Comprehensive Physiology Wiley Online Library

Receptor Identification and Characterization

Full Article on Wiley Online Library



Abstract

The sections in this article are:

1 Identification of Receptors Based on Functional Properties
1.1 Schild Analysis
2 Identification of Receptors with Radioligand Binding
2.1 Choosing the Biological Preparation
2.2 Choosing a Radioligand
2.3 Separating Bound From Free Radioligand
2.4 Irrelevant Binding
2.5 Establishing That the Radioligand Binding Detected Reflects Interaction With the Physiologically Significant Receptor
2.6 Analysis of Radioligand Binding Data Based on the Law of Mass Action
2.7 Fractional Occupancy
2.8 Assumptions Inherent in the Law of Mass Action
3 Saturation Binding Studies
3.1 Defining Nonspecific Binding in a Radioligand Binding Assay
4 Defining The Specificity of Radioligand Binding Using Competitive Binding Analysis
4.1 Analyzing Competitive Binding Data
4.2 Calculating the Ki From the IC50
4.3 Homologous Competitive Binding Curves Permit an Assessment of Both Kd and Bmax
5 Kinetic Analysis of Radioligand Binding Experiments
5.1 Studies of Radioligand Dissociation
5.2 Studies of Radioligand Association
6 Complex Binding Phenomena
6.1 Competitive Binding Experiments With Two (or More) Receptor Sites
6.2 Saturation Binding Experiments With Two Receptor Sites
6.3 Comparing One‐ and Two‐Site Models
6.4 The Slope Factor of a Competitive Binding Curve
6.5 Using Dissociation Experiments to Investigate Complex Binding
6.6 Distinguishing Between Independent Receptor Subtypes and Negative Cooperativity
6.7 Evaluating Allosteric Phenomena
6.8 Obtaining Independent Data to Clarify the Biological Origin of Complex Binding Phenomena
6.9 Agonist Binding and the Ternary Complex
7 Summary
Figure 1. Figure 1.

Dose‐response curves. With many full agonists, such as a, occupation of a small fraction of receptors elicits a nearly maximal response. A partial agonist can either be an agent such as b, where higher occupancy is required to elicit a response, or an agent such as c, which does not elicit the maximum response even at very high concentrations. Stephenson 20 would have referred to agonist a as a “strong” agonist.

Figure 2. Figure 2.

Schild regression. In the presence of a competitive antagonist (dashed line) the dose‐response curve for an agonist is shifted to the right. The agonist can still elicit the same maximal response, but it takes a higher concentration. A demonstrates the definition of the dose ratio. It is the concentration of agonist required to elicit a response in the presence of antagonist divided by the concentration of agonist needed to elicit that same response in the absence of antagonist. B: a Schild regression. Each circle represents a dose ratio of agonist in the presence of a different concentration of antagonist. If the drug is a competitive antagonist, the slope of the Schild regression line must equal 1.0. The intercept when log (dose ratio — 1) equals zero (that is, when dose ratio = 2) is the Kd of the antagonist. The term pA2 sometimes is used to refer to the Kd for a competitive antagonist; pA2 = ‐log Kd for the antagonist.

Figure 3. Figure 3.

Saturation binding. A: total and nonspecific binding at equilibrium as a function of increasing concentrations of radioligand. The difference, also shown, is specific binding. B: specific binding only. Note the definition of Bmax (the maximal amount of binding extrapolated to infinite radioligand concentration) and Kd (the concentration of radioligand that occupies half the receptors at equilibrium). C: the same data, but plotting the concentration of radioligand on a logarithmic axis. The solid part of the curve in C corresponds to the curve in B; the dotted portion is extrapolated. D: the same data transformed into a Scatchard plot.

Figure 4. Figure 4.

Competitive binding experiments. A: definition of the IC50 as the concentration of unlabeled drug that competes for half of specific radioligand binding at equilibrium. B: competition by three different drugs. The IC50 is lowest for drug a, so drug a has the highest affinity. The order of potency is a>b>c. C: an 81‐fold increase in the concentration of unlabeled drug makes the curve descend from 90% to 10% occupancy. If the curve is shallower or steeper than this, then the radioligand and competitor do not compete for binding to a single class of binding sites with unchanging affinity for radioligand and competitor.

Figure 5. Figure 5.

Dissociation kinetics. A: before the first time point on the graph, binding was allowed to occur, perhaps to equilibrium. At time zero on the graph, dissociation was initiated by infinite dilution or by addition of an excess of unlabeled drug. Under either condition, the radioligand cannot rebind after dissociating from the receptor; consequently the total amount of binding decreases over time. The half‐life (t1/2) is the time when half of the specific binding has dissociated. B: the same data on a log plot, which linearizes the dissociation data.

Figure 6. Figure 6.

Association kinetics. Radioligand binding was added at time zero, and the graph shows the increase in specific binding thereafter.

Figure 7. Figure 7.

Complex competitive binding curves. A: the meaning of the slope factor. A fractional slope factor (such as −0.5) means that the curve is shallow. A normal competitive binding curve (single binding site, no cooperativity) has a slope factor of −1.0, as shown in the solid curve. B: competition for two different binding sites. The radioligand binds identically to both sites, but the competitor has a tenfold difference in affinity. Although the curve is shallow, it is not obviously biphasic.

Figure 8. Figure 8.

Saturation binding to two sites. A: binding to two distinct receptor types indicated by dotted and dashed curves. The two sum to the solid curve. Experimental data will follow the solid curve, and computer analysis is needed to find the Bmax and Kd of the two components. B: the same data as represented on a Scatchard plot.

Figure 9. Figure 9.

Distinguishing between multiple subtypes (A) and negative cooperativity (B) by dissociation kinetic strategies. If the binding sites are independent (no cooperativity, right panel) the rate of dissociation will be identical when dissociation is initiated by infinite dilution (solid curve) or addition of excess unlabeled drug (dashed curve). If the binding sites are cooperative, then the dissociation curve will differ depending on how dissociation was initiated. When monitoring of the dissociation phase is initiated by adding an excess of unlabeled drug, all of the binding sites are always occupied by either labeled or unlabeled drug. With infinite dilution, there is no rebinding and the number of occupied sites decreases over time. After radioligand dissociates from one site, the affinity at another site will increase and the observed rate of dissociation will slow.

Figure 10. Figure 10.

Evaluating allosteric regulation. Radioligand is allowed to bind to a site, and then dissociation begins when an excess of unlabeled agent is added. Addition of an allosteric modified can accelerate or slow the rate of dissociation.

Figure 11. Figure 11.

The extended ternary complex model. The left panel shows the simple ternary complex model. Receptors can be coupled or uncoupled from G. Agonist or hormone (H) binding facilitates the interaction of R with R. The right panel shows the extended ternary complex model. Receptors exist in interconvertible R and R* states. Binding of agonist promotes transition to the R* state. Only the R* state can interact with G proteins. (see refs. 9, 15.



Figure 1.

Dose‐response curves. With many full agonists, such as a, occupation of a small fraction of receptors elicits a nearly maximal response. A partial agonist can either be an agent such as b, where higher occupancy is required to elicit a response, or an agent such as c, which does not elicit the maximum response even at very high concentrations. Stephenson 20 would have referred to agonist a as a “strong” agonist.



Figure 2.

Schild regression. In the presence of a competitive antagonist (dashed line) the dose‐response curve for an agonist is shifted to the right. The agonist can still elicit the same maximal response, but it takes a higher concentration. A demonstrates the definition of the dose ratio. It is the concentration of agonist required to elicit a response in the presence of antagonist divided by the concentration of agonist needed to elicit that same response in the absence of antagonist. B: a Schild regression. Each circle represents a dose ratio of agonist in the presence of a different concentration of antagonist. If the drug is a competitive antagonist, the slope of the Schild regression line must equal 1.0. The intercept when log (dose ratio — 1) equals zero (that is, when dose ratio = 2) is the Kd of the antagonist. The term pA2 sometimes is used to refer to the Kd for a competitive antagonist; pA2 = ‐log Kd for the antagonist.



Figure 3.

Saturation binding. A: total and nonspecific binding at equilibrium as a function of increasing concentrations of radioligand. The difference, also shown, is specific binding. B: specific binding only. Note the definition of Bmax (the maximal amount of binding extrapolated to infinite radioligand concentration) and Kd (the concentration of radioligand that occupies half the receptors at equilibrium). C: the same data, but plotting the concentration of radioligand on a logarithmic axis. The solid part of the curve in C corresponds to the curve in B; the dotted portion is extrapolated. D: the same data transformed into a Scatchard plot.



Figure 4.

Competitive binding experiments. A: definition of the IC50 as the concentration of unlabeled drug that competes for half of specific radioligand binding at equilibrium. B: competition by three different drugs. The IC50 is lowest for drug a, so drug a has the highest affinity. The order of potency is a>b>c. C: an 81‐fold increase in the concentration of unlabeled drug makes the curve descend from 90% to 10% occupancy. If the curve is shallower or steeper than this, then the radioligand and competitor do not compete for binding to a single class of binding sites with unchanging affinity for radioligand and competitor.



Figure 5.

Dissociation kinetics. A: before the first time point on the graph, binding was allowed to occur, perhaps to equilibrium. At time zero on the graph, dissociation was initiated by infinite dilution or by addition of an excess of unlabeled drug. Under either condition, the radioligand cannot rebind after dissociating from the receptor; consequently the total amount of binding decreases over time. The half‐life (t1/2) is the time when half of the specific binding has dissociated. B: the same data on a log plot, which linearizes the dissociation data.



Figure 6.

Association kinetics. Radioligand binding was added at time zero, and the graph shows the increase in specific binding thereafter.



Figure 7.

Complex competitive binding curves. A: the meaning of the slope factor. A fractional slope factor (such as −0.5) means that the curve is shallow. A normal competitive binding curve (single binding site, no cooperativity) has a slope factor of −1.0, as shown in the solid curve. B: competition for two different binding sites. The radioligand binds identically to both sites, but the competitor has a tenfold difference in affinity. Although the curve is shallow, it is not obviously biphasic.



Figure 8.

Saturation binding to two sites. A: binding to two distinct receptor types indicated by dotted and dashed curves. The two sum to the solid curve. Experimental data will follow the solid curve, and computer analysis is needed to find the Bmax and Kd of the two components. B: the same data as represented on a Scatchard plot.



Figure 9.

Distinguishing between multiple subtypes (A) and negative cooperativity (B) by dissociation kinetic strategies. If the binding sites are independent (no cooperativity, right panel) the rate of dissociation will be identical when dissociation is initiated by infinite dilution (solid curve) or addition of excess unlabeled drug (dashed curve). If the binding sites are cooperative, then the dissociation curve will differ depending on how dissociation was initiated. When monitoring of the dissociation phase is initiated by adding an excess of unlabeled drug, all of the binding sites are always occupied by either labeled or unlabeled drug. With infinite dilution, there is no rebinding and the number of occupied sites decreases over time. After radioligand dissociates from one site, the affinity at another site will increase and the observed rate of dissociation will slow.



Figure 10.

Evaluating allosteric regulation. Radioligand is allowed to bind to a site, and then dissociation begins when an excess of unlabeled agent is added. Addition of an allosteric modified can accelerate or slow the rate of dissociation.



Figure 11.

The extended ternary complex model. The left panel shows the simple ternary complex model. Receptors can be coupled or uncoupled from G. Agonist or hormone (H) binding facilitates the interaction of R with R. The right panel shows the extended ternary complex model. Receptors exist in interconvertible R and R* states. Binding of agonist promotes transition to the R* state. Only the R* state can interact with G proteins. (see refs. 9, 15.

References
 1. Barker, E. L., R. S. Westphal, D. Schmidt, and E. Sanders‐Bush. Constitutively active 5‐hydroxytryptamine 2c receptors reveal novel inverse agonist activity of receptor ligands. J. Biol. Chem. 260: 11607–11690, 1994.
 2. Cheng, Y. and W. H. Prusoff. Relationship between the inhibition constant (Ki) and the concentration of an inhibitor that causes a 50% inhibition (I50) of an enzymatic reaction. Biochem. Pharmacol. 22: 3099–3108, 1973.
 3. Cohen, J. B., M. Weber, J ‐P. Changeaux. Effects of local anesthetics and calcium on the interaction of cholinergic ligands with the nicotinic receptor protein from Torpedo marmorata. Mol. Pharmacol. 10: 904–932, 1974.
 4. DeLean, A., A. A. Hancock, and R. J. Lefkowitz. Validation and statistical analysis of the computer modeling method for quantitative analysis of radioligand binding data for mixtures of pharmacological subtypes. Mol. Pharmacol. 21: 5–16, 1982.
 5. Garcia, M. L., V. F. King, P. K. S. Siegl, J. P. Reuben, and G. J. Kaczorowski. Binding of Ca++ entry blockers to cardiac sarcolemmal membrane vesicles: Characterization of diltiazem‐binding sites and their interaction with dihydropyridine and aralkylamine receptors. J. Biol. Chem. 261: 8146–8157, 1986.
 6. Horstman, D. A., S. Brandon, A. L. Wilson, C. A. Guyer, E. J. Cragoe, Jr., and L. E. Limbird. An aspartate conserved among G protein‐–coupled receptors confers allosteric regulation of α2‐adrenergic receptors by sodium. J. Biol. Chem. 365 (35): 21590–21595, 1990.
 7. Kenakin, R. P., R. A. Bond, and T. I. Bonner. Definition of pharmacological receptors. Pharm. Rev. 44: 351–362, 1992.
 8. Klotz, I. M. and D. L. Hunston. Mathematical models for ligand‐receptor binding: Real sites, ghost sites. J. Biol. Chem. 259: 10060–10062, 1984.
 9. Lefkowitz, R. J., S. Cotecchia, P. Samama, and T. Costa. Constitutive activity of receptors coupled to guanine nucleotide regulatory proteins. Trends Pharmacol. Sci. 14: 303–307, 1993.
 10. Limbird, L. E. Cell Surface Receptors: A Short Course on Theory and Methods, Second Edition. Boston: Kluwer Academic Publishers, 1996.
 11. Monaghan, D. T., H. J. Olverman, L. Nguyen, and J. C. Watkins. Two classes of N‐methyl‐d‐aspartate recognition sites and differential distribution and differential regulation by glycine. Proc. Natl. Acad. Sci. USA 85: 9836–9840, 1988.
 12. Prezeau, L., J. Gomez, S. Ahern, S. Mary, T. Galvez, J. Bockaert, and J ‐P. Pin. Changes in the carboxyl‐terminal domain of metabotropic glutamate receptor 1 by alternative splicing generates receptors with differing agonist‐independent activity. Mol. Pharmacol. 49: 422–429, 1996.
 13. Rodbard, D. A graphic method for the determination and presentation of binding parameters in complex systems. Anal. Biochem. 20: 525–532, 1973.
 14. Rosenthal, H. E. A graphic method for the determination and presentation of binding parameters in complex systems. Anal. Biochem. 20: 525–532, 1967.
 15. Samama, P., S. Cotecchia, T. Costa, and R. J. Lefkowitz. A mutation‐induced activated state of the β2‐adrenergic receptor: extending the ternary complex model. J. Biol. Chem. 268: 4625–4636, 1993.
 16. Scatchard, G. The attractions of proteins for small molecules and ions. Ann. N.Y. Acad. Sci. 51: 660–672, 1949.
 17. Schild, H. O. pA, a new scale for the measurement of drug antagonism. Brit. J. Pharm. 2: 189–206, 1947.
 18. Simons, S. S., Jr. and W. B. Pratt. Glucocorticoid receptor thiols and steroid‐binding activity. Methods Enzymol. 257: 406–422, 1995.
 19. Sklar, L. Signal transduction and ligand‐receptor dynamics in the human neutrophil. Transient response and occupancy‐response relations at the formyl peptide receptor. J. Biol. Chem. 260: 1461–1467, 1985.
 20. Stephenson, R. P. A modification of receptor theory. Br. J. Pharm. 11: 379–393, 1956.
 21. Swillens, S. Interpretation of binding curves obtained with high receptor concentrations; practical aid for computer analysis. Mol. Pharmacol. 47: 1197–1203, 1985.

Contact Editor

Submit a note to the editor about this article by filling in the form below.

* Required Field

How to Cite

Lee E. Limbird, Harvey Motulsky. Receptor Identification and Characterization. Compr Physiol 2011, Supplement 20: Handbook of Physiology, The Endocrine System, Cellular Endocrinology: 49-67. First published in print 1998. doi: 10.1002/cphy.cp070104