the system used. This relates to the concept of receptor reserve whereby maximum agonist effects can be achieved at low levels of receptor occupancy (binding)—for example, 10 percent occupancy may be enough to produce a maximum response and therefore there is a 90 percent receptor reserve. Receptor reserve depends on both the receptor number ([Rtot]) and the efficiency of stimulus-response coupling as well as the intrinsic efficacy of the agonist. Hence, noncompetitive antagonists will have differing capabilities to depress the maximal response to the same agonist in different systems. The same will be true for different agonists in the same system. The potency of noncompetitive antagonists can be estimated using various models but as a “rule of thumb” the pA2 (−log[B]+log (r−1)), as defined above for competitive antagonists, gives a reasonably accurate estimate of the antagonist affinity (pKB) when measured at low levels of agonist response [14]. See Figure 2.14B.

Figure 2.14. Noncompetitive Antagonism in Functional Assays.

Simulations showing the effect of a noncompetitive antagonist on responses to the same agonist in a system with high receptor reserve (A) or low receptor reserve (B). Increasing concentrations of the antagonist (3, 10, 30 nM) cause more marked depression of the agonist maximum effect in the low reserve system. Data was simulated using a form of the Operational Model of agonism that assumes that antagonist binding precludes binding of the agonist [32]. The model parameters used were Em=100, n=1, τ=100 (high reserve) or τ=3 (low reserve), pKA=5.0, pKB=9.0. Estimates of the antagonist affinity (pKB) can be made by fitting data directly to this model or approximated as pA2=−log[B]+log(r−1) when a concentration-ratio (r) measured at low response levels is used (B).

All of the modes of antagonism described above are orthosteric; that is, the antagonist blocks access of the agonist to its binding site through steric hindrance. Allosteric antagonists in contrast bind to their own site on the receptor to induce a change in conformation of the receptor, which in turn alters the affinity or efficacy of the receptor for the agonist [33,34]. It is now clear that allosteric ligands can both increase and decrease the affinity and efficacy of other ligands, so allosteric modulators is a more appropriate term. Indeed, perhaps the best known therapeutically used allosteric modulators are the benzodiazepines, which increase the conductance of the GABAA receptor. One of the key properties of allosteric modulators is their saturability of effect, which can be evidenced in functional experiments such as Schild analysis where a curvilinear plot results (Figure 2.15). Similarly, in Cheng-Prusoff type analyses, such antagonists will produce less than 100 percent inhibition of the agonist response. This behavior results from the fact that while the allosterically modified receptor may have diminished affinity (and/or efficacy) for the agonist, the agonist can still produce receptor activation in the presence of the modulator. As is evident from Figure 2.15B, use of concentration-ratios (r) at low antagonist concentrations can yield reasonably accurate estimates of compound affinity.

Figure 2.15. Allosteric Antagonism.

(A) The effects of acetylcholine (Ach) on the electrically evoked contractions of the guinea pig left atrium in the absence (■) or presence of the allosteric modulator gallamine at the following concentrations: 10 μM (▴), 30 μM (▾), 100 μM (♦), 300 μM (□), and 500 μM (●). (B) The Schild plot of the data shown in (A). The solid line (slope=1) denotes the behavior expected for a competitive antagonist, whereas the dashed line shows the best fit linear regression (and associated slope factor) through the points. The curve through the points and associated parameter estimates represent the fit to an allosteric model. The estimated pKB was 6.03 and the α value of 5.3×10−3 equates to a gallamine-induced decrease in the affinity of ACh of 189-fold.

Reproduced with permission from Christopoulos and Kenakin, 2002 [35].

Allosteric modulation offers a number of potential advantages over orthosteric antagonists. First, they can modify (i.e., reduce or increase by a small amount) endogenous agonist signals without completely blocking them, thus allowing fine-tuning of responses. Second, there is the potential to increase the duration of allosteric effect by loading the receptor compartment with large concentrations of modulator. Such large concentrations will have no further effect than to prolong the saturated allosteric effect (i.e., the saturability of the allosteric ligand can be used to limit effect but increase duration). Another potential advantage of allosterism is increased selectivity. Orthosteric antagonists often have limited selectivity across receptor subtypes. For example, most muscarinic receptor antagonists exhibit poor selectivity between the five known subtypes (M1–M5), presumably because they are competing with acetylcholine for very similar recognition sites. However, the surrounding protein structure of the receptors are sufficiently different to offer the potential for selective stabilization of receptor conformations by allosteric modulators. These potential advantages of allosteric modulators remain largely theoretical as very few such agents have to date reached the market. Nevertheless, the approval of the CCR5 antagonist Maraviroc (Selzentry®) for the treatment of human immunodeficiency virus (HIV) infection demonstrated the feasibility of this approach. This compound inhibits HIV entry by binding to a receptor site distinct from where the viral gp120 envelope protein binds [36,37].

Finally, although the discussion above focuses on receptors, allosteric modulation of enzyme function is a well-known phenomenon. The availability of binding sites distinct from those for the substrate again offers the potential for increased selectivity. For example, compounds designed to bind to an allosteric site in a particular protein kinase are likely to have improved selectivity over compounds targeting the ATP binding site.

Read full chapter

URL: //www.sciencedirect.com/science/article/pii/B978012417205000002X

Cellular and Molecular Toxicology

J.P.V. Heuvel, in Comprehensive Toxicology, 2010

2.03.4.4.4 Irreversible and pseudoirreversible antagonists

Some types of antagonists produce irreversible antagonism of agonist-mediated responses. This is most commonly observed when an antagonist produces a covalent modification in a receptor. Irreversible antagonists will yield curves that are identical to noncompetitive antagonist if the response is proportional to the number of receptors occupied. (However, remember spare receptors.) Note that if the k−1 is sufficiently slow, competitive antagonists may appear to be irreversible. Noncompetitive, irreversible, and pseudoirreversible antagonism share a common property in that they are nonsurmountable antagonists, that is, increased agonist does not cause maximal response.

Read full chapter

URL: //www.sciencedirect.com/science/article/pii/B9780080468846002037

Oxidised ATP

Charles Kennedy, in xPharm: The Comprehensive Pharmacology Reference, 2007

Pre-Clinical Research

Oxidized ATP was originally used to affinity label nucleotide binding sites in enzymes as it interacts with unprotonated lysine residues to form covalent bonds. It was subsequently shown to also be an irreversible antagonist at recombinant P2X1, P2X2Evans et al (1995) and P2X7Surprenant et al (1996) receptors. The antagonism developed slowly and was non-competitive. The effect at P2X7 receptors likely underlies the ability of oxidized ATP to inhibit ATP-induced permeabilization of macrophages El-Moatassim and Dubyak (1993), Murgia et al (1993) and human lymphocytes Wiley et al (1994). In most of these studies, the antagonism was irreversible.

Read full chapter

URL: //www.sciencedirect.com/science/article/pii/B9780080552323629885

Receptors | Opioid Receptors☆

Hui Zheng, ... Horace H. Loh, in Encyclopedia of Biological Chemistry (Third Edition), 2021

Structures of Opioid Receptors

As classical members of GPCR superfamily with seven transmembrane domains, it was difficult to generate the crystal structures of opioid receptors. Therefore, the crystal structures of opioid receptors with sufficient resolution become available only as late as 2012.

By using an irreversible antagonist β-FNA, which binds deeply within a large solvent-exposed pocket, the crystal structure of the mouse µ-opioid receptor was generated at 2.8 Å resolution. Most ligands of GPCRs do not bind as deep as the β-FNA as demonstrated in previously reported structures of GPCRs. In addition, µ-opioid receptor crystallizes as a symmetrical dimer through transmembrane domain 5 and 6 (Manglik et al., 2012).

Also in 2012, by using the selective antagonist JDTic, the crystal structure of human κ-opioid receptor was generated at 2.9 Å resolution. Parallel dimers were also observed in the crystal structure of human κ-opioid receptor (Wu et al., 2012). Also in the same year, by using the selective antagonist, naltrindole, the crystal structure of the δ-opioid receptor was generated (Granier et al., 2012). The crystal structures of the three opioid receptors provide additional insights into conserved and ligand-specific elements. The lower part of the binding pocket is highly conserved among opioid receptors, whereas the upper part contains divergent residues that confer subtype selectivity.

Read full chapter

URL: //www.sciencedirect.com/science/article/pii/B9780128194607001122

Drug Antagonism

Terry P. Kenakin, in Pharmacology in Drug Discovery and Development (Second Edition), 2017

Slow Dissociation Kinetics and Noncompetitive Antagonism

For true competitive kinetics to be operative, the antagonist that has been preequilibrated with the receptors must dissociate quickly enough for the agonist present in the receptor compartment to bind according to mass action (Fig. 4.9A). If this does not occur, the antagonist will occupy an inordinately high percentage of the receptors and antagonism will dominate. This percentage of receptors could be high enough to prevent the agonist from producing a maximal response, thus an insurmountable effect on the agonist dose–response curve will result (Fig. 4.9B). This is sometimes also referred to as “pseudo-irreversible” antagonism, because the antagonist is essentially irreversibly bound to the receptor within the time frame relevant for the production of response by the agonist. This often results in a depressed maximal response in the agonist dose–response curve. The degree of maximal response depression depends on the number of receptors needed to induce response in the cell preparation, that is, the magnitude of the receptor reserve (see chapter: Predicting Agonist Effect). If the cell is extremely sensitive to the agonist, such that only a small fraction of the receptor population needs to be activated to produce tissue maximal response, then even an irreversible antagonist may shift the curve to the right with little depression of maximum, that is, if only 7% of the receptors are needed for maximal response, then irreversibly blocking 80% of the receptors will still enable the agonist to produce a maximal response. On the other hand, if the cell is not very sensitive to the agonist and 100% of the receptors are needed for the production of maximal response, an irreversible antagonist will immediately produce a depression of the maximal response to the agonist. These different scenarios are shown in Fig. 4.10. In cases of true noncompetitive orthosteric blockade where the maximal response to the agonist is depressed by all concentrations of antagonist, the pIC50 can be obtained as for competitive antagonists (Fig. 4.8) and the resulting pIC50 is essentially equal to the pKB (see Fig. 4.11). This obviates the use of a correction factor such as that provided by Eq. (4.5).

Figure 4.9. The kinetics of reequilibration of agonists, antagonists and receptors. (A) For simple competitive antagonist systems where there is sufficient time for reequilibration between receptors. The reduction in antagonist receptor occupancy (dotted line) rapidly adjusts as the agonist binds to receptors (solid line). (B) For a slowly dissociating antagonist (broken line), the agonist binding is biphasic, characterized by an initial rapid phase (where the agonist binds to open receptors) and a slow phase whereby the agonist must deal with a slowly dissociating antagonist. The gray rectangle represents the window of opportunity to measure agonist response in the presence of the antagonist; if this is less than the time required for complete reequilibration of agonist, antagonist and receptors then the agonist receptor occupancy will be less than would be defined by true competitive interactions.

Figure 4.10. The effects of a noncompetitive antagonist in two different tissue systems. The top left panel is a system possessing a receptor reserve for the agonist (100% response can be obtained by occupancy of 70% of the receptors). Therefore, noncompetitive (pseudo-irreversible) blockade of receptors will result in dextral displacement of the agonist dose–response curve with no depression of maximal response until concentrations of antagonist that block >70% are present. The response system to the right has no reserve for the agonist (note the linear relationship between percentage agonist occupancy and response). Under these circumstances, the noncompetitive antagonist will produce depression of the maximal response to the agonist at all concentrations.

Figure 4.11. Effects of a noncompetitive antagonist on an agonist dose–response curve (left panel); open circles show the effects of the antagonist on a selected dose of agonist that produces 80% maximal response in the absence of the antagonist. Right panel shows the responses to that same selected dose of agonist as a function of the concentrations of competitive antagonist producing the blockade. It can be seen that an inverse sigmoidal curve results (pIC50 curve). The value of the abscissa corresponding to the half-maximal ordinate value of this curve is the IC50, namely the molar concentration of antagonist that produces half-maximal inhibition of the defined agonist response. The same pIC50 curve would be obtained for any level of response.

So far, two kinetic extremes of orthosteric antagonist action have been discussed; a rapidly dissociating antagonist that allows true competitiveness between agonist and antagonist to result and a slowly dissociating antagonist (essentially irreversible) that causes the antagonist to eliminate a certain fraction of the receptor from consideration of agonism. There are a large number of systems, referred to as “hemi-equilibria,” that fall in between these extremes. Here, the antagonist partially dissociates from the receptors in the presence of the agonist causing only the high concentrations of agonist (those requiring high receptor occupancies) to be subjected to irreversible blockade. Under these circumstances, the dose–response curves to the agonist are shifted to the right by the antagonist but may have a truncated (essentially “chopped”) maximal response (see Fig. 4.12). In these cases, the potency of the antagonist can be estimated with a pA2 and/or Schild analysis on the parallel shifted portions of the curves.

Figure 4.12. Antagonist hemi-equilibrium. Shown are the effects of increasing concentrations of a competitive antagonist in a system where there is insufficient time for complete reequilibration between agonist, antagonist and receptors within the time-frame available to measure response to the agonist in the presence of the antagonist. This type of system is characterized by a depression of maximal response to a new steady-state level below that of the control curve but greater than zero. The antagonist potency can be estimated with Schild analysis or through estimation of the pA2 (shaded rectangle).

The main reason for noting the patterns of antagonism produced by orthosteric antagonists with various rates of offset kinetics is to determine the correct method of estimating antagonist potency. As discussed previously, all antagonism can be estimated with a pIC50, but there are circumstances where this empirical quantity differs from the correct pKB (notably with simple competitive antagonists). The pA2 can also be a useful estimate, but a full analysis of a range of concentrations of antagonists in a Schild analysis is needed to ensure identity of the pA2 with the pKB. Finally, a direct fit of curves for a simple competitive antagonist can yield an estimate of pKB (Fig. 4.7). For true noncompetitive (insurmountable) antagonism, where the maximal response to the agonist is depressed, directly fitting experimental data requires a more explicit model than that available for simple competitive antagonism (ie, Eq. 4.4). Specifically, a noncompetitive model based on the Black–Leff operational model (see chapter: Predicting Agonist Effect) can be used according to the equation shown below (Derivations and Proofs, Appendix B.7):

(4.6)Response=[A]nτnEm [A]nτn+(( [A]+KA)([B]/K B+1))n

where KA and τ are the affinity and efficacy of the agonist [A], Em is the maximal response of the system, n is a slope fitting parameter, and KB is the equilibrium dissociation constant for the antagonist [B]. An example of the application of Eq. (4.6) to fitting data is shown in Fig. 4.13.

Figure 4.13. Dose–response curves obtained in the presence of increasing concentrations of a noncompetitive antagonist fit to the model for noncompetitive antagonism (Eq. 4.6).

While the pattern of curves for antagonism observed in vitro is relevant to the quantification of antagonist potency, it is not as important in therapeutic use. Specifically, the production of competitive (surmountable) versus noncompetitive antagonism can simply be a function of the way the antagonist is studied. The key factor here is the time allowed for the collection of agonist response in the presence of antagonist. If this time is short and insufficient for reequilibration of agonist, antagonist, and receptors to occur, then noncompetitive antagonism will result. For example, a truncated time for response collection is operative in measuring transient calcium release, a rapid process. Thus, many antagonists produce noncompetitive antagonism in this type of assay. The very same antagonist may produce surmountable simple competitive antagonism (parallel shift to the right of curves with no depression of maximum) in another assay that allows a longer time for agonist, antagonist and receptor to reequilibrate. Reporter assays, which incubate the ligands for 24 hours while gene expression takes place, are one such assay; thus, a given antagonist may be noncompetitive in a calcium transient assay and surmountably competitive in a reporter assay (see Fig. 4.14).

Figure 4.14. Effect of different time periods to measure agonist response (in the presence of a slowly dissociating orthosteric antagonist) shown in top panel. Panel on the left shows observed antagonism when the period for measurement of response is too short for complete reequilibration (noncompetitive antagonism is observed). Panel on the right shows the same system when sufficient time for the slowly dissociating antagonist is allowed for proper reequilibration between agonist, antagonist and receptors. Under these circumstances, simple competitive antagonism is observed.

The ultimate aim of these experiments is to utilize antagonists in vivo for therapeutic advantage. Under these circumstances, the antagonist will enter a system already activated by agonist (much as in a pIC50 experiment) and produce a reduction in the basal response. Once the antagonist diffuses out of the receptor compartment (through whole body clearance; see chapter: Pharmacokinetics II: Distribution and Multiple Dosing), the basal response will be regained providing the steady-state function of the system has not otherwise changed. Since it would be postulated that most physiological systems will operate in a region sensitive to changes in the endogenous agonist concentration (lower region of the dose–response curve), the effects of antagonists on the high concentration end of the agonist dose–response curve probably are not relevant to whole body physiology. Fig 4.15 shows the in vivo effect of an antagonist on a basal agonist response. Fig. 4.16 shows the in vivo responses of a simple competitive (surmountable) and a noncompetitive (insurmountable) antagonist; it can be seen that the in vivo effects are nearly identical. Therefore, the determination of competitive versus noncompetitive mechanisms in vitro is important only from the point of view of using the correct model to measure the pKB.

Figure 4.15. Effects of an antagonist in vivo where a level of basal activity is present. The dose–response curve to the endogenous agonist will be shifted to the right and/or depressed by the antagonist; the panel on the right shows the observed response in vivo. As the antagonist enters the receptor compartment and binds, the elevated basal response is depressed; as the antagonist washes out of the receptor compartment, the response returns.

Figure 4.16. In vivo effects (see Fig. 4.15) of a competitive versus a noncompetitive antagonist on endogenous elevated basal response.

Irreversible Antagonists

If the antagonist has no appreciable rate of offset from the receptor (but simply irreversibly binds to the receptor never to dissociate), then a KB cannot be measured because a steady-state of a constant fraction of occupied receptor cannot be attained. This is because, as long as the antagonist is not depleted from the medium, it will continue to inactivate the receptors until there are no more receptors left in the tissue to inactivate. In essence, the onset of an irreversible antagonist is a chemical reaction that runs to completion when either the antagonist or receptors are depleted. Since KB=rate of offset/rate of onset, then KB→0 as equilibration with antagonist progresses (the rate of offset is essentially=0). Fig. 4.17 shows two antagonists producing apparent noncompetitive antagonism as they are equilibrated with the tissue for increasing periods of time. Unlike the reversible antagonist which takes the tissue to a submaximally blocked state (depressed maximum dose–response curve which does not change with further equilibration time), the truly irreversible antagonist blocks the response until a dose–response relationship for the agonist cannot be attained. When noncompetitive receptor antagonism is observed it should be confirmed that submaximal states of inhibition can be attained and that a KB can be measured to differentiate pseudo-irreversible antagonism from truly irreversible antagonism.

Figure 4.17. Irreversible receptor blockade. Top panel shows diminution of agonist receptor occupancy with a reversible noncompetitive antagonist (solid line) and an irreversible antagonist (broken line). Six time points (noted with the numbered circles) are chosen and the dose–response curve to the agonist shown in the presence of the reversible (left panel) and irreversible (right panel) antagonists. The circled numbers next to the dose–response curves refer to the timepoints for antagonist onset. While the reversible antagonist comes to a steady state by timepoints 4–6, the irreversible antagonist continues to depress response until no response to the agonist can be obtained. The lack of nonzero steady-state denotes the irreversibility of the antagonism.

Read full chapter

URL: //www.sciencedirect.com/science/article/pii/B9780128037522000041

Which of the following is the chemical alteration of drug molecules into metabolites by body cells?

Drug metabolism is the chemical alteration of a drug by the body.

Which body surface absorbs drugs most readily?

The small intestine tends to be the location of greatest absorption potential for most drugs due to its large surface area, the presence of both active and passive absorption mechanisms, and near neutral pH.

What is a type of cellular movement that allows intake of large molecules?

Endocytosis is a type of active transport that moves particles, such as large molecules, parts of cells, and even whole cells, into a cell. There are different variations of endocytosis, but all share a common characteristic: the plasma membrane of the cell invaginates, forming a pocket around the target particle.

What is the cellular material that interacts with the drugs?

The drug-receptor complex acts on specific regions of the genetic material deoxyribonucleic acid (DNA) in the cell nucleus, resulting in an increased rate of synthesis for some proteins and a decreased rate for others.

Toplist

Neuester Beitrag

Stichworte