实验方法> 生物信息学技术> 数据库>Measuring Protein Interactions by Optical Biosensors

Measuring Protein Interactions by Optical Biosensors

关键词: measuring protein interactions来源: 互联网

  • Abstract
  • Table of Contents
  • Figures
  • Literature Cited

Abstract

 

In recent years, optical evanescent wave biosensors have been used to characterize protein?protein interactions, including determination of equilibrium binding constants and bimolecular rate constants. This surface binding technique can provide information about chemical on?rate constants, the lifetimes of complexes formed, and the time course of the signal. This unit provides a thorough, well?illustrated discussion of the principles of optical biosensors, experimental design, ligand immobilization, experimental data analysis, and common obstacles and possible solutions.

        GO TO THE FULL PROTOCOL: PDF or HTML at Wiley Online Library Table of Contents

  • Strategic Planning
  • Immobilization Protocols
  • Binding Experiments and Data Analysis
  • Common Experimental Obstacles
  • Summary
  • Acknowledgment
  • Literature Cited
  • Figures
  • Tables

        GO TO THE FULL PROTOCOL: PDF or HTML at Wiley Online Library Materials

 

GO TO THE FULL PROTOCOL: PDF or HTML at Wiley Online Library Figures

  •   Figure 17.6.1 Schematic presentation of a typical optical biosensor experiment. Light is coupled into a structure that allows generation of surface‐confined electromagnetic waves (e.g., surface plasmons in a gold film, or modes of a planar waveguide), which are sensitive to the refractive index of the solution in the range of the evanescent field, n surf . Typical penetration depths of the sensitive volume into the solution are in the order of 100 nm. Ligands are attached to the sensor surface, as indicated by half‐circles. (A ) When analytes (full circles) are introduced into the solution above the surface, reversible interactions with the ligand lead to association events at a rate governed by K on c A c L,free (solid arrows), and dissociation events at a rate c complex K off (dashed arrows). In the association phase, association events outnumber dissociation events until a steady state is reached. (B ) When the surface is washed with running buffer in the absence of analyte, only dissociation events take place. (C ) Signal obtained from probing the refractive index n surf during the sequential application of the configurations depicted in A and B, given in arbitrary units. Following the association phase (A) and the dissociation phase (B), usually a regeneration procedure is applied for removing all remaining analyte from the surface before a new experimental cycle takes place.
    View Image
  •   Figure 17.6.2 (A ) The binding of a bivalent analyte can take place through the interaction with a single ligand molecule (left), or with two ligand molecules (right), depending on the local ligand density. If an unoccupied immobilized ligand molecule is within an accessible distance to an already singly bound analyte molecule, then the on‐rate constant of the second interaction is strongly enhanced because of the entropic colocalization constraints. The overall dissociation rate constant of the double attached analyte is substantially reduced (usually by orders of magnitude), because it requires the virtually simultaneous dissociation of both interactions, which is much more improbable than a single dissociation event. The overall binding kinetics obtained from multivalent analytes is highly complex. (B ) If the configuration is reversed, with the bivalent binding partner immobilized to the sensor surface, then both valencies act independently (left), and (in the absence of cooperativity of the sites) lead to binding kinetics that are identical to that of the monovalent interactions (right).
    View Image
  •   Figure 17.6.3 Schematic presentation of an ideal 1:1 pseudo first‐order reaction. Shown is a superposition of the binding progress curves that should be obtained in a series of experiments at different analyte concentrations. The data are given in percent of the signal at maximal binding, but the units of the signal do not affect the results of the data analysis. The arrows on the curves at the higher concentrations indicate that steady‐state binding is attained, whereas, for the lower curves, longer association times would be required as part of a steady‐state analysis.
    View Image
  •   Figure 17.6.4 (A ) Time course of an equilibrium titration experiment. Arrows indicate the time‐points of the step‐wise increase in the analyte concentration. (B ) Steady‐state binding analysis according to Equation 17.6.5. In a plot of the steady‐state signal versus the base‐10 logarithm of the analyte concentration, the binding isotherm exhibits a typical sigmoid shape. The inflection point at 50% saturation determines K D , and the width of the curve is characterized by ∼10% saturation at 1/10 K D and ∼90% saturation at 10 K D (dotted lines). Visible inspection of the data in this representation already allows a robust parameter estimation.
    View Image
  •   Figure 17.6.5 Mass transport effects on the surface binding process. (A ) In the association phase, insufficient transport does not fully replenish the free analyte in a zone near the sensor surface (depletion zone). The inability to maintain the concentration c A in the depletion zone leads to a limited surface binding rate. (B ) The corresponding effect of mass transport limitation in the dissociation phase is that the surface is insufficiently washed and dissociated analyte is insufficiently removed. This generates a zone near the surface in which free analyte (after dissociation from the ligand) can rebind to empty ligand sites before diffusing into the bulk flow. This zone is indicated as retention zone. (C ) Introduction of an excess of the soluble form of the ligand as a competitor into the buffer of the dissociation phase helps prevent rebinding to the surface. Soluble ligand can bind to dissociated analyte before rebinding takes place, and the soluble complex can diffuse into the buffer flow. This can allow the measurement of K off free of mass transport effects. (D ) Time course of dissociation during a mass transport induced rebinding situation (as depicted in panel B), and after introduction of a soluble competitor into the washing buffer (arrow). In the presence of the competitor, the signal reflects the chemical off‐rate constant K off , while in the rebinding situation, the apparent rate governing the signal is reduced.
    View Image
  •   Figure 17.6.6 Exemplary data of the interaction of TCR with superantigen (for details of this interaction see Andersen et al., ). (A ) Elution profile of the TCR from size‐exclusion chromatography, indicating the fractions used for kinetic analysis of the interaction with immobilized superantigen, as shown in (B ). It should be noted that fraction 1 (bold solid line), fraction 2 (dashed line), and fraction 3 (solid line) all exhibit multiphasic binding, with different relative magnitudes of the slower component. It should also be noted that despite the lower concentration of fraction 1, which results in the lowest response in the association phase, the signal in the dissociation phase is highest. This slower kinetic component in the binding progress curve reflects different amounts of aggregates bound to the sensor surface. The aggregates have a slower dissociation because of their multivalent attachment. For comparison, the same interaction is shown in the reverse orientation, with immobilized TCR and soluble superantigen (dotted line). (C ) Sequence of association‐dissociation curves at different superantigen concentrations, binding to immobilized TCR. In this orientation, the association and the dissociation is much faster, virtually monophasic, and the binding is completely reversible, which provides further evidence that the slow kinetic components introduced in the different fractions of the soluble TCR sample in Panel B are artifacts. For quantitative analysis, the signal measured at a nonfunctionalized surface (dotted line) is subtracted in order to remove the bulk refractive index contribution of the analyte. (D ) A plot of the net steady‐state binding signal versus superantigen concentration allows the measurement of the affinity of the interaction free from artifacts due to TCR aggregation that would be introduced in the configuration of Panel B.
    View Image
  •   Figure 17.6.7 Effects of the ligand density on the binding kinetics for interactions with relative low affinity (A ) and medium affinity (B ). Each panel shows the interaction of immobilized superantigen with single‐chain TCR, observed at high ligand surface density (1700 RU, solid lines) and low ligand surface density (700 RU, dashed lines). For easier comparison, the signal obtained at the lower density surface was scaled proportionally. In panel A, a slow phase of binding in the association phase and a residual binding (possible slow dissociation of multivalently bound aggregates) is introduced at high ligand density, under otherwise identical conditions. For the interaction in panel B, the chemical off‐rate constant is smaller than for the low‐affinity interaction shown above. Nevertheless, from comparison of the binding curves at different ligand surface densities (under otherwise identical conditions) it is obvious that an increased ligand density has significant effects on the surface binding kinetics. This observation could be due to aggregation or to mass transport limitations, both of which are more likely at higher ligand densities.
    View Image

Videos

Literature Cited

   Andersen, P.S., Lavoie, P.M., Sekaly, R.P., Churchill, H., Kranz, D.M., Schlievert, P.M., Karjalainen, K., and Mariuzza, R.A. 1999. Role of the T cell receptor alpha chain in stabilizing TCR‐superantigen‐MHC class II complexes. Immunity 10:473‐483.
   Buckle, P.E., Davies, R.J., Kinning, T., Yeung, D., Edwards, P.R., Pollard‐Knight, D., and Lowe, C.R. 1993. The resonant mirror: A novel optical sensor for direct sensing of biomolecular interactions. Part II: Applications. Biosens. Bioelectron. 8:355‐363.
   Davis, S.J., Ikemizu, S., Wild, M.K., and van der Merwe, P.A. 1998. CD2 and the nature of protein interactions mediating cell‐cell recognition. Immunol. Rev. 163:217‐236.
   Edwards, P.R., Gill, A., Pollard‐Knight, D.V., Hoare, M., Buckle, P.E., Lowe, P.A., and Leatherbarrow, R.J. 1995. Kinetics of protein‐protein interactions at the surface of an optical biosensor. Anal. Biochem. 231:210‐217.
   Edwards, P.R., Maule, C.H., Leatherbarrow, R.J., and Winzor, D.J. 1998. Second‐order kinetic analysis of IAsys biosensor data: Its use and applicability. Anal. Biochem. 263:1‐12.
   Garland, P.B. 1996. Optical evanescent wave methods for the study of biomolecular interactions. Q. Rev. Biophys. 29:91‐117.
   Gershon, P.D. and Khilko, S. 1995. Stable chelating linkage for reversible immobilization of oligohistidine tagged proteins in the Biacore surface plasmon resonance detector. J. Immunol. Methods. 183:65‐76.
   Glaser, R.W. and Hausdorf, G. 1996. Binding kinetics of an antibody against HIV p24 core protein measured with real‐time biomolecular interaction analysis suggest a slow conformational change in antigen p24. J. Immunol. Methods. 189:1‐14.
   Hall, D.R. and Winzor, D.J. 1997. Use of a resonant mirror biosensor to characterize the interaction of carboxypeptidase A with an elicited monoclonal antibody. Anal. Biochem. 244:152‐160.
   Hermanson, G.T. 1996. Bioconjugate Techniques. Academic Press, San Diego.
   Khilko, S.N., Jelonek, M.T., Corr, M., Boyd, L.F., Bothwell, A.L.M., and Margulies, D.H. 1995. Measuring interactions of MHC class I molecules using surface plasmon resonance. J. Immunol. Methods. 183:77‐94.
   Knoll, W. 1998. Interfaces and thin films as seen by bound electromagnetic waves. Annu. Rev. Phys. Chem. 49:569‐638.
   Leckband, D.E. 1997. The influence of protein and interfacial structure on the self‐assembly of oriented protein arrays. Adv. Biophys. 34:173‐190.
   Leckband, D.E., Kuhl, T., Wang, H.K., Herron, J., Muller, W., and Ringsdorf, H. 1995. 4‐4‐20 anti‐fluorescyl IgG Fab′ recognition of membrane bound hapten: Direct evidence for the role of protein and interfacial structure. Biochemistry 36:11467‐11478.
   Lukosz, W. 1991. Principles and sensitivities of integrated optical and surface plasmon sensors for direct affinity sensing and immunosensing. Biosens. Bioelectronics. 6:215‐225.
   Malmborg, A.C. and Borrebaeck, C.A. 1995. Biacore as a tool in antibody engineering. J. Immunol. Methods. 183:7‐13.
   Margulies, D.H., Corr, M., Boyd, L.F., and Khilko, S.N. 1993. MHC Class I/peptide interactions: Binding specificity and kinetics. J. Mol. Recognit. 6:59‐69.
   Margulies, D.H., Plaksin, D., Khilko, S.N., and Jelonek, M.T. 1996. Studying interactions involving the T‐cell antigen receptor by surface plasmon resonance. Curr. Opin. Immunol. 8:262‐270.
   Minton, A.P. 1995. Confinement as a determinant of macromolecular structure and reactivity. 2. Effects of weakly attractive interactions between confined macrosolutes and confining structures. Biophys. J. 68:1311‐1322.
   Minton, A.P. 1998. Molecular crowding: Analysis of effects of high concentrations of inert cosolutes on biochemical equilibria and rates in terms of volume exclusion. Methods Enzymol. 295:127‐149.
   Muller, K.M., Arndt, K.M., and Plückthun, A. 1998. Model and simulation of multivalent binding to fixed ligands. Anal. Biochem. 261:149‐158.
   Myszka, D.G., Jonsen, M.D., and Graves, B.J. 1998. Equilibrium analysis of high affinity interactions using Biacore. Anal. Biochem. 265:326‐30.
   Nieba, L., Krebber, A., and Plückthun, A. 1996. Competition Biacore for measuring true affinities: Large differences from values determined from binding kinetics. Anal. Biochem. 234:155‐165.
   Ober, R.J. and Ward, E.S. 1999a. The influence of signal noise on the accuracy of kinetic constants measured by surface plasmon resonance experiments. Anal. Biochem. 273:49‐59.
   Ober, R.J. and Ward, E.S. 1999b. The choice of reference cell in the analysis of kinetic data using Biacore. Anal. Biochem. 271:70‐80.
   O'Shannessy, D.J. and Winzor, D.J. 1996. Interpretation of deviations from pseudo‐first‐order kinetic behavior in the characterization of ligand binding by biosensor technology. Anal. Biochem. 236:275‐283.
   O'Shannessy, D.J., Brigham‐Burke, M., and Peck, K. 1992. Immobilization chemistries suitable for use in the Biacore surface plasmon resonance detector. Anal. Biochem. 205:132‐136.
   O'Shannessy, D.J., Brigham‐Burke, M., Soneson, K.K., Hensley, P., and Brooks, I. 1993. Determination of rate and equilibrium binding constants for macromolecular interactions using surface plasmon resonance: Use of nonlinear least squares analysis methods. Anal. Biochem. 212:457‐468.
   O'Shannessy, D.J., O'Donnell, K.C., Martin, J., and Brigham‐Burke, M. 1995. Detection and quantitation of hexa‐histidine‐tagged recombinant proteins on western blots and by a surface plasmon resonance biosensor technique. Anal Biochem. 229:119‐124.
   Plant, A.L., Brigham‐Burke, M., Petrella, E.C., and O'Shannessy, D.J. 1995. Phospholipid/alkanethiol bilayers for cell‐surface receptor studies by surface plasmon resonance. Anal. Biochem. 226:342‐348.
   Ramsden, J.J., Bachmanova, G.I., and Archakov, A.I. 1996. Immobilization of proteins to lipid bilayers. Biosens. Bioelectronics 11:523‐528.
   Schuck, P. 1996. Kinetics of ligand binding to receptor immobilized in a polymer matrix, as detected with an evanescent wave biosensor. I. A computer simulation of the influence of mass transport. Biophys. J. 70:1230‐1249.
   Schuck, P. 1997a. Use of surface plasmon resonance to probe the equilibrium and dynamic aspects of interactions between biological macromolecules. Annu. Rev. Biophys. Biomol. Struct. 26:541‐566.
   Schuck, P. 1997b. Reliable determination of binding affinity and kinetics using surface plasmon resonance biosensors. Curr. Opin. Biotechnol. 8:498‐502.
   Schuck, P. and Minton, A.P. 1996a. Minimal requirements for internal consistency of the analysis of surface plasmon resonance biosensor data. Trends Biochem. Sci. 252:458‐460.
   Schuck, P. and Minton, A.P. 1996b. Analysis of mass transport limited binding kinetics in evanescent wave biosensors. Anal. Biochem. 240:262‐272.
   Schuck, P., Millar, D.B., and Kortt, A.A. 1998. Determination of binding constants by equilibrium titration with circulating sample in a surface plasmon resonance biosensor. Anal. Biochem. 265:79‐91.
   Schuster, S.C., Swanson, R.V., Alex, L.A., Bourret, R.B., and Simon, M.I. 1993. Assembly and function of a quaternary signal transduction complex monitored by surface plasmon resonance. Nature 365:343‐347.
   Sigal, G.B., Bamdad, C., Barberis, A., Strominger, J., and Whitesides, G.M. 1996. A self‐assembled monolayer for the binding and study of histidine‐tagged proteins by surface plasmon resonance. Anal. Chem. 68:490‐497.
   Silhavy, T.J., Szmelcman, S., Boos, W., and Schwartz, M. 1975. On the significance of the retention of ligand by protein. Proc. Natl. Acad. Sci. USA. 72:2120‐2124.
   Stein, T. and Gerisch, G. 1996. Oriented binding of a lipid‐anchored cell adhesion protein onto a biosensor surface using hydrophobic immobilization and photoactive crosslinking. Anal. Biochem. 237:252‐259.
   van der Merwe, P.A. and Barclay, A.N. 1996. Analysis of cell‐adhesion molecule interactions using surface plasmon resonance. Curr. Opin. Immunol. 8:257‐261.
   Yarmush, M.L., Patankar, D.B., and Yarmush, D.M. 1996. An analysis of transport resistances in the operation of Biacore; implications for kinetic studies of biospecific interactions. Mol. Immunol. 33:1203‐1214.
Key References
   Davis et al, 1998. See above.
   Contains a detailed description of analyte aggregation effects on the measured sur
推荐方法

Copyright ©2007 ANTPedia, All Rights Reserved

京ICP备07018254号 京公网安备1101085018 电信与信息服务业务经营许可证:京ICP证110310号